id
stringlengths 30
36
| source
stringclasses 1
value | format
stringclasses 1
value | text
stringlengths 5
878k
|
---|---|---|---|
no-problem/9911/astro-ph9911498.html | ar5iv | text | # Are bright gamma-ray bursts a fair sample ?
## 1 Introduction
During the 90’s the observations of the Burst and Transient Source Experiment (BATSE) on board the Compton Gamma-Ray Observatory provided a wealth of data on the properties of gamma-ray bursts at soft gamma-ray energies. The interpretation of these data was however complicated by our lack of knowledge of GRB distances. This situation changed dramatically in 1997 with the discovery of afterglows at X-ray wavelength by BeppoSAX, which led to the discovery of visible afterglows and to the first distance determinations. In this paper we show that the availability of burster distances sheds a new light on the interpretation of GRB properties measured at $`\gamma `$-ray energies.
The redshifts measured since 1997 (Table 1) have exposed the very broad dispersion of GRBs in luminosity. With these new observations in mind, we discuss here the possibility that it is the burster intrinsic luminosity, and not the distance to the source, which determines the burst brightness measured at the earth. In Section 2, we show that this hypothesis is supported by the distribution of GRB luminosities presently available. In Section 3, we explain that it also naturally explains some well known (statistical) properties of the gamma-ray bursts. The consequences of this hypothesis on our understanding of the GRB population are discussed in Section 4.
We now define our use of the words brightness and luminosity. We call the burst intensity measured at the earth brightness. The most common measures of brightness are the peak flux (in units of ph cm<sup>-2</sup> s<sup>-1</sup>) and the fluence (in units of erg cm<sup>-2</sup>). We call the burst energy emitted at the source luminosity. The most common measures of luminosity are the peak luminosity (in units of ph s<sup>-1</sup>) and the total luminosity (in units of erg). In the absence of information on the beaming factor of the gamma-ray emission, the peak and total luminosities are computed under the assumption that the source is radiating isotropically; if the $`\gamma `$ emission is beamed toward us, the total energy radiated by the source could be much smaller. In order to keep this paper simple we deal with a single measure of the burst brightness (the fluence) and the corresponding measure of luminosity (the total luminosity). We have checked that the use of the peak flux does not change our conclusions.
## 2 The Brightness Luminosity Correlation of GRBs
The first measures of GRB redshifts have exposed the broad range of intrinsic luminosities of these sources and their comparatively small range of distances. In order to provide a more quantitative view of this statement, we show in Table 2 various estimates of the dispersion of GRBs in distance and in luminosity.
The Table 2 strongly suggests that the parameter which primarily determines the burst brightness measured at the earth is not the distance of the source, but its intrinsic luminosity. This situation is the opposite of the standard candle hypothesis. In the following we call it the Brightness Luminosity Correlation hypothesis (or BLUC). Such a situation can only happen if the bursters have a particular spatial distribution which is discussed in Section 4.3.
It is clear, however, that the number of redshifts which have been measured so far is too small to draw definite conclusions. A few tens of redshifts spanning the whole range of GRB brightnesses will probably be needed to transform what we still consider as a hypothesis into a firmly established GRB property. Nevertheless we consider that, despite these uncertainties, the BLUC hypothesis has enough impact on our understanding of GRBs to deserve a discussion of its consequences. This is done in the following sections.
## 3 Brightness dependant properties of GRBs
The brightness dependance of GRB properties has been extensively studied during the 90’s as a way to unravel cosmological effects (e.g. spectral redshift or time dilation). The rationale behind this work was the concept that faint GRBs were more distant on average, and that they should consequently be more affected by the expansion of the universe. These studies have disclosed two important properties of GRBs, the so-called Hardness-Intensity Correlation (or HIC) and the Time Dilation (TD).
The BLUC conjecture, on the contrary, states that faint GRBs are not due to more distant sources but to sources which are intrinsically less luminous. This leads to a different interpretation of the Hardness-Intensity Correlation and of the Time Dilation which we discuss now.
### 3.1 The Hardness-Intensity Correlation
The Hardness-Intensity correlation is the observation that bright GRBs have on average harder energy spectra than faint GRBs. This property has been discussed by several authors in various contexts (e.g. Mallozzi et al. 1995; Dezalay et al. 1997 and ref. therein). Within the context of BLUC, the Hardness-Intensity Correlation simply reflects an underlying correlation between the luminosity of the source and its spectral hardness. This effect is indeed expected within the framework of cosmological models which invoke a plasma expanding at ultra-relativistic velocities, with a Lorentz factor ($`\mathrm{\Gamma }`$) of several hundred. The relativistic expansion of the emitting plasma multiplies the energy of the photons by a factor $`\mathrm{\Gamma }`$ while it increases the source luminosity by a huge factor (of the order of $`\mathrm{\Gamma }^3`$). The combination of these two effects naturally produces a correlation between the average photon energy and the luminosity of the source; if the burst brightness reflects the radiated luminosity (as postulated by the BLUC conjecture) this correlation is observed as HIC.
### 3.2 The Time Dilation
Time Dilation is the observation that the timescales in the time histories of faint bursts are typically longer than those measured in bright GRBs. The reality of this effect and its interpretation have been subject to ample discussions (e.g. Lestrade et al. 1993, Norris et al. 1994, Band 1994, Mitrofanov et al. 1996, Lee and Petrossian 1997, Stern et al. 1997 and ref. therein). In the context of the BLUC hypothesis, TD means that the timescales are longer in the light curves of intrinsically subluminous bursts.
In the absence of a detailed model of the GRB prompt emission there is no straightforward interpretation of this feature (unlike for the Hardness-Intensity Correlation). We note, however, that Ramirez-Ruiz & Fenimore (1999) have recently found that the faint peaks within a gamma-ray burst last longer than the more intense ones. This feature seems to supports the fact that low luminosity emission has longer characteristic timescales.
### 3.3 The No-Host problem
Another property of GRBs which has been discussed over the last years is the so-called No-Host problem, which is based on deep observations of the error boxes of several bright historical GRBs. A detailed analysis of these error boxes (obtained by triangulation over the last 30 years) shows that they do not contain bright galaxies. If we assume that GRBs are hosted by normal galaxies, the apparent magnitude of the brightest galaxy in each error box can be used to derive a lower limit on the typical distance scale of those bright GRBs. The no-host problem arises when one tries to extrapolate the distance derived for the brightest events to the population of faint GRBs. Schaefer (1999) shows that if faint GRBs are a distant version of bright bursts and if they are hosted by normal galaxies, they must be placed at very large distances (z $``$ 6). An alternative explanation proposed by Schaefer, is that GRBs do not reside in normal host galaxies.
The BLUC conjecture offers a third way to solve this problem. If the brightness of GRBs is dominated by their intrinsic luminosity, faint GRBs are not distant versions of the brightest events. They are instead bursts which are intrinsically less luminous but which have essentially the same distance scale (see below). The BLUC hypothesis thus allows all GRBs to reside in normal galaxies.
## 4 Discussion
This section is devoted to a brief analysis of the consequences that BLUC would have on our understanding of the GRB population if future redshift measurements confirm it.
### 4.1 What is an average GRB ?
As emphasized in the title, the BLUC hypothesis implies that bright GRBs are not representative of the bulk of the population. They are intrinsically more luminous, with harder spectra and cannot be used to infer the properties of average GRBs. It seems thus better to use faint or intermediate GRBs to derive the typical characteristics of the population (duration, energy of the peak of the SED…).
### 4.2 The interpretation of the curve log(N)-log(S)
Within the framework of BLUC the power law distribution of bright GRBs is not the consequence of the spatial distribution of nearby sources but a direct measure of the luminosity distribution of gamma-ray bursts. In the internal shocks paradigm, this distribution is closely related to the distribution of the Lorentz factors of the emitting plasma. In this context it looks like an interesting coincidence that this slope equals $`3/2`$ which is precisely the value expected for sources homogeneously distributed in a Euclidean space.
The break in the intensity distribution occurs when the luminosity function is fully sampled for nearby bursters. The interpretation of the curve Log(N)-Log(S) in the context of BLUC presents many other interesting properties which we plan to discuss in a future paper (Atteia et al., in preparation). In a more general way, BLUC provides a natural explanation of the fact that burst subclasses appear to have different intensity distribution (e.g. Belli 1997, Pendleton et al. 1998, Tavani 1998). Since the brightness distribution reflects the luminosity distribution, it is not surprising that GRB subclasses selected according to their temporal or spectral properties display different luminosity (hence brightness) distributions.
### 4.3 The GRB distribution in distance
If the BLUC conjecture is correct, the distance of a GRB has little impact on its observed brightness. The only way to achieve such a situation is to consider bursters which are restricted to a limited range of distances. This means that the bulk of the burster population occupies a shell-like volume around us with the more distant GRBs being only a few times farther than the nearby ones (while sources which are simply bounded in space which can have a very broad range of distances). This seems to indicate that most GRBs occured at a particular epoch of the life of the universe. In the context of the current ideas on the origin of GRBs, which relate them to violent stellar explosions, the BLUC conjecture thus appears compatible with the existence of a relatively well defined period of enhanced stellar formation.
Another way to express this situation is to say that GRBs belonging to different classes of brightness have essentially the same distribution in distance. An amusing consequence is that modest GRB detectors (like PVO or ULYSSES) do sample the whole volume containing the GRBs, but for the brighest ones only. More importantly, this formulation provides an effective way to check the BLUC hypothesis via its prediction that faint and bright bursts must have the same range of redshifts. The availability of a few tens of redshifts in the next few years with BeppoSAX and HETE-2 should confirm or discard this conjecture. Should BLUC be confirmed, the redshifts already measured provide a good idea of the extent of the GRB distribution in distance.
###### Acknowledgements.
The author thanks J-P. Lestrade and R. Mochkovitch for valuable comments. The author is also grateful to the BATSE team for making the Current BATSE GRB Catalog available at http://www.batse.msfc.nasa.gov/batse/grb/catalog/current/. |
no-problem/9911/cond-mat9911132.html | ar5iv | text | # Cooper Pairing in Ultracold 40K Using Feshbach Resonances
## Abstract
We point out that the fermionic isotope <sup>40</sup>K is a likely candidate for the formation of Cooper pairs in an ultracold atomic gas. Specifically, in an optical trap that simultaneously traps the spin states $`|9/2,9/2`$ and $`|9/2,7/2`$, there exists a broad magnetic field Feshbach resonance at $`B=196_{21}^{+9}`$ that can provide the required strong attractive interaction between atoms. An additional resonance, at $`B=191_{10}^{+5}`$ gauss, could generate p-wave pairing between identical $`|9/2,7/2`$ atoms. A Cooper-paired degenerate Fermi gas could thus be constructed with existing ultracold atom technology.
Recently an ultracold gas of fermionic <sup>40</sup>K atoms was cooled to the quantum degenerate regime . This achievement opens a new chapter in the story of ultracold matter, complementary to the Bose-Einstein condensation work that has been going on for over four years now. The degenerate Fermi gas (DFG) is expected to exhibit novel behavior in its thermodynamics , collision dynamics , and scattering of light . Perhaps the most intriguing prospect for the DFG is the potential to observe a pairing of the fermions, leading to a derived superfluid state, analogous to the Cooper pairing of electrons in a superconductor .
To make such a pairing work requires an effective attraction between colliding atom pairs in the gas. For bosons, an attractive interaction corresponds to a negative value of the s-wave scattering length. For fermions, however, the Pauli exclusion principle prohibits s-wave scattering of atoms in identical spin states. This leaves only p-wave collisions as a pairing mechanism, but the resulting interactions are energetically suppressed and are generally considered to give experimentally unattainable pairing transition temperatures . On the other hand, a recent proposal has suggested that p-wave interactions may be enhanced by the application of very large dc electric fields, which could be generated by powerful CO<sub>2</sub> lasers .
A second possibility would be to use two different spin states of a fermionic atom, thus restoring s-wave collisions as a pairing mechanism. In this context <sup>6</sup>Li appears to be an attractive candidate , since it possesses a large negative s-wave scattering length $`a_s=2160`$ $`a_0`$, in units of the Bohr radius $`a_0`$ . \[In this paper we distinguish s-wave and p-wave scattering lengths with the subscripts “$`s`$” and “$`p`$”. To avoid confusion with standard notations, we indicate singlet and triplet explicitly in superscripts, as in Eq. (3), below.\] In this case the critical temperature for Cooper pairing is approximately
$$T_c\frac{E_F}{k_B}\mathrm{exp}\left(\frac{\pi }{2k_F|a_s|}\right),$$
(1)
where $`E_F`$ and $`k_F`$ are the Fermi energy and momentum, respectively. As pointed out in Ref. , for experimentally realizable Fermi energies $`E_F/k_B600`$ nK there would result $`T_c15`$ nK for <sup>6</sup>Li. Thus any alkali atom with a similarly large, negative scattering length should be a viable candidate for Cooper pairing.
The purpose of this paper is to consider the prospects for Cooper pairing in <sup>40</sup>K, in both s-waves and p-waves, based on a magnetic-field Feshbach resonance that can be used to tune its scattering length. This atom has been trapped and cooled in several labs . The ability to tune scattering lengths resonantly using magnetic fields is now a proven technology. To date, this resonant tuning has been observed in Na , Rb , and Cs . That these resonances are in fact useful tools for manipulation of ultracold gases has been amply demonstrated recently in an experiment that used them to Bose-condense the otherwise uncondensible <sup>85</sup>Rb isotope .
Although the “required” scattering length to ensure formation of Cooper pairs will depend strongly on experimental circumstances, we can estimate a reasonable set of parameters using the guidelines laid out in Ref. . A first requirement is that the resulting Cooper-paired state be mechanically stable, which for a two-component gas with number densities $`n_1`$ and $`n_2`$ requires
$$n_1n_2a_s^6\left(\frac{\pi }{48}\right)^2.$$
(2)
When the scattering length violates this condition, two kinds of instability may occur: if $`a_s<0`$, at least one component collapses into a dense, probably solid state; whereas if $`a_s>0`$, the two components will phase-separate . Note that with a tunable $`a_s`$ these instabilities can be probed experimentally in <sup>40</sup>K.
If we assume equal densities of $`n_1=n_2=10^{14}`$ cm<sup>-1</sup>, then Eq. (2) imposes the restriction $`|a_s|<1700`$ $`a_0`$. Conservatively, we will adopt a target value of $`a_s=1000`$ $`a_0`$ in the following. In this case, for a Fermi temperature of $`T_F600`$ nK (compare Ref. ) we would find a Cooper pairing temperature of $`T_c25`$ nK in <sup>40</sup>K. Moreover, we are interested in the stability of this $`T_c`$ against variations in the magnetic field strength. Let us require that $`T_c`$ remain constant to within a small fraction, say $`10\%`$. Eq. (1) then tells us that we must maintain $`a_s`$ constant to within $`3\%`$. We will see below that this criterion should be relatively easy to meet for the resonance described.
To compute Feshbach resonances in <sup>40</sup>K, we employ the standard close-coupled Hamiltonian for ultracold alkali-atom scattering . As usual, meaningful results can be obtained from this Hamiltonian only if it is fine-tuned with the help of experimental data. In this case we will employ the constraints imposed by a recent analysis of photoassociation spectroscopy of the $`0_g^{}`$ state of <sup>39</sup>K<sub>2</sub> . This analysis reveals a <sup>39</sup>K triplet scattering length (in $`a_0`$) of
$$a_s^{\mathrm{triplet}}(39)=170.045(C_6\overline{C}_6)\pm 25,$$
(3)
with $`\overline{C}_6=3800`$ atomic units . This parametrization allows for an uncertainty in the $`C_6`$ coefficient that determines the long-range van der Waals attraction between the atoms. The experiment itself provides no direct information on the value of $`C_6`$. The result in Eq. (3) is consistent with a complementary analysis of the $`1_u`$ state of <sup>39</sup>K, which gives $`60`$ $`a_0`$ $`<a_s^{\mathrm{triplet}}(39)<`$ $`15`$ $`a_0`$ .
Rescaling by the appropriate reduced mass, Eq. (3) implies for <sup>40</sup>K a nominal triplet scattering length of $`a_s^{\mathrm{triplet}}(40)=176`$ $`a_0`$. This result is consistent with the values obtained in a direct collisional measurement in <sup>40</sup>K . Finally, we take the singlet scattering length to be $`a_s^{\mathrm{singlet}}(40)=105`$ $`a_0`$ . This value is fairly well constrained by the existing data; moreover, the results of this paper depend only weakly on its exact value.
There remains the issue of the value of $`C_6`$ to employ in the calculations. The results of Marinescu et al. cover the fairly broad range $`C_6=3800\pm 200`$ atomic units . The accuracy of this result is limited by uncertainties in the atomic data used in the calculation. By contrast, a new high-precision calculation by Derevianko et al. predicts a much narrower range of $`C_6=3987\pm 15`$ . This improvement is largely due to Derevianko et al.’s accurate calculation of atomic structure, which freed them from experimental uncertainties. Their track record is impressive: for Na and Rb , their predictions are within experimental uncertainty of inferred values of $`C_6`$. This result lends credence to their value of $`C_6`$ for potassium, which we will adopt here. In this case the largest uncertainty in potassium scattering lengths arise from the $`\pm 25`$ in Eq. (3), rather than from $`C_6`$. Taking this uncertainty into account, and rescaling the mass, the <sup>40</sup>K triplet s-wave scattering length is given by $`a_s^{\mathrm{triplet}}(40)=176_{27}^{+77}`$ $`a_0`$.
Perhaps the most appealing candidate spin states in which to seek a Feshbach resonance would be the magnetically trappable states $`|fm=`$ $`|9/2,9/2`$ and $`|9/2,7/2`$, which are already trapped in the JILA experiment . However, as reported in Ref. , no such resonance exists. There may be resonances for nearby spin states, but these should be very narrow ($`\mathrm{\Delta }B1`$ gauss) and probably not useful for Cooper pairing.
There is, however, a broad resonance in collisions between the states $`|9/2,9/2`$ and $`|9/2,7/2`$, as illustrated in Fig. 1. This resonance, lying between 175 and 205 gauss, is easily accessible experimentally. Moreover its broad width implies that the scattering length can be tuned quite accurately. The inset to Fig. 1 focuses on the region near $`a_s=1000`$ $`a_0`$. To maintain this value of the scattering length to within $`3\%`$ (i.e., to maintain $`T_c`$ constant to within $`10\%`$, as discussed above) would require holding $`B`$ steady to within $`0.1`$ gauss. Since the two states are strong-field seekers, they cannot be trapped in the usual magnetic traps that have traditionally been used for BEC studies. Nevertheless, the two states could be held in an optical trap. These traps have recently attained great stability, with lifetimes exceeding 300 seconds . Note also that an optical trap ensures that the magnetic field can be made uniform across the entire trap, so that all atoms would experience the same pairing interaction. Evaporative cooling may be possible is these traps, as well .
These particular spin states are also appealing in terms of their stability against collisional losses. At the ultralow temperatures of interest here, p-wave collisions are strongly suppressed, meaning that there are virtually no losses due to collisions between atoms in the same spin state. Inelastic collisions that produce $`|9/2,5/2`$ states are also energetically forbidden, since the energy of this state lies 2.3 mK higher in energy than the $`|9/2,7/2`$ state at the magnetic fields considered. There would then remain only the collision process
$$|9/2,9/2+|9/2,7/2|9/2,9/2+|9/2,9/2.$$
(4)
This collision cannot occur in a spin-exchange process, which must conserve the sum $`m_a+m_b=8`$ of the magnetic quantum numbers. Nor can it proceed by the spin-spin dipolar interaction . This is because an incident s-wave can only couple to a d-wave final state in this processes, but d-waves are forbidden for identical final spin states. Thus the mixture we envision is virtually immune to two-body loss processes.
This leaves us with the possibility for three-body loss processes, where two bodies recombine into a molecule with the other carrying away the binding energy. These processes can generally contribute to heating, trap loss, or contamination with unwanted molecular states. They have been observed to exert a strong influence on Bose-Einstein condensates, especially near Feshbach resonances . The age of quantitative calculation of three-body recombination has just begun . Nevertheless, we can argue that these losses, too, are suppressed in this system. Roughly this is because any three-body collision in a two-component Fermi gas must involve two identical atoms. Again invoking the exclusion principle, these atoms must have a nonzero relative angular momentum, which effectively keeps them apart, suppressing the collision. Following the more careful hyperspherical treatment of the type in Ref. , this would lead to a threshold law where the three-body recombination rate vanishes at low $`E`$ as $`E^{1/2}`$, in contrast to the $`E`$-independent rate expected for bosons.
Finally, we return to the subject of possible p-wave Cooper pairing, similar to that envisioned in Ref. . Non-s-wave pairing is already known in superconductors and in superfluid <sup>3</sup>He. However, the ability to produce this pairing in a dilute, weakly interacting atomic gas, and moreover to control the strength of coupling, would enable detailed experimental and theoretical study, as has already been the case for dilute Bose condensates. In this case the pairing temperature, analogous to Eq. (1), is given by
$$T_c\frac{E_F}{k_B}\mathrm{exp}\left(\frac{\pi }{2(k_F|a_p|)^3}\right).$$
(5)
Here $`a_p`$ stands for the “p-wave scattering length,” defined by
$$a_p^3=\underset{k0}{lim}\frac{\delta _p(k)}{k^3},$$
(6)
where $`\delta _p(k)`$ is the p-wave scattering phase shift and $`k`$ is the wave number. The cubic dependence on $`a_p`$ of the exponential in (5) places more severe restrictions on $`a_p`$ than in the s-wave case. For example, for $`T_F=600`$ nK, setting $`a_p=1000`$ $`a_0`$ in (5) would yield a critical temperature of only $`T_c=0.002`$ nK, whereas for $`a_p=1500`$ we would get $`T_c=14`$ nK. In the latter case, if we again require that $`T_c`$ be constant to within $`10\%`$, we find that $`a_p`$ must be constant to within $`1\%`$.
For <sup>40</sup>K the naturally occurring value of the triplet p-wave scattering length is $`a_p^{\mathrm{triplet}}(40)=100`$ $`a_0`$, which is far too small to be of use. Fortunately, there are Feshbach resonances in this case, too. Generally speaking, these resonances lie at approximately the same values of magnetic field as the s-wave resonances, since each s-wave bound state that can resonate is accompanied by a p-wave bound state at a nearby energy. For example, in p-wave collisions of $`|9/2,7/2`$ and $`|9/2,5/2`$, there are extremely narrow resonances, as for the s-wave case.
We therefore again seek resonances in states with negative values of $`m_f`$. In particular, for collisions in a gas of pure $`|9/2,7/2`$ atoms, we find a fairly broad resonance at a position of $`B=191_{10}^{+5}`$ gauss, as illustrated in Fig. 2. This resonance has nearly the same shape as the familiar s-wave resonances, but with an additional inflection when $`a_p0`$, arising from the cube-root dependence of $`a_p`$ on $`\delta _p`$. This resonance is also somewhat narrower in magnetic field than the s-wave resonance reported above. In this case holding $`a_p1500`$ $`a_0`$ constant to within one percent requires holding the magnetic field constant to within perhaps 0.001 gauss.
In a given experiment the desired values of scattering lengths may differ from the sample values we have considered. In this case, it is useful to present approximate fitting formulas for the resonances, computed for the nominal interaction potentials. For s-waves, this fit is
$$a_s164\frac{1260}{(B196.2)}.$$
(7)
The p-wave fit, for $`a_p<0`$ and very near resonance, is
$$a_p600\frac{21}{(B191.02)}.$$
(8)
In each case, the scattering length is in $`a_0`$ and the field $`B`$ is in gauss.
In conclusion, these magnetic-field Feshbach resonances make possible a variable Cooper-pairing interaction in ultracold <sup>40</sup>K gases. Such interactions will enable detailed studies of s- and p-wave superfluid states, including their instabilities, at a level not possible before. Significantly, to implement these resonances requires no technology beyond what is currently available.
I thank D. Jin and C. Greene for useful discussions. This work was supported by the National Science Foundation. |
no-problem/9911/hep-lat9911038.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Lattice gauge fixing is unavoidable to compute non perturbatively the propagators of the fundamental fields appearing in the QCD Lagrangian . It is also necessary in non-perturbative renormalization schemes which use gauge dependent matrix elements to renormalize composite operators, and it becomes a fundamental technical ingredient in the so called non gauge invariant quantizations of chiral theories .
Up to now, the Landau gauge is the only covariant gauge for which an efficient numerical algorithm is available.
A few years ago, an unconventional method to implement a generic covariant gauge on the lattice was proposed . In principle this procedure takes into account the contributions from each gauge orbit with appropriate weight. With this method a gauge dependence of the gluon propagator, in particular at zero momentum, had been found . On the other hand the implementation of this method for physical lattices can be numerically demanding.
Here we follow a more conservative procedure to fix a generic covariant gauge on the lattice that stems from ref. . It is based, as in the Landau case, on the minimization of a functional $`H_A[G]`$ chosen in such a way that its absolute minima correspond to a gauge transformation $`G`$ satisfying the appropriate gauge condition (see section 2). In the continuum it is easy to show that this procedure is equivalent to the Faddeev-Popov quantization for covariant gauges, i.e. the two procedures lead to the same matrix elements for a generic gauge dependent operator. As usual the Faddeev-Popov factor can be written as a Gaussian integral of local Grassman variables, the resulting effective action is invariant under the BRST transformations and the correlation functions of the operators satisfy the appropriate Slavnov-Taylor identities. On the compact lattice the presence of Gribov copies is an obstruction to define partition functions with exact BRST invariance as it has been shown by H. Neuberger a few years ago . The Neuberger’s argument concerns a fundamental theoretical point whose relevance for the measurement of gauge dependent operators on the lattice must be clarified. Nevertheless, we believe that new numerical gauge fixing algorithms represent a concrete tool to improve the understanding of the lattice gauge properties.
In the following we propose a new numerical gauge fixing procedure which generalizes the usual Landau gauge fixing to generic covariant gauge. The numerical implementation of our procedure on the lattice uses a straight-forward generalization of the standard Landau algorithm, i.e. a steepest descent iterative algorithm minimizes a discretization of the new functional $`H_A[G]`$. In order to have an efficient and feasible gauge fixing procedure, the discretized functional $`H_U[G]`$ must be chosen carefully. In fact a naïve discretization of this functional, due to its complicated structure, would lead to an algorithm which either does not converge or takes too much computer time .
The main purpose of this paper is to show that a simple discretization $`H_U[G]`$ exists and leads to a minimization algorithm which converges with a good efficiency. The application of this algorithm to the non-perturbative evaluation of gauge dependent Green functions like the quark and gluon propagator will clarify the gauge dependence of the fitted parameters.
In order to show the feasibility of our method, we have computed the gluon propagator at different values of the gauge parameter at small volumes, finding a sensitive gauge dependence.
The plan of the paper is as follows. In section 2 we review the method proposed to fix a generic covariant gauge on the lattice. In section 3 we describe a very simple and efficient discretization of the gauge fixing functional. In section 4 we give the details of our numerical simulations and report our main numerical results.
## 2 Covariant Gauge Fixing
In this section we set the notation and briefly formulate the covariant gauge fixing method in the continuum. In the following we will neglect the problem of Gribov copies, assuming that a gauge section in the space of gauge fields intersects all gauge orbits once and only once. In the Landau gauge the expectation value of a gauge dependent operator is given by
$$𝒪=\delta A_\mu \delta \eta \delta \overline{\eta }𝒪e^{S(A)S_{ghost}(\eta ,\overline{\eta },A)}\delta (_\mu A_\mu ),$$
and the gauge fixing condition can be enforced non-perturbatively by minimizing Gribov’s functional
$$F_A[G]A^G^2=\text{Tr}\left(A_\mu ^GA_\mu ^G\right)d^4x.$$
(1)
The Landau gauge is readily extended to a general covariant gauge-fixing condition of the form
$$_\mu A_\mu ^G(x)=\mathrm{\Lambda }(x),$$
(2)
where $`\mathrm{\Lambda }(x)=\lambda ^a(x)\frac{T^a}{2}`$ belongs to the Lie algebra of the group and $`\text{Tr}(T^aT^b)=2\delta ^{ab}`$. Since gauge-invariant quantities are not sensitive to changes of gauge condition, it is possible to average over $`\mathrm{\Lambda }(x)`$ with a Gaussian weight
$$𝒪=\delta \mathrm{\Lambda }e^{\frac{1}{\alpha }{\scriptscriptstyle d^4xTr[\mathrm{\Lambda }^2]}}\delta A_\mu \delta \eta \delta \overline{\eta }𝒪e^{S(A)S_{ghost}(\eta ,\overline{\eta },A)}\delta (_\mu A_\mu \mathrm{\Lambda }),$$
(3)
obtaining the standard formula
$$𝒪=\delta A_\mu \delta \eta \delta \overline{\eta }𝒪e^{S(A)S_{ghost}(\eta ,\overline{\eta },A)}e^{\frac{1}{\alpha }{\scriptscriptstyle d^4xTr[(_\mu A_\mu )^2]}}.$$
This formula is adopted as a definition for the expectation value of gauge dependent operators. The invariance under BRST transformations of the effective action in eq. (2) leads to the Slavnov-Taylor identities among different correlation functions. In particular the longitudinal part of the gluon propagator has to be equal to the free one, i.e.
$$\frac{2}{N^21}\text{Tr}_\mu A_\mu (x)_\nu A_\nu (y)\alpha \delta (xy),$$
(4)
where $`N`$ is the number of colors. Following the usual technique, the gauge-fixing condition (2) is obtained non perturbatively by minimizing the new functional
$$H_A[G]d^4x\text{Tr}\left[(_\mu A_\mu ^G\mathrm{\Lambda })(_\nu A_\nu ^G\mathrm{\Lambda })\right],$$
(5)
which obviously reaches its absolute minima ($`H_A[G]=0`$) when eq. (2) is satisfied. Therefore in this case the Gribov copies of the equation (2) are associated with different absolute minima of eq. (5). All the stationary points of $`H_A[G]`$ correspond to the following gauge condition
$$D_\nu _\nu (_\mu A_\mu ^G\mathrm{\Lambda })=0,$$
(6)
where $`D_\nu `$ is the covariant derivative. Hence, due to the complexity of the new functional, its minimization could find “spurious” solutions which correspond to zero modes of the operator $`D_\nu _\nu `$ and do not satisfy the gauge condition in eq. (2). Of course the numerical minimization of the discretized version of eq. (5) can reach relative minima (spurious solutions) with $`H_A[G]0`$ which are not distinguishable from the absolute minima. Hence this could simulate the effect of an enlarged set of numerical Gribov copies. Within the limitations of this preliminary study we do not find any practical difference between the use of the new functional with respect to the standard Landau one (see results given in Fig. 2 below).
On the lattice, the expectation value of a gauge dependent operator $`𝒪`$ in a generic covariant gauge is
$$𝒪=\frac{1}{Z}𝑑\mathrm{\Lambda }e^{\frac{1}{\alpha }_xTr[\mathrm{\Lambda }^2]}𝑑U𝒪(U^{G_\alpha })e^{\beta S(U)},$$
(7)
which is the straight-forward discretization of Eq. (3), where $`\mathrm{\Lambda }`$ is dimensionless on the lattice. $`S(U)`$ is the Wilson lattice gauge invariant action and $`G_\alpha `$ is the gauge transformation that minimizes the discretized version of the functional (5). On the lattice, the correct adjustment to the measure is included in eq. (7) by evaluating the operator over the gauge rotated links. Therefore it is not necessary to introduce ghost fields but it is mandatory to fulfill the gauge fixing condition numerically in order to get $`G_\alpha `$.
## 3 The Driven Discretization
In the Wilson discretization of gauge theories, the fundamental fields are the links $`U_\mu `$ which act as parallel transporters of the theory. Hence the lattice fields $`A_\mu `$ are derived quantities which tend to the continuum gluon field as the lattice spacing vanishes. As a consequence on the lattice it is possible to choose different definitions of $`A_\mu `$ formally equal up $`O(a)`$ terms. In quantum field theory this ambiguity is well understood because any pair of operators, differing from each other by irrelevant terms, will tend to the same continuum operator. This feature, checked in perturbation theory, has been verified numerically at the non-perturbative level in ref. , where it has been shown that different definitions of the gluon field give rise to Green’s functions proportional to each other, guaranteeing the uniqueness of the continuum gluon field.
The freedom to choose the lattice definition of $`A_\mu `$ can be used to build discretized functionals which lead to efficient gauge-fixing algorithms. In the standard Landau gauge fixing, for example, the discretization of the functional (1) is given by
$$F_U[G]=\frac{1}{VTa^2g^2}Tr\underset{x,\mu }{}\left[U_\mu ^G(x)+U_\mu ^G(x)2I\right],$$
(8)
where $`V`$ is the 3-dimensional volume and $`T`$ the time size of the lattice. This formula corresponds, only up to $`O(a)`$ terms, to the naïve discretization of eq. (1) that is obtained from the standard lattice definition of the gluon field
$$A_\mu (x)\left[\frac{U_\mu (x)U_\mu ^{}(x)}{2iag}\right]_{Traceless}.$$
(9)
On the other hand $`F_U[G]`$ has the important property that it depends only linearly on $`G(\overline{x})`$, when the iterative algorithm visits the lattice point $`\overline{x}`$. This feature would be spoiled if $`F_U[G]`$ were defined assuming literally the naïve discretization (9) in eq. (1).
In order to study the convergence of the algorithm, two quantities are usually monitored as a function of the number of iteration steps: $`F_U[G]`$ itself and
$$\theta _F=\frac{1}{VT}\underset{x}{}Tr[\mathrm{\Delta }_F\mathrm{\Delta }_F^{}],$$
(10)
where
$`\mathrm{\Delta }_F(x)=\left[X_F(x)X_F^{}(x)\right]_{Traceless}{\displaystyle \frac{\delta F_U[G]}{\delta ϵ}},`$ (11)
being $`G=e^{iϵ^a\frac{T^a}{2}}`$ and
$`X_F(x)`$ $`=`$ $`{\displaystyle \underset{\mu }{}}\left(U_\mu (x)+U_\mu ^{}(x\mu )\right).`$ (12)
$`\mathrm{\Delta }_F(x)`$ is proportional to the first derivative of $`F_U[G]`$ and reaches zero as the functional is extremized. Therefore the quality of the gauge fixing is determined by the parameter $`\theta _F`$ which corresponds to the continuum quantity $`d^4x\text{Tr}(_\mu A_\mu )^2`$.
A naïve discretization of $`H_A[G]`$ will generate a quadratic dependence on $`G`$, which could prevent the convergence of the algorithm. This obstacle has been overcome by taking advantage of the freedom to choose the gluon field definition. In fact, as in the Landau case, it has been possible to find a discretization of $`H_A[G]`$ (”driven discretization”) that depends linearly on $`G(\overline{x})`$ and corresponds, up to $`O(a)`$ terms, to the continuum limit, eq. (5). This aim can be reached by choosing each different term of $`H_U[G]`$ in order to guarantee the local linear dependence on $`G(\overline{x})`$, instead of being the algebraic consequence of a particular $`A_\mu `$ definition. We propose the following compact form of $`H_U[G]`$:
$$H_U[G]=\frac{1}{VTa^4g^2}Tr\underset{x}{}J^G(x)J^G(x),$$
(13)
where
$`J(x)`$ $`=`$ $`N(x)ig\mathrm{\Lambda }(x),`$
$`N(x)`$ $`=`$ $`8I+{\displaystyle \underset{\nu }{}}\left(U_\nu ^{}(x\nu )+U_\nu (x)\right).`$ (14)
It is easy to see that locally $`H_U[G]`$ transforms linearly in $`G(\overline{x})`$ and its continuum limit is the functional (5). $`H_U[G]`$ is positive semidefinite and, unlike the Landau case, it is not invariant under global gauge transformations.
The functional $`H_U[G]`$ can be minimized using the same numerical technique adopted in the Landau case. In order to study the convergence of the algorithm, two quantities can be monitored as a function of the number of iteration steps: $`H_U[G]`$ itself and
$$\theta _H=\frac{1}{VT}\underset{x}{}Tr[\mathrm{\Delta }_H\mathrm{\Delta }_H^{}],$$
(15)
where
$`\mathrm{\Delta }_H(x)`$ $`=`$ $`\left[X_H(x)X_H^{}(x)\right]_{Traceless}{\displaystyle \frac{\delta H_U[G]}{\delta ϵ}}`$ (16)
and
$`X_H(x)`$ $`=`$ $`{\displaystyle \underset{\mu }{}}\left(U_\mu (x)J(x+\mu )+U_\mu ^{}(x\mu )J(x\mu )\right)`$ (17)
$``$ $`8J(x)72I+igN(x)\mathrm{\Lambda }(x).`$
$`\mathrm{\Delta }_H`$ is the driven discretization of the eq. (6); it is proportional to the first derivative of $`H_U[G]`$ and, analogously to the continuum, it is invariant under the transformations $`\mathrm{\Lambda }(x)\mathrm{\Lambda }(x)+C`$, where $`C`$ is a constant matrix belonging to the $`SU(3)`$ algebra. During the minimization process $`\theta _H`$ decreases to zero and $`H_U[G]`$ becomes constant. The quality of the convergence is measured by the final value of $`\theta _H`$.
## 4 Numerical Simulations and Results
We have generated 50 $`SU(3)`$ thermalized link configurations using the Wilson action with periodic boundary conditions at $`\beta =6.0`$ for $`8^4`$ and $`8^3\times 16`$ volumes. Following the prescription contained in eq. (7), for each Monte Carlo configuration and for each lattice site we have extracted a matrix $`\mathrm{\Lambda }(x)`$ according to a Gaussian distribution at a fixed $`\alpha `$ value. The gauge-fixing code implements an iterative overrelaxed minimization algorithm for $`F_U[G]`$ and $`H_U[G]`$. We have monitored the quantities $`F_U`$ and $`\theta _F`$ for the standard Landau gauge-fixing algorithm and $`H_U[G]`$ and $`\theta _H`$ for the new one after every lattice sweep. We have found that the value of the overrelaxing parameter $`\omega `$ adopted for the Landau algorithm $`\omega =1.72`$ is a good choice also for the new gauge fixing.
In Fig. 1 we report the values of $`\theta _H`$ for different $`\alpha `$ values as a function of the gauge fixing sweeps for a typical thermalized configuration at $`\beta =6.0`$ and $`VT=8^4`$.
Each sweep of the new algorithm takes $`15\%`$ more computer time with respect to the Landau case. For $`\alpha =0`$, which in this new procedure corresponds to the Landau gauge and is obtained by taking $`\mathrm{\Lambda }=0`$, the number of sweeps necessary to fix the gauge is $`1.2`$ times that of the standard algorithm. In the case of $`\alpha 0`$, the number of sweeps to minimize $`H_U[G]`$ with a given precision increases when $`\alpha `$ decreases.
Once the configuration have been rotated in a given gauge, we have tested our procedure computing the following two point correlation functions
$`𝒜_0𝒜_0(t)`$ $``$ $`{\displaystyle \frac{1}{V^2}}{\displaystyle \underset{𝐱,𝐲}{}}\text{Tr}A_0(𝐱,t)A_0(𝐲,0),`$ (18)
$`𝒜_i𝒜_i(t)`$ $``$ $`{\displaystyle \frac{1}{3V^2}}{\displaystyle \underset{i}{}}{\displaystyle \underset{𝐱,𝐲}{}}\text{Tr}A_i(𝐱,t)A_i(𝐲,0),`$ (19)
$`𝒜𝒜(t)`$ $``$ $`{\displaystyle \frac{1}{4V}}{\displaystyle \underset{\mu ,\nu }{}}{\displaystyle \underset{𝐱}{}}\text{Tr}_\mu A_\mu (𝐱,t)_\nu A_\nu (𝐱,0),`$ (20)
where $`\mu `$ and $`\nu `$ run from 1 to 4, $`i`$ from 1 to 3 and the trace is over the color indices. In eqs. (18)-(20) the definition (9) is adopted for the gluon field on the lattice and $`_\mu `$ indicates the usual backward derivative. The correlators in eqs. (18) and (19) are relevant to the investigation of the QCD gluon sector, while we use the correlation in eq. (20) to check the Slavnov-Taylor identity for the longitudinal component of the gluon propagator. The statistical errors for the correlation functions have been estimated by the jacknife method. A thorough study of these operators will be presented in a forthcoming paper .
In Fig. 2 the Green functions computed with the standard Landau gauge fixing $`𝒜_i𝒜_i_F`$ and $`𝒜_i𝒜_i_H`$ evaluated with the new algorithm with $`\alpha =0`$, are reported.
The remarkable agreement shows that the new algorithm reproduces very well the results of the standard one within an overall scale factor close to 1. Moreover, the possible spurious solutions of the new gauge condition apparently do not affect this operator at least at $`\alpha =0`$.
Although in this paper we did not perform a systematic study of the Gribov copies for this gauge fixing procedure, a rough preliminary search shows, for $`\alpha =0`$, the same pattern of copies of the Landau case.
In Fig. 3 the data for the correlation (18) are reported. For $`\alpha =0`$ they show the well known flat behaviour forced by the Landau gauge.
For $`\alpha 0`$ the data show an almost flat curve with an enhancement at $`t=1`$ probably due to the effect of contact terms. In this case the flatness stems from the fact that $`d^3x_0A_00`$ because the average value of $`\mathrm{\Lambda }`$ on each time slice is negligible.
In Fig. 4 our results for the correlations $`𝒜_i𝒜_i_H`$ for different $`\alpha `$ values as a function of the time slice $`t`$ are reported for a set of $`50`$ SU(3) configurations at $`\beta =6.0`$ and $`VT=8^4`$.
Although in the case of small volume and a small number of configurations, the gauge dependence of the gluon propagator is clearly shown. For increasing $`\alpha `$ values, the time dependence of the gluon propagator becomes flatter since it approaches a limit where the gauge fixing effects disappear. It is also interesting to note that the $`\alpha `$ dependence of the gluon propagator shown in Fig. 4 does not seem to be re-absorbed by an overall scaling factor. This plot shows the feasibility of this procedure to study the gauge dependence of physically interesting correlators.
We complete the numerical checks of the gauge-fixing we are proposing, showing the numerical behaviour of a particular Slavnov-Taylor identity. This identity is a direct consequence of our way to implement the covariant gauge fixing and it must be understood as a numerical check of the procedure. This check measures the correlator of the operator $`𝒜`$ with itself, and if this operator has been well gauge-fixed, it is related to the autocorrelator of the $`\mathrm{\Lambda }`$’s and then it must be a delta function. In fact, for a finite volume $`VT`$ with periodic boundary conditions, the naïve Slavnov-Taylor identity gives for the correlator of eq. (20) the following relation:
$$𝒜𝒜(t)\alpha (\delta _{t,0}\frac{1}{VT})+𝒪(a).$$
(21)
Eq. (21) shows that the shape of the correlation function (20) is expected to be proportional to a $`\delta `$-function up to the square of the $`_\mu A_\mu (x)`$ renormalization constant, apart from volume and $`𝒪(a)`$ effects. In order to verify eq. (21), in Fig. 5 we plot the correlation functions (20) versus the time $`t`$. We believe that the good agreement between these data and eq. (21) is an indication of the absence of possible spurious solutions induced by our method. The small signal at $`t=2,8`$ is likely due to the residual correlation induced by the enhancement in the point $`t=1`$.
## 5 Conclusions
In this paper we have described an efficient method to fix non-perturbatively a generic covariant gauge on the lattice. This procedure is equivalent to the Faddeev-Popov quantization for covariant gauges. Therefore it can be used to compare the lattice renormalized correlation functions with the same quantities computed in continuum perturbation theory.
We have tested the algorithm on different SU(3) lattices for different values of the gauge parameter $`\alpha `$. Our numerical data show that the statistical fluctuations of measured quantities is comparable to the standard Landau gauge fixing algorithm.
We have computed the correlation functions relevant for the investigations of the gluon propagator at a few $`\alpha `$ values and we have found a sensitive dependence on the gauge parameter.
In collaboration with the Boston University Center for Computational Science, we are applying this method to the study of the gluon propagator on physical lattices.
## Acknowledgments
We warmly thank Massimo Testa for many fruitful discussions on this subject.
L.G. thanks Claudio Rebbi for interesting discussions on this subject.
We thank the Center for Computational Science of Boston University where part of this computation has been done.
L. G. was supported in part under DOE grant DE-FG02-91ER40676.
S. P. has been partially supported by the INFN-MIT “Bruno Rossi” Exchange Program while part of this paper was written. He thanks all the members of the MIT Center for Theoretical Physics for their warm hospitality. |
no-problem/9911/astro-ph9911162.html | ar5iv | text | # The Spectral Appearance of Primeval Galaxies
## 1. The zoo of high–redshift galaxies
The search for primeval galaxies has been one of the long–term programs of observational cosmology, along with the development of sensitive detectors, and is currently receiving an exciting boost with the new generation of 8–metre class telescopes. However, the search itself is somewhat hampered by the fuzziness of the concept of “primeval galaxy”. This is generally understood as being “a galaxy which is captured at the epoch of its formation”. Since there are several competing theories about what a forming galaxy should look like, it is readily possible that we are missing part of the process of galaxy formation because we do not know yet what to search for.
More specifically, two general paradigms have been proposed in the last twenty years or so. In the picture of monolithic collapse, galaxies form at a given epoch $`z_{for}`$, when the physical conditions of the universe are favourable, and evolve at different rates which are fixed by the initial conditions. In the picture of hierarchical collapse, there is nothing like a given redshift of galaxy formation. Larger galaxies form from the merging of smaller ones, which on their turn have formed from the merging of still smaller lumps, and so on. The beginning of the process took place at some early redshift $`z30`$ (when the first objects can cool) and is still going on now. As a result, the “epoch of galaxy formation” can be defined e.g. as the epoch when the first stars formed, when 50 % of the stars have formed, or when the morphology was fixed after the last major merging event. This hierarchical galaxy formation is now modelled in the context of hierarchical clustering where dark matter completely rules gravitational collapse.
The models of galaxy formation have to reproduce the wide variety of objects which are now observed at high redshift, probably after strong observational biasses that are not fully understood. These objects are generally selected according to photometric criteria. We now have Luminous Blue Compact Galaxies (LBCGs; $`z1`$ and $`I_{AB}<22.5`$), Lyman–Break Galaxies (LBGs; $`z3`$ or $`z4`$, and $`I_{AB}<25`$), Extremely Red Objects (EROs; $`RK>6`$ and $`K<19.5`$), Lyman Alpha Galaxies (LAGs; $`z4`$ or $`z5`$, and $`EW(\mathrm{Ly}\alpha )>100`$ Å), faint submm sources that are likely to be the high–$`z`$ counterparts of the local “Ultraluminous Infrared Galaxies” discovered by iras (high–$`z`$ ULIRGs; $`z>1`$ and $`L_{IR}>10^{12}`$ $`L_{}`$), Damped Lyman $`\alpha `$ Absorbers (DLAs; $`N_H>10^{22}`$ atom cm<sup>-2</sup>), etc. In order to link the predictions of the models of galaxy formation with these observations, it is necessary to model the spectral energy distributions (SEDs) on the widest wavelength range, from the rest–frame UV to the rest–frame submm.
## 2. The case for high–redshift dust
As a matter of fact, several pieces of observational evidence are now converging to show that there is a significant amount of extinction in high–redshift galaxies. Consequently, the absorption and emission processes due to dust cannot be neglected in assessing the luminosity budget of forming galaxies. Among these pieces of evidence, we can quote:
(i) The discovery of the Cosmic IR Background at a level 10 times higher than the no–evolution predictions based on the iras luminosity functions, and twice as high as the Cosmic Optical Background (Puget et al. 1996, Guiderdoni et al. 1997, Fixsen et al. 1998, Hauser et al. 1998, Lagache et al. 1999).
(ii) The IR and submm counts that have broken the CIRB into its brightest contributors at 15 $`\mu `$m (ISOCAM down to $`0.1`$ mJy, Aussel et al. 1999, Elbaz et al. 1999), 175 $`\mu `$m (ISOPHOT down to $`100`$ mJy, Kawara et al. 1998, Puget et al. 1999), and 850 $`\mu `$m (SCUBA down to 2 mJy, Smail et al. 1997, Hughes et al. 1998, Barger et al. 1998, 1999a, Eales et al. 1999). Although the poor spatial resolution of the observations makes the identification of the optical counterparts somewhat difficult, the first results of the spectroscopic follow–ups seem to show that some of the sources are at $`z>1`$ (Smail et al. 1998, Lilly et al. 1999, Barger et al. 1999b).
(iii) At least some of the EROs (e.g. HR10 at $`z=1.44`$, Cimatti et al. 1998a) are high–redshift dusty objects.
(iv) The extinction in LBCGs at $`z1`$ corresponds to a factor 3 extinction on the rest–frame flux at 2800 Å (Flores et al. 1999).
(v) The extinction in LBGs at $`z3`$ and 4 is high ($`0.1E(BV)0.5`$), with a trend of a larger extinction for the brightest objects (Steidel et al. 1999, Meurer et al. 1999). This corresponds to a factor 5 extinction on the rest–frame flux at 1600 Å.
(vi) Dust is present in high–redshift radiogalaxies and QSOs, up to $`z=4.69`$ (see e.g. Hughes et al. 1997, Cimatti et al. 1998b). The optical spectrum of a gravitationally–lensed galaxy at $`z=4.92`$ already shows a reddening factor amounting to $`0.1<E(BV)<0.3`$ (Soifer et al. 1998).
(vii) The correlation of dust with metallicity has still to be understood, since, for instance, the extremely metal–poor galaxy SBS0335-052 (1/30 of solar) has a significant IR emission (Thuan et al. 1999).
## 3. Spectral energy distributions of young galaxies
We hereafter present a model of synthetic SEDs, called stardust, which contains up–to–date spectrophotometric modelling coupled to chemical evolution and dust absorption. The model allows one to make consistent predictions for the SEDs of galaxies in a very broad wavelength range. Details can be found in Devriendt et al. (1999; hereafter DGS).
The model uses the so–called “isochrone scheme” and includes a compilation of the Geneva tracks (see e.g. Charbonnel et al. 1996, and references therein) for various metallicities $`Z`$, and masses 0.8 $`M/M_{}`$ 120. For stars less massive than $`1.7M_{}`$, some of the old tracks stop at the Giant Branch tip. More recent grids of models, based on Geneva tracks and covering the evolution of low mass stars (0.8 to 1.7 $`M_{}`$) from the Zero–Age Main Sequence up to the end of the Early–AGB, are included for $`Z=`$ 0.001 and $`Z=`$ 0.02. For the late stages of other metallicities (Horizontal Branch, Early–AGB), we either interpolate or extrapolate $`\mathrm{log}L_{bol}`$, $`\mathrm{log}T_{eff}`$ and $`\mathrm{log}t`$ versus $`\mathrm{log}Z`$ from the available tracks. The final stages of the stellar evolution (Thermally–Pulsing–AGB and Post–AGB) are not included. We use the grid of theoretical fluxes from Kurucz (1992) which covers all metallicities from $`\mathrm{log}Z/Z_{}=+1.0`$ to $`\mathrm{log}Z/Z_{}=5.0`$, and 61 temperatures from 3500 K to 50 000 K. Each spectrum spans a wavelength range between 90 $`\mathrm{\AA }`$ and 160 $`\mu `$m, with a mean resolution of 20 $`\mathrm{\AA }`$. For the coldest stars (K and M-type stars) with $`T`$ 3750 K, Kurucz’s models fail to reproduce the observed spectra. Therefore, we prefer to use models coming from different sources (see DGS).
Fig 1 gives the passive evolution of a Simple Stellar Population with Salpeter IMF (slope $`x=1.35`$ from $`m=0.1`$ to $`120`$ $`M_{}`$). Fig 2 compares our stardust model with other models available in the literature (see DGS for details) : gissel 1998 (Bruzual and Charlot, 1993), pégase (Fioc and Rocca–Volmerange, 1997), and our stardust. Though the stellar data are different, the agreement is generally very good in the UV/visible, except when Post—AGB are dominant (that is, in old stellar populations without star formation activity). In particular, the UV to SFR ratio that is used to derive the cosmic SFR from rest–frame 1600 Å or 2800 Å observations is remarkably similar in the three models.
Our model computes chemical evolution and takes into account the effect of metallicity on the stellar tracks and stellar spectra in a consistent way. In addition, the metallicity of the gas is followed. Under simple assumptions on the variation of the extinction curve with metallicity (based on the study of the Milky Way, The Large Magellanic Cloud, and the Small Magellanic Cloud), and on the geometry and relative distributions of dust and stars (homogeneous mix in an oblate spheroid), transfer can be easily solved, and the amount of luminosity absorbed by dust is estimated. The SEDs in the IR/submm are then computed to reproduce the correlation of iras colours with total IR luminosity $`L_{IR}`$, following Guiderdoni et al. (1997, 1998). The model reproduces the obscuration curve of Calzetti et al. (1994) observed in a sample of nearby starbursts. It also reproduces the correlation of IR to 2200 Å flux with the UV slope $`\beta `$ of the SEDs (around 2200 Å) observed by Meurer et al. (1995).
Finally, the optical and IR/submm spectra are connected to give evolving synthetic SEDs from the far UV (90 Å) to the submm ($`1`$ mm). In the centimetre and metre range, a radio component can be added under assumptions on the slope and on the correlation of radio fluxes with IR fluxes. The SEDs depend on three parameters (in addition to the IMF) : the SFR timescale $`t_{}`$, the age $`t`$, and a concentration parameter $`f_H`$ that describes the size of the gaseous disk ($`f_H=1`$ is for normal spirals, and a radial collapse by a factor 10 corresponds to $`f_H=100`$). Fig 3 gives typical spectra for a spiral and a forming galaxy, without and with extinction (respectively with $`f_H=1`$ and $`f_H=100`$, this latter value being typical of an ULIRG).
## 4. From local to high–redshift ULIRGs
Fig 4 gives the fits of a sample of local galaxies obtained with our theoretical SEDs and a $`\chi ^2`$ procedure. The sample gathers Virgo Cluster spirals (Boselli et al. 1998) and ULIRGs (Rigopoulou et al. 1996) that have a sufficient number of photometric points at optical, NIR, MIR, FIR, and submm wavelengths (and, for some of them, in the radio). Nine spectra have been ordered in fig 4 according to their $`L_{IR}`$, in a sequence that parallels the compilation of Sanders & Mirabel (1996). There is some degeneracy between the parameters $`t_{}`$ and $`t`$ of the fits, but their combination corresponds to similar SFRs. This is illustrative of the ability of the model to capture the characteristic features of the objects.
This can be used to study the efficiency of the UV and IR fluxes to trace the underlying SFR. As shown in fig 5, the UV fluxes do scale with SFR for optically–thin galaxies (slope $`1`$), but the proportionality breaks down for ULIRGs. Without any information on the optical thickness of the high–redshift galaxies, the rest–frame UV fluxes could lead to erroneous estimates. In contrast, the IR fluxes trace SFRs on several orders of magnitudes, with a slope 0.95, and independently of the optical thickness. The origin of this behaviour is not clear yet. It is probably due to the fact that there are basically two regimes : either the galaxy is optically thin, and the IR roughly corresponds to a constant fraction of the UV, or it is optically thick, and all the UV flux is absorbed by dust and released in the IR.
These galaxies can also be used to predict the photometric properties at higher redshift. Fig 6 and 7 give the observer–frame optical magnitudes and IR/submm fluxes versus redshift.
## 5. Semi–analytic and hybrid models of galaxy formation
These synthetic SEDs can subsequently be implemented into semi–analytic models of galaxy formation. The interest of such an approach is that the distribution of the SFR timescales $`t_{}`$ can be computed from the distribution of dynamical timescales $`t_{dyn}`$ of the host haloes, under the assumption $`t_{}=\beta t_{dyn}`$, where $`\beta `$ is an efficiency parameter which is fixed observationally (see e.g. Kennicutt 1998). The ages $`t`$ of the stellar populations are computed from the formation redshifts and the redshifts at which the galaxies are observed. The distribution of the gas column densities is also computed from the size distribution of disks (obtained from the size distribution of haloes after conservation of angular momentum). The ULIRGs correspond to a further radial collapse by a factor 10 ($`f_H=100`$).
Following Guiderdoni et al. (1997, 1998), Devriendt & Guiderdoni (1999) proposed results of faint galaxy counts that are produced by a simple, semi–analytic prescription using the peaks formalism. The usual recipes are implemented for gas cooling, dissipative collapse, and stellar feedback. It is assumed that dust is heated by starbursts, and that the Star Formation Rate is $`SFR(t)=M_{gas}/t_{}`$, with $`t_{}\beta t_{dyn}`$. The spectra are taken from stardust. This paper gives predictions of faint counts at optical, IR, and submm wavelengths, for a variety of cosmologies. It turns out that the ionizing flux escaping from high–redshift galaxies estimated with this model is unable to ionize the intergalactic medium (Devriendt et al. 1998).
## 6. Conclusions and prospects
These synthetic SEDs and the spectral templates of fig 4 are useful tools to analyse the panchromatic view on high–redshift, young and “primeval” galaxies. The spectral templates are available upon request to the authors. The implementation of such spectra into semi–analytic models is a promising way to predict the statistical properties of a host of objects that are now observed at high redshift, through various spectral windows, and with various selection criteria.
However, the modelling of the SFR history of galaxies in the context of hierarchical galaxy formation is still incomplete. In particular, the ULIRGs appear to be key objects to understand the formation of bulges. A more realistic treatment of such objects requires the monitoring of the merging rates, and involves keeping track of the spatial and dynamical information in models of galaxy formation. This can be done, for instance, with the new generation of the so–called “hybrid” models in which the merging history trees are directly built from the outputs of cosmological N–body simulations. The implementation of stardust into such models is in progress.
## References
Aussel, H., Cesarsky, C.J., Elbaz, D., Starck, J.L., 1999, A&A, 342, 313
Barger, A.J., Cowie, L.L., Sanders, D.B., Fulton, E., Taniguchi, Y., Sato, Y., Kawara, K., Okuda, H., 1998, Nature, 394, 248
Barger, A.J., Cowie, L.L., Sanders, D.B., 1999a, ApJ, 518, L5
Barger, A.J., Cowie, L.L., Smail, I., Ivison, R.J., Blain, W., Kneib, J.P., 1999b, AJ, 117, 2656
Boselli, A., Lequeux, J., Sauvage, M., Boulade, O., Boulanger, F., Cesarsky, D., Dupraz, C., Madden, S., Viallefond, F., Vigroux, L., 1998, A&A, 335, 53
Bruzual, A.G., Charlot, S., 1993, ApJ, 405, 538
Calzetti, D., Kinney, A.L., Storchi–Bergmann, T., 1994, ApJ, 429, 582
Charbonnel, C., Meynet, G., Maeder, A., Schaerer, D., 1996, A&AS, 115, 339
Cimatti, A., Andreani, P., Rottgering, H., Tilanus, R., 1998, Nature, 1998a, 392
Cimatti, A., Freudling, W., Rottgering, H., Ivison, R.J., Mazzei, P., 1998b, A&A, 329, 399
Devriendt, J.E.G., Sethi, S., Guiderdoni, B., Nath, B., 1998, MNRAS, 298, 708
Devriendt, J.E.G., Guiderdoni, B., Sadat, R., 1999, A&A, 350, 381
Devriendt, J.E.G., Guiderdoni, B., 1999, submitted
Eales, S., Lilly, S., Gear, W., Dunne, L., Bond, J.R., Hammer, F., Le Fèvre, O., Crampton, D., 1999, ApJ, 515, 518
Elbaz D., Aussel H., Cesarsky C.J., Desert F.X., Fadda D., Franceschini A., Harwit, M., Puget J.L., Starck J.L., 1999, in The Universe as seen by ISO, P. Cox & M.F. Kessler (eds), 1998, UNESCO, Paris, ESA Special Publications series (SP-427)
Fioc, M., Rocca–Volmerange, B., 1997, A&A, 326, 950
Fixsen, D.J., Dwek, E., Mather, J.C., Bennett, C.L., Shafer, R.A., 1998, ApJ, 508, 123
Flores, H., Hammer, F., Thuan, T.X., Cesarsky, C., Désert, F.X., Omont, A., Lilly, S.J., Eales, S., Crampton, D., Le Fèvre., O., 1999, ApJ, 517, 148
Guiderdoni, B., Bouchet, F.R., Puget, J.L., Lagache, G., Hivon, E., 1997, Nature, 390, 257
Guiderdoni, B., Hivon, E., Bouchet, F.R., Maffei, B., 1998, MNRAS, 295, 877
Hauser, M.G., Arendt, R., Kelsall, T., Dwek, E., Odegard, N., Welland, J., Freundenreich, H., Reach, W., Silverberg, R., Modeley, S., Pei, Y., Lubin, P., Mather, J., Shafer, R., Smoot, G., Weiss, R., Wilkinson, D., Wright, E., 1998, ApJ, 508, 25
Hughes, D., Dunlop, J.S., Rawlings, S., 1997, MNRAS, 289, 766
Hughes, D., Serjeant, S., Dunlop, J., Rowan–Robinson, M., Blain, A., Mann, R.G., Ivison, R., Peacock, J., Efstathiou, A., Gear, W., Oliver, S., Lawrence, A., Longair, M., Goldschmidt, P., Jenness, T., 1998, Nature, 394, 241
Kawara, K., Sato, Y., Matsuhara, H., Taniguchi, Y., Okuda, H., Sofue, Y., Matsumoto, T., Wakamatsu, K., Karoji, H., Okamura, S., Chambers, K.C., Cowie, L.L., Joseph, R.D., Sanders, D.B., 1998, A&A, 336, L9
Kennicutt, R.G., 1998, in Starbursts: Triggers, Nature and Evolution, B. Guiderdoni & A. Kembhavi (eds), EDP Sciences/Springer–Verlag
Kurucz, R., 1992, IAU Symposium 149, 225
Lagache, G., Abergel, A., Boulanger, F., Désert, F.X., Puget, J.L., 1999, A&A, 344, 322
Lilly, S.J., Eales, S.A., Gear, W.K.P., Hammer, F., Le Fèvre, O., Crampton, D., Bond, J.R., Dunne, L., 1999, ApJ, 518, 641
Meurer, G.R., Heckman, T.M., Leitherer, C., Kinney, A. Robert, C., Garnett, D.R., 1995, AJ, 110, 2665
Meurer, G.R., Heckman, T.M., Calzetti, D., 1999, ApJ, 521, 64
Puget, J.L., Abergel, A., Bernard, J.P., Boulanger, F., Burton, W.B., Désert, F.X., Hartmann, D., 1996, A&A, 308, L5
Puget, J.L., Lagache, G., Clements, D.L., Reach, W.T., Aussel, H., Bouchet, F.R., Cesarsky, C., Désert, F.X., Dole, H., Elbaz, D., Franceschini, A., Guiderdoni, B., Moorwood, A.F.M., 1999, A&A, 345, 29
Rigopoulou, D., Lawrence, A., Rowan–Robinson, M., 1996, MNRAS, 278, 1049
Smail, I., Ivison, R.J., Blain, A.W., 1997, ApJ, 490, L5
Smail, I., Ivison, R.J., Blain, A.W., Kneib, J.P., 1998, ApJ, 507, L21
Soifer, B.T., Neugebauer, G., Franx, M., Matthews, K., Illingworth, G.D., 1998, ApJ, 501, L171
Steidel, C.C., Adelberger, K.L., Giavalisco, M., Dickinson, M., Pettini, M., 1999, ApJ, 519, 1
Thuan, T.X., Sauvage, M., Madden, S., 1999, ApJ, 516, 783 |
no-problem/9911/nucl-th9911043.html | ar5iv | text | # Low-𝒎_⊥ 𝝅⁺-𝝅⁻ Asymmetry Enhancement from Hadronization of QGP
\[
## Abstract
We show that in sudden hadronization of QGP a non-equilibrium value of the pion phase space occupancy parameter $`\gamma _q>1`$ is expected in order to accommodate the entropy excess in QGP and the process of gluon fragmentation. When $`\gamma _q`$ is near to its maximum allowed value, pion overabundance is shown to arise at low $`m_{}`$, where the charged pion asymmetry is also amplified. These effects are considered and we show that their magnitude suffices to explain pertinent experimental data obtained in 160–200$`A`$ GeV Pb- and S- induced reactions.
\]
Highly relativistic collisions of heavy nuclei are under experimental and theoretical study of which the primary aim is the discovery of the deconfined quark-gluon plasma (QGP) phase , and the understanding of its transformation into hadrons. We shall demonstrate how the measurement of low $`m_{}`$ central rapidity abundance of charged pion yields, $`dN(\pi ^{})/dm_{}`$ and $`dN(\pi ^+)/dm_{}`$, provides a sensitive way to probe the outcome of hadronization of QGP. We will, in particular, show that the relative deformation of low-$`m_{}`$ pion spectra provides a measure for chemical (particle) non-equilibrium overabundance phenomenon.
Thermal equilibration (i.e., equilibration of the energy distribution) of final state hadrons is seen in particle spectra to be a well working hypothesis. This, however, does not imply chemical (i.e., abundance) equilibration. Since chemical hadron abundance equilibrium is approached on the same time scale as the duration of the heavy ion collisions, it has been proposed to analyze the abundances of final state hadrons allowing chemical non-equilibrium phase space occupancies . This has been applied initially only to describe the abundance of strange quarks , which were early on understood to be not necessarily produced rapidly enough . More recently, it has been suggested that chemical equilibration of light quarks and gluons is indeed not assured at RHIC energies and beyond . On the other hand, at SPS, the excess of pion multiplicity has been observed and understood to result from excess entropy content of the primordial high entropy phase formed in the interaction . This leads naturally to the hypothesis of light quark overabundance generated, e.g., by gluon fragmentation in hadronization of the QGP high entropy phase. This can be most easily described by counting light valence quark abundance in hadrons , as was done for strangeness.
Indeed, when the multi-particle production processes, in 158–200$`A`$ GeV S– and Pb–Pb collisions carried out at CERN-SPS, have been analyzed using the methods of the statistical Fermi model , light quark overabundance and the associated chemical non-equilibrium was found. Overall, the results of this analysis show several features that combined do make a strong case that at least the strange hadronic particles seen at CERN-SPS, are emerging nearly directly from sudden hadronization of deconfined QGP phase of hadronic matter . Moreover, the recognition that the hadron abundances are better described with oversaturation of the quark phase space allows a notable reduction of the chemical freeze-out temperature, $`T_f=145\pm 5`$ MeV, accompanied by a reduction of the baryochemical potential, $`\mu _B=200`$–210 MeV, is consistent with the underlying dynamical picture. What remains open is the issue if the predominant hadronic fraction, pions, also follow the sudden hadronization model, a point we address here.
This general discussion introduced already a few statistical parameters of the Fermi model of well established meaning and we will complete the list now. Aside of the chemical freeze-out temperature $`T_f`$, further chemical parameters control hadron abundances: we note the light quark fugacity $`\lambda _u`$ and $`\lambda _d`$ and the phase space occupancy of light quarks $`\gamma _q=\gamma _u=\gamma _d`$. We recall that $`\gamma _q`$ controls overall abundance of quark pairs, while $`\lambda _i`$ control the difference between quarks and anti-quarks of given flavor. The difference between $`\lambda _i`$ and $`\gamma _i`$ is that, for quarks and anti-quarks the same factor $`\gamma _i`$ applies, while the antiparticle fugacity is inverse of the particle fugacity. The phase space occupancy $`\gamma _q`$ is determined to exceed unity significantly, once introduced: results obtained switching on progressively more parameters of the Fermi model show considerable improvement in the capability of the model to describe the data .
The proper statistical physics foundation of $`\gamma _i`$ is obtained considering the maximum entropy principle: it has been determined that while the limit $`\gamma _i1`$ maximizes the specific chemical entropy, this maximum is not very pronounced, indicating that a system with dynamically evolving volume will, in general, find more effective paths to increase entropy, than offered by the establishment of the absolute chemical equilibrium .
It is usual to introduce the geometric average of the light quark fugacities, related to the baryochemical potential:
$$\lambda _q^2\lambda _d\lambda _u,\mu _qT\mathrm{ln}\lambda _q,\mu _B3\mu _q.$$
(1)
Similarly, one introduces the chemical potentials for up and down quarks:
$$\mu _dT\mathrm{ln}\lambda _d,\mu _uT\mathrm{ln}\lambda _u,\delta \mu \mu _d\mu _u.$$
(2)
The difference in the number of net up ($`u\overline{u}`$) and net down ($`d\overline{d}`$) quarks participating in the dense matter fireball is given by the initial condition, i.e., which nuclei collide. In a hadron gas, $`\delta \mu /\mu _q`$ varies as function of chemical freeze-out temperature as shown in \[8, Fig. 1\].
The relative number of particles is controlled by fugacity and phase space occupancies of the constituents. The composite hadronic particle chemical occupancy and fugacity is expressed by constituent contributions:
$$\lambda _i=\underset{ji}{}\lambda _j,\gamma _i=\underset{ji}{}\gamma _j.$$
(3)
The abundances of particles produced in QGP disintegration is most conveniently described by considering the phase space distribution of particles: the abundance of a hadron $`h`$, with $`j`$-components, freezing out from the source is
$$N_h\gamma _he^{E_h/T},\gamma _h\underset{jh}{}\gamma _j\lambda _j.$$
(4)
For example, the effective chemical fugacities for $`\pi ^{}`$ composed of light quark-antiquark pair of a differing flavor is
$$\gamma _\pi ^{}=\gamma _q^2\frac{\lambda _d}{\lambda _u},\gamma _{\pi ^+}=\gamma _q^2\frac{\lambda _u}{\lambda _d},$$
(5)
while for neutral pions we simply have:
$$\gamma _{\pi ^0}=\gamma _q^2.$$
(6)
The case of pions is, however, exceptional in another way: the SPS data analysis shows that the pion yield is governed by a large fugacity, and thus it is imperative to revert in the Fermi model to use Bose distribution function:
$`{\displaystyle \frac{d^3N_\pi }{d^3p}}={\displaystyle \frac{1}{\gamma _\pi ^1e^{E_\pi /T}1}},E_\pi =\sqrt{m_\pi ^2+p^2}.`$ (7)
We see that the range of values for $`\gamma _\pi `$ is bounded from above by the Bose singularity:
$$\gamma _\pi <\gamma _\pi ^c=e^{m_\pi /T}.$$
(8)
When $`\gamma _\pi \gamma _\pi ^c`$ the lowest energy state (in the continuum limit with $`p0`$ ) will acquire macroscopic occupation and a pion condensate is formed. Such a condensate ‘consumes’ energy without consuming entropy of the primordial high entropy QGP phase and we cannot expect this to occur in hadronization of a high entropy phase into a low entropy confined phase. However, a sudden hadronization process will naturally have the tendency to approach the limiting value $`\gamma _\pi \gamma _\pi ^c`$ in order to more efficiently connect the deconfined and the confined phases, since, as we show in Fig. 1, the entropy density is nearly twice as high at $`\gamma _\pi \gamma _\pi ^c`$ than at $`\gamma _\pi =1`$. We note that all thermodynamic quantities of interest here do not develop a singularity in the limit $`\gamma _\pi \gamma _\pi ^c`$.
We thus consider as a possible and indeed likely QGP hadronization mechanism, the conversion of the excess of entropy of the deconfined state into oversaturation of pion phase space. This mechanism can replace the more conventional approaches that also described the ‘absorption’ of the excess QGP entropy:
1) formation of a mixed phase which allows the volume of the fireball of hadronic matter to grow in the hadronization; the excess of QGP in entropy is converted into the greater volume of lower entropy density hadronic gas.
2) reheating which allows the entropy excess to be accommodated in the momentum ‘volume’ of the phase space.
Of course, both these effects can coexist, and also they can, in principle, coexist with our proposed third mechanism, the chemical non-equilibrium. In fact, which exact path is taken is only found in a microscopic description which allows for all these effects with appropriate relative rates. A model study of the dynamics of the hadronization process shows that the reaction mechanism here proposed is possible , though a full understanding of the microscopic hadron phase transformation remains at present beyond our capability. However, aside of the fact that the data analysis heavily favors our proposal, we note that the conventional processes are by far more complex and thus more difficult to realize on the short time scale $`𝒪(10^{22}\text{ s})`$ available in heavy ion collisions, as compared to a rather straight forward gluon fragmentation, accompanied by recombination of quarks into pions, which is the probable microscopic origin of the large chemical nonequilibrium value of $`\gamma _q`$.
Let us return now to the study of pion abundances: the slight chemical asymmetry originating in the larger neutron number than proton number in heavy nuclei, leads to light quark abundance asymmetry, $`N_d>N_u`$, and requires that one of the charged pion fugacities ($`\pi ^{}`$) increases (towards the condensation point), while the other charged pion ($`\pi ^+`$) fugacity is moved further away from it. The result of this is clearly visible in Eq. (7) when combined with Eqs. (5,6): as function of energy this slight change has the effect of enhancing the yield of low energy $`\pi ^{}`$ over that of $`\pi ^+`$. There is a much more significant effect when we are near to $`\gamma _\pi ^c`$ than it is the case for $`\gamma _q=1`$ which was explored earlier . In the following, we will address in detail the statistical parameters as they are determined for the 158–200$`A`$ GeV interactions and presented in table I. We note that for the Pb–Pb system, the value of $`\lambda _s=1.10`$ compensates the Coulomb effect distortion of the phase space of strange quarks in QGP. Thus, it is a measure of the magnitude of the Coulomb potential at hadronization, and hence of the freeze-out volume . We so estimate the volume of the fireball at chemical freeze-out to be $`V2100\text{fm}^3`$, and this value is used to check if the absolute yields of pions, and in particular here negative hadrons ($`\pi ^{},`$ K<sup>-</sup>, $`\overline{p}`$), are consistent with measurement of the experiment NA49 . We find within our approach that among the 200 ‘negatives’ at central rapidity $`\mathrm{\Delta }y=1`$, there are (roughly) about 145 directly produced $`\pi ^{}`$, 20 $`\overline{p}`$ and $`K^{}`$, and only about 30 $`\pi ^{}`$ that arise from resonance decays.
Fig. 2 presents a case study of the effect on the thermal pion $`m_{}`$ spectra. We integrate the longitudinal momenta in the range of rapidity $`0.5<y<0.5`$ and consider the difference in charged pion yield divided by the sum,
$$D_q^\pi (m_{})\frac{N(\pi ^{})N(\pi ^+)}{N(\pi ^{})+N(\pi ^+)},$$
(9)
as function of $`m_{}`$, for small $`m_{}m_\pi <0.3`$ GeV, showing the result on logarithmic scale. The lowest line shows the small effect of isospin asymmetry in Pb–Pb collisions when the pion phase space is governed by the equilibrium chemical condition $`\gamma _q=1`$; this result corresponds, for the range of our statistical parameters here applicable, to the results reported in Ref. . As we move up, for each of the lines in Fig. 2, we increment $`\gamma _q=1`$ by 0.1 and the thick and top line corresponds to the value implied by the analysis of the experimental results, $`\gamma _q=1.6`$. The magnitude of the effect is now quite surprising (remember logarithmic scale), and most importantly, we do see a very clear sensitivity to the value of $`\gamma _q`$. While the ‘modulator’ of this sensitivity is the isospin asymmetry $`\lambda _d/\lambda _u`$, this quantity is in principle measured by the results shown in Fig. 2, since for $`m_{}m_\pi >0.3`$ GeV we have:
$$D_q^\pi (m_{})|_{>0.3\text{GeV}}\frac{(\lambda _d/\lambda _u)^21}{(\lambda _d/\lambda _u)^2+1}.$$
(10)
It must be stressed that the result shown in Fig. 2 must be compared to experiment with some caution, since it is just a case study aiming to show the extraordinary sensitivity of the $`D_q^\pi (m_{})`$ observable to the quantity $`\gamma _q`$. In order to be able to interpret actual data, one has to study the diluting pion component from resonance decay and the effect of collective flow. It can be hoped that the flow effect is small, and indeed cancelling in the pion yield ratio. The contribution of resonance decays turns out also to be rather small in the range of statistical parameters considered in view of the small chemical freeze-out temperature.
Given the magnitude of the $`\gamma _q`$-effect, we also studied in Fig. 3 the spectrum of neutral pions, which is practically indistinguishable from the average of the charged pion spectra, $`(\pi ^++\pi ^{})/2`$. The solid line shown in Fig. 3 are not renormalized, and (on logarithmic scale) the difference between the curves for $`m_{}m_\pi >0.3`$ GeV is the result of the varying yield coefficient $`\gamma _q^2`$. More importantly, for $`m_{}m_\pi <0.2`$ GeV with growing $`\gamma _q`$ we see a strong up-bending of the pion spectrum. Dashed lines in Fig. 3 show results for relatively small $`\mathrm{\Delta }y=0.02`$ and demonstrate that there is just a minimal spectral deformation due to a finite central $`\mathrm{\Delta }y=1`$ integral.
In order to understand the experimental pion spectra it is necessary to consider the $`\gamma _q`$-effect, combined with the flow effect and the hadron disintegration feeding of pion spectra. We recall here that aside of direct influence on primordial pions, $`\gamma _q`$ also impacts the relative yield of primordial mesons compared to the yield of secondary mesons from resonance decays. Thus omission to consider $`\gamma _q`$-effect can mislead the data interpretation. We recall here a recent analysis of the $`\pi ^0`$ spectra which has reached somewhat unorthodox conclusions about the freeze-out conditions as seen in Ref. \[18, Table 1\]. We believe that allowance for chemical non-equilibrium, specifically the introduction of $`\gamma _q\gamma _q^c`$ should alter the conclusions of the WA98 collaboration. Even though they study the pion spectrum at rather large $`m_{}`$, $`\gamma _q`$ impacts the cocktail mix of direct with decay pions. There is another detailed fit which also restricted the single particle pion spectra to be exactly at chemical equilibrium. However, these authors do note that in order to describe the pion yield, they need a pion fugacity $`\gamma _q^{100}=1.36`$, a value which may (barely) allow their tacit assumption $`\gamma _q(T=100\text{MeV})=1`$, made when they interpret particle spectra. At the higher value of chemical freeze-out temperature a range of $`\gamma _q`$ is expected that in our opinion requires reevaluation of their results. On the other hand, the parallel analysis of Schlei and collaborators , which did not attempt a precise fit, and did not describe the low $`m_{}`$ pion spectra, and which obtains a similar chemical freeze-out temperature as used here, appears to us to remain valid.
It would be interesting to see what effect would arise after explicit allowance for a pion fugacity is made in these studies. Since we do not model here the effect of collective flow as was done in Refs. , we cannot ourselves answer this question in full. However, there is data from NA44 on charged pion ratios. Considering:
1) the relatively small dilution of primordial pion spectra by pions from resonance decays, expected for the range of chemical freeze-out parameters shown in table I, and
2) the cancellation of the flow deformation of the spectra in spectral ratios of particles of the same mass,
we have evaluated the expected $`\gamma _q`$-effect for the charged pion ratio results presented by NA44 collaboration , as shown in Fig. 4. We note that we slightly overpredict at small $`m_{}`$ these results, an effect that is required given the inclusion of hadron resonance decay feeding into the experimental data.
Several attempts have been made before to explain these NA44 data. The experimental group presented calculational results obtained with RQMD(v1.08) , which could not account for the data. Even though NA44 considered in their numerical model also hadron resonance decay effect, a scheme has been proposed to create the observed pion asymmetry from strangeness asymmetry : in presence of finite baryon number there are more $`s`$-carrier baryons than antibaryons and more $`\overline{s}`$-carrier mesons than $`s`$-carrier mesons. Of course, such a description can easily be falsified by study of strange hadron abundances. In thermal models, for $`\gamma _q=1`$, the NA44 asymmetry cannot be explained for the widest range of reasonable statistical parameters, a fact we have realized long ago, and which is implicit in other work . These studies have thus lead to the believe that Coulomb field effects in single particle spectra are the source of pion asymmetry and a large number of authors has explored this hypothesis . It is quite possible that in part the effect of the pion asymmetry is due to the Coulomb effect, and in part explained by the $`\gamma _q`$-effect here described, considering that the value $`\gamma _q=1.6`$ obtained from the study of hadron abundances has a significant statistical error.
The Coulomb effect work is based on the picture of charged pions acted upon by the Coulomb field after formation at a classically well determined point. The pion asymmetry is a result of the hypothesis that the pion formation weight $`W^\pm `$ is actually equal for both $`\pi ^+`$ and $`\pi ^{}`$ at equal local momentum:
$$\frac{d^3W^\pm }{d^3p}\frac{1}{e^{\sqrt{p(R_n)^2+m_\pi ^2}/T}1},$$
(11)
where $`R_n`$ is a static or dynamic formation surface. In the WKB approximation:
$$p^{}(x)=\sqrt{(E_\pi V(x))^2m_\pi ^2}.$$
(12)
We believe that considerable improvement in these studies of the Coulomb effect is possible. In our view, the Coulomb field influences the quantum density of states obtainable from the scattering phase $`\eta `$. This introduces in addition to the statistical density matrix weight $`\text{Tr}e^{\beta H}`$ an additional density of states weights $`d\eta ^\pi ^{}/dE,d\eta ^{\pi ^+}/dE`$, which due to the presence of attractive (for $`\pi ^{}`$) and repulsive (for $`\pi ^+`$) Coulomb potential is Coulomb deformed at asymptotic pion momenta values that sample the physical size of the potential. We note that for $`R_n=8`$$`10`$ fm this effect is expected to occur for $`E_\pi m_\pi <\pi ^2/2R_n^2m_\pi 10`$$`20`$ MeV in static case. In quantum formulation the presence of (Coulomb) potential alters locally the reduced wave length $`\lambda ^{}=\mathrm{}/p_\pi (x)`$, but not the asymptotic energy $`E_\pi `$ of a quantum state. The work most closely related to this approach takes Coulomb wave-functions (thus better than the WKB illustration made above) but adds ad hoc density of state hypothesis, see Ref \[23, Eq. (2)\], using here the density of matter in configuration space folded with the Coulomb wave. The pion asymmetry is arising as result of substitution of Eq. (12) in Eq. (11).
Our findings have interesting bearing on the forthcoming experimental RHIC results. The data from the experiment PHOBOS should in our opinion also allow to observe the $`\gamma _q`$-effect, both in soft pion spectra, and in pion asymmetry, provided that the non-equilibrium occupancy, $`\gamma _q1.6`$, which we associated with the QGP hadronization, also occurs at RHIC — as we expect . The analysis of these data should follow the pattern outlined here: from the slight inequality of the abundance at moderate $`m_{}`$, according to Eq. (10), the value of chemical asymmetry can be deduced and this is used to determine the value of $`\gamma _q`$, given the shape of the low-$`m_{}`$ spectral asymmetry following from Eq. (7). We note that Coulomb analysis also predicts a similar effect at RHIC energies .
We have shown that the phase space overoccupancy $`\gamma _q1.6`$ of pions arising in the sudden hadronization process of the deconfined quark-gluon plasma enhances the low-$`m_{}`$ pion spectra, and amplifies the abundance ratio of charged pions. We have quantitatively demonstrated the great sensitivity of these effects to the value of $`\gamma _q>1`$ and have shown how this $`\gamma _q`$-effect helps understand SPS soft pion results of NA44 and WA98 collaborations. We conclude that the reported excess of soft pions and the asymmetry of soft charged pion spectra provides evidence for sudden and explosive hadronization (glue fragmentation) of the entropy rich deconfined phase. This finding about the mechanism of hadronization corroborates the conclusions we obtained studying high $`m_{}`$ strange hadron spectra.
Acknowledgments:
Supported in part by a grant from the U.S. Department of Energy, DE-FG03-95ER40937 . LPTHE, Univ. Paris 6 et 7 is: Unité mixte de Recherche du CNRS, UMR7589. |
no-problem/9911/hep-lat9911025.html | ar5iv | text | # 𝐼=2 Pion Scattering Length with Wilson Fermions presented by N. Ishizuka
## 1 Introduction
Lattice calculations of scattering lengths of the two-pion system is an important step for understanding of strong interactions beyond the hadron mass spectrum. For the $`I=0`$ process, which is difficult due to the presence of disconnected contributions, only one group carried out the calculation . For the $`I=2`$ process, on the other hand, a number of calculations has been carried out with the Staggered and the Wilson fermions . These calculations reported results in agreement with the prediction of current algebra or lowest-order chiral perturbation theory(CHPT) . It is known, however, that this prediction differs from the experimental value over $`1.4\sigma `$, and that the higher order effects of CHPT are small .
The past lattice calculations were made on coarse lattices with small sizes, and the continuum extrapolation was not taken. In this article we report on our high statistics calculation of the $`I=2`$ pion scattering length aiming to improve on these points. This work is carried out in quenched lattice QCD employing the standard plaquette action for gluons with the Wilson fermions. The number of configurations (and lattice size) are $`187(16^3\times 64)`$, $`120(24^3\times 64)`$, and $`100(32^3\times 80)`$ for $`\beta =5.9`$, $`6.1`$, and $`6.3`$. The pion mass covers the range $`450900\mathrm{M}\mathrm{e}\mathrm{V}`$.
## 2 Method
The energy eigenvalue of a two-pion system in a finite periodic box $`L^3`$ is shifted by finite-size effect. Lüscher presented a relation between the energy shift $`\mathrm{\Delta }E`$ and the $`S`$-wave scattering length $`a_0`$ given by
$$\mathrm{\Delta }E\frac{m_\pi L^2}{4\pi ^2}=T+AT^2+BT^3+O(T^4),$$
(1)
where $`T=a_0/(\pi L)`$. Since $`T`$ takes a small value, typically $`10^2`$, in our simulation we can neglect the higher order terms $`O(T^4)`$. The constants $`A`$ and $`B`$ are geometrical values $`A=8.9136\mathrm{}`$ and $`B=62.9205\mathrm{}`$.
The energy shift $`\mathrm{\Delta }E`$ can be obtained from the ratio $`R(t)=G(t)/D(t)`$, where
$`G(t)=\pi ^+(t)\pi ^+(t)W^{}(t_1)W^{}(t_2)`$ (2)
$`D(t)=\pi ^+(t)W^{}(t_1)\pi ^+(t)W^{}(t_2).`$ (3)
In order to enhance signals we use wall sources (denoted by $`W^{}`$) and fix gauge configurations to the Coulomb gauge. The two wall sources are placed at different time slices $`t_1`$ and $`t_2`$ to avoid contaminations from Fierz-rearranged terms in the two-pion state that would occur for the choice $`t_1=t_2`$. In this work we set $`t_2=t_1+1`$ and $`t_1=8`$, $`10`$, $`13`$ for $`\beta =5.9`$, $`6.1`$, $`6.3`$. Quark propagators are solved with the Dirichlet boundary condition imposed in the time direction and the periodic boundary condition in the space directions. In region $`t>>t_1,t_2`$ the ratio behaves as $`R(t)Z(1\mathrm{\Delta }E(tt_1)+O(t^2))`$.
As an example, the ratio $`R(t)`$ at $`\beta =6.3`$ and $`\kappa =0.1513`$ corresponding to $`m_\pi =433(4)\mathrm{MeV}`$ is plotted in Fig. 1. The signal is very clear and $`Z1`$. This means the overlap of our wall sources with the two-pion state is sufficiently large. In general there are higher order terms $`O(t^2)`$ in $`R(t)`$, but we cannot resolve them in Fig. 1. Making a linear fitting in the range $`t=2762`$, we obtain $`(a\mathrm{\Delta }E)=5.97(60)\times 10^3`$. Solving equation (1) with this value we obtain $`T=1.73(15)\times 10^2`$, which corresponds to $`a_0=0.525(45)(1/\mathrm{GeV})`$.
## 3 Result
In Fig. 2 we compare our new results for $`a_0/a_0^{\mathrm{CHPT}}`$ with those of old calculations, where $`a_0^{\mathrm{CHPT}}`$ is the prediction of current algebra : $`a_0^{\mathrm{CHPT}}=m_\pi /(16\pi f_\pi ^2)`$. For the decay constant $`f_\pi `$ we use the value at finite $`m_\pi `$ at finite lattice spacing referred in each study. This ratio has been commonly employed to make a comparison of current algebra and lattice calculations with different quark actions and parameters. The open symbols refer to results with the Staggered fermions and filled ones are those of the Wilson fermions. The legends give $`\beta `$, the spatial lattice size $`L`$, and collaborations of the studies ( K : Kuramashi et.al. , SGK : Sharpe et.al. , GPS : Gupta et.al. , Ours : our calculation ). We also plot the experimental value at $`m_\pi =140\mathrm{M}\mathrm{e}\mathrm{V}`$. We find that our results are inconsistent with old results, especially with those of the Staggered fermions.
A possible cause of the discrepancy is the systematic error of determination of $`f_\pi `$ needed to calculate $`a_0^{\mathrm{CHPT}}`$. In Fig. 3 we compare our results with old calculations in terms of $`a_0/m_\pi `$. The same symbols as those in Fig. 2 are used. The lattice results including ours are almost consistent with each other. Also they appear to be in more agreement with the experiment than with the prediction of current algebra.
We note that the calculation of $`f_\pi `$, being determined by the amplitude of correlation function of pion and axial vector current, is quite difficult. Various systematic errors may well enter in their determinations. Further the mass dependence of $`f_\pi `$ is not small and $`a_0/a_0^{\mathrm{CHPT}}`$ is very sensitive to it. For these reasons we analyze $`a_0/m_\pi `$ below.
From chiral symmetry $`a_0/m_\pi `$ behaves as
$$a_0/m_\pi =A+Bm_\pi ^2+Cm_\pi ^2\mathrm{log}(m_\pi ^2/\mathrm{\Lambda }^2)+O(m_\pi ^4).$$
(4)
For the Wilson fermions we should consider another term $`1/m_\pi ^2`$ that arises from breaking of chiral symmetry. Golterman and Bernard also proposed the same term pointing out that it can appear from quenching effect . However, these effects are very small in our simulation as we do not observe a rapid variation of $`a_0/m_\pi `$ expected from such a term in Fig. 3. Further the chiral logarithm term, $`m_\pi ^2\mathrm{log}(m_\pi ^2/\mathrm{\Lambda }^2)`$, is also small.
In Fig. 4 our results for $`a_0/m_\pi `$ in the chiral limit obtained by a linear fitting in $`m_\pi ^2`$ are plotted, together with the experimental value and the prediction of current algebra at $`m_\pi =140\mathrm{M}\mathrm{e}\mathrm{V}`$. We observe a very clear linear dependence in the lattice spacing $`a`$. By a linear extrapolation, we then obtain
$`a_0/m_\pi =1.91(25)(1/\mathrm{GeV}^2)`$ (5)
$`a_0m_\pi =0.0374(49),`$ (6)
in the continuum limit.
This result is consistent with the experimental value: $`a_0/m_\pi =1.43(61)(1/\mathrm{GeV}^2)`$ ($`a_0m_\pi =0.028(12)`$). The difference from the prediction of the current algebra given by $`a_0/m_\pi =2.3(1/\mathrm{GeV}^2)`$ ($`a_0m_\pi =0.045`$) is about $`1.5\sigma `$. Since scaling violation is not small, and our data points are far from the continuum limit, as seen in Fig. 4, further calculations nearer to the continuum limit is desirable. In addition studies with the Staggered fermions should be repeated in a systematic manner for comparison with present results.
This work is supported by the Supercomputer Project No.45 (FY1999) of High Energy Accelerator Research Organization (KEK), and also in part by the Grants-in-Aid of the Ministry of Education ( Nos. 09304029, 10640246, 10640248, 10740107, 10740125, 11640294, 11740162 ). K-I.I is supported by the JSPS Research Fellowship. |
no-problem/9911/cond-mat9911344.html | ar5iv | text | # Spinodal decomposition of off-critical quenches with a viscous phase using dissipative particle dynamics in two and three spatial dimensions
## I Introduction
Over the last few years, the dissipative particle dynamics (DPD) model of complex fluids has received considerable attention. It has matured from its somewhat arbitrary initial formulation into a model with a solid theoretical basis. Furthermore, it has been applied with considerable success to a large number of computer simulations of complex fluid systems such as colloidal suspensions, polymeric fluids, spinodal decomposition of binary immiscible fluids, and amphiphilic fluids. DPD also looks promising for simulating multiphase flows and flow in porous media, and is now considered a useful technique alongside the other complex fluid algorithms: molecular dynamics, lattice-gas automata, and techniques based on the lattice-Boltzmann equation. No single technique can yet be applied to all situations, and each has different strengths and weaknesses. While molecular dynamics is in principle the most accurate microscopic approach, in practice it is too slow in both its quantum (Car–Parinello) and classical forms because of its excessive detail. Discrete mesoscopic methods developed from lattice-gas automata have had some success, but they too have problems, such as lacking Galilean invariance. The traditional approach of continuum fluid dynamics has met with limited success in modeling behavior on the length and time scales characteristic of complex fluids.
In this paper we investigate phase separation from both symmetric (critical) and off-critical quenches in binary immiscible fluids of generally differing viscosity using the hydrodynamically-correct DPD model. Our motivation is to probe and extend knowledge of the behavior of differing-viscosity fluids, of application, for example, to the action of detergents and the extraction of oil from reservoir rocks. Phase separation in equal-viscosity binary immiscible fluid systems has been simulated using a variety of techniques, including DPD ; molecular dynamics ; Monte Carlo ; cell dynamical systems without hydrodynamics and with Oseen tensor hydrodynamics ; time-dependent Ginzburg–Landau models with and without hydrodynamics ; lattice-gas automata ; and lattice-Boltzmann techniques . Spinodal decomposition of differing-viscosity immiscible fluids has previously been simulated in two dimensions by a time-dependent Ginzburg–Landau model without hydrodynamics . We discuss the effect of the proportion of each phase on the scaling behavior in both two and three spatial dimensions. We also examine the behavior of equal-viscosity fluids, comparing our results to similar lattice-gas simulations in two dimensions .
After describing our fluid model in the following section, we discuss the expected temporal development of the characteristic size of the separating domains in Section III. We then describe our method for calculating the characteristic domain size and its rate of growth in Section IV. This is followed in Sections V and VI by information on the simulations performed and a discussion of the results, and by some conclusions in Section VII. Finally, as an Appendix to this paper, we make a few comments on the high-performance computing aspects of this work.
## II The Fluid Model
In 1992, Hoogerbrugge and Koelman proposed dissipative particle dynamics as a novel particulate model for the simulation of complex fluid behavior. DPD was developed in an attempt to capture the best aspects of molecular dynamics and lattice-gas automata. It avoids the lattice-based problems of lattice-gas automata, yet maintains an elegant simplicity and larger scale that keeps the model much faster than molecular dynamics. This simplicity also makes DPD highly extensible, such as for including the interactions of complex molecules or modeling flow in an arbitrary number of spatial dimensions. The key features of the model are that the fluid is grouped into packets, termed “particles”, and that mass and momentum are conserved but energy is not. Particles are normally interpreted as representing a coarse-graining of the fluid, so that each particle contains many molecules . Since the intrinsic time scale of DPD represents the correlated motion of mesoscopic packets of atoms or molecules, it is typically orders of magnitude larger than the time scale of molecular dynamics . Particle positions and momenta are real variables, and are not restricted to a grid.
Español and Warren’s analysis showed that the original DPD model does not satisfy detailed balance, so the equilibrium states (if they exist) cannot be simply characterized. Detailed balance is the condition equating the rates of forward and backward transition probabilities in a dynamical system, and is a sufficient (but not necessary) condition guaranteeing that the system has a (Gibbsian) equilibrium state . Español and Warren formulated a Fokker–Planck equation and equivalent set of stochastic differential equations which lead to an analogous continuous-time model,
$$\{\begin{array}{cc}d𝐩_i=\underset{ji}{}𝐅_{ij}dt=\underset{ji}{}\left[𝐅_{ij}^Cdt+𝐅_{ij}^Ddt+𝐅_{ij}^RdW_{ij}\right]\hfill & \\ d𝐱_i=\frac{𝐩_i}{m_i}dt.\hfill & \end{array}$$
(1)
In these equations, $`𝐩_i`$, $`𝐱_i`$, and $`m_i`$ denote the momentum, position, and mass of particle $`i`$. $`𝐅_{ij}^C`$ is a conservative force acting between particles $`i`$ and $`j`$, while $`𝐅_{ij}^D`$ and $`𝐅_{ij}^R`$ are the dissipative and random forces. $`dW_{ij}=dW_{ji}`$ are independent increments of a Wiener process. By Itô calculus
$$dW_{ij}dW_{kl}=\left(\delta _{ik}\delta _{jl}+\delta _{il}\delta _{jk}\right)dt,$$
(2)
so $`dW_{ij}`$ is an infinitesimal of order $`\frac{1}{2}`$ and $`dW_{ij}`$ can be written $`\theta _{ij}\sqrt{dt}`$, where $`\theta _{ij}=\theta _{ji}`$ is a random variable with zero mean and unit variance . Detailed balance is satisfied by this continuous-time version of DPD with an appropriate choice for the form of the forces , and so equilibrium states are guaranteed to exist and be Gibbsian. To ensure that the associated fluctuation–dissipation theorem holds, Español and Warren suggested the forces assume the following forms :
$$𝐅_{ij}^C=\alpha \omega _{ij}\widehat{𝐞}_{ij},$$
(3)
$$𝐅_{ij}^D=\gamma \omega _{ij}^2\left(\widehat{𝐞}_{ij}𝐯_{ij}\right)\widehat{𝐞}_{ij},$$
(4)
and
$$𝐅_{ij}^R=\sigma \omega _{ij}\widehat{𝐞}_{ij},$$
(5)
where $`𝐯_{ij}=(𝐩_i/m_i)(𝐩_j/m_j)`$ is the difference in velocities of particles $`j`$ and $`i`$, $`\widehat{𝐞}_{ij}`$ is the unit vector pointing from particle $`j`$ to particle $`i`$, and $`\omega _{ij}`$ is a weighting function depending only on the distance separating particles $`i`$ and $`j`$. The constants $`\alpha `$, $`\gamma `$, and $`\sigma `$ are chosen to reflect the relative importance of the conservative, dissipative (viscous), and random components in the fluid of interest. As a consequence of detailed balance and the fluctuation–dissipation theorem, $`\gamma `$ and $`\sigma `$ are related to Boltzmann’s constant $`k_B`$ and the equilibrium temperature $`T`$ by
$$\frac{\sigma ^2}{\gamma }=2k_BT.$$
(6)
In order to remain as close as possible to the original DPD model, Español and Warren chose the friction weight function to be
$$\omega _{ij}=1\frac{r_{ij}}{r_c}$$
(7)
within the constant cutoff length $`r_c>0`$ and zero otherwise, where $`r_{ij}`$ is the distance between particles $`i`$ and $`j`$. Adding Eqs. (3)–(5) together, the total force is
$$𝐅_{ij}=\left[\alpha \gamma \omega _{ij}\left(\widehat{𝐞}_{ij}𝐯_{ij}\right)+\frac{\sigma \theta _{ij}}{\sqrt{dt}}\right]\omega _{ij}\widehat{𝐞}_{ij}.$$
(8)
Immiscible fluid mixtures exist because individual molecules attract similar and repel dissimilar molecules. The most common example of such behavior arises in mixtures of oil and water below a critical temperature. The nonpolar, hydrophobic molecules of oil attract one another through short-range van der Waals forces, while the polar water molecules have complex, long-range hydrophilic attractions which are dominated by electrostatic interactions, including hydrogen bonds. At the atomistic level employed in molecular dynamics, such interactions demand a detailed treatment. However, to obtain mesoscopic and macroscopic level descriptions using DPD, the microscopic model can be drastically simplified.
In order to model immiscible fluids, the simplest modification to the one-component dissipative particle dynamics algorithm is to introduce a new variable, called the “color” by analogy with Rothman–Keller . For example, we could choose red to represent oil and blue to represent water. When two particles of different color interact we increase the conservative force, thereby increasing the repulsion. That is,
$$\alpha \alpha _{ij}=\{\begin{array}{cc}\alpha _0\hfill & \text{if particles }i\text{ and }j\text{ are the same color}\hfill \\ \alpha _1\hfill & \text{if particles }i\text{ and }j\text{ are different colors}\hfill \end{array}$$
(9)
where $`\alpha _0`$ and $`\alpha _1`$ are constants with $`0\alpha _0<\alpha _1`$. As a consequence of mass and momentum conservation, the Navier–Stokes equation is obeyed within a single-phase DPD fluid and within regions of homogeneity of each of the two binary immiscible fluid phases . Likewise, detailed balance is also preserved, at least in the limit of continuous time .
The immiscible fluids described above are identical to each other. However, it is often the case that the two fluids in a mixture will differ physically. For example, oil and water typically have different viscosities. To model binary fluids with differing viscosity we again adopt the simplest approach: labeling one of the two phases (colors) as more viscous. When two particles of the same viscous color interact, the dissipative (viscous) force is increased; in order to keep the temperature constant, we must correspondingly decrease the random force according to Eq. (6), i.e.,
$$\gamma \gamma _{ij}=\{\begin{array}{cc}\gamma _0\hfill & \text{if either particle }i\text{ or }j\text{ is not the viscous color}\hfill \\ \gamma _1\hfill & \text{if both particles }i\text{ and }j\text{ are the viscous color.}\hfill \end{array}$$
(10)
$$\sigma \sigma _{ij}=\{\begin{array}{cc}\sigma _0\hfill & \text{if either particle }i\text{ or }j\text{ is not the viscous color}\hfill \\ \sigma _1\hfill & \text{if both particles }i\text{ and }j\text{ are the viscous color,}\hfill \end{array}$$
(11)
where
$$\frac{\sigma _0^2}{\gamma _0}=\frac{\sigma _1^2}{\gamma _1}=2k_BT.$$
(12)
Our previous study of finite difference methods applicable to simulations with non-conservative forces indicated that the finite-difference algorithm suggested by Groot and Warren is a good choice for DPD. Their method is a variation on the familiar velocity-Verlet algorithm, adding a momentum estimate before the force evaluation:
$`𝐱_i(t+\mathrm{\Delta }t)`$ $`=`$ $`𝐱_i(t)+{\displaystyle \frac{\mathrm{\Delta }t}{m_i}}\left[𝐩_i(t)+{\displaystyle \frac{\mathrm{\Delta }t}{2}}𝐟_i(t)\right]`$ (13)
$`𝐩_i(t+\mathrm{\Delta }t)`$ $`=`$ $`𝐩_i(t)+{\displaystyle \frac{\mathrm{\Delta }t}{2}}𝐟_i(t)`$ (14)
$`𝐟_i(t+\mathrm{\Delta }t)`$ $`=`$ $`{\displaystyle \underset{ji}{}}𝐅_{ij}(t+\mathrm{\Delta }t)`$ (15)
$`𝐩_i(t+\mathrm{\Delta }t)`$ $`=`$ $`𝐩_i(t)+{\displaystyle \frac{\mathrm{\Delta }t}{2}}\left[𝐟_i(t)+𝐟_i(t+\mathrm{\Delta }t)\right],`$ (16)
where $`\mathrm{\Delta }t`$ is the time step size and $`𝐟_i(t)`$ is the force acting on particle $`i`$ at time $`t`$. The DPD units of length, mass, and time are specific to the particular set of model parameters, and the exact relationship between these parameters and a real fluid in a particular situation can be determined by considering the dimensionless groups relevant to the motion of that system, such as the Mach, Reynolds, and Weber numbers.
Many further modifications to the model have been suggested and implemented by others. Some of the most interesting include energy conservation, colloidal particles, and polymers. In very recent work, it has been shown that it is possible to derive a modified version of DPD directly from the underlying molecular dynamics .
## III Temporal Behavior of the Characteristic Domain Size
A central quantity in the study of growth kinetics is the characteristic domain size $`R(t)`$. For binary systems in the regime of sharp domain walls, this is usually thought to follow algebraic growth laws of the form
$$R(t)t^\beta .$$
(17)
For symmetric quenches without hydrodynamic interactions, dynamical scaling theory and experiment indicate that the scaling exponent $`\beta =\frac{1}{3}`$. If flow effects are relevant, typically
$$\beta =\{\begin{array}{cc}\frac{1}{2}\hfill & \text{for }RR_h\text{ (diffusive)}\hfill \\ \frac{2}{3}\hfill & \text{for }RR_h\text{ (inertial hydrodynamic)}\hfill \end{array}$$
(18)
in two dimensions, and
$$\beta =\{\begin{array}{cc}\frac{1}{3}\hfill & \text{for }RR_d\text{ (diffusive)}\hfill \\ 1\hfill & \text{for }R_dRR_h\text{ (viscous hydrodynamic)}\hfill \\ \frac{2}{3}\hfill & \text{for }RR_h\text{ (inertial hydrodynamic),}\hfill \end{array}$$
(19)
in three dimensions, where $`R_h=\eta ^2/(\rho \kappa )`$ is the hydrodynamic length and $`R_d=\sqrt{\eta D}`$ is the diffusive length, expressed in terms of the absolute (dynamic) viscosity $`\eta `$, density $`\rho `$, surface tension coefficient $`\kappa `$, and diffusion coefficient $`D`$. These scaling laws follow directly from dimensional analysis of the macroscopic fluid dynamics equations (so-called model H, or Cahn–Hilliard coupled to Navier–Stokes hydrodynamics) in the appropriate regimes . When $`RR_d`$ in three dimensions, diffusive effects dominate and we expect the domains to form via the Lifshitz–Slyozov–Wagner (LSW) evaporation–condensation mechanism . When $`RR_h`$ in both two and three dimensions, we expect the growth mechanism to be surface tension driven by hydrodynamic flow, balanced by inertial effects . If $`R_dRR_h`$ in three dimensions, we expect viscous hydrodynamic effects to dominate, as predicted by Siggia ; in this regime the surface tension drives the transport of the fluid along the interface, which is possible only if both phases are continuous . This third growth regime cannot occur in two dimensions, and so we expect to see diffusive growth for $`RR_h`$ . Due to our choice of model parameters and the small size of our simulations, we do not expect to probe the viscous or inertial hydrodynamic regimes.
Simulations in two spatial dimensions are especially useful to emphasize the importance of correct hydrodynamics: simulations which do not conserve momentum typically give $`\beta =\frac{1}{3}`$ for the diffusive regime ($`RR_h`$. Simulation methods with correct hydrodynamic interactions typically give the expected result of $`\beta =\frac{1}{2}`$ (see e.g., Refs. ). It is worth noting that momentum conservation is not thought necessary to model spinodal decomposition in binary metallic alloys, since phase separation occurs by the migration of atoms to neighboring vacancies on the crystalline lattice . Simulations based on lattice-Boltzmann techniques also typically display $`\beta =\frac{1}{3}`$. These observations are supported by a renormalization group approach which has shown that thermal fluctuations cause Brownian motion-driven coalescence and play a crucial role in causing $`\beta `$ to assume the value of $`\frac{1}{2}`$ ; although some lattice-Boltzmann techniques include these fluctuations , most do not.
There is some confusion in the literature over which scaling exponents should be observed for off-critical quenches, and whether or not the algebraic scaling law \[Eq. (17)\] should in fact be obeyed for any off-critical binary immiscible fluid. Several authors have reported the coexistence of multiple length scales . It is likewise uncertain as to what behavior should be observed from binary immiscible fluids of differing viscosity . Certainly, one growth mechanism we expect for off-critical quenches of both equal and differing viscosity is the LSW evaporation–condensation mechanism, for which $`\beta =\frac{1}{3}`$ .
It should also be noted that there are still some experimental and theoretical challenges in unraveling the behavior of systems in which both the order parameter and the momentum are locally conserved . Experimentally, for example, it is difficult if not impossible to study two-dimensional fluid systems. As far as numerical studies are concerned, it is important to recognize that three-dimensional simulations are particularly demanding on all the aforementioned techniques, and so definitive results are harder to come by than in two dimensions.
Indeed, Cates’s group in Edinburgh has recently reported somewhat conflicting results from three-dimensional studies of binary immiscible fluid separation for equal viscosity fluids using dissipative particle dynamics and lattice-Boltzmann methods . While the former suggests the persistence of non-universal length scales, hypothesized to be due to the intrusion of a “molecular” or discretization length scale, this is not supported by the latter, from which finite-size effects were more rigorously excluded. Note, however, that whereas the lattice-Boltzmann scheme was based on a Landau free-energy approach, and is essentially no more than a finite-difference solution of the continuum model H equations, the macroscopic equations to which the spinodal DPD scheme corresponds have not yet been derived. This makes it unclear whether the two systems being simulated are really one and the same. Both of these studies emphasize the importance of observing dynamical scaling over decades of time before drawing conclusions as to the true nature of the scaling regime. One notable result of their work is the clarification of the relative time scales over which computer simulation techniques typically operate. For the simulations they discuss, molecular dynamics time scales are on the order of $`10^2`$ in dimensionless units, lattice-gas automata and Langevin dynamics probe time scales on the order of $`10^2`$ through $`10^4`$, DPD typically $`10^3`$ to $`10^5`$, and methods based on the lattice-Boltzmann equation $`10^1`$ through $`10^8`$.
Grant and Elder have recently argued that $`\beta \frac{1}{2}`$ in any asymptotic scaling regime because the Reynolds number $`Re=\rho RV/\eta `$ (where $`V`$ is a characteristic velocity) cannot diverge, and in fact must remain less than a critical value $`Re_{\mathrm{cr}}`$ to avoid the onset of turbulence and possible turbulent remixing of the fluids. The conclusion they draw is that the $`\beta =\frac{2}{3}`$ scaling regime must be transient . However, their analysis neglects mention of the relative strength of the interface, quantified for example by the Weber number $`N_{\mathrm{We}}=\rho RV^2/\kappa `$. If $`N_{\mathrm{We}}`$ is small at the onset of turbulence we would expect turbulent remixing to be delayed, or perhaps even postponed indefinitely allowing $`Re`$ to diverge. In any case, the separation dynamics would likely be affected by $`ReRe_{\mathrm{cr}}`$; for example, the fluid viscosity in a turbulent region is roughly proportional to $`Re`$ .
## IV Estimating the Characteristic Domain Size
In order to characterize the phase separation kinetics within a binary immiscible fluid, we need a practical tool to measure the characteristic domain size corresponding to the state of the system at a given point in time. The use of the static structure function for this purpose is widespread. However, we have noticed bizarre “early time” behavior when using the static structure function to characterize simulations of highly off-critical quenches. Specifically, the characteristic domain size would sometimes change suddenly by an order of magnitude. This abrupt change in behavior did not correspond to anything observable in the time evolution of the positions of the DPD particles. Such anomalous behavior is likely due to the large fluctuations prevalent in the static structure function for small simulations of highly-mixed binary fluids, for which the characteristic domain size is very small.
The radial distribution function $`g(r)`$ is a well-established tool for the analysis of single-phase fluids , and indicates the likelihood of finding two particles separated by a distance $`r`$. For binary fluids we can also use the same-phase and differing-phase distribution functions . The same-phase distribution function $`g_{00}(r)`$ describes the likelihood of finding two particles of the same phase separated by a distance $`r`$, and the differing-phase distribution function $`g_{01}(r)`$ describes the likelihood of finding two particles of differing phase separated by a distance $`r`$. From the peaks of the differing-phase radial distribution function, we can estimate the characteristic separation of particles of differing phase (i.e., the characteristic domain size). Consequently, we decided to calculate the distribution between particles of different phase (color),
$$g_{01}(r)=\frac{mn(r,\mathrm{\Delta }r)}{\rho N\varphi [1\varphi ]V(r,\mathrm{\Delta }r)},$$
(20)
where $`m=m_ii`$ is the mass of each particle (assumed all identical), $`n(r,\mathrm{\Delta }r)`$ is the number of particle pairs of differing phase with separation between $`r\mathrm{\Delta }r/2`$ and $`r+\mathrm{\Delta }r/2`$, $`N`$ is the total number of particles, $`\varphi (0,1)`$ is the fraction of particles of one phase (the more viscous phase if the two phases are of different viscosity), and $`V(r,\mathrm{\Delta }r)`$ is the volume of the spherical shell of radius $`r`$ and thickness $`\mathrm{\Delta }r`$ (from $`r\mathrm{\Delta }r/2`$ to $`r+\mathrm{\Delta }r/2`$). It is apparent from Eq. (20) that $`g_{01}(r)`$ can only be calculated for $`rL/2`$ in a simulation with a periodic box size of $`L`$; this should not be interpreted as an undesirable limitation as finite size effects are normally significant for $`R>L/2`$ . A value of $`\mathrm{\Delta }r=L/2000`$ gives a reasonable compromise between noise and resolution for the size of simulation we discuss in this paper.
The principal difficulty lies in analyzing the results automatically, as we need to calculate $`g_{01}(r)`$ at many time steps within each simulation in order to display the domain growth over time. We chose to calculate the median of that portion of the smoothed $`g_{01}(r)`$ curve extending above a threshold value $`1+3s`$, where $`s`$ is the standard deviation of the smoothed curve when it has effectively a constant unit value, estimated from the last tenth of the smoothed $`g_{01}(r)`$ curve at $`t=1`$. We used a fourth order Savitzky–Golay smoothing filter to smooth $`g_{01}(r)`$ over $`r`$ at each point in time with a symmetric window size chosen to reduce the noise while leaving significant features untouched (41 points in two dimensions and 101 points in three). We chose the cutoff threshold of $`1+3s`$ so as not to bias the median by the size of the periodic simulation box; likewise, we chose the median in preference to the mean (first moment) because the median is less biased by outliers, such as those 2.5% of noisy points which effectively have unit value but extend above the threshold. Both the median and mean give poor estimates of the characteristic domain size as it approaches $`L/2`$, the limit of calculable $`g_{01}(r)`$. When this situation is detected, we use the global maximum of $`g_{01}(r)`$ to estimate the domain size instead; this is a continuous transition for symmetrical peaks.
Once we have calculated the characteristic domain size for a series of $`g_{01}(r)`$ curves taken at different times from a single simulation, we begin a search for linear sections in the plot of $`\mathrm{log}_{10}R`$ vs. $`\mathrm{log}_{10}t`$. This has also been automated, using analysis of variance to decide whether a given section of the plot is linear or cubic. We used the analytic expressions for the least-squares fits with moments taken about the means to minimize the effect of roundoff error. We keep only the longest linear sections, subject to there not being any significant gaps in the coverage of the log–log plot. An ensemble average of a number of simulations yields a plot of the scaling exponent $`\beta `$ vs. $`\mathrm{log}_{10}t`$, in which long horizontal (zero gradient) sections represent algebraic growth. This procedure allows more accurate determination of the scaling exponent than a visual comparison of $`\mathrm{log}_{10}R`$ vs. $`\mathrm{log}_{10}t`$ to a straight line of a particular slope, and provides a statistically valid determination of whether or not the growth is truly algebraic over a particular time period. Finite size effects for $`RL/2`$ normally result in non-algebraic growth or unusually low scaling exponents; either case is easily detected by our method. A further advantage is the automation we have described, permitting large ensemble averages with minimal effort.
We should note that estimating the characteristic domain size using the median can occasionally lead to discontinuities. If these discontinuities are large enough they will cause a break between linear sections. Small discontinuities may be spanned by a single line, which would affect the slope ($`\beta `$) and be detrimental to the results. However, large discontinuities will not be spanned, and so will not affect the slope.
The results we obtained using these techniques are considerably better than those obtained with the static structure function, especially for highly-mixed fluids where the pair distribution function is not drowned by fluctuations to the extent that the static structure function is. Furthermore, the pair distribution function is more intuitive to analyze than the static structure function, which aids our interpretation of the results.
## V Simulation Results
For the simulations of fluids with differing viscosity, we chose one phase to be an order of magnitude more viscous than the other, i.e., $`\gamma _1=10\gamma _0`$. Before simulating this new complex fluid, it is advisable to verify that increasing the parameter $`\gamma `$, while keeping the temperature constant, does indeed increase the viscosity. For the lower-viscosity fluid we chose the model parameters shown in Table I. The absolute (dynamic) viscosity of a homogeneous DPD fluid can be estimated theoretically from the continuous-time viscosity, ignoring the effect of the conservative forces (i.e., $`\alpha `$=0) ; the continuous-time viscosity of this lower-viscosity fluid is $`\eta =2.8`$.
In order to verify this estimate, we performed a series of simulations of steady shear using Lees–Edwards periodic boundary conditions. We performed a total of 63 simulations with a time step of $`\mathrm{\Delta }t=0.1`$ (in our DPD time units), each from a different random initial configuration and each allowed to settle to steady-state shear before measurement began. We studied systems of both 1600 and 6400 particles, six simulations at each of nine different shear rates for the former and three simulations at each of three distinct shear rates for the latter. As the results from the larger simulations gave a mean viscosity nearly identical to that of the smaller ones, we can conclude that finite size effects do not bias the viscosity of the smaller, faster simulations. We calculated the velocity profile for each set of parameters, and found it to be statistically indistinguishable from linear in every case.
Analyzing these simulations led to a conclusion of $`\eta =1.94\pm 0.01`$. Others have also found discrepancies between theory and simulation, particularly regarding the kinematic contribution to viscosity , so the difference between the simulated viscosity and the continuous-time viscosity is not entirely surprising. Since molecular dynamics simulations containing only conservative forces give a finite viscosity, we would be surprised if the theoretical estimate did not differ from our calculations.
In order to measure the viscosity of the more viscous fluid, we set up a series of 1600 particle simulations of a homogeneous DPD fluid as described above, using a total of 30 simulations. We varied the shear rate in the range that gave the best results in the previous simulations. The model parameters differed in that $`\gamma `$ was a factor of ten larger while $`\sigma `$ was a factor of $`\sqrt{10}`$ larger to keep the temperature constant \[see Eq. (12)\]; these parameters are shown in Table II. We decreased the time step to $`\mathrm{\Delta }t=0.01`$ due to the increased magnitude of the dissipative and stochastic forces. Analyzing these results led to the conclusion that $`\eta =17.2\pm 0.3`$, which confirms the increase in viscosity. This increase is close to a factor of ten, so that for similar model parameters it is reasonable to conclude that $`\gamma `$ is approximately proportional to viscosity.
It is possible to measure the surface tension of a binary immiscible fluid by integrating the pressure tensor across a flat interface, or by verifying Laplace’s law for a series of equilibrium bubbles of varying radii . Laplace’s law was verified for a DPD binary immiscible fluid in our previous simulations , and so surface tension measurements were omitted from the present study. These calculations would, however, allow identification of the unit of time for comparison with other simulation techniques (see e.g., Refs. ). Theoretical estimates for surface tension are also available, but are of similar accuracy to the viscosity estimate above.
For the simulation of fluids of differing viscosity, we must also choose the relatively small time step size of $`\mathrm{\Delta }t=0.01`$ in order to ensure the stability of the algorithm as a result of the increased magnitude of the dissipative forces. This has the consequence of making these simulations computationally much more expensive than equal-viscosity simulations. In this context, it is worth commenting on the virtues of using DPD to perform simulations of differing-viscosity fluids compared with other models. Sappelt and Jäckle use an approach based on the Cahn–Hilliard equation (so-called model B without noise) with a concentration-dependant mobility so do not include hydrodynamics, unlike our approach with DPD. The other mesoscale techniques which could be used to model these fluids include lattice-gas automata and methods based on the lattice-Boltzmann equation. As with DPD, there is a high computational price to the lattice-gas approach, which requires adjustment to the collisional outcomes of the look-up tables. Lattice-Boltzmann (more correctly, lattice-BGK) models can be used, based for example on the Swift–Osborn–Yeomans free energy functional approach , but there is a similar computational price, although with the additional complication of poorly understood numerical instabilities. However, DPD offers the simplest algorithmic implementation which is thermodynamically consistent.
Our differing-viscosity simulations used the model parameters shown in Table III. Each simulation had 6400 particles, and ran for 50,000 time steps. We chose to use only 6400 particles in our simulations so that individual simulations would be quick enough to be run multiple times, allowing us to consider the effect of changing the phase fraction and viscosity, and to increase the confidence in our results and calculate accurate error estimates with ensemble averaging. Simulations with more particles would have given better resolution of small scale features (relative to the system size) and postponed the finite size effects, at the price of increased computational demands. Our resources would only have allowed a few simulations of the size and length used by Jury et al. (1,000,000 particles), making studies of the sort we describe in this paper impossible.
We calculated the pair distribution function at 64 logarithmically-spaced points in time during the course of each simulation, starting from time $`t=1`$ and finishing at $`t=500`$ (in our DPD time units). In both two and three dimensions, we performed ten simulations at each of nine different fractions of the viscous phase, ranging from $`\varphi =0.1`$ (10% viscous phase) to $`\varphi =0.9`$ (90% viscous phase). We show the time evolution of a single simulation at each value of the viscous fraction in Figs. 14 for both two and three dimensions. We represent the positions of the DPD particles in two dimensions by a gray scale map, in which the particle positions are weighted by Eq. (7) and assigned an intensity of gray based upon the proportion of each phase in this localized-average. Figures 3 and 4 display the three-dimensional surface of the interface between the two immiscible phases, as defined by there being equal proportions of each fluid in the localized-average. In these four figures the gradual development of domains can be seen. We examined the results of the domain size analysis for each simulation in plots of $`\mathrm{log}_{10}R`$ vs. $`\mathrm{log}_{10}t`$, and further examined each ensemble average in plots of $`\beta `$ vs. $`\mathrm{log}_{10}t`$; the latter yielded the mean values of $`\beta `$ and corresponding 68% confidence intervals for significant sections of algebraic growth. We show these results for two and three dimensions in Tables IV and V, and Figs. 5 and 6. The range of $`\mathrm{log}_{10}t`$ we give in the tables should be taken as a rough guide, since there is often a high degree of correlation between the start and end times of a particular linear section and its growth exponent within the broad category of late time. We should emphasize that we are using the term “late time” as a relative category in this paper, to distinguish these results from those obtained at very early times. Due to the model parameters and small size of these simulations, we do not expect to probe the viscous or inertial hydrodynamic regimes (see Eqs. (17)–(19), and compare with Refs. ).
We also constructed a series of equal-viscosity simulations in both two and three dimensions. These had $`\mathrm{\Delta }t=0.1`$, 6400 particles, and stopped at $`t=1000`$. The model parameters were the same as in the differing-viscosity simulations, with the exception that both phases were identical to the less-viscous phase of the differing-viscosity simulations (i.e., $`\gamma =\gamma _0`$ and $`\sigma =\sigma _0`$). We calculated the pair distribution function at 64 different points from $`t=1`$ to $`t=1000`$, with ten simulations at each of eight different minority phase fractions from $`\varphi =0.05`$ to $`\varphi =0.5`$ in two dimensions, and ten simulations at each of five different fractions from $`\varphi =0.1`$ to $`\varphi =0.5`$ in three dimensions. We show the time evolution of a single simulation at each value of the minority fraction in Figs. 7 and 8 for two and three dimensions respectively. In these two figures the gradual development of domains can be seen, and is greatly slowed for small minority fraction. We show the scaling exponents for two and three dimensional equal-viscosity fluids in Tables VI and VII, and Figs. 9 and 10, where we mirror the scaling exponents about $`\varphi =0.5`$ to show the full range of minority phase fraction. Comparison with Figs. 7 and 8 qualitatively confirms the same behavior.
## VI Discussion
A feature common to all these simulations is the lack of scale-invariance at very early times, until approximately $`t=10`$ to $`t=50`$ (in our DPD time units) depending on the exact composition of the fluid. This is apparent in the $`g_{01}(r)`$ curves as multiple peaks of similar magnitude, but could not be seen in the frequency domain shown by the static structure function. Figure 11 shows an example of the differing-phase radial distribution function for a three-dimensional equal-viscosity simulation ($`\varphi =0.5`$) at $`t=13.9`$; the crosses represent the actual data and the curve represents the smoothed data. This is one of the simulations we show in Fig. 8. The multiple peaks typically evolve by changing their height and relative weight, but not their position. This is noticeably different from our “late time” behavior of the distribution function, where a single peak gradually advances its position while broadening but retaining its height and general shape. It is because of this “early time” behavior that it was decided not to use solely the global maximum of $`g_{01}(r)`$ to estimate the characteristic domain size.
At these “early times” in all of the simulations, we observed algebraic growth with a very small exponent, roughly $`0.126\pm 0.003`$ in two dimensional equal-viscosity simulations, $`0.062\pm 0.009`$ in two dimensional differing-viscosity simulations (except $`\varphi =0.9`$ (90% viscous phase), for which $`\beta =0.24\pm 0.07`$), and $`0.0246\pm 0.0008`$ for simulations in three dimensions. This corresponds to the region of breakdown of scale invariance we described at the beginning of this section. These “early time” exponents are unaffected by viscous or minority phase fraction in all of the simulations, except the differing-viscosity simulations in two dimensions. Here $`\beta `$ decreases with increasing $`\varphi `$ (viscous phase fraction), probably due to the domain-growth arresting effect of increasing viscosity. We observed a remarkable (though short) regime of $`\beta =0.24\pm 0.07`$ for $`\varphi =0.9`$ (90% viscous phase), for which we have no adequate explanation. Others normally discard similar early time regimes without comment or as an “early stage” or “transient” regime . However, there is growing evidence for the coexistence of multiple domain sizes and hence a breakdown in universality, at least in certain phase-ordering domains .
For “late time” domain growth in the two-dimensional differing-viscosity simulations (see Fig. 5), we observed a fairly constant value of $`\beta `$ for $`\varphi =0.2`$ (20% viscous phase) through $`\varphi =0.6`$ (60% viscous phase), decreasing both for $`\varphi =0.1`$ and very slightly for $`\varphi =0.7`$ and $`\varphi =0.8`$, then decreasing sharply at $`\varphi =0.9`$. This asymmetry is consistent with the variation of the “early time” exponent, in that an increasingly viscous fluid is expected to develop domains more gradually. At increasingly rarefied fractions ($`\varphi =0.1`$ and $`\varphi =0.9`$), domain growth is retarded by the increased isolation of the droplets. The “late time” growth exponent throughout is effectively $`\frac{1}{3}`$, which suggests that the presence of fluids of differing viscosity interferes with the normal $`\beta =\frac{1}{2}`$ growth mechanism in two dimensions. The growth exponent of $`\frac{1}{3}`$ is expected from the LSW evaporation–condensation mechanism . This is in some ways analogous to the effect obtained by deliberately breaking momentum conservation in symmetric quenches, as described previously . Our domains are considerably less circular at all viscous fractions than those observed by Sappelt and Jäckle (compare Figs. 1 and 2 with Ref. ). This is likely due to the lack of hydrodynamic interactions in their model and to the greater difference in viscosity between their two phases. As in their simulations, our two-phase structure for fluids of differing viscosity is not very different from the structure for fluids of equal viscosity. Moreover, our simulations do not reveal any new insights regarding interfacial structure.
In three dimensions, the “late time” domain growth of differing-viscosity fluids displays nearly the opposite behavior, with the scaling exponent increasing as the viscous fraction reaches its extremes. This could be explained by the increased fluid mobility in simulations with an extra spatial dimension, as the majority phase is completely connected and so the domain growth could occur according to the $`\beta =\frac{2}{3}`$ mechanism, which is surface tension driven by hydrodynamic flow, balanced by inertial effects. This is qualitatively substantiated by inspection of Figs. 3 and 4, where a larger degree of connectivity of the majority phase can be observed at the extremes of viscous fraction than for $`\varphi 0.5`$. However, a more obvious mechanism for the domain growth would be the $`\beta =\frac{1}{3}`$ LSW evaporation–condensation mechanism ; it may be a combination of these two mechanisms that leads to our observed $`\frac{1}{3}<\beta <\frac{2}{3}`$ growth. A slight asymmetry in the “late time” growth exponent is also evident, with domain growth proceeding more slowly with increasingly viscous fluids of non-extreme viscous fraction.
For equal-viscosity fluids in two dimensions we observed the expected $`\beta =\frac{1}{2}`$ for symmetric quenches ($`\varphi =0.5`$. As we reduced the minority phase fraction, we observed a steady decrease in the scaling exponent until $`\beta =\frac{1}{3}`$ is reached at the extremes. This confirms the results of other workers that increasingly off-critical quenches retard the domain growth, while providing support for the observed slowdown of growth at the extremes of viscous fraction in differing-viscosity fluids in two dimensions, which we commented on above.
In three dimensions, the domain growth of equal-viscosity fluids appears largely unaffected by varying the minority phase fraction. Although there may be some increase in $`\beta `$ at the extremes of $`\varphi `$ (as seen in the differing-viscosity fluid in three dimensions), this is difficult to confirm definitely because of the large variation in rate of growth observed for the simulations with $`\varphi =0.1`$, and hence correspondingly large confidence interval. The scaling exponent throughout is close to $`\beta =\frac{1}{3}`$. Whereas Jury et al. were intending to probe the viscous or inertial hydrodynamic regimes with their DPD simulations of equal-viscosity fluids in three dimensions, we aimed only to probe length scales below $`R_d`$ and $`R_h`$ \[see Eqs. (17)–(19)\]. As such, our results are fully consistent with theirs. Our exclusion of finite size effects is more rigorous than theirs, and although not as extreme as that advocated by Kendon et al. our method gives statistical confidence that these domains are scaling algebraically. Both Jury et al. and Kendon et al. were able to cover the time domain more fully in three-dimensional equal-viscosity symmetric quenches only at the cost of performing a large number of computationally very intensive and very expensive simulations.
## VII Conclusions
In this paper, we have described simulations of the domain growth and phase separation of hydrodynamically-correct binary immiscible fluids of differing and equal viscosity as a function of minority phase concentration in both two and three spatial dimensions. Due to our choice of model parameters and the small size of our simulations, we did not expect to probe the viscous or inertial hydrodynamic regimes. In three dimensions, we found that the characteristic domain size scales as $`t^{1/3}`$ for simulations of differing and equal-viscosity fluids developing from symmetric and slightly off-critical quenches. For highly off-critical quenches we observe an increase in the scaling exponent. In two dimensions, we also observe $`t^{1/3}`$ in simulations of differing-viscosity fluids developing from symmetric and slightly off-critical quenches, although we observe a decrease in the scaling exponent for highly off-critical quenches. In equal-viscosity fluids in two dimensions, we observe $`t^{1/2}`$ for symmetric quenches and a roughly linear decrease to $`t^{1/3}`$ for highly off-critical quenches; these results are in agreement with similar lattice-gas simulations in two dimensions .
Obtaining meaningful results for ensemble averages of highly off-critical binary immiscible fluids was only made feasible by our automation of the calculation of the characteristic domain size by the pair correlation function. It also made possible the identification of a regime of breakdown of scale invariance at very early times, which was not noticeable in our original analysis using the static structure function. Further simulations aimed to probe the viscous and inertial hydrodynamic regimes \[see Eqs. (17)–(19)\] would be a useful addition to this work, as would simulations aimed to cover longer periods of time; however, both would require substantially increased computational work.
## Acknowledgments
Many helpful discussions were had with Peter Bladon, Bruce Boghosian, Alan Bray, Mike Cates, Pep Español, Simon Jury, Colin Marsh, John Melrose, and Julia Yeomans; extra thanks is due to Matthew Probert and Matt Segall. We would also like to thank NATO for a grant which has supported this work in part, the EPSRC (U.K.) E7 Grand Challenge in Colloidal Hydrodynamics for providing access to the Cray T3D at the Edinburgh Parallel Computing Centre, the CSAR service at the University of Manchester for access to their Cray T3E and SGI Origin2000 through the EPSRC High Performance Computing for Complex Fluids grant, and the Cambridge High Performance Computing Facility for access to their Hitachi SR2201. KEN gratefully acknowledges financial support from NSERC (Canada) and the ORS Awards Scheme (U.K.).
## High-Performance Computing
In this appendix we provide a few comments regarding the running of DPD simulations on high performance computers. We usually had easy access to single-processor workstations, with a large variety of types and speeds of processors. Multi-processor machines allowing parallel execution of simulations are much less common and are more difficult to obtain access to, although they have become more common during this research project. We used both the Cray T3D of the Edinburgh Parallel Computing Centre (EPCC) and the Hitachi SR2201 of the Cambridge High Performance Computing Facility (HPCF) for computing the results described in this paper; the former consisted of 512 processor nodes and the latter consists of 256 nodes.
The implementation of the dissipative particle dynamics algorithm is very similar to that of conventional molecular dynamics algorithms . For example, we divide the periodic spatial domain (the simulation cell) into a regular array of equally sized link-cells, such that each side of the rectangular domain has an integer number of cells and each cell is at least $`r_c`$ across. Each link-cell consists of a dynamically allocated array of particles, and pointers to the neighboring cells. Individual particles consist of the position–momentum vector pair and a color index.
For each time step we iterate through the particles in each link-cell, calculating the force acting on each particle as it interacts with the particles in the same and neighboring link-cells. Since the DPD force acts between pairs of particles, we must ignore half of the neighboring cells to avoid duplication. When considering a different particle pair, we compare the square of the separation distance with $`r_c^2`$, skipping to the next particles if the pair is out of range. We then compute the new position and velocity as determined by the finite-difference algorithm (see Eq. (13) and Ref. ).
We may write the complete state of the system to file, and we can perform other calculations thereafter, for example to determine the temperature and pressure of the system. We used the freely-available Gnu-make utility to dictate the compilation process, since the decision structures it contains make it simple to write programs portable to a large range of architectures. We created a comprehensive, automated test suite to make it easy to verify that optimizations of the calculations did not accidentally change the results of the computations.
Given constant $`r_c`$ and number density $`n=\rho /m`$, the DPD algorithm scales linearly (in both computation time and memory size) with increasing number of particles ($`N`$), and is limited by computation time on all but the smallest machines. The main simulations we performed for this paper consisted of 6400 particles, and it is on the parallel and serial performance of this size of simulation that we will make most of the following comments. Details of the performance of this size of simulation in two dimensions are shown in Tables VIII and IX. These tables give the elapsed time per node in seconds and relative parallel efficiency for the first 1000 time steps, including data for a variety of computers and partition sizes. These data are for code compiled with the highest level of optimization, including some small reductions in floating-point accuracy. Table VIII describes the computers used to calculate the results in this paper, while Table IX describes the computers to which we have recently been allowed access, such as the Computer Services for Academic Research (CSAR) Cray T3E and SGI Origin2000 in Manchester.
A typical simulation of 50,000 time steps takes 2.5 hours on a 350 MHz Intel Pentium II PC, the fastest single-processor machine to which we had common access. This same simulation would take 2.9 hours on a 16-node partition of the T3D at the EPCC. However, to minimize fragmentation of the machine, jobs using up to 32 nodes were limited to a total execution time of 30 minutes. One possibility was to break up the run into 30 minute portions, but this introduces additional overhead and complications; however, new jobs start instantly because they need not be queued. A better option was to run jobs on a 64-node partition, task farming four 16-node jobs to run simultaneously. There was a 12 hour limit to 64–512 node jobs (6 hours during the week), but there was often a long wait in the queues. If the efficient usage of billed time was a significant concern, sixteen 2-node jobs would complete in 8.0 hours. However, during the week this meant restarting halfway through and waiting in the queue again. Similar comments apply to the Hitachi SR2201, although its queues were limited to 8 hours maximum. The extra administrative overheads involved and the billed usage means that we usually concentrated computation on the serial workstations. However, parallel execution becomes more attractive with larger simulations.
The parallel efficiency of DPD with 6400 particles is good only for a modest number of processor nodes. This is particularly true of the more modern parallel machines such as the Cray T3E and SGI Origin2000, which are proportionally faster in processing than communicating when compared with their older counterparts. Much better parallel efficiency has been observed with larger simulations. The Cray T3D shows an unusual increase in efficiency when going from a serial calculation to a 2-node parallel calculation with 6400 particles; this could be explained by any number of hardware-specific arguments. We should note that the results for the Origin2000 include the effect of sharing the machine with other users, unlike all the other machines whose results appear in Tables VIII and IX, for which each node was dedicated to our calculations.
We decided to write the main simulation program in C/C++ as opposed to Fortran. This choice was made because C/C++ were believed to be the most appropriate languages for dealing with DPD simulations which consist of a large amount of book-keeping wrapped around fairly simple computations. C/C++ and Fortran are highly portable to different computer architectures, and although well-written Fortran is more efficient on vector machines, for almost all other situations they are of similar speed, given equally good compilers. The use of vector machines (such as the Hitachi SR2201) was not anticipated when this work on DPD began several years ago. Furthermore, it was not believed that the basic algorithm would vectorize well, due to the short vector length in typical computations. Large programs are easier to maintain in C/C++ than in Fortran, although the increasingly well-supported Fortran 90 and 95 make the difference less significant.
Finally, we comment on our findings in tuning the message passing interface (MPI) calls for the Cray T3D. In our simulations, it was found that blocking calls (sends and receives) were faster than non-blocking calls and were easier to use correctly. Furthermore, better scaling was achieved by sending the size of a variable-size message in a separate message rather than probing incoming messages to determine their size. Finally, using derived data types to remove unneeded data from messages was slower than sending everything. |
no-problem/9911/astro-ph9911441.html | ar5iv | text | # Theoretical models for classical Cepheids: V. Multiwavelength relations
## 1 Introduction
Since the Leavitt (1912) discovery of the relationship between period and apparent magnitude for Cepheids in the Magellanic Clouds, these variables have been playing a fundamental role in the determination of distances to Local Group galaxies, in the calibration of various secondary distance indicators and finally in the evaluation of the Hubble constant $`H_0`$. A considerable amount of observational studies have been devoted to the calibration of their characteristic Period-Luminosity ($`PL`$) and Period-Luminosity-Color ($`PLC`$) relations, e.g. via the Baade-Wesselink method (Ripepi et al. 1997; Gieren et al. 1998 and references therein), or making use of Cepheids with distances from Galactic open cluster main-sequence fitting (Turner et al. 1998 and references therein), or adopting the Large Magellanic Cloud (LMC) distance found in other independent ways (see Walker 1999).
The most important question to be answered is whether the bolometric $`PL`$ and $`PLC`$ are universal, as early suggested by Iben & Renzini (1984), or depending on the chemical abundances of the variables. This is a fundamental point since if there is a metallicity effect on the relations in the various photometric bands and the galaxies whose distances we are deriving have metallicities different from the calibrating Cepheids, then a significant correction could be necessary. In recent times, the occurrence of a metallicity effect has become more and more evident, with the observational tests of this effect suggesting that metal-rich Cepheids are brighter than metal-poor Cepheids at fixed period (see Kochanek 1997; Beaulieu et al. 1997; Sasselov et al. 1997; Kennicutt et al. 1998). As for theoretical estimates, models based on linear pulsation calculations suggest a very low effect (Chiosi et al. 1993; Saio & Gautschy 1998; Alibert et al. 1999), whereas nonlinear, nonlocal and time-dependent convective pulsating models suggest that both the zero point and the slope of the predicted $`PL`$ relations depend on metallicity, with the amplitude of the metallicity effect decreasing at the longer wavelengths (Bono et al. 1999b \[Paper II\]). Moreover, these models show that metal-rich Cepheids are on average intrinsically fainter than the metal-poor ones, at fixed period.
It is worth noticing that all the observational efforts rely on the assumption that the slope of the $`PL`$ relation for different passbands is that found for LMC Cepheids, leading us to suspect that forcing the slope of the multiwavelength relations to be constant may introduce some systematic errors in the attempt of disentangling reddening from metallicity effects. On the other hand, at variance with the linear-nonadiabatic approach which supplies only the blue edge of the instability strip <sup>1</sup><sup>1</sup>1Linear red edge estimates by Chiosi et al. (1993) and by Alibert et al. (1999) were fixed, more or less tentatively, at the effective temperature where the growth rate attains its maximum value. However, the nonlinear results suggest that such an assumption is far from being adequate (Bono et al. 1999c \[Paper III\])., the models presented in Paper II provide fine constraints on both blue and red limits of the pulsation, thus avoiding dangerous ad hoc assumptions on the width of the instability strip (see also Tanvir 1999). These models give also the pulsation amplitude and the predicted mean magnitudes of the pulsator.
The theoretical $`PL`$ and $`PLC`$ relations presented in Paper II deal with the intensity-weighted $`<M_B>`$, $`<M_V>`$ and $`<M_K>`$ magnitudes. In order to have a full insight into the metallicity effect, and considering that the HST observations of extragalactic Cepheids consider the $`I`$ band, in this paper we present the full set of theoretical relations (i.e. Period-Luminosity, Period-Color, Color-Color and Period-Luminosity-Color) in the $`BVRIJK`$ bands. The predicted relations are reported in Sect. 2, while Sect. 3 gives brief comparisons of our theoretical results against observed data for calibrating Cepheids. Some final remarks close the paper.
## 2 Predicted relations
The nonlinear convective pulsating models adopted in this paper are computed with four values of the stellar mass ($`M/M_{}=`$5, 7, 9, 11) and three chemical compositions ($`Y`$=0.25, $`Z`$=0.004; $`Y`$=0.25, $`Z`$=0.008; $`Y`$=0.28, $`Z`$=0.02), taken as representative of Cepheids in the Magellanic Clouds and in the Galaxy. The basic assumptions on the input physics, computing procedures and the adopted mass-luminosity relation have been already discussed in Bono et al. (1998), Bono et al. (1999a \[Paper I\]) and Paper III and will not be repeated. Here we wish only to remark that the whole instability strip moves toward lower effective temperatures as the metallicity increases and, as a consequence, that the bolometric magnitude of the pulsators increases as the metal abundance increases (see Fig. 2 in Paper II).
### 2.1 Period-Luminosity
From the bolometric light curves of our pulsating models and adopting the grid of atmosphere models provided by Castelli et al. (1997a,b), we first derive the predicted intensity-weighted mean magnitude $`<M_\lambda >`$ of pulsators for the $`BVRIJK`$ passbands <sup>2</sup><sup>2</sup>2These atmosphere models do not allow the calculation of synthetic $`H`$ magnitudes.. The fundamental pulsators plotted in Fig. 1, where the different lines depict the blue and red limits of the pulsation region at the various compositions, show that the metal abundance effects on the predicted location and width of the instability strip are significantly dependent on the photometric passband. In particular, the models confirm early results (e.g., see Laney & Stobie 1994; Tanvir 1999 and reference therein) that infrared magnitudes are needed to reduce both the intrinsic scatter and the metal abundance dependence of the $`PL`$ relation.
This matter has been already discussed in Paper II on the basis of the predicted $`PL_V`$ and $`PL_K`$ relations derived under the assumption of a uniformly populated instability strip. However, since the $`PL`$ is a “statistical” relation which provides the average of the Cepheid magnitudes $`\overline{M_\lambda }`$ at a given period, we decide to estimate the effect of a different pulsator population. Thus, following the procedure outlined by Kennicutt et al. (1998), for each given chemical composition we populate the instability strip<sup>3</sup><sup>3</sup>3The location of the instability strip is constrained by the predicted blue and red edges. with 1000 pulsators and a mass distribution as given by $`dn/dm=m^3`$ over the mass range 5-11$`M_{}`$. The corresponding luminosities are obtained from the mass-luminosity relation derived from canonical evolutionary models, i.e. with vanishing efficiency of convective core overshooting (see Bono & Marconi 1997; Paper II; Paper III).
Fig. 2 shows the resulting $`\mathrm{log}PM_\lambda `$ distribution of fundamental pulsators with the three selected metallicities. We derive that the pulsator distribution becomes more and more linear by going toward the infrared and that, aiming at reducing the intrinsic scatter of $`PL`$, the pulsator distributions in the $`\mathrm{log}PM_B`$, $`\mathrm{log}PM_V`$ and $`\mathrm{log}PM_R`$ planes are much better represented by a quadratic relation. The dashed lines in Fig. 2 show the quadratic least square fit ($`\overline{M_\lambda }=a+b\mathrm{log}P+c\mathrm{log}P^2`$), while the solid lines refer to the linear approximation ($`\overline{M_\lambda }=a+b\mathrm{log}P`$). The values for the coefficients $`a`$, $`b`$ and $`c`$ of each $`PL_\lambda `$ relation are listed in Table 1 (quadratic solution) and Table 2 (linear solution), together with the rms dispersion ($`\sigma `$) of $`M_\lambda `$ about the fit. However, we wish to remark that the present solutions refer to a specific pulsator distribution and that different populations may modify the results. As a matter of example, if the longer periods (log$`P`$1.5) are rejected in the final fit, then the predicted linear $`PL`$ relations become steeper and the intrinsic dispersion in the $`BVR`$ bands is reduced (see Table 3).
As a whole, each predicted $`PL_\lambda `$ relation seems to become steeper at the lower metal abundances, with the amplitude of the effect decreasing from $`B`$ to $`K`$ filter. Moreover, the slope and the intrinsic dispersion of the predicted $`PL`$ relation at a given $`Z`$ decrease as the filter wavelength increases (see Fig. 3 for the linear relations), in close agreement with the observed trend (e.g. see Fig. 6 in Madore & Freedman 1991).
In closing this section, let us observe that the adopted way to populate the instability strip allows $`PL`$ relations only for the “static” magnitudes (the value the star would have were it not pulsating), while the observations deal with the mean magnitudes averaged over the pulsation cycle \[($`m_\lambda `$) if magnitude-weighted and $`<m_\lambda >`$ if intensity-weighted\]. It has been recently shown (Caputo et al. 1999 \[Paper IV\]) that the differences among static and mean magnitudes and between magnitude-weighted and intensity-weighted averages are always smaller than the intrinsic scatter of the $`PL`$ relation ($`0.1`$ mag for V magnitudes and $`0.01`$ mag for K magnitudes). Here, given the marginal agreement of the coefficients in Table 1, Table 2 and Table 3 with the values given in Paper IV, we conclude that the additional effect on the intrinsic scatter of the $`PL`$ relations, as due to different Cepheid populations, is negligible, provided that a statistically significant sample of variables is taken into account.
### 2.2 Period-Color and Color-Color
The three methods of deriving the mean color over the pulsation cycle use either ($`m_j`$-$`m_i`$), the average over the color curve taken in magnitude units, or $`<m_jm_i>`$, the mean intensity over the color curve transformed into magnitude, or $`<m_j>`$-$`<m_i>`$, the difference of the mean intensities transformed into magnitude, performed separately over the two bands. In Paper IV it has been shown that there are some significant differences between static and synthetic mean colors, and that the predicted ($`m_j`$-$`m_i`$) colors are generally redder than $`<m_j>`$-$`<m_i>`$ colors, with the difference depending on the shape of the light curves, in close agreement with observed colors for Galactic Cepheids. As a matter of example, the predicted difference $`(BV)[<B><V>]`$ ranges from 0.02 mag to 0.08 mag, whereas $`(VK)[<V><K>]`$ is in the range of 0.014 mag to 0.060 mag.
Fig. 4 shows the pulsator synthetic mean colors as a function of the period for the three different metallicities. Since, given the finite width of the instability strip, also the $`PC`$ relation is a “statistical” relation between period and the average of the mean color indices $`CI`$, we follow the procedure of Sect. 2.1. It is evident from Fig. 4 that the pulsator distribution shows a quadratic behavior, as well as that the intrinsic dispersion of $`PC`$ relations allows reddening estimates within $`0.07`$ mag, on average. However, as far as they could help, we present the predicted relations $`\overline{CI}=A+B`$log$`P`$ at the various compositions. The relations are plotted in Fig. 4 and the $`A`$ and $`B`$ coefficients are listed in Table 4 together with the rms deviation of the color about the fit ($`\sigma `$). Note that also the slope of the predicted $`PC`$ relations depends on the metal abundance, increasing with increasing $`Z`$.
As for the color-color ($`CC`$) relations, Fig. 5 shows that the spurious effect due to the finite width of the instability strip is almost removed. For this reason, from our pulsating models we derive a set of color-color ($`CC`$) relations correlating $`<B>`$-$`<V>`$ with $`<V>`$-$`<R>`$, $`<V>`$-$`<I>`$, $`<V>`$-$`<J>`$ and $`<V>`$-$`<K>`$, but similar relations adopting ($`m_j`$-$`m_i`$) or $`<m_jm_i>`$ colors can be obtained upon request. The predicted $`CC`$ linear relations are given in Table 5, while Fig. 5 shows the pulsator distribution in the color-color plane. One may notice that the intrinsic scatter of the theoretical relations is very small, with the rms dispersion of $`<B>`$-$`<V>`$ about the fit smaller than 0.01 mag.
### 2.3 Period-Luminosity-Color
Since the pioneering paper by Sandage (1958), Sandage & Gratton (1963) and Sandage & Tammann (1968), it is well known that if the Cepheid magnitude is given as a function of the pulsator period and color, i.e. if the Period-Luminosity-Color relation is considered, then the tight correlation among the parameters of individual Cepheids is reproduced (see also Laney & Stobie 1986; Madore & Freedman 1991; Feast 1995 and references therein).
In Paper II it has been shown that the intrinsic scatter of the $`PLC`$ relations in the visual ($`<M_V>`$, $`<B>`$-$`<V>`$) and infrared ($`<M_V>`$, $`<V>`$-$`<K>`$) is $`0.04`$ mag. Furthermore, we showed in Paper IV that the systematic effect on the predicted visual $`PLC`$ relations, as due to the adopted method of averaging magnitudes and colors over the pulsation cycle, are always larger than the intrinsic scatter of the relation. For this reason in this paper we present only the predicted $`PLC`$ relations $`<M_V>=\alpha +\beta \mathrm{log}P+\gamma [<M_V><M_\lambda >]`$, but similar relations for magnitude-weighted values can be obtained upon request.
From the least square solutions through the fundamental models we derive the coefficients $`\alpha `$, $`\beta `$ and $`\gamma `$ presented in Table 6, together with the residual dispersion ($`\sigma `$) of $`<M_V>`$ about the fit. Figs. 6-8 illustrate the remarkably small scatter of the $`PLC`$ relations (see Fig. 1 for comparison). Moreover, one may notice that adopting $`BV`$ color, the metal-rich Cepheids are brighter than metal-poor ones with the same period and color (see lower panel of Fig. 6), whereas the opposite trend holds with $`VK`$ color (see Fig. 8). As a “natural” consequence, the predicted relationship with $`VI`$ color (lower panel of Fig. 7) turns out to be almost independent of the metal abundance.
In order to complete the theoretical framework for classical Cepheids, we have finally considered the Wesenheit quantities $`W`$ (Madore 1982) which are often used to get a reddening-free formulation of the $`PL`$ relation. With $`A_\lambda `$ giving the absorption in the $`\lambda `$-passband, one has
$$W(B,V)=V[A_V/(A_BA_V)](BV)$$
$$W(B,V)=V_0+A_V[A_V/(A_BA_V)][(BV)_0+(A_BA_V)]$$
$$W(B,V)=V_0[A_V/(A_BA_V)](BV)_0,$$
$$W(V,R)=R[A_R/(A_VA_R)](VR)$$
$$W(V,R)=R_0+A_R[A_R/(A_VA_R)][(VR)_0+(A_VA_R)]$$
$$W(V,R)=R_0[A_R/(A_VA_R)](VR)_0,$$
$$etc\mathrm{}.$$
Table 7 gives the coefficients of the theoretical reddening-free $`PL`$ \[hereinafter $`WPL`$\] relations derived from our fundamental models. Note that present results adopt $`A_V/E_{BV}=3.1`$, $`A_R/E_{VR}=5.29`$, $`A_I/E_{VI}=1.54`$, $`A_J/E_{VJ}=0.39`$ and $`A_K/E_{VK}=0.13`$ from the Cardelli et al. (1989) extinction model, but formulations using different ratios of total to selective absorption can be obtained upon request.
Figs. 9-10 show the theoretical Wesenheit quantities as a function of the period, together with the predicted relations. From a comparison of Table 7 with Table 2 one derives that the intrinsic scatter of $`W(B,V)PL`$ and $`W(V,R)PL`$ is significantly lower than the dispersion of $`PL`$, while no significant improvement occurs with $`W(V,I)PL`$ and somehow larger dispersions are found $`W(V,J)PL`$ and $`W(V,K)PL`$.
Moreover, the comparison between Figs. 9-10 and Figs. 6-8 discloses the deep difference between $`PLC`$ and $`WPL`$ relations (see also Madore & Freedman 1991). The former ones are able to define accurately the properties of individual Cepheids within the instability strip, whereas the latter ones are thought to cancel the reddening effect. As a consequence, provided that the variables are at the same distance and have the same metal abundance, the scatter in observed $`PLC`$ relations should depend on errors in the adopted reddening, whereas the scatter in observed $`WPL`$ relations is a residual effect of the finite width of the strip.
## 3 Comparison with Galactic Cepheids
The principal aim of this paper is to present a wide homogeneus theoretical scenario for Cepheid studies. The analysis of observed pulsational properties is out of our present intentions, but it seems necessary to verify the reliability of our models against well studied variables. For this purpose, we take into consideration the sample of calibrating Galactic Cepheids studied by Gieren et al. (1998, \[GFG\]) and by Laney & Stobie (1994, \[LS\]). To the intensity mean magnitudes $`BVJK`$ magnitudes we add the $`<I>`$ magnitudes as derived from $`<V><I>(VI)`$=-0.03 mag, where the magnitude-weighted (V-I) colors are by Caldwell & Coulson (1987). Such a constant correction, suggested by GFG, is confirmed (see Fig. 11) from our synthetic mean colors. Note that, according to GFG, we excluded SV Vul, GY Sge and S Vul for problems with a variable period and EV Sct, SZ Tau and QZ Nor for uncertainty with regard to the pulsation mode.
Before analyzing the calibrating Galactic Cepheids, let us briefly consider the recent results on Magellanic Clouds Cepheids provided by the OGLE II microlensing survey (Udalski et al. 1999). The final $`PL_V`$ and $`PL_I`$ relations given by these authors for Cepheids with $`\mathrm{log}P1.5`$, are $`M_V`$=-1.18-2.765log$`P`$ and $`M_I`$=-1.66-2.963log$`P`$, with a LMC distance modulus $`\mu _{LMC}`$=18.22 mag. The data in Table 3 with $`Z`$=0.008 yield $`M_V`$=-1.37-2.75log$`P`$ and $`M_I`$=-1.95-2.98log$`P`$, namely a predicted slope in close agreement with OGLE measurements but a zero-point which is brighter by roughly 0.24 mag, suggesting $`\mu _{LMC}`$=18.46 mag, a value which fits the recent upward revision of the LMC distance derived from field red clump stars (Zaritsky 1999).
As for the $`W(V,I)PL`$ relation presented by Udalski et al. (1999), we find that its slope (-3.28) is in fair agreement with the results (-3.17) of our models with $`Z`$=0.008 (see also the lower panel of Fig. 12). Conversely, the OGLE $`PLC_I`$ relation discloses a significative difference with respect to our predicted relations. As shown in the upper panel of Fig. 12, it seems that the OGLE color-term is not able to fully remove the effect of the finite width of the instability strip (as a matter of comparison, see the lower panel of Fig. 7). At the moment we cannot explain such a disagreement and we expect the release of LMC and SMC data to test the full set of our predictions.
Passing to the Galactic Cepheids, Fig. 13 and Fig. 14 show the predicted linear $`PL`$ relations with $`Z`$=0.02 (dotted line) in comparison with the calibrating Cepheids (dots) from LS and GFG, respectively. Note that both the reddening and true distance modulus given by the authors are adopted. We find a fair agreement in the infrared, whereas in the $`V`$ and $`I`$ bands the observed Cepheids with the longer periods appear brighter than the predicted relations. This could suggest that these luminous variables have lower metallicities than the currently adopted value $`Z`$=0.02 (see Fry & Carney 1997) or be near the blue edge of the instability strip. One could also suspect that the predicted slopes with $`Z`$=0.02 are smaller than the actual ones. This seems supported by Fig. 15 where the steeper $`PL_V`$ relation provided by Udalski et al. (1999) is taken into account.
However, in order to remove the effects of the finite width of the instability strip, let us apply our predicted $`PLC`$ relations to the calibrating Cepheids. Figs. 16-17 disclose that predictions and observed data are now in a better agreement. Meanwhile, there are some evidence that the calibrating Galactic Cepheids lie on the average below the theoretical $`PLC`$ relations with $`Z`$=0.02. In other words, our predicted relations yield somehow larger true distance moduli than those given by LS and GFG. Specifically, the left panels of Fig. 18 show that the difference between LS and GFG data and the results of our predicted $`PLC_{VI}`$ relation with $`Z`$=0.02, by adopting the reddening values listed by these authors, have a significant period dependence. This trend cannot be ascribed to our color and period term since a quite similar behavior is present in the right panels which refer to the $`PLC_{VI}`$ relation given by Udalski et al. (1999). Notwithstanding the different values of the period and color terms, one finds a similar trend, with the OGLE relation suggesting smaller distances than those listed by LS and GFG.
On the other hand, we show in Fig. 19 that the $`BVI`$ reddenings <sup>4</sup><sup>4</sup>4We adopt the ratios of total to selective absorption in the $`VI`$ bands from Caldwell & Coulson (1987) and Laney & Stobie (1993). derived from the predicted color-color relation with $`Z`$=0.02 are slightly lower than those given by LS and GFG. The origin of the discrepancy could be found in Fig. 20, where our predicted color-color relation with $`Z`$=0.02 (dotted line) is plotted together with the two quadratic relations given by Dean et al. (1978) for the intrinsic color-color locus of the Galactic Cepheids (solid line and dashed line). Since observational studies generally adopt the solution depicted with the solid line, the reason for which our predicted $`CC`$ relation yields slightly smaller $`BVI`$ reddenings is explained. This could provide the key to understand the difference in the true distance moduli.
## 4 Summary
From nonlinear, nonlocal and time-dependent convective pulsating models we have derived a set of homogeneus $`PL`$, $`PC`$, $`CC`$, $`PLC`$ and $`WPL`$ relations in the $`BVRIJK`$ passbands at varying chemical composition. We find that the predicted relations are, in various degrees, metallicity dependent, suggesting that the adoption of universal relations for Cepheid studies should be treated with caution.
The slope of the predicted relations with $`Z`$=0.004-0.008 matches the results of LMC and SMC Cepheids collected during the OGLE II microlensing survey (Udalski et al. 1999), but our zero points suggest a larger LMC distance, in agreement with the suggestions by Zaritsky (1999).
As for the comparison of the predicted relations at $`Z`$=0.02 with the observed data of calibrating Galactic Cepheids, we find a fair agreement even though the slope of our $`PL_V`$ and $`PL_I`$ relations seems shallower than observed. This is the consequence of the significant metallicity dependence in the predicted relations and we believe that a careful analysis (of larger samples of Galactic Cepheids) which takes into account the actual metallicity dispersion is required to settle the question.
Special requests for other theoretical relations can be addressed to M. Marconi.
###### Acknowledgements.
We deeply thank the referee (Dr. Bersier) fore several valuable comments which improved the first version of the paper. Financial support for this work was provided by the Ministero dell’Università e della Ricerca Scientifica e Tecnologica (MURST) under the scientific project “Stellar Evolution” (Vittorio Castellani, coordinator). |
no-problem/9911/cond-mat9911334.html | ar5iv | text | # Comments on New representations of the Hecke algebra and algebraic Bethe Ansatz for an integrable generalized spin ladder, by H.-Q. Zhou, H. Frahm and M.D. Gould, cond-mat/9911072
## Abstract
The authors of cond-mat/9911072 claim to introduce “new representations of the Hecke algebra.” These representations are shown to be the XXC models introduced two years ago in solv-int/9712008, and repeatedly studied and referred to in subsequent papers. They are in fact representations of the Temperley-Lieb algebra. Several other remarks are made and mistakes are pointed out.
1) The authors of claim to have recently constructed a “novel class of representations of the Hecke algebra.” It is shown here that these representations were already known. They were discovered two years ago and named XXC models .
Start from the matrix $`\stackrel{ˇ}{R}(x)`$ defined by (2) and (4) in :
$$\stackrel{ˇ}{R}(x)=\stackrel{ˇ}{R}xq^2\stackrel{ˇ}{R}^1,$$
(1)
where
$`\stackrel{ˇ}{R}`$ $`=`$ $`q{\displaystyle \underset{\alpha A,\beta B}{}}\left(X^{\alpha \beta }X^{\beta \alpha }+X^{\beta \alpha }X^{\alpha \beta }\right)+\left(q^2+1\right)\left({\displaystyle \frac{1}{2}}{\displaystyle \underset{\alpha A}{}}X^{\alpha \alpha }\right)\left({\displaystyle \frac{1}{2}}{\displaystyle \underset{\alpha A}{}}X^{\alpha \alpha }\right)`$ (2)
$`+{\displaystyle \frac{q^21}{2}}{\displaystyle \underset{\alpha A}{}}\left(X^{\alpha \alpha }IIX^{\alpha \alpha }\right)+\left({\displaystyle \frac{1}{4}}q^2+{\displaystyle \frac{3}{4}}\right)II`$
and the index sets $`A`$ and $`B`$ take values from $`\{1,2,\mathrm{},m\}`$ and $`\{m+1,m+2,\mathrm{},n\}`$, respectively. The $`n\times n`$ matrix $`X^{\alpha \beta }`$ has only one non-vanishing element, a 1 at row $`a`$ and column $`b`$. The matrix $`X^{\alpha \beta }`$ is denoted by $`E^{\alpha \beta }`$ below.
Now consider the operators defined in . Let $`n`$, $`n_1`$ and $`n_2`$ be three positive integers such that $`n_1+n_2=n`$, and $`A`$, $`B`$ be two disjoint sets whose union is the set of basis states of $`C^n`$, with card$`(A)=n_1`$ and card$`(B)=n_2`$. Let also
$`P^{(3)}`$ $`=`$ $`{\displaystyle \underset{aA}{}}{\displaystyle \underset{\beta B}{}}\left(E^{\beta a}E^{a\beta }+E^{a\beta }E^{\beta a}\right)`$ (3)
$`P^{(1)}`$ $`=`$ $`P^{(+)}+P^{()}={\displaystyle \underset{aA}{}}{\displaystyle \underset{\beta B}{}}\left(E^{aa}E^{\beta \beta }+E^{\beta \beta }E^{aa}\right)`$ (4)
$`P^{(2)}`$ $`=`$ $`{\displaystyle \underset{a,a^{^{}}A}{}}E^{aa}E^{a^{^{}}a^{^{}}}+{\displaystyle \underset{\beta ,\beta ^{^{}}B}{}}E^{\beta \beta }E^{\beta ^{^{}}\beta ^{^{}}}`$ (5)
The $`n_1n_2`$ parameters $`x_{a\beta }`$ of were taken all equal to one. Latin indices belong to $`A`$ while Greek indices belong to $`B`$.
With obvious identifications and $`mn_1`$ and $`n_2=nm`$, a simple and short calculation allows one to find $`\stackrel{ˇ}{R}^1`$, and to rewrite (1) as follows:
$$\stackrel{ˇ}{R}(x)=q(1x)P^{(3)}+(1q^2)(xP^{(+)}+P^{()})+(1xq^2)P^{(2)}$$
(6)
With the usual passage to additive variable, one obtains the XXC models in their asymmetric form i.e. their Hecke algebra form . A gauge transformation (a special type of similarity transformation) gives the ‘deformed free-fermion’ and Temperley-Lieb forms .
2) The XXC models are not just representations of Hecke algebra, they are more accurately representations of the Temperley-Lieb algebra . They have an underlying $`sl(2)`$ structure. The natural generalizations to $`sl(m+1)`$ are the multiplicity $`A_m`$ models, which include the XXC models . They are representations of the Hecke algebra.
3) The matrix (1), at $`x=1`$, is more accurately equal to $`(1q^2)I`$, and not $`I`$.
4) The Hamiltonian (5) in is exactly the XXC Hamiltonian (18) in . (The boundary terms do not contribute under periodic boundary conditions.) See also (22) in . The $`J`$-term was added by hand, and commutes with all the conserved quantities of the model .
5) In equation (13) of , $`\mathrm{\Lambda }^{(1)}`$ is independent of its arguments. After equation (13): “Unfortunately, it ($`\tau ^{(1)}=tr_{}[P_1\mathrm{}P_M]`$) can not be diagonalized directly in terms of the algebraic Bethe Ansatz.” Rather fortunately, any such unit-shift operator is automatically diagonalized by algebraic Bethe Ansatz (for periodic boundary conditions) every time $`\stackrel{ˇ}{R}(1)`$ is proportional to the identity operator (i.e. $`\stackrel{ˇ}{R}`$ is regular). This is one of the most basic aspects of the algebraic Bethe Ansatz in the framework of the QISM. It is enough to consider any regular $`\stackrel{ˇ}{R}`$-matrix with 3 states per site. Examples include the spin-1 $`sl(2)`$ matrix and the $`sl(3)`$ matrix, both trigonometric or rational.
6) Footnote 1: The thermodynamics may be affected by the increased degeneracy of all the states. The phases of certain roots of unity with order proportional to the number of sites, may affect the finite-size corrections. All this requires careful checks.
7) The “mapping” of is misleading and meaningless. Most Bethe Ansatz equations look alike and are determined by the Dynkin diagram of the Lie algebra and weight of the representation at hand. It necessary to complement them with the eigenvalues and eigenvectors to which they refer, and to take into account the degeneracies and the specific features of the model. The mapping to the Perk-Shultz models is false.
Case in point: The Bethe Ansatz equations (16) in are simply wrong. |
no-problem/9911/hep-ph9911530.html | ar5iv | text | # QCD sum rules on the light cone and 𝐵→𝜋 form factors
## 1 Motivation: CP violation and the CKM matrix
Within the Standard model CP violation is parameterized by a complex phase in the CKM-matrix $`V_{CKM}`$. The various entries of $`V_{CKM}`$ are known with different accuracy: The best known matrix elements are $`V_{ud}`$ and $`V_{us}`$, the first one is obtained from the comparison of super-allowed nuclear $`\beta `$-decay with $`\mu `$-decay, the latter one from the decay $`K\pi \overline{l}\nu `$. Among the matrix elements which are fairly well known are $`V_{cd}`$ (obtained from single charm production in deep-inelastic $`\nu N`$-scattering and from semileptonic decays of charmed mesons), $`V_{cs}`$ (which can be obtained from the decay $`D\overline{K}\overline{l}\nu `$) and $`V_{cb}`$. The latter one can be measured either inclusively, using $`BD`$ transitions or exclusively in the semileptonic $`BD`$ transition. Among the matrix elements which are least well known are $`V_{ub}`$, $`V_{td}`$, $`V_{ts}`$ and $`V_{tb}`$. The matrix elements involving the top quark may be obtained from $`B_d`$-$`\overline{B}_d`$ mixing ($`V_{td}`$), $`B_s`$-$`\overline{B}_s`$ mixing ($`V_{ts}`$) and single top production ($`V_{tb}`$). The remaining one, $`V_{ub}`$, can be obtained either from the lepton spectrum in inclusive $`BX\overline{l}\nu `$ decays or from exclusive semileptonic $`B`$-decays.
In order to extract the entries of the $`V_{CKM}`$ matrix from the experimental data one needs certain input information from theory. In many cases the theoretical calculations have to rely on non-perturbative methods for QCD. Among the techniques which have been used are: Chiral perturbation theory (for the extraction of $`V_{us}`$), heavy quark effective theory (for $`V_{cb}`$), lattice calculations (for $`V_{ub}`$, $`V_{cs}`$ and $`V_{td}`$), QCD sum rules (for $`V_{ub}`$ and $`V_{cs}`$ ) as well as quark models (for $`V_{ub}`$ and $`V_{cs}`$). Among these the first four enjoy the property that they are based on first principles, whereas quark models are, as the name already indicates, phenomenological models, whose systematic errors are difficult to quantify.
In this talk we will focus on the extraction of $`|V_{ub}|`$ from the exclusive semileptonic decay $`B\pi \overline{l}\nu `$, using QCD sum rules techniques. In the next section we explain the basic principles underlying the sum rule technique. QCD sum rules are based on the operator product expansion (OPE). We mention briefly the differences between a short-distance expansion and an expansion on the light-cone. In the last section we will focus on the sum rule for the decay $`B\pi \overline{l}\nu `$ and give numerical results.
## 2 The QCD sum rule technique
The derivation of a QCD sum rule involves the following six steps:
Step 1: Take a correlation function. For example, in order to calculate the B-meson decay constant $`f_B`$ using QCD sum rules, one would start from the correlation function
$`\mathrm{\Pi }(q^2)`$ $`=`$ $`i{\displaystyle d^4xe^{iqx}0|T(\overline{u}(x)i\gamma _5b(x),\overline{b}(0)i\gamma _5u(0))|0}.`$ (1)
Step 2: Write a dispersion relation for $`\mathrm{\Pi }(q^2)`$:
$`\mathrm{\Pi }(q^2)`$ $`=`$ $`{\displaystyle \frac{1}{\pi }}{\displaystyle 𝑑s\frac{\text{Im}\mathrm{\Pi }(s)}{sq^2}}+\text{subtractions}`$ (2)
Step 3: Express the absorbative part of $`\mathrm{\Pi }(q^2)`$ in terms of hadronic quantities. In our example above this would involve $`f_B`$.
Step 4: Calculate the correlation function in the asymptotic region using QCD and operator product expansion.
Step 5: Use quark-hadron duality to relate the hadronic representation from step 3 to the absorbative part of the QCD calculation from step 4.
Step 6: The sum rule can be improved by applying a Borel transformation to both sides.
These steps provide a “cooking recipe” for QCD sum rules. A few comments: The QCD calculation in step 4 is based on the operator product expansion. A nonlocal composite operator like $`j(x)j(0)`$ (in the example above we have $`j(x)=\overline{u}(x)i\gamma _5b(x)`$) is expanded into a series of well-defined local operators $`𝒪_n`$:
$`j(x)j(0)`$ $`=`$ $`{\displaystyle \underset{n}{}}C_n𝒪_n`$ (3)
This separates soft and hard physics: The Wilson coefficients $`C_n`$ contain the information on short distance physics and can be calculated using perturbation theory. The matrix elements of the local operators $`𝒪_n`$ parameterize the long distance physics. They are universal non-perturbative quantities. There are two version of the operator product expansion: The short distance expansion and the expansion on the light-cone. In the former one the various local operators are classified by their dimensions, where as in the latter one the classification goes by twist (dimension minus spin).
The sum rule can be improved by applying the Borel operator to both the hadronic representation and the QCD calculation. The Borel operator is given by
$`\widehat{B}`$ $`=`$ $`\underset{Q^2\mathrm{},n\mathrm{},Q^2/n=M^2}{lim}{\displaystyle \frac{1}{(n1)!}}\left(Q^2\right)^n\left({\displaystyle \frac{d}{dQ^2}}\right)^n.`$ (4)
The Borel transformation gives rise to an exponential supression of the higher resonances and the continuum in the hadronic representation. The power corrections in the QCD calculations are supressed by factors of $`(1/M^2)^n`$. An upper limit on $`M^2`$ is obtained by requiring that the contributions from the higher resonances and the continuum should not be too large. A lower limit on $`M^2`$ is obtained from requiring that terms supressed by powers of $`1/M^2`$ should be subdominant. It is important to check that this defines a window, in which the final results are insensitive to the variation of the Borel parameter $`M^2`$.
## 3 $`B\pi \overline{l}\nu `$ and sum rules on the light cone
The relevant hadronic matrix element is parameterized by two form factors $`f^+`$ and $`f^{}`$ :
$`\pi (q)|\overline{u}\gamma _\mu b|B(p+q)`$ $`=`$ $`2f^+(p^2)q_\mu +\left(f^+(p^2)+f^{}(p^2)\right)p_\mu `$ (5)
If we neglect lepton masses only the form factor $`f^+`$ gives a contribution to the decay width. In order to obtain $`f^+`$ from QCD sum rules we start from the correlation function
$`F_\mu (p,q)`$ $`=`$ $`i{\displaystyle d^4xe^{ipx}\pi (q)|T(\overline{u}(x)\gamma _\mu b(x),\overline{b}(0)im_b\gamma _5d(0))|0}`$ (6)
$`=`$ $`F(p^2,(p+q)^2)q_\mu +\stackrel{~}{F}(p^2,(p+q)^2)p_\mu .`$
It can be shown that a short distance expansion is only useful in the soft pion limit ($`q0`$). A better approach is provided by the expansion around the light cone $`x^2=0`$ with operators of increasing twist. In our case the leading twist-2 contribution is given by
$`\pi (q)|\overline{u}(x)\gamma _\mu \gamma _5d(0)|0`$ $`=`$ $`iq_\mu f_\pi {\displaystyle \underset{0}{\overset{1}{}}}𝑑ue^{iuqx}\phi _\pi (u).`$ (7)
$`\phi _\pi (u)`$ is known as the twist-2 pion light-cone wave function. The correlation function $`F`$ can now be written as a convolution of the pion wave function with a hard scattering amplitude $`T(p^2,(p+q)^2,u)`$:
$`F(p^2,(p+q)^2)`$ $`=`$ $`f_\pi {\displaystyle \underset{0}{\overset{1}{}}}𝑑u\phi _\pi (u)T(p^2,(p+q)^2,u)`$ (8)
The hard scattering amplitude $`T`$ can be calculated within perturbation theory, whereas the light-cone wave function contains the non-perturbative information. At NLO both the hard scattering amplitude $`T`$ and the wave function $`\phi _\pi `$ depend on a factorization scale $`\mu `$. The evolution of $`\phi _\pi `$ with this scale can be calculated perturbatively . The Brodsky-Lepage evolution kernel for the pion wave function is analog to the Altarelli-Parisi kernel describing the evolution of parton densities. It turns out that it is convenient to express the wave function in terms of Gegenbauer polynomials. To leading order these Gegenbauer polynomials are eigenfunctions of the evolution kernel. However this is no longer true at next-to-leading order and mixing effects have to be taken into account. At present the twist 2 contributions have been calculated to next-to-leading order , and twist 3 and twist 4 to leading order . Fig. 1 shows the result for the $`B\pi `$ form factor obtained from the light-cone sum rule in comparison to lattice results.
The theoretical error is estimated from a variation of the Borel parameter, the $`b`$-quark mass, the threshold parameter $`s_0`$, the quark condensate density and the normalization scale. Furthermore our insufficient knowledge of the shape of the light-cone wave functions, unknown higher order perturbative corrections or higher twist effects contribute as well to the theoretical error. Our final result is fitted to the parameterization
$`f^+(p^2)`$ $`=`$ $`{\displaystyle \frac{f^+(0)}{\left(1\frac{p^2}{m_B^{}^2}\right)\left(1\alpha \frac{p^2}{m_B^{}^2}\right)}}`$ (9)
with
$`f^+(0)=0.27\pm 0.05,`$ $`\alpha =0.35\pm 0.01.`$ (10)
The form factor $`f^+(p^2)`$ enters the partial decay width
$`{\displaystyle \frac{d\mathrm{\Gamma }}{dp^2}}`$ $`=`$ $`{\displaystyle \frac{G_F^2|V_{ub}|^2}{24\pi ^3}}\left(E_\pi ^2m_\pi ^2\right)^{3/2}\left(f^+(p^2)\right)^2.`$ (11)
Integration yields
$`\mathrm{\Gamma }(B^0\pi ^{}e^+\nu _e)`$ $`=`$ $`(6.7\pm 2.8)|V_{ub}|^2\text{ps}^1.`$ (12)
Comparing this with the experimental value
$`\mathrm{\Gamma }(B^0\pi ^{}e^+\nu _e)`$ $`=`$ $`(1.15\pm 0.34)10^4\text{ps}^1`$ (13)
one obtains
$`|V_{ub}|`$ $`=`$ $`0.0041\pm 0.0005\pm 0.0006,`$ (14)
where the first error corresponds to the current experimental uncertainty and the second error to the estimated theoretical uncertainty. One may expect more precise data from the forecoming experiments of Babar and Belle. On the theoretical side, an improved knowledge of the pion light-cone wave function and perturbative corrections to twist 3 would reduce the theoretical error.
## References |
no-problem/9911/astro-ph9911516.html | ar5iv | text | # Untitled Document
Energy Sources of Soft Gamma-Ray Repeaters
K. S. Cheng<sup>1</sup> and Z. G. Dai<sup>2</sup>
<sup>1</sup>Department of Physics, University of Hong Kong, Hong Kong, China
<sup>2</sup>Department of Astronomy, Nanjing University, Nanjing 210093, China
ABSTRACT
Quiescence and burst emission and relativistic particle winds of soft gamma-ray repeaters (SGRs) have been widely interpreted to result from ultrastrongly magnetized neutron stars. In this magnetar model, the magnetic energy and gravitational energy of the neutron stars are suggested as the energy sources of all the emission and winds. However, Harding, Contopoulos & Kazanas (1999) have shown that the magnetic field should be only $`3\times 10^{13}`$ G in order to match the characteristic spin-down timescale of SGR $`180620`$ and its SNR age. Here we argue that if the magnetic field is indeed so weak, the previously suggested energy sources seem problematic. We further propose a plausible model in which SGR pulsars are young strange stars with superconducting cores and with a poloidal magnetic field of $`3\times 10^{13}`$ G. In this model, the movement of the flux tubes not only leads to crustal cracking of the stars, giving rise to deconfinement of crustal matter to strange matter, but also to movement of internal magnetic toruses with the flux tubes. The former process will result in burst and quiescence emission and the latter process will produce steady relativistic winds which will power the surrounding supernova remnants.
Subject headings: gamma-ray: bursts – magnetic fields – dense matter – stars: neutron
1. INTRODUCTION
The soft gamma-ray repeaters (SGRs) are a small, enigmatic class of high-energy transient sources, which differ from classical gamma-ray bursts by their durations (typically $`0.11`$ s), soft spectra with characteristic energies of $`3050`$ keV and their repetition. The other properties of four known SGRs include: (1) all of them are associated with supernova remnants (SNRs). SGR $`052566`$ appears to be associated with SNR N49 in the Large Magellanic Cloud (Evans et al. 1980; Cline et al. 1982). The second burster, SGR $`180620`$, which produced $`110`$ bursts during a 7-yr span (Laros et al. 1987) and recently became active again (Kouveliotou et al. 1994), appears to be coincident with SNR G$`10.00.3`$ (Murakami et al. 1994), confirming an earlier suggestion (Kulkarni & Frail 1993). The age of this SNR was estimated to be $`10^4`$ yr based on angular diameter versus surface brightness argument (Kulkarni et al. 1994). The third burster, SGR $`1900+14`$, is associated with SNR G$`42.8+0.6`$ (Vasisht et al. 1994), whose age is also $`10^4`$ yr. The fourth burster, SGR $`162741`$, was recently discovered to be associated with SNR G$`337.00.1`$ (Hurley et al. 1999; Woods et al. 1999). From these SGR-SNR associations, the burst peak luminosities can be estimated to be a few orders of magnitude higher than the standard Eddington luminosity for a stellar-mass star. For example, SGR $`180620`$ produced bursts with $`10^4`$ times the Edditington luminosity (Fenimore, Laros & Ulmer 1994). (2) In addition to short bursts of soft gamma-ray photons, the persistent X-ray emission has been detected from SGRs (Murakumi et al. 1994; Vasisht et al. 1994; Rothschild, Kulkarni & Lingenfelter 1994). The luminosities of the persistent X-ray emission are $`7\times 10^{35}\mathrm{erg}\mathrm{s}^1`$ for SGR $`052566`$, $`3\times 10^{35}\mathrm{erg}\mathrm{s}^1`$ for SGR $`180620`$, and $`10^{35}\mathrm{erg}\mathrm{s}^1`$ for SGR $`1900+14`$. (3) Recently, a period of $`P=7.47`$ s and its derivative $`\dot{P}=8.3\times 10^{11}\mathrm{s}\mathrm{s}^1`$ have been detected from SGR $`180620`$ in quiescent emission (Kouveliotou et al 1998), and a period of $`P=5.16`$ s and its derivative $`\dot{P}=6.0\times 10^{11}\mathrm{s}\mathrm{s}^1`$ have been discovered from SGR $`1900+14`$ in quiescent emission (Kouveliotou et al. 1999). All of these observations clearly show that SGRs are young pulsars. Furthermore, if the period derivatives are driven by magnetic dipole radiation, it can be shown (Pacini 1969) that the dipolar magnetic field is given by $`B_p=3.2\times 10^{19}(P\dot{P})^{1/2}`$ G, which would yield dipolar magnetic fields of $`8\times 10^{14}`$ and $`5\times 10^{14}`$ G for SGR $`180620`$ and SGR $`1900+14`$ respectively. Therefore, the SGR pulsars are magnetars, “neutron stars” with magnetic fields $`10^{14}`$ G. Such stars were first proposed by Duncan & Thompson (1992), Usov (1992) and Paczyński (1992).
However, the above estimate of dipolar magnetic fields leads to characteristic spin-down ages much smaller than the SNR ages. This difficulty can be alleviated by introducing relativistic particle outflows from SGRs. The existence of such a wind has been inferred indirectly by X-ray and radio observations of the synchrotron nebula G$`10.00.3`$ around SGR $`180620`$ (Murakami et al. 1994; Kulkarni et al. 1994). Thompson & Duncan (1996) estimated that the particle luminosity from SGR $`180620`$ is of the order of $`10^{37}\mathrm{erg}\mathrm{s}^1`$. Such an energetic wind will also affect the spin-down torque of the pulsar by distorting the dipole field structure near the light cylinder (Thompson & Blaes 1998). Furthermore, Harding, Contopoulos & Kazanas (1999) have found that if SGR $`180620`$ puts out a continuous particle wind of $`10^{37}\mathrm{erg}\mathrm{s}^1`$, then the pulsar age is consistent with that of the surrounding SNR, but the derived surface dipole magnetic field is only $`3\times 10^{13}`$ G, in the range of normal radio pulsars.
It has been widely thought that ultra-strong magnetic fields are an origin of SGR quiescence and burst emission and relativistic particle winds (Thompson & Duncan 1995, 1996). As analyzed in Section 2, however, a magnetic field of $`3\times 10^{13}`$ G may be too weak to be considered as an energy source of SGR quiescence emission and relativistic particle winds. Furthermore, the rotational energy, gravitational energy and crustal strain energy of SGR pulsars are not yet suitable. Following Cheng & Dai (1998) and Dai & Lu (1998), in Section 3 we will propose a model in which SGR pulsars are young, magnetized strange stars with superconducting cores. We argue that this model can provide an explanation for all the observed properties of SGRs including steady winds with luminosities of $`10^{37}\mathrm{erg}\mathrm{s}^1`$. In the final section, we will discuss some differences between anomalous X-ray pulsars (AXPs) and SGRs in our model.
2. PREVIOUSLY SUGGESTED ENERGY SOURCES
Observationally, SGRs have both quiescence and burst emission and steady winds. In the following we estimate the energies of the emission and wind from SGR $`180620`$. First, assuming that the luminosity of the wind is $`L_w10^{37}\mathrm{erg}\mathrm{s}^1`$, we obtain the total observed wind energy $`E_w=L_wt_{\mathrm{SNR}}3\times 10^{48}\mathrm{ergs}`$. Second, the total energy of the persistent X-ray emission is given by $`E_x=L_xt_{\mathrm{SNR}}6\times 10^{46}\mathrm{ergs}`$, where $`L_x`$ is the persistent X-ray luminosity ($`2\times 10^{35}\mathrm{erg}\mathrm{s}^1`$). Third, the total observed energy of SGR bursts, assuming isotropic emission, can be estimated by $`E_{b,\mathrm{tot}}=E_b(t_{\mathrm{SNR}}/\tau _{\mathrm{int}})3\times 10^{46}(E_b/10^{41}\mathrm{ergs})(t_{\mathrm{int}}/10^6\mathrm{s})^1`$, where $`E_b`$ is the typical energy of a burst ($`10^{41}\mathrm{ergs}`$) and $`\tau _{\mathrm{int}}`$ is the interval timescale of SGR bursts ($`10^6`$ s).
Theoretically, there are four energy sources for the persistent X-ray and burst emission and the wind. The first energy source is the rotational energy of the pulsar $`E_{\mathrm{rot}}4\times 10^{44}(P/7.47\mathrm{s})^2\mathrm{ergs}`$. Second, assuming a uniform poloidal field configuration in the interior, the total magnetic energy is $`E_B3\times 10^{44}[B_p/(3\times 10^{13}\mathrm{G})]^2\mathrm{ergs}`$. Moreover, the numerical studies of Heyl & Kulkarni (1998) show that a magnetic field with $`3\times 10^{13}`$ G doesn’t obviously decay even in $`10^6`$ yr, implying that this magnetic energy cannot be varied in the SGR age. The third energy source is the gravitational energy of the pulsar. It is well known that the available gravitational energy for a rotating star ($`\mathrm{\Delta }E_G`$) is only the difference in the gravitational energy between this star and a nonrotating (spherical) star for the same baryon mass. Assuming that the SGR pulsar is a slowly rotating Maclaurin spheroid, we easily demonstrate $`\mathrm{\Delta }E_G=5E_{\mathrm{rot}}2\times 10^{45}(P/7.47\mathrm{s})^2\mathrm{ergs}`$. The final energy source is the crustal strain energy. Assuming that the SGR pulsar is a neutron star, we obtain the strain energy (Baym & Pines 1971): $`E_{\mathrm{strain}}2\times 10^{45}ϵ^4\mathrm{ergs}\mathrm{\Delta }E_G`$, where $`ϵ`$ is the eccentricity of the star.
Comparing the theoretical energy sources with the observed energies of the persistence and burst emission and the wind, we find that the previously suggested energies are much smaller than required by observations. For example, the magnetic energy is about four orders of magnitude smaller than the wind energy. Therefore, we conclude that these energy sources are too weak to be considered as origins of SGRs. Furthermore, even if a magnetar-strength field of $`10^{14}`$ G, as Harding et al. (1999) argued in the case of an episodic wind with small duty cycle, is assumed, this conclusion remains correct. What are energy sources of SGRs?
3. OUR ENERGY SOURCES
We now propose a plausible model for SGRs, in which SGR pulsars are young, magnetized strange stars with superconducting cores. The structure of strange stars has been widely studied (for a recent review see Cheng, Dai & Lu 1998). An interesting possible signature for the existence of strange stars has been found in a few low-mass X-ray binaries (Stergioulas, Kluzniak & Bulik 1999), in which the kHz quasi-periodic oscillation phenomena were recently observed (Zhang et al. 1998). Another strange star candidate is an unusual hard X-ray burster, GRB J1744$``$28 (Cheng et al. 1998).
It is well known that supernova explosions are very likely to produce neutron stars. Because of hypercritical accretion, the neutron stars may subsequently accrete sufficient mass ($`0.5M_{}`$) to convert to massive strange stars (Cheng & Dai 1996; Dai & Lu 1999; Wang et al. 1999). Since the density profile of a strange star is much different from that of a neutron star for the same baryon mass, differential rotation may occur in the interior of the newborn strange star. Dai & Lu (1998) have argued that such a differentially rotating strange star could lead to a classical gamma-ray burst. The basic idea of their argument is: In a differentially rotating strange star, internal poloidal magnetic field will be wound up into a toroidal configuration and linearly amplified as one part of the star rotates about the other part. Only when it increases up to a critical field, $`B_f2\times 10^{17}`$ G, will the toroidal field be sufficiently buoyant to overcome fully the stratification in the composition of the strange star core. And then the buoyant magnetic torus will be able to float up to and break through the stellar surface. Reconnection of the surface magnetic field will produce a quickly explosive event as a peak of a gamma-ray burst. This idea is similar to that of Kluzniak & Ruderman (1998) who discussed the neutron star case. Here we further suggest that after the gamma-ray burst many magnetic toruses with $`B_\varphi <B_f`$ (toroidal field configuration) could remain in the interior of the strange star in a timescale of $`10^4`$ yr.
After its birth, a strange star must start to cool due to neutrino emission. As with a neutron star, the strange star core may become superconducting when its interior temperature is below the critical temperature. Bailin & Love (1984) found that the superconducting transition temperature in strange matter is about 400 keV. Therefore, a strange star with age of $`10^4`$ yr after its supernova birth must have a core temperature much lower than the superconducting transition temperature. The interior temperature of the strange star decreases as $`T10^8(t/\mathrm{yr})^{1/4}`$ K, so $`T10^7`$ K when $`t10^4`$ yr. The quark superconductor is likely to be marginally type-II with zero temperature critical field $`B_c10^{17}`$ G (Bailin & Love 1984; Benvenuto, Vucetich & Horvath 1991; Chau 1997). Furthermore, Chau (1997) argued that after the quark superconductor appears in the strange star, the coupling between quantized vortex lines and (poloidal) magnetic flux tubes in the strange star is so strong that when the vortex lines are moving outward due to spinning down of the star, the magnetic flux tubes are also moving outward with them. According to this argument, Cheng & Dai (1998) proposed a plate tectonic model for strange stars which is, in principle, similar to that proposed by Ruderman (1991) for neutron stars. In this model, when the star spins down due to magnetic dipole radiation and wind emission, the vortex lines move outward and pull the flux tubes with them. However, since the terminations of the flux tubes are anchored in the base of the highly conducting crystalline crust, the flux tubes will produce sufficient tension to crack the crust and pull parts of the broken platelet into the strange quark matter. The time interval between two successive cracking events is estimated to be (Cheng & Dai 1998)
$`\tau _{\mathrm{int}}`$ $``$ $`10^6\left({\displaystyle \frac{B_c}{10^{17}\mathrm{G}}}\right)^1\left({\displaystyle \frac{B_p}{3\times 10^{13}\mathrm{G}}}\right)^1\left({\displaystyle \frac{\theta _s}{0.03}}\right)`$ (1)
$`\times \left({\displaystyle \frac{\mu }{10^{27}\mathrm{dyn}\mathrm{cm}^2}}\right)\left({\displaystyle \frac{l}{10^4\mathrm{cm}}}\right)\left({\displaystyle \frac{R}{10^6\mathrm{cm}}}\right)^1\left({\displaystyle \frac{t_{\mathrm{SNR}}}{10^4\mathrm{yr}}}\right)\mathrm{s},`$
where $`\theta _s`$ and $`\mu `$ are the shear angle (estimated below) and the shear modulus ($`10^{27}`$ $`\mathrm{dyn}\mathrm{cm}^2`$) at the base of the crust of the strange star respectively, and $`l`$ is the crustal thickness ($`10^4`$ cm). The melting temperature of the crust is $`T_m10^3(\rho _b/\mathrm{g}\mathrm{cm}^3)^{1/3}Z^{5/3}\mathrm{K}10^9\mathrm{K}`$, where $`\rho _b`$ is the mass density at the base of the crust ($`4\times 10^{11}\mathrm{g}\mathrm{cm}^3`$) and $`Z`$ is the charge number of nuclei ($`Z=26`$ for iron) (Shapiro & Teukolsky 1983). At age of $`10^4`$ yr, the interior temperature of the strange star $`T10^7\mathrm{K}0.1T_m`$ and thus $`\theta _s10^110^2`$ (Ruderman 1991). We see that the time given by equation (1) is consistent with the typical time interval between SGR bursts.
Because each baryon can release the deconfinement energy of $`30`$ MeV (the accurate value is dependent upon the quantum chromodynamics parameters), the total amount of energy release is estimated as
$$\mathrm{\Delta }E_b3\times 10^{42}\left(\frac{\eta }{0.1}\right)\left(\frac{M_{\mathrm{cr}}}{10^5M_{}}\right)\left(\frac{l}{10^4\mathrm{cm}}\right)^2\left(\frac{R}{10^6\mathrm{cm}}\right)^2\mathrm{ergs},$$
(2)
where $`\eta `$ is the fractional mass in the cracking area $`l^2`$ which is dragged into the core (Cheng & Dai 1998). At least half of this amount will be carried away by thermal photons with the typical energy $`kT30`$ MeV. In the presence of a strong magnetic field ($`3\times 10^{13}`$ G), these thermal photons will convert into electron/positron pairs when $`[E_\gamma /(2m_ec^2)]B\mathrm{sin}\mathrm{\Theta }/B_q1/15`$, where $`E_\gamma `$ is the photon energy, $`B_q=m_e^2c^3/(\mathrm{}e)=4.4\times 10^{13}`$ G, and $`\mathrm{\Theta }`$ is the angle between the photon propagation direction and the direction of the magnetic field (Ruderman & Sutherland 1975). The energies of the resulting pairs will be lost via synchrotron radiation. The characteristic synchrotron energy is given by $`E_{\mathrm{syn}}1.5\gamma _e^2\mathrm{}eB\mathrm{sin}\mathrm{\Theta }/(m_ec)3.0\mathrm{MeV}`$, where $`\gamma _e`$ is the Lorentz factor of the pairs ($`30`$). These synchrotron photons will be converted into secondary pairs because the optical depth for photon-photon pair production is much larger than one. The Lorentz factor of the secondary pairs is about 3.0. Thus, we obtain a cooling distribution of mildly relativistic pairs, whose self-absorbed synchrotron emission has been shown to provide excellent fits to the spectral data of SGR bursts (Liang & Fenimore 1995). In addition, after the cracking event, roughly half of the resulting thermal energy from deconfinement of normal matter to strange quark matter will be absorbed by the stellar core, and thus the surface radiation luminosity at thermal equilibrium has been found to be consistent with the observed persistent X-ray luminosity (Cheng & Dai 1998).
As the quantized vortex lines move outward during the stellar spinning down and pull the magnetic flux tubes together, the magnetic toruses are also pulled toward the equatorial region with the flux tubes due to the interaction between them (Chau, Cheng & Ding 1992). The upper limit of $`B_\varphi `$ is $`B_f`$; its lower limit can be given as follows. The magnetic toruses must are sufficiently buoyant to overcome fully the stratification in the composition of the crust, requiring $`B_\varphi (8\pi \rho _bc_s^2)^{1/2}5\times 10^{15}`$ G, where $`c_s`$ is the speed of sound of the crust ($`c_s2\times 10^9\mathrm{cm}\mathrm{s}^1`$). The density of the flux tubes is given by $`n=B_p/\mathrm{\Phi }_v4\times 10^{20}\mathrm{cm}^2`$, where $`\mathrm{\Phi }_v`$ is the magnetic flux of each tube ($`8\times 10^8\mathrm{G}\mathrm{cm}^2`$) (Chau 1997). Now we define a timescale: $`\mathrm{\Delta }t=1/(v\sqrt{n})2\times 10^5\mathrm{s}`$, where $`v`$ is the speed of flux tubes ($`3\times 10^6\mathrm{cm}\mathrm{s}^1`$) (Cheng & Dai 1998). The physical meaning for $`\mathrm{\Delta }t`$ is that there must be a flux tube to move toward to the surface in the equatorial region in this timescale, implying that there must be magnetic toruses which are pulled simultaneously to the surface in the equatorial region and which break the surface. The reconnection of magnetic toruses leads to an episodic wind. Because $`\mathrm{\Delta }t`$ is much smaller than $`\tau _{\mathrm{int}}`$, too many episodic winds will constitute a steady wind. The magnetic energy including in each episodic wind can be estimated as $`e_w=V_bB_\varphi ^2/(8\pi )`$ where $`V_b`$ is the volume of a torus. It is important to note that the magnetic torus is emerging from the quark matter core to the surface and subsequently releasing this energy via reconnection in the equatorial region. Such an energy must be contaminated by crustal baryons whose mass is at most $`m_wM_{\mathrm{cr}}V_b/(4\pi R^2l)`$. In fact, the mass of the contaminating baryons should be a fraction ($`\xi `$) of $`m_w`$. Thus, the Lorentz factor of the steady wind is
$$\mathrm{\Gamma }\frac{e_w}{\xi m_wc^2}30\left(\frac{\xi }{0.1}\right)^1\left(\frac{M_{\mathrm{cr}}}{10^5M_{}}\right)^1\left(\frac{B_\varphi }{10^{17}\mathrm{G}}\right)^2\left(\frac{R}{10^6\mathrm{cm}}\right)^2\left(\frac{l}{10^4\mathrm{cm}}\right).$$
(3)
The luminosity of the wind can be estimated as
$$L_w\frac{B_\varphi ^2}{8\pi }\frac{\chi R^3}{t_{\mathrm{SNR}}}10^{37}\left(\frac{\chi }{10^2}\right)\left(\frac{B_\varphi }{10^{17}\mathrm{G}}\right)^2\left(\frac{t_{\mathrm{SNR}}}{10^4\mathrm{yr}}\right)^1\mathrm{erg}\mathrm{s}^1,$$
(4)
where $`\chi R^3`$ is the volume of all the flux tubes ($`\chi R`$ is the characteristic length of the velocity gradient and is assumed to be $`l`$). This relativistic wind will inject into and power the surrounding SNR.
4. DISCUSSION
Quiescence and burst emission and relativistic particle winds of SGRs have been widely interpreted to result from ultrastrongly magnetized neutron stars (Thompson & Duncan 1995, 1996). In such a magnetar model, the SGR bursts are due to readjustment of the magnetic field, possibly accompanied by cracking of the neutron-star crust, and the persistent X-ray emission is due to decay of the magnetic field, while the winds are due to thermal radiation from hot spots and Alfven wave emission (Thompson & Blaes 1998). It is very clear that the energy sources of all the emission and winds are the magnetic energy and gravitational energy of the neutron star. However, Harding et al. (1999) have shown that the magnetic field must be up to $`3\times 10^{13}`$ G in order to match the characteristic spin-down timescale for SGR $`180620`$ and its surrounding SNR age. Here we have argued that if the magnetic field is indeed so low, the previously suggested energy sources seem problematic because they are much smaller than the observed energies for the persistent X-ray and burst emission and the wind. We have further proposed another plausible model in which SGR pulsars are young strange stars with superconducting cores and with magnetic fields of $`3\times 10^{13}`$ G following Cheng & Dai (1998) and Dai & Lu (1998). In our model, the movement of the flux tubes not only leads to crustal cracking, giving rise to deconfinement of crustal matter to strange matter, but also to movement of internal magnetic toruses with the flux tubes. As have shown in the above section, the former process will result in burst and quiescence emission and the latter process will produce relativistic winds which will power the surrounding SNR.
Another group of sources having periods and period derivatives similar to SGRs are the AXPs, pulsating X-ray sources with periods in the range $`612`$ s and period derivatives in the range of $`10^{12}10^{11}\mathrm{s}\mathrm{s}^1`$ (Gotthelf & Vasisht 1998). These sources have shown only strong quiescent X-ray emission with no bursting behavior. Moreover, if the souces are highly magnetic pulsars, their characteristic ages ($`P/2\dot{P}`$) are in excellent agreement with the SNR ages, implying that the sources have no wind emission. Why are there such obvious differences between SGRs and AXPs? We suggest that AXPs be neutron stars with magnetic fields of $`10^{15}`$ G but without any toroidal magnetic field. In the case of slowly rotating neutron stars, crustal cracking cannot produce an observed burst because of too low gravitational energy release, and there is no wind in the absence of toroidal magnetic fields.
This work was supported by a RGC grant of Hong Kong government and the National Natural Science Foundation of China.
REFERENCES
Bailin, B., & Love, A. 1984, Phys. Rep., 107, 325
Baym, G., & Pines, D. 1971, Ann. Phys., 66, 816
Benvenuto, O. G., Vucetich, H., & Horvath, J. E. 1991, Nucl. Phys. B, 24, 160
Chau, H. F. 1997, ApJ, 479, 886
Chau, H. F., Cheng, K. S., & Ding, K. Y. 1992, ApJ, 399, 213
Cheng, K. S., & Dai, Z. G. 1996, Phys. Rev. Lett., 77, 1210
Cheng, K. S., & Dai, Z. G. 1998, Phys. Rev. Lett., 80, 18
Cheng, K. S., Dai, Z. G., & Lu, T. 1998, Int. J. Mod. Phys. D, 7, 139
Cheng, K. S., Dai, Z. G., Wei, D. M., & Lu, T. 1998, Science, 280, 407
Cline, T. L. et al. 1982, ApJ, 255, L45
Dai, Z. G., & Lu, T. 1998, Phys. Rev. Lett., 81, 4301
Dai, Z. G., & Lu, T. 1999, ApJ, 519, L155
Duncan, R. C., & Thompson, C. 1992, ApJ, 392, L2
Evans, W. D. et al. 1980, ApJ, 237, L7
Fenimore, E. E., Laros, J. G., & Ulmer, A. 1994, ApJ, 432, 742
Gotthelf, E. V., & Vasisht, G. 1998, New Astron., 3, 293
Harding, A. K., Contopoulos, I., & Kazanas, D. 1999, ApJ, 525, L125
Heyl, J. S., & Kulkarni, S. R. 1998, ApJ, 506, L61
Hurley, K. et al. 1999, ApJ, 519, L143
Laros, J. G. et al. 1987, ApJ, 320, L111
Liang, E., & Fenimore, E. E. 1995, ApJ, 451, L57
Kluzniak, W., & Ruderman, M. A. 1998, ApJ, 505, L113
Kouveliotou, C. et al. 1994, Nature, 368, 125
Kouveliotou, C. et al. 1998, Nature, 393, 235
Kouveliotou, C. et al. 1999, ApJ, 510, L115
Kulkarni, S. R., & Frail, D. A. 1993, Nature, 365, 33
Kulkarni, S. R. et al. 1994, Nature, 368, 129
Murakami, T. et al. 1994, Nature, 368, 127
Pacini, F. 1969, Nature, 221, 624
Paczynski, B. 1992, Acta Phys., 41, 145
Rothschild, R. E., Kulkarni, S. R., & Lingenfelter, R. E. 1994, Nature, 368, 432
Ruderman, M. A. 1991, ApJ, 366, 261
Ruderman, M. A., & Sutherland, P. G. 1975, ApJ, 196, 51
Shapiro, S. L., & Teukolsky, S. A. 1983, Black Holes, White Dwarfs and Neutron Stars (John Wiley & Sons), P. 90
Stergioulas, N., Kluzniak, W., & Bulik, T. 1999, A&A, in press (astro-ph/9909152)
Thompson, C., & Blaes, O. 1998, Phys. Rev. D, 57, 3219
Thompson, C., & Duncan, R. C. 1995, MNRAS, 275, 255
Thompson, C., & Duncan, R. C. 1996, ApJ, 473, 322
Usov, V. V. 1992, Nature, 357, 452
Vasisht, G., Kulkarni, S. R., Frail, D. A., & Greiner, J. 1994, ApJ, 431, L35
Wang, X. Y., Dai, Z. G., Lu, T., Wei, D. M., & Huang, Y. F. 1999, astro-ph/9910029
Woods, P. M. et al. 1999, ApJ, 519, L139
Zhang, W. et a. 1998, ApJ, 500, L171 |
no-problem/9911/astro-ph9911517.html | ar5iv | text | # Galactic Positron Production from Supernovae
## Abstract
ABSTRACT The energy deposition into the ejecta of type Ia supernovae is dominated at late times by the slowing of positrons produced in the $`\mathrm{\beta }^+`$ decays of <sup>56</sup>Co. Fits of model-generated light curves to observations of type Ia supernovae suggest that a significant number of positrons escape the ejecta and annihilate as a delayed emission. In this work, the isotopic yields of the $`\mathrm{\beta }^+`$ decay unstable nuclei, <sup>56</sup>Co, <sup>44</sup>Ti & <sup>26</sup>Al are combined with the delayed annihilation fractions of each isotope for types Ia, Ib, II supernovae to generate an estimate of the positron production rate due to supernovae. It is shown that SN produced positrons can explain a sizeable fraction of the galactic positron annihilation radiation, as measured at 511 keV by the CGRO/OSSE, SMM and TGRS instruments.
1 NRC/NRL Resident Research Associate, Naval Research Lab, Code 7650, Washington DC 20375
2 Clemson University, Clemson, SC 29631
1. INTRODUCTION
Supernovae (SNe), both thermonuclear and core collapse, have been suggested to be sources of galactic positrons through the $`\beta ^+`$ decays of <sup>56</sup>Co, <sup>44</sup>Sc & <sup>26</sup>Al (Clayton 1973). The annihilation of these positrons can occur promptly (defined here as within a decade of the SN event) or can be delayed (defined here as $``$10<sup>3</sup> years or more after the SN event). The annihilation of interest in this paper is the delayed emission where the integrated contributions from many SNe combine to produce a diffuse emission. The three isotopes, <sup>56</sup>Co, <sup>44</sup>Sc & <sup>26</sup>Al, have very different mean lifetimes and thus have different factors that determine what fraction of the total decay positrons annihilate as a delayed emission. In this paper we define the parameters relevant to quantifying the SN contribution to galactic positrons. We then estimate these parameters, comparing our values with previous estimates, and emphasizing which estimates have observational support. In particular, we show fits to the late light curves of type Ia SNe which suggest that a significant fraction of <sup>56</sup>Co positrons escape the SN ejecta to contribute a sizeable portion of the observed annihilation radiation.
Implicit in this entire study is the assumption that all SNe are approximated by the types Ia, Ib & II and that each type is homogeneous to the extent that a handful of models are adequate to describe the positron yield of the entire class. It is further assumed that the explosion kinematics and nucleosynthesis of current SN modeling is accurate enough that parameter ranges shown truly span the range of solutions. The goal of this paper is to estimate the SN contribution to galactic positron annihilation. To perform this estimate, SN rates used in this study are infered from other galaxies. It is possible that the Galaxy has anomalous SN rates (as would occur if there has been a starburst at the galactic center within the last 10<sup>6</sup> years) or that there is leakage of positrons out of the Galaxy. Future 511 keV and 1.8 MeV maps (as well as maps of other nuclear decay lines) will be used invert this problem and determine the recent galactic SN history. The current study is a consistency check as much as it is a solution to galactic positron production.
2. POSITRON YIELD PARAMETERS
The positron production rate per isotope, per SN type, is the product of the isotopic yield times the $`\beta ^+`$ decay branching ratio, times the fraction of decay positrons which annihilate as a delayed emission, times the SN rate of that SN type. This product is then summed by isotope and by SN type to arrive at the total positron production rate. Assuming that the only relevant isotopes are <sup>56</sup>Co, <sup>44</sup>Sc & <sup>26</sup>Al, and that three SN types must be considered (types Ia, Ib & II), quantifying the SN positron production rate reduces to 9 isotopic yields, 9 delayed annihilation fractions, 3 branching ratios & 3 SN rates. The branching ratios and other $`\beta ^+`$ decay parameters for the three decays are relatively well-known and are shown in Table 1. The most complete estimates of the other 21 parameters was performed in a paper by Chan & Lingenfelter (1993) (hereafter CL). They simulated positron transport through SN models of all three types arriving at a number of conclusions. (1) For the $`\beta ^+`$ decay of <sup>26</sup>Al, the mean lifetime is long-enough that <sup>26</sup>Al is a delayed emission source independent of positron transport physics. For <sup>26</sup>Al, the delayed fraction can be set equal to 1 for all three SN types. (2) The independent observation of 1.8 MeV line emission due to the de-excitation of <sup>26</sup>Mg produced in the <sup>26</sup>Al $``$ <sup>26</sup>Mg decay can estimate the <sup>26</sup>Al contribution (though not specifically the SN-produced <sup>26</sup>Al contribution) to galactic positron annihilation. (3) For <sup>56</sup>Co & <sup>44</sup>Sc decay positrons, the alignment and/or strength of the magnetic field must be considered when estimating the delayed annihilation fraction. For <sup>44</sup>Sc, the effect of the magnetic field varies from a $``$1% effect (type Ia), to a 66% effect (type II). For <sup>56</sup>Co in type Ia SNe, the delayed annihilation fraction changes even more dramatically. (4) If the magnetic field characteristics are favorable to positron escape, then the decay of <sup>56</sup>Co produced in type Ia SNe may be the dominant contributor of galactic positrons. If not, then the dominant contributor of galactic positrons is from <sup>44</sup>Sc decays.
The critical parameter is thus the delayed annihilation fraction for <sup>56</sup>Co decays in type Ia SNe, which depends upon the magnetic field characteristics of the SN ejecta. CL assumed that the field strength is strong enough to confine positrons to the field lines, and estimated the positron survival for two opposite geometries. The first geometry assumes that the field is frozen into the ejecta and is combed to become essentially radial as the ejecta expands. Positrons spiral along the field lines experiencing mirroring/beaming, with a fraction of the positrons escaping the SN ejecta. The escaping positrons then enter a lower density medium and annihilate on longer timescales. This scenario will be refered to as the “radial” scenario ($`l`$=$`\mathrm{}`$ in CL terminology). The second geometry assumes that the field remains turbulent throughout the expansion of the SN ejecta. Positrons are trapped in the same location (in mass coordinates) as they are emitted. This scenario is refered to as the “trapping” scenario ($`l`$=0 in CL terminology), and the fraction of positrons that survive 10<sup>3</sup> years in the dense ejecta is much lower than the escape fraction in the radial scenario. Colgate et al. (1980) argued for a third, “weak field” scenario, where the field strength is inadequate to confine the positrons. In this scenario the positrons follow photon-like, straight line trajectories. Simulations have shown that the escape fractions and energy deposition rates for this scenario are approximately equal to the radial scenario (a result suggested by Colgate et al.). Throughout the remainder of this paper, only the terms radial and trapping will be used, but it is implied that radial represents radial or weak field solutions.
3. SN Ia LIGHT CURVES & POSITRON ESCAPE
CL calculated escape fractions from type Ia SNe for radial and trapping field geometries, but did not attempt to determine which scenario occurs in nature. They refered to the opposing conclusions of Colgate et al. (1980) and Axelrod (1980). Colgate et al. showed that energy deposition rates featuring positron escape could suitably explain the B band light curves of SN 1937C & SN 1972E. Axelrod argued for an alternative explanation, that no positron escape was evidenced, rather that the apparent deficit in the luminosity (relative to 100% positron trapping) was due to emission of an increased fraction of the deposited energy in unobservable infrared energy bands. This phenomenon was refered to as an “infrared catastrophe” and has since been observed in the core collapse supernova, SN 1987A. Recent papers have been only somewhat successful in clarifying the picture. Cappellaro et al. (1997) and Ruiz-Lapuente et al. (1997) fit model-generated bolometric light curves to observations, both studies concluding that positron escape is favored in some, but not all cases. Fransson et al. (1996), after including a network of reaction rates, simulated multi-band light curves for the SN Ia model, DD4 (assuming positron trapping). They generated light curves which featured an “infrared catastrophe”, but which are in conflict with the light curves of SN 1972E (and all other late observations of SN Ia). None of these papers performed a detailed treatment of the positron transport, adapting photon transport codes instead.
Milne et al. (1999) explicitly treated both the gamma-photon and positron transport, assuming radial and trapping field geometries, and allowing for a range of ionizations (ranging from 1% ionization to triple ionization). A characteristic model-generated bolometric light curve of the model W7 is shown in Figure 1. Before $``$200<sup>d</sup>, positron lifetimes are short for all field and ionization assumptions, the light curves are identical with the zero lifetime (or instantaneous) light curve (In). With time, the SN ejecta expands and rarefies, increasing the positron lifetimes. In the radial scenario (R), the expansion leads to positron escape, and thus a deficit of energy deposition. As free electrons are more efficient at slowing positrons than are bound electrons, the low 1% ionization light curve features the maximum deficit from instantaneous annihilation. In the trapping scenario (T), the expansion permits non-zero lifetimes, but the majority of the energy is later deposited. At late times, the deposition of this stored energy makes the trapping light curves brighter than either the radial or the instantaneous light curves. For low 1% ionization, longer positron lifetimes lead to more energy being stored. The delayed deposition of this stored energy leads to brighter light curves at later times for low ionization solutions.
Bolometric light curves were generated for 22 SN Ia models. These models were selected to span Chandrasekhar, sub-Chandrasekhar and merger scenarios. The separation between radial and trapping light curves was found to exceed either the effects of model type or the level of ionization. These model-generated light curves were then fitted to the 10 SN Ia best observed at late times.<sup>1</sup><sup>1</sup>1Discussions regarding the validity of fitting model-generated energy deposition rates to observed band photometry and/or “uvoir” bolometric light curves are given in Milne et al. (1999). Of these 10 SNe, eight were considered either normal- or super-luminous. Of these 8, seven were suitably fitted with positron escape according to the radial field scenario. Five of the seven were better fitted with positron escape than with positron trapping. None were better fitted with positron trapping. The four best examples are shown in Figure 2, all 10 SNe are shown in Milne et al. (1999). The preference of positron escape is apparent in SNe 1992A (both bolometric (not shown) and V band, fit with DD23C (Hőflich et al. 1998)) & 1990N (fit with W7DN (Yamaoka et al. 1992)) without further explanation. For SN 1937C (fit with DET2 (Hőflich, Khokhlov, & Wheeler 1995)), B band data was used, thus fitting began after 120<sup>d</sup> at which time color evolution has been empirically determined to cease in SN Ia. For SN 1991T (fit with HECD (Kumagai 1997)), the very late emission is dominated by a light echo, discovered spectrally by Schmidt et al. (1996) and imaged by HST (Boffi et al. 1998). Though the light curves for the range of ionizations (filled curves) are shown with a constant luminosity light echo included, the dashed and dot-dashed lines show the low ionization radial and trapping curves with no light echo. It is apparent that the trapping curves are already too bright without the additional emission from an echo, the radial curve suitably explains the data before 480 days. This suggests a light echo that “turns on” after 480 days, which would occur if the light echo is due to the peak light sweeping through an off-axis cloud. The asymmetry of the HST image would appear to permit that interpretation.
The conclusion of Milne et al. (1999) is that the late light curves of normal- and super-luminous type Ia SNe can be suitably explained with positron escape. Though the sample is neither large enough, nor well-enough observed to rule out trapping in all cases, the lack of counter-examples suggest escape to be the dominant result. Many of the SNe used in that study were also used in the studies of Cappellaro et al. (1997), Ruiz-Lapuente et al. (1997) & Fransson et al. (1996). Milne et al. (1999) did not find positron trapping to be favored for any SN, in mild disagreement with some of the conclusions of these earlier studies. Based upon these observations, the positron escape fractions claimed in Milne et al. (1999) will be inserted into the <sup>56</sup>Co-SN Ia portion of the positron production rate calculations rather than the trapping-radial range suggested by CL. The determination of the escape of <sup>56</sup>Co positrons from type Ia SNe was based upon the existence of a “positron phase”, an epoch during which the energy deposition is dominated by the slowing of positrons. This “positron phase” does not occur in core collapse SNe, due to the more efficient trapping of gamma-photons. The dominance of <sup>56</sup>Co decay photons transitions to the dominance of <sup>57</sup>Co decay photons without an epoch of positron dominance. Since no determination of the field characteristics of types Ib and II SNe are possible, the trapping-radial ranges claimed by CL are used in section 5.
4. ISOTOPIC YIELDS
The SN Ia delayed annihilation fraction may prove to be the most critical parameter for quantifying positron production from SNe, but twenty other parameters must also be estimated. The other eight delayed annihilation fractions were adequately approximated by CL, the nine isotopic yields and three SN rates remain undetermined.
The isotopic yields of <sup>56</sup>Ni (the parent isotope of <sup>56</sup>Co) are the best constrained. In all three SN types, energy deposition from the <sup>56</sup>Ni and subsequent <sup>56</sup>Co decays power the light curves. This allows fits of model-generated bolometric light curves to observations, combined with distance estimates to the host galaxy, to constrain the <sup>56</sup>Ni yields. Type Ia SNe produce the most <sup>56</sup>Ni per SN event, ranging from 0.3 -0.9 M in various SN Ia models. Diehl (1997) suggests the typical <sup>56</sup>Ni yield to be 0.5 M. As shown by Milne et al. (1999), the larger nickel yield in Chandrasekhar mass models partially compensates for the lower escape fraction (relative to sub-Chandrasekhar mass models), leading to the positron yields being virtually independent of the progenitor mass. The SN Ia <sup>56</sup>Ni yield estimate is also influenced by the existence of super- and sub-luminous classes. An additional result of Milne et al. (1999) is the acceptance of both Chandrasekhar mass and sub-Chandrasekhar mass explanations for normally- and super-luminous SNe Ia, but rejection of all currently proposed models for sub-luminous SNe Ia. CL discussed a “Ip” model class as an explanation for sub-luminous SNe Ia. The Ip model class featured large positron yields. This work will not include that class due to the findings of Milne et al. (1999).
The <sup>56</sup>Ni yields from type II SNe are well-constrained due to the observations of the nearby SN 1987A. That SN produced 0.08 M of <sup>56</sup>Ni, the fiducial value used by Dermer & Skibo (1997). The type Ib <sup>56</sup>Ni yield is taken from CL, ranging from 0.08 -0.28 M. Diehl (1997) estimates the type II/Ib <sup>56</sup>Ni yield to be 0.1 M. The type II/Ib <sup>56</sup>Ni yields are of little importance due to the low delayed annihilation fractions for these SNe.
The isotopic yields of <sup>44</sup>Ti are obtained from the outputs of SN models (with fewer observational constraints) and are poorly constrained. Timmes et al. (1996) show a range of <sup>44</sup>Ti yields for core collapse models, claiming 3 x 10<sup>-5</sup> M (type II) & 6 x 10<sup>-5</sup> M (type Ib) to be typical values, but later tripling these yields in a positron production calculation. Nomoto (1997) suggests that type II SNe produce 5 x 10<sup>-5</sup> M of <sup>44</sup>Ti, in agreement with the value used by Diehl (1997) for types II/Ib. Meyer et al. (1995) estimates the <sup>44</sup>Ti yield from a 25 M SN to be 1.5 x 10<sup>-4</sup> M. The <sup>44</sup>Ti $``$ <sup>44</sup>Sc $``$ <sup>44</sup>Ca decay has been directly observed in the Cas A SNR (type II or Ib) via the 1.157 MeV de-excitation line of <sup>44</sup>Ca and the 68 keV and 78 keV de-excitation lines of <sup>44</sup>Sc. From measurements taken by the CGRO/COMPTEL, the CGRO/OSSE and the Rossi X-Ray Timing Explorer High Energy X-Ray Timing Experiment, the <sup>44</sup>Ti yield is estimated to be 2.2 x 10<sup>-4</sup> M (The et al. 1998, using the 85<sup>y</sup> <sup>44</sup>Ti mean lifetime from Ahmad et al. 1997). The SN Ia <sup>44</sup>Ti yields are equally uncertain. The Chandrasekhar mass model, W7 (Nomoto, Theilemann & Yokoi 1984), produces $``$1.8 x 10<sup>-5</sup> M of <sup>44</sup>Ti. The low mass portion of sub-Chandrasekhar mass models produce more <sup>44</sup>Ti, ranging from (2-40) x 10<sup>-4</sup> M (Woosley & Weaver 1994), though these low mass examples cannot explain 100% of SN Ia events due to nucleosynthesis considerations. Dermer & Skibo (1997) use the CL results, scaling the <sup>44</sup>Ti yield with the <sup>56</sup>Ni yields. This leads to (3-8) x 10<sup>-5</sup> M, 1.7 x 10<sup>-4</sup> M & $``$ 2.0 x 10<sup>-4</sup> M per SN for types Ia, Ib, II respectively.
The isotopic yield of <sup>26</sup>Al in type Ia SNe is very low ($``$ 10<sup>-6</sup> -Diehl 1997). In core collapse SNe, estimates of the yields range from 2 x 10<sup>-4</sup> M (for SN II/Ib -Diehl 1997) to (0.3 -20) x 10<sup>-5</sup> M (for SN II -Prantzos 1996). Timmes & Woosley (1997) suggest the mass function averaged <sup>26</sup>Al yield per core collapse SN event is 7.7 x 10<sup>-5</sup> M.
4. SUPERNOVA RATES
Estimation of SN rates in the Galaxy is dependent upon the collective SN rates from other spiral galaxies.<sup>2</sup><sup>2</sup>2Estimates of the galactic SN rate from the historical record, from surveys of supernova remnants, and from galactic nucleosynthesis are considered to be less reliable than assuming the Galaxy to be an Sb galaxy and assuming that the SN rates scale with L<sup>B</sup>. The SN rates from SN surveys are given in units of 10<sup>10</sup> L$`{}_{}{}^{B}{}_{}{}^{}`$ and scaled to 2.3 x 10<sup>10</sup> L$`{}_{}{}^{B}{}_{}{}^{}`$. The differences between suggested rates is due to different algorithms accounting for selection effects. The SN Ia rate is relatively well-constrained, estimates are 0.4$`\pm `$0.1 (Cappellaro et al. 1997), 0.28 (Tammann et al. 1994), 0.5 (Hatano et al. 1997). The SN Ib rate is approximately equal to the SN Ia rate, but the estimates are more varied. Estimates are 0.2$`\pm `$0.1 (Cappellaro et al. 1997), 0.8 (Hatano et al. 1997), 0.3 (Tammann et al. 1994). The SN II rate is the largest of the three and the most varied. Estimates are 1.2$`\pm `$0.6 (Cappellaro et al. 1997), 3.8 (Hatano et al. 1997), 1.5 (Tammann et al. 1994).
5. POSITRON PRODUCTION RATES
Shown in Table 2 are the parameters relevant to estimating the galactic positron production rate from SNe. In many cases, the most favorable positron yields per SN event, as shown in column (6), are realized only by anomalous SNe, and would have large recurrence times. As seen in columns (6) and (7), SN Ia <sup>56</sup>Co decays yield the most positrons per SN and per second. The <sup>56</sup>Co decay contributions from SN Ib & II are zero due to rejecting the 100% mixing scenario allowed by CL. The <sup>44</sup>Ti decay contributions (per second) are not negligible. The favorable <sup>44</sup>Ti yields cannot be claimed for all three SN types simultaneously, as that would lead to anomalously large <sup>44</sup>Ca/<sup>56</sup>Fe production. The largest <sup>44</sup>Ti decay contribution would result if a considerable fraction of all SN Ia events are low-mass sub-Chandrasekhar explosions, and core collapse SNe possess radial (or weak) magnetic fields. The <sup>26</sup>Al decay contribution is dominated by core collapse SNe. The positron yield per second is better estimated by including results of measurements of the 1.8 MeV line emission. In 82% of <sup>26</sup>Al $``$ <sup>26</sup>Mg decays, both a positron and a 1.8 MeV line photon are produced. Core collapse SNe have been suggested to account for as much as 100% of galactic <sup>26</sup>Al (Timmes et al. 1997). Assuming the entire emission to emanate from core collapse SNe at the distance of the galactic center ($``$ 8 kpc distant), reported 1.8 MeV central radian fluxes of 3 x 10<sup>-4</sup> ph cm<sup>-2</sup> s<sup>-1</sup> (Knodlseder et al. 1999) translates to 1.9 x 10<sup>42</sup> e+ s<sup>-1</sup>, approximately equal to the <sup>44</sup>Ti decay production rate.
The total positron production rates are 1.0 x 10<sup>43</sup> e+ s<sup>-1</sup> for <sup>56</sup>Co decays, 2.4 x 10<sup>42</sup> e+ s<sup>-1</sup> for <sup>44</sup>Ti decays, and 1.9 x 10<sup>42</sup> e+ s<sup>-1</sup> for <sup>26</sup>Al decays. These values are in general agreement with estimates by Timmes et al. (1996), who derived the values 1.6 x 10<sup>43</sup> e+ s<sup>-1</sup>, 4.5 x 10<sup>42</sup> e+ s<sup>-1</sup>, and 2.9 x 10<sup>42</sup> e+ s<sup>-1</sup> respectively. Notably different is the <sup>44</sup>Ti yield. In that study, it was erroneously suggested that two positrons are generated in 95% of <sup>44</sup>Ti $``$ <sup>44</sup>Sc $``$ <sup>44</sup>Ca decays (a single positron is generated). That study also tripled the <sup>44</sup>Ti yields in all three SN types to permit their chemical evolution model to match the solar <sup>44</sup>Ca abundance, but used the lower delayed annihilation fraction (34% rather than the value 99% used in this study). The net difference is that the <sup>44</sup>Ti yields in Timmes et al. (1996) are roughly double the values claimed here. The difference in <sup>26</sup>Al decay yield may be due to a different 1.8 MeV total flux.
Assuming a constant SN rate over the last 10<sup>6</sup> years, steady-state positron production/annihilation and no leakage of positrons from the Galaxy, the above production rate can be compared with the 511 keV line flux. A positron production rate of 1.4 x 10<sup>43</sup> e+ s<sup>-1</sup> would generate a 511 keV flux of 1.1 x 10<sup>-3</sup> ph cm<sup>-2</sup> s<sup>-1</sup> if emitted from the galactic center ($``$ 8 kpc distant) and annihilating with a positronium fraction of 0.95. That is about 1/3 - 1/2 of the total 511 keV flux suggested by CGRO/OSSE, SMM & TGRS measurements (Milne et al. 1999b). Correcting this value to a distributed emission has not been performed.<sup>3</sup><sup>3</sup>3Timmes et al. (1997) quoted a positron production rate of 1.5 x 10<sup>-3</sup> e+ s<sup>-1</sup> from 511 keV measurements, that value is too low by a considerable fraction.
6. DISCUSSION
The positrons produced in the $`\beta ^+`$ decays of <sup>56</sup>Co, <sup>44</sup>Ti & <sup>26</sup>Al have been shown to potentially contribute a large fraction of the total galactic positron production rate. This work shows observational support for the the suggestion made by Chan & Lingenfelter (1993) that significant numbers of <sup>56</sup>Co decay positrons can escape from the ejecta of type Ia SNe. Observations of supernovae and supernova remnants, combined with improved nucleosynthesis modeling has led to improved constraints upon the various parameters that determine SN positron production.
A future step in modeling the galactic positron production rate from SNe will be to compare spatial distributions with CGRO/OSSE measurements of 511 keV positron-electron annihilation radiation. Type II/Ib SNe are seen only in the disks of spiral galaxies, type Ia SNe occur in both bulges and disks. More progress must be made towards understanding both the intrinsic B/D ratios of type Ia SNe in spiral galaxies, and the B/D ratio measured in the 511 keV emission before useful comparisons can be made.
An observation that would solidify the picture of SN positron production would be the detection of 511 keV emission from nearby SNe. In type Ia SNe, detection of emission above the level expected for <sup>44</sup>Ti decays would be evidence of positron escape. In all SN types, the intensity and distribution of the emission would trace the diffusion of positrons from the remnant into the Galaxy. Additionally, the detection of prompt positron annihilation in SNe would better constrain isotopic yields and delayed annihilation fractions. The late light curves of SNe Ia suggest positron escape, but more observations are required to accurately quantify the positron yields. If all these observations are made, the hypothesis of SN production of galactic positrons will be demonstrated and positron annihilation radiation measurements can be used to probe the recent SN history in the Galaxy.
REFERENCES
Ahmad, I. et al. 1997, Phys. Rev. Lett., 80, 2550
Axelrod, T.S. 1980, Ph.D.thesis, Univ. California at Santa Cruz
Boffi, F.R. et al. 1998, BAAS, 192, 6.07
Cappellaro, E., et al. 1997a, A & A, 322, 431
Cappellaro, E., et al. 1997b, A & A, 328, 203
Chan, K.-W., Lingenfelter, R. 1993, ApJ, 405, 614
Clayton, D.D. 1973, Nature, 244:139, 137
Colgate, S., Petschek, A.G., Kreise, J.T. 1980, ApJ, 237, L81
Dermer, C.D., Skibo, J.G. 1997, ApJ, 487, L57
Diehl, R. 1997, in Proceedings of the 2nd INTEGRAL Workshop, 9
Fransson, C., Houck, J., Kozma, C. 1996, in IAU Colloq. 145, Supernovae and Supernova Remnants, ed. R. McCray & Z. Wang (Cambridge: Cambridge University Press), 41
Hatano, K., Fisher, A., Branch, D. 1997, 290, 360
Hőflich, P., Khokhlov, A., Wheeler, J.C. 1995, ApJ, 444, 831
Hőflich, P., Wheeler, J.C., Theilemann, F.-K. 1998, ApJ, 495, 617
Knodlseder, J., et al. 1999, A & A, 344, 68
Kumagai, S. 1997, private communication
Lira, P. et al. 1998, AJ, 115, 234
Meyer, B.S., Weaver, T.A., Woosley, S.E. 1995, Meteoritics, 30, 325
Milne, P.A., The, L.-S., Leising, M.D. 1999, ApJS, 124, 503
Milne, P.A., et al. 1999b, these Proceedings, ,
Nomoto, K., Theilemann, F.-K., Yokoi, K. 1984, ApJ, 286, 644
Nomoto, K., et al. 1997, Nuclear Physics A, 621, 467
Prantzos, N., Diehl, R. 1996, Phys. Rep., 267, 1
Ruiz-Lapuente, P., Spruit, H. 1997, ApJ, 500, 360
Schaefer, B.E. 1994, ApJ, 426, 493
Schmidt, B.P. et al. 1994, ApJ, 434, L19
Sunzteff, N.B. 1996, in IAU Colloq. 145, Supernovae and Supernova Remnants, ed. R. McCray & Z. Wang (Cambridge: Cambridge University Press), 41
Tammann, G.A., Lőffler, W., Schrőder, A. 1994, ApJS, 92, 487
The, L.-S., et al. 1998, ApJ, 504, 500
Timmes, F.X., et al. 1996, ApJ, 464, 322
Timmes, F.X., Diehl, R., Hartmann, D.H. 1997, ApJ, 479, 760
Timmes, F.X., Woosley, S.E. 1997, ApJ, 481, L81
Yamaoka, H., et al. 1992, ApJ, 393, L55 |
no-problem/9911/math9911078.html | ar5iv | text | # Secant Varieties and Birational Geometry
## 1. Introduction
In this paper we continue the geometric construction of a sequence of flips associated to an embedded projective variety begun in \[V2\]. We give hypotheses under which this sequence of flips exists, and state some conjectures on how positive a line bundle on a curve must be to satisfy these hypotheses. These conjectures deal with the degrees of forms defining various secant varieties to curves and seem interesting outside of the context of the flip construction.
As motivation, we have the work of A. Bertram and M. Thaddeus. In \[T1\] this sequence of flips is constructed in the case of smooth curves via GIT, in the context of the moduli space of rank two vector bundles on a smooth curve. An understanding of this as a sequence of log flips is given in \[B3\], and further examples of sequences of flips of this type, again constructed via GIT, are given in \[T2\],\[T3\]. Our construction, however, does not use the tools of Geometric Invariant Theory and is closer in spirit to \[B1\],\[B2\].
In Section 2, we review the constructions in \[B1\] and \[T1\] and describe the relevant results from \[V2\]. In Section 3 we discuss the generation of $`SecX`$ by cubics. In particular, we show (Theorem 3.2) that large embeddings of varieties have secant varieties that are at least set theoretically defined by cubics. We also offer some general conjectures and suggestions in this direction for the generation of higher secant varieties.
The construction of the new flips is somewhat more involved than that of the first in \[V2\]. We give a general construction of a sequence of birational transformations in Section 4, and we describe in detail the second flip in Section 5.
We mention that some of the consequences of these constructions and this point of view are worked out in \[V3\].
Notation: We will decorate a projective variety $`X`$ as follows: $`X^d`$ is the $`d^{th}`$ cartesian product of $`X`$; $`S^dX`$ is $`Sym^dX=X^d/S_d`$, the $`d^{th}`$ symmetric product of $`X`$; and $`^dX`$ is $`Hilb^d(X)`$, the Hilbert Scheme of zero dimensional subschemes of $`X`$ of length $`d`$. Recall (Cf. \[Go\]) that if $`X`$ is a smooth projective variety then $`^dX`$ is also projective, and is smooth if and only if either $`dimX2`$ or $`d3`$.
Write $`Sec_k^{\mathrm{}}X`$ for the (complete) variety of $`k`$-secant $`\mathrm{}`$-planes to $`X`$. As this notation can become cluttered, we simply write $`Sec^{\mathrm{}}X`$ for $`Sec_{\mathrm{}+1}^{\mathrm{}}X`$ and $`SecX`$ for $`Sec_2^1X`$. Note also the convention $`Sec^0X=X`$. If $`V`$ is a $`k`$-vector space, we denote by $`(V)`$ the space of 1-dimensional quotients of $`V`$. Unless otherwise stated, we work throughout over the field $`k=`$ of complex numbers. We use the terms locally free sheaf (resp. invertible sheaf) and vector bundle (resp. line bundle) interchangeably. Recall that a line bundle $``$ on $`X`$ is nef if $`.C0`$ for every irreducible curve $`CX`$. A line bundle $``$ is big if $`^n`$ induces a birational map for all $`n0`$.
Acknowledgments: I would like to thank Aaron Bertram, Sheldon Katz, Zhenbo Qin, and Jonathan Wahl for their helpful conversations and communications.
## 2. Overview of Stable Pairs and the Geometry of SecX
Fix a line bundle $`\mathrm{\Lambda }`$ on a fixed smooth curve $`X`$, and denote by $`M(2,\mathrm{\Lambda })`$ the moduli space of semi-stable rank two vector bundles $`E`$ with $`^2E=\mathrm{\Lambda }`$. There is a natural rational map, the Serre Correspondence
$$\mathrm{\Phi }:(\mathrm{\Gamma }(X,K_X\mathrm{\Lambda })^{})M(2,\mathrm{\Lambda })$$
given by the duality $`\mathrm{Ext}^1(\mathrm{\Lambda },𝒪)H^1(X,\mathrm{\Lambda }^1)H^0(X,K_X\mathrm{\Lambda })^{}`$, taking an extension class $`0𝒪E\mathrm{\Lambda }0`$ to $`E`$. One has an embedding $`X(\mathrm{\Gamma }(X,K_X\mathrm{\Lambda })^{})`$ (at least in the case $`d=c_1(\mathrm{\Lambda })3`$) and $`\mathrm{\Phi }`$, defined only for semi-stable $`E`$, is a morphism off $`Sec^kX`$ where $`k=\left[\frac{d1}{2}\right]`$ \[B2\]. This map is resolved in \[B1\] by first blowing up along $`X`$, then along the proper transform of $`SecX`$, then along the transform of $`Sec^2X`$ and so on until we have a morphism to $`M(2,\mathrm{\Lambda })`$.
A different approach is taken in \[T1\]. There, for a fixed smooth curve $`X`$ of genus at least $`2`$ and a fixed line bundle $`\mathrm{\Lambda }`$, the moduli problem of semi-stable pairs $`(E,s)`$ consisting of a rank two bundle $`E`$ with $`^2E=\mathrm{\Lambda }`$, and a section $`s\mathrm{\Gamma }(X,E)\{0\}`$, is considered. This, in turn, is interpreted as a GIT problem, and by varying the linearization of the group action, a collection of (smooth) moduli spaces $`M_1,M_2,\mathrm{},M_k`$ ($`k`$ as above) is constructed. As stability is an open condition, these spaces are birational. In fact, they are isomorphic in codimension one, and may be linked via a diagram
where there is a morphism $`M_kM(2,\mathrm{\Lambda })`$. The relevant observations are first that this is a diagram of flips (in fact it is shown in \[B3\] that it is a sequence of log flips) where the ample cone of each $`M_i`$ is known. Second, $`M_1`$ is the blow up of $`(\mathrm{\Gamma }(X,K_X\mathrm{\Lambda })^{})`$ along $`X`$, $`\stackrel{~}{M_2}`$ is the blow up of $`M_1`$ along the proper transform of the secant variety, and all of the flips can be seen as blowing up and down various higher secant varieties. Finally, the $`M_i`$ are isomorphic off loci which are projective bundles over appropriate symmetric products of $`X`$.
Our approach is as follows: The sequence of flips in Thaddeus’ construction can be realized as a sequence of geometric constructions depending only on the embedding of $`X^n`$. An advantage of this approach is that the smooth curve $`X`$ can be replaced by any smooth variety. Even in the curve case, our approach applies to situations where Thaddeus’ construction does not hold (e.g. for canonical curves with $`\mathrm{Cliff}X>2`$). In \[V2\], we show how to construct the first flip using only information about the syzygies among the equations defining the variety $`X^n`$. We summarize this construction here.
###### Definition 2.1.
Let $`X`$ be a subscheme of $`^n`$. The pair $`(X,F_i)`$ satisfies condition $`\mathbf{(}K_d\mathbf{)}`$ if $`X`$ is scheme theoretically cut out by forms $`F_0,\mathrm{},F_s`$ of degree $`d`$ such that the trivial (or Koszul) relations among the $`F_i`$ are generated by linear syzygies.
We say $`(X,V)`$ satisfies $`(K_d)`$ for $`VH^0(^n,𝒪(d))`$ if $`V`$ is spanned by forms $`F_i`$ satisfying the above condition. We say simply $`X`$ satisfies $`(K_d)`$ if there exists a set $`\{F_i\}`$ such that $`(X,F_i)`$ satisfies $`(K_d)`$, and if the discussion depends only on the existence of such a set, not on the choice of a particular set.
As $`(K_2)`$ is a weakening of Green’s property $`(N_2)`$\[G\], examples of varieties satisfying $`(K_2)`$ include smooth curves embedded by complete linear systems of degree at least $`2g+3`$, canonical curves with $`\mathrm{Cliff}X3`$, and sufficiently large embeddings of arbitrary projective varieties.
To any projective variety $`X^{s_0}`$ defined (as a scheme) by forms $`F_0,\mathrm{},F_{s_1}`$ of degree $`d`$, there is an associated rational map $`\phi :^{s_0}^{s_1}`$ defined off the common zero locus of the $`F_i`$, i.e. off $`X`$. This map may be resolved to a morphism $`\stackrel{~}{\phi }:\stackrel{~}{^{s_0}}^{s_1}`$ by blowing up $`^n`$ along $`X`$, or equivalently by projecting from the closure of the graph $`\overline{\mathrm{\Gamma }_\phi }^{s_0}\times ^{s_1}`$. We have the following results on the structure of $`\stackrel{~}{\phi }`$:
###### Theorem 2.2.
\[V2, 2.4-2.10\] Let $`(X,F_i)`$ be a pair that satisfies $`(K_d)`$. Then:
1. $`\phi :^{s_0}X^{s_1}`$ is an embedding off of $`Sec_d^1X`$, the variety of $`d`$-secant lines.
2. The projection of a positive dimensional fiber of $`\stackrel{~}{\phi }`$ to $`^{s_0}`$ is either contained in a linear subspace of $`X`$ or is a linear space intersecting $`X`$ in a $`d`$-tic hypersurface.
If, furthermore, $`X`$ does not contain a line then $`\stackrel{~}{\phi }`$ is an embedding off the proper transform of $`Sec_d^1X`$. $`\mathrm{}`$
###### Theorem 2.3.
\[V2, 3.8\] Let $`(X,V)`$ satisfy $`(K_2)`$ and assume $`X^{s_0}`$ is smooth, irreducible, contains no lines and contains no quadrics. Then:
1. The image of $`\stackrel{~}{SecX}=\stackrel{~}{Sec_2^1X}`$ under $`\stackrel{~}{\phi }`$ is $`^2X`$.
2. $`=\stackrel{~}{\phi }_{}(𝒪_{\stackrel{~}{SecX}}(H))`$ is a rank two vector bundle on $`^2X`$, where $`𝒪_{\stackrel{~}{^{s_0}}}(H)`$ is the proper transform of the hyperplane section on $`^{s_0}`$.
3. $`\stackrel{~}{\phi }:\stackrel{~}{SecX}^2X`$ is the $`^1`$-bundle $`_{^2X}^2X`$. $`\mathrm{}`$
This implies $`\stackrel{~}{SecX}`$, and hence $`\stackrel{~}{M_2}=\mathrm{Bl}_{\stackrel{~}{SecX}}(\stackrel{~}{^{s_0}})`$, are smooth. To complete the flip, we construct a base point free linear system on $`\stackrel{~}{M_2}`$, and take $`M_2`$ to be the image of the associated morphism. Denoting $`\stackrel{~}{SecX}=()`$, the sheaf $`=\stackrel{~}{\phi }_{}(N_{()/\stackrel{~}{^{s_0}}}^{}𝒪_{()}(1))`$ is locally free of rank $`n2dimX1`$ on $`^2X`$. Write $`()=_{^2X}()`$ and rename $`\stackrel{~}{\phi }`$ as $`\stackrel{~}{\phi _1}^+`$:
###### Theorem 2.4.
\[V2, 4.13\] Let $`(X,V)`$ satisfy $`(K_2)`$ and assume $`X^{s_0}`$ is smooth, irreducible, contains no lines and contains no plane quadrics. Then there is a flip as pictured below with:
1. $`\stackrel{~}{^{s_0}}`$, $`\stackrel{~}{M_2}`$, and $`M_2`$ smooth
2. $`\stackrel{~}{^{s_0}}()M_2()`$, hence if $`codim((),\stackrel{~}{^{s_0}})2`$ then $`\mathrm{Pic}\stackrel{~}{^{s_0}}\mathrm{Pic}M_2`$
3. $`h_1`$ is the blow up of $`M_2`$ along $`()`$
4. $`\pi `$ is the blow up of $`\stackrel{~}{^{s_0}}`$ along $`()`$
5. $`\stackrel{~}{\phi _1}^{}`$, induced by $`𝒪_{M_2}(2HE)`$, is an embedding off of $`()`$, and the restriction of $`\stackrel{~}{\phi _1}^{}`$ is the projection $`()^2X`$
6. $`\stackrel{~}{\phi _1}^+`$, induced by $`𝒪_{\stackrel{~}{^{s_0}}}(2HE)`$, is an embedding off of $`()`$, and the restriction of $`\stackrel{~}{\phi _1}^+`$ is the projection $`()^2X`$
To continue this process following Thaddeus, we need to construct a birational morphism $`\stackrel{~}{\phi _2}^+:M_2^{s_2}`$ which contracts the transforms of $`3`$-secant $`2`$-planes to points, and is an embedding off their union. The natural candidate is the map induced by the linear system $`𝒪_{M_2}(3H2E)`$. We discuss two different reasons for this choice that will guide the construction of the entire sequence of flips. Section 3 addresses the question of when this system is globally generated. Note that we abuse notation throughout and identify line bundles via the isomorphism $`\mathrm{Pic}\stackrel{~}{^{s_0}}\mathrm{Pic}M_k`$.
The first reason is quite naive: Just as quadrics collapse secant lines because their restriction to such a line is a quadric hypersurface, so too do cubics vanishing twice on a variety collapse every $`3`$-secant $`^2`$ because they vanish on a cubic hypersurface in such a plane. Similarly, to collapse the transform of each $`k+1`$-secant $`^k`$ via a morphism $`\stackrel{~}{\phi _k}^+:M_k^{s_k}`$, the natural system is $`𝒪_{M_k}((k+1)HkE)`$.
Another reason is found by studying the ample cones of the $`M_i`$. Note that the ample cone on $`\stackrel{~}{^{s_0}}(=M_1)`$ is bounded by the line bundles $`𝒪_{\stackrel{~}{^{s_0}}}(H)`$ and $`𝒪_{\stackrel{~}{^{s_0}}}(2HE)`$. Both of these bundles are globally generated, and by Theorems 2.2 and 2.3, they each give birational morphisms whose exceptional loci are projective bundles over Hilbert schemes of points of $`X`$ ($`^1XX`$ and $`^2X`$ respectively).
On $`M_2`$, the ample cone is bounded on one side by $`𝒪_{M_2}(2HE)`$. This gives the map $`\stackrel{~}{\phi _1}^{}:M_2^{s_1}`$ mentioned in Theorem 2.4; in particular it is globally generated, the induced morphism is birational, and its exceptional locus is a projective bundle over $`^2X`$. On the other side, the ample cone contains a line bundle of the form $`𝒪_{M_2}((2m1)HmE)`$ (\[V2, 4.9\]). In fact, if $`X`$ is a smooth curve embedded by a line bundle of degree at least $`2g+5`$, it is shown in \[T1\] that the case $`m=2`$ suffices, i.e. that the ample cone is bounded by $`𝒪_{M_2}(2HE)`$ and $`𝒪_{M_2}(3H2E)`$. Therefore, it is natural to look for conditions under which $`𝒪_{M_2}(3H2E)`$ is globally generated. Thaddeus further shows that under similar positivity conditions, the ample cone of $`M_k`$ is bounded by $`𝒪_{M_k}(kH(k1)E)`$ and $`𝒪_{M_k}((k+1)HkE)`$.
Noting the fact that $`h_1^{}𝒪_{M_2}(3H2E)=𝒪_{\stackrel{~}{M_2}}(3H2E_1E_2)`$, it is not difficult to see (using Zariski’s Main Theorem) that this system will be globally generated if $`SecX^{s_0}`$ is scheme theoretically defined by cubics, because a cubic vanishing twice on a variety must also vanish on its secant variety. Unfortunately, there are no general theorems on the cubic generation of secant varieties analogous to quadric generation of varieties. We address this question in the next section.
## 3. Cubic Generation of Secant Varieties
Example 3.1 Some examples of varieties whose secant varieties are ideal theoretically defined by cubics include:
1. $`X`$ is any Veronese embedding of $`^n`$ \[Ka\]
2. $`X`$ is the Plücker embedding of the Grassmannian $`𝔾(1,n)`$ for any $`n`$ \[H, 9.20\].
3. $`X`$ is the Segre embedding of $`^n\times ^m`$ \[H, 9.2\]. $`\mathrm{}`$
We prove a general result:
###### Theorem 3.2.
Let $`X^{s_0}`$ satisfy condition $`(K_2)`$. Then $`Sec(v_d(X))`$ is set theoretically defined by cubics for $`d2`$.
###### Proof.
We begin with the case $`d=2`$, the higher embeddings being more elementary.
Let $`Y=v_2(X)`$, $`V=v_2(^{s_0})^N`$, and $`H`$ the linear subspace of $`^N`$ defined by the hyperplanes corresponding to all the quadrics in $`^{s_0}`$ vanishing on $`X`$. Then $`Y=VH`$ as schemes and we show, noting that $`SecV`$ is ideal theoretically defined by cubics, that $`SecY=SecVH`$ as sets.
Note that the map $`\phi _1:^{s_0}^{s_1}`$ can be viewed as the composition of the embedding $`v_2:^{s_0}^N`$ with the projection from $`H`$, $`^N^{s_1}`$.
Let $`pSecVH`$. If $`pV`$, then $`pY=VH`$ hence $`pSecY`$.
Otherwise, any secant line $`L`$ to $`V`$ through $`p`$ intersects $`V`$ in a length two subscheme $`Z`$. $`Z`$ considered in $`^{s_0}`$ determines a unique line in $`^{s_0}`$ whose image in $`^N`$ is a plane quadric $`QV`$ spanning a plane $`M`$. If $`HQ=Z^{}Y`$ is non-empty then $`Z^{}\{p\}HM`$, hence either $`H`$ intersects $`M`$ in a line $`L^{}`$ through $`p`$ or $`MH`$. In the first case $`L^{}`$ is a secant line to $`Y`$, in the second $`QY`$. In either situation $`pSecY`$.
All that remains is the case $`HM=\{p\}`$ and $`HQ`$ is empty. However in this case the line $`L`$, and hence the scheme $`Z=LQ`$ is collapsed to a point by the projection. As the rational map $`\phi `$ is an embedding off $`SecX`$, this implies $`Z`$ lies on the image of a secant line to $`X^{s_0}`$. As a length two subscheme of $`^{s_0}`$ determines a unique line, $`Q`$ must be the image of a secant line to $`X^{s_0}`$ contradicting the assumption that $`HQ`$ is empty.
For $`d>2`$, note that the projection from $`H`$ is an embedding off $`VH`$ (this can be derived directly from Theorem 2.2 or see \[V1, 3.3.1\]). Therefore, if $`H`$ intersects a secant line, the line lies in $`H`$, hence is a secant line to $`Y`$. ∎
Example 3.3 As Green’s $`(N_2)`$ implies $`(K_2)`$, this shows that the secant varieties to the following varieties are set theoretically defined by cubics:
1. $`X`$ a smooth curve embedded by a line bundle of degree $`4g+6+2r`$, $`r0`$.
2. $`X`$ a smooth curve with $`\mathrm{Cliff}X>2`$, embedded by $`K_X^r`$, $`r2`$.
3. $`X`$ a smooth variety embedded by $`\left(K_XL^{(dimX+3+\alpha )}\right)^r`$, $`\alpha 0,r2`$, $`L`$ very ample.
4. $`X`$ a smooth variety embedded by $`L^{2r}`$ for all $`r0`$, $`L`$ ample. $`\mathrm{}`$
Remark 3.4 Notice that in the case $`d=2`$ of Proposition 3.2, the cubics that at least set theoretically define the secant variety satisfy $`(K_3)`$. This is because:
1. The ideal of the secant variety of $`v_2(^{s_0})`$ is generated by cubics, and the module of syzygies is generated by linear relations \[JPW, 3.19\]. Hence $`Sec(v_2(^{s_0}))`$ satisfies $`(K_3)`$.
2. It is clear from the definition that if $`X^n`$ satisfies $`(K_d)`$, then any linear section does as well. $`\mathrm{}`$
Example 3.5 If $`X^n`$ is a smooth quadric hypersurface, then $`v_2(X)`$ is given by the intersection of $`v_2(^n)`$ with a hyperplane $`H`$. Furthermore, the intersection of $`Sec(v_2(^n))`$ with $`H`$ is a scheme $`S`$ with $`S_{red}Sec(v_2(X))`$. Therefore, a general smooth quadric hypersurface has $`Sec(v_2(X))Sec(v_2(^n))H`$ as schemes, hence $`Sec(v_2(X))`$ satisfies $`(K_3)`$. $`\mathrm{}`$
We record here a related conjecture of Eisenbud, Koh, and Stillman as well as a partial answer proven by M.S. Ravi:
###### Conjecture 3.6.
\[EKS\] Let $`L`$ be a very ample line bundle that embeds a smooth curve $`X`$. For each $`k`$ there is a bound on the degree of $`L`$ such that $`Sec^kX`$ is ideal theoretically defined by the $`(k+2)\times (k+2)`$ minors of a matrix of linear forms.
###### Theorem 3.7.
\[R\] If $`degL4g+2k+3`$, then $`Sec^kX`$ is set theoretically defined by the $`(k+2)\times (k+2)`$ minors of a matrix of linear forms.
These statements provide enough evidence to make the following basic:
###### Conjecture 3.8.
Let $`L`$ be an ample line bundle on a smooth variety $`X`$, $`k1`$ fixed. Then for all $`n0`$, $`L^n`$ embeds $`X`$ so that $`Sec^kX`$ is ideal theoretically defined by forms of degree $`k+2`$, and furthermore satisfies condition $`(K_{k+2})`$.
Remark 3.9 If $`X`$ is a curve with a $`5`$-secant $`3`$-plane, then any cubic vanishing on $`SecX`$ must vanish on that $`3`$-plane. Hence $`SecX`$ cannot be set theoretically defined by cubics. This should be compared to the fact that if $`X`$ has a trisecant line, then $`X`$ cannot be defined by quadrics. In particular, this shows that Green’s condition $`(N_2)`$ is not even sufficient to guarantee that their exists a cubic vanishing on $`SecX`$. For example, if $`X`$ is an elliptic curve embedded in $`^4`$ by a line bundle of degree $`5`$, then $`SecX`$ is a quintic hypersurface. Therefore, any uniform bound on the degree of a linear system that would guarantee $`SecX`$ is even set theoretically defined by cubics must be at least $`2g+4`$. $`\mathrm{}`$
We can use earlier work to give a more geometric necessary condition for $`SecX`$ to be defined as a scheme by cubics. Specifically, in \[V2, 3.7\] it is shown that the intersection of $`\stackrel{~}{SecX}`$ with the exceptional divisor $`E`$ of the blow up of $`^{s_0}`$ along $`X`$ is isomorphic to $`\mathrm{Bl}_\mathrm{\Delta }(X\times X)`$. This implies that if $`\pi :\stackrel{~}{SecX}SecX`$ is the blow up along $`X`$, then $`\pi ^1(p)\mathrm{Bl}_p(X),pX`$. In fact, it is easy to verify that if $`X`$ is embedded by a line bundle $`L`$, then $`\pi ^1(p)\mathrm{Bl}_p(X)\mathrm{\Gamma }(X,L_p^2)`$ where $`\mathrm{\Gamma }(X,L_p^2)`$ is identified with the fiber over $`p`$ of the projectivized conormal bundle of $`X^{s_0}`$. Now, if $`SecX`$ is defined as a scheme by cubics, then the base scheme of $`𝒪_{\stackrel{~}{^{s_0}}}(3H2E)`$ is precisely $`\stackrel{~}{SecX}`$. The restriction of this series to $`\mathrm{\Gamma }(X,L_p^2)`$ is thus a system of quadrics whose base scheme is $`\mathrm{Bl}_p(X)`$. In other words, if $`X`$ is a smooth variety embedded by a line bundle $`L`$ that satisfies $`(K_2)`$ and if $`SecX`$ is scheme theoretically defined by cubics, then for every $`pX`$ the line bundle $`L𝒪(2E_p)`$ is very ample on $`\mathrm{Bl}_p(X)`$ and $`\mathrm{Bl}_p(X)\mathrm{\Gamma }(\mathrm{Bl}_p(X),L𝒪(2E_p))`$ is scheme theoretically defined by quadrics.
In the case $`X`$ is a curve, this implies that a uniform bound on $`degL`$ that would imply $`SecX`$ is defined by cubics must be at least $`2g+4`$, the same bound encountered in Remark 3.8. The construction in \[B1\] shows similarly that any uniform bound that would imply $`Sec^kX`$ is defined by $`(k+2)`$-tics must be at least $`2g+2+2k`$. We combine these observations with the degree bounds encountered in the constructions of \[T1\] and \[B1\] to form the following:
###### Conjecture 3.10.
Let $`X`$ be a smooth curve embedded by a line bundle $`L`$. If $`deg(L)2g+2k`$ then $`Sec^{k1}X`$ is defined as a scheme by forms of degree $`k+1`$. If $`deg(L)2g+2k+1`$ then $`Sec^{k1}X`$ satisfies condition $`(K_{k+1})`$. $`\mathrm{}`$
## 4. The General Birational Construction
Suppose that $`X`$ satisfies $`(K_2)`$, is smooth, and contains no lines and no plane quadrics. Suppose further that $`SecX`$ is scheme theoretically defined by cubics $`C_0,\mathrm{},C_{s_2}`$, and that $`SecX`$ satisfies $`(K_3)`$. Under these hypotheses, we construct a second flip as follows: We know that $`𝒪_{M_2}(3H2E)`$ is globally generated by the discussion above; hence this induces a morphism $`\stackrel{~}{\phi _2}^+:M_2^{s_2}`$ which agrees with the map given by the cubics $`\phi _2:^{s_0}^{s_2}`$ on the locus where $`M_2`$ and $`^{s_0}`$ are isomorphic. By Theorem 2.2, $`\stackrel{~}{\phi _2}^+`$ is a birational morphism. We wish first to identify the exceptional locus of $`\stackrel{~}{\phi _2}^+`$. It is clear that $`\stackrel{~}{\phi _2}^+`$ will collapse the image of a $`3`$-secant $`2`$-plane to a point, hence the exceptional locus must contain the transform of $`Sec^2X`$. However by Theorem 2.2, we know that the rational map $`\phi _2`$ is an embedding off $`Sec_3^1(SecX)`$, the trisecant variety to the secant variety. This motivates the following
###### Lemma 4.1.
Let $`X^n`$ be an irreducible variety. Assume either of the following:
1. $`Sec^kX`$ is defined as a scheme by forms of degree $`2k+1`$.
2. $`X`$ is a smooth curve embedded by a line bundle of degree at least $`2g+2k+1`$.
Then $`Sec^kX=Sec_{k+1}^1(Sec^{k1}X)`$ as schemes.
###### Proof.
First, choose a $`(k+1)`$-secant $`k`$-plane $`M`$. $`M`$ then intersects $`Sec^{k1}X`$ in a hypersurface of degree $`k+1`$, hence every line in $`M`$ lies in $`Sec_{k+1}^1(Sec^{k1}X)`$. As $`Sec^kX`$ is reduced and irreducible, $`Sec^kXSec_{k+1}^1(Sec^{k1}X)`$ as schemes.
For the converse, assume the first condition is satisfied. Choose a line $`L`$ that intersects $`Sec^{k1}X`$ in a scheme of length at least $`k+1`$. It is easy to verify that $`Sec^kX`$ is singular along $`Sec^{k1}X`$, hence every form that vanishes on $`Sec^kX`$ must vanish $`2k+2`$ times on $`L`$. By hypothesis, however, $`Sec^kX`$ is scheme theoretically defined by forms of degree $`2k+1`$, hence each of these forms must vanish on $`L`$.
The sufficiency of the second condition follows from Thaddeus’ construction and \[B3, §2,(i)\]. ∎
This implies that if $`Sec^kX`$ satisfies $`(K_{k+2})`$ and if $`Sec^kX=Sec_{k+1}^1(Sec^{k1}X)`$, then the map $`\phi _{k+1}:^{s_0}^{s_{k+1}}`$ given by the forms defining $`Sec^kX`$ is an embedding off of $`Sec^{k+1}X`$. We use Theorem 2.2 to understand the structure of these maps via the following two lemmas:
###### Lemma 4.2.
If the embedding of a projective variety $`X^n`$ is $`(2k+4)`$-very ample, then the intersection of two $`(k+2)`$-secant $`(k+1)`$-planes, if nonempty, must lie in $`Sec^kX`$ (in fact, it must be an $`\mathrm{}+1`$ secant $`^{\mathrm{}}`$ for some $`\mathrm{}k`$). In particular, $`Sec^{k+1}X`$ has dimension $`(k+2)dimX+k+1`$.
###### Proof.
The first statement is elementary: Assume two $`(k+2)`$-secant $`(k+1)`$-planes intersect at a single point. If the point is not on $`X`$, then there are $`2k+4`$ points of $`X`$ that span a $`(2k+2)`$-plane, which is impossible by hypothesis. Hence the intersection lies in $`Sec^0X=X`$. A simple repetition of this argument for larger dimensional intersections gives the desired result. The statement of the dimension follows immediately; or see \[H, 11.24\]. ∎
###### Lemma 4.3.
Let $`X^{s_0}`$ be an irreducible variety whose embedding is $`(2k+4)`$-very ample. Assume that $`Sec^kX`$ satisfies $`(K_{k+2})`$, and that $`Sec^{k+1}X=Sec_{k+2}^1(Sec^kX)`$ as schemes. Let $`\mathrm{\Gamma }`$ be the closure of the graph of $`\phi _{k+1}`$ with projection $`\pi :\mathrm{\Gamma }^{s_0}`$. If $`a`$ is a point in the closure of the image of $`\phi _{k+1}`$ and $`F_a\mathrm{\Gamma }`$ is the fiber over $`a`$ then $`\pi (F_a)`$ is one of the following:
1. a reduced point in $`^{s_0}Sec^{k+1}X`$
2. a $`(k+2)`$-secant $`(k+1)`$-plane
3. contained in a linear subspace of $`Sec^kX`$
###### Proof.
The first and third possibilities follow directly from Theorem 2.2.
For the second, note that a priori $`\pi (F_a)`$ could be any linear space intersecting $`Sec^kX`$ in a hypersurface of degree $`k+2`$. However, Lemma 4.2 and the hypothesis that $`Sec^{k+1}X=Sec_{k+2}^1(Sec^kX)`$ immediately imply that any such linear space must be $`k+1`$ dimensional; hence a $`(k+2)`$-secant $`(k+1)`$-plane. ∎
With these results in hand we present the general construction.
Let $`Y_0`$ be an irreducible projective variety and suppose $`\beta _i:Y_0Y_i,1ij`$, is a collection of dominant, birational maps. Define the dominating variety of the collection, denoted $`_{(0,1,\mathrm{},j)}`$, to be the closure of the graph of
$$(\beta _1,\beta _2,\mathrm{},\beta _j):Y_0Y_1\times Y_2\times \mathrm{}\times Y_j$$
Denote by $`_{(a_1,a_2,\mathrm{},a_r)}`$ the projection of $`_{(0,1,\mathrm{},j)}`$ to $`Y_{a_1}\times Y_{a_2}\times \mathrm{}\times Y_{a_r}`$. Note that $`_{(a_1,a_2,\mathrm{},a_r)}`$ is birationally isomorphic to $`_{(b_1,b_2,\mathrm{},b_k)}`$ for all $`0a_r,b_kj`$. Note further that if the $`\beta _i`$ are all morphisms then $`_{(0,1,\mathrm{},j)}Y_0`$, in other words only rational maps contribute to the structure of the dominating variety.
Definition/Notation 4.4 We say $`X^{s_0}`$ satisfies condition $`(K_2^j)`$ if $`Sec^iX`$ satisfies condition $`(K_{2+i})`$ for $`0ij`$; hence $`X`$ satisfies $`(K_2^0)`$ if and only if $`X`$ satisfies $`(K_2)`$, $`X`$ satisfies $`(K_2^1)`$ if and only if $`X`$ satisfies $`(K_2)`$ and $`SecX`$ satisfies $`(K_3)`$, etc. $`\mathrm{}`$
If $`X^{s_0}`$ satisfies condition $`(K_2^j)`$, then each rational map $`\phi _i:^{s_0}^{s_i}`$ is birational onto its image for $`1ij+1`$, and assuming the conclusion of Lemma 4.1 each $`\phi _i`$ is an embedding off $`Sec^iX`$. Therefore $`_{(i)}`$ is the closure of the image of $`\phi _i`$, $`_{(0,i)}`$ is the closure of the graph of $`\phi _i`$, and in the notation of Theorem 2.4 $`_{(0,1,2)}\stackrel{~}{M_2}`$ and $`_{(1,2)}M_2`$. Note $`_{(0)}=^{s_0}`$.
###### Lemma 4.5.
$`_{(0,1,2,\mathrm{},i)}`$ is the blow up of $`_{(0,1,2,\mathrm{},i1)}`$ along the proper transform of $`Sec^{i1}X`$, $`1ij+1`$.
###### Proof.
This is immediate from the definition (or see \[V1, 3.1.1\]). ∎
Remark 4.6 The spaces constructed in \[B1\] are of the type $`_{(0,1,2,\mathrm{},k)}`$. The spaces $`\stackrel{~}{M_k}`$ and $`M_k`$ constructed in \[T1\] are $`\stackrel{~}{M_k}_{(k2,k1,k)}`$ and $`M_k_{(k1,k)}`$. $`\mathrm{}`$
Our goal is to understand explicitly the geometry of this web of varieties generalizing Theorem 2.4. In the next section we describe in detail the structure of the second flip. As each subsequent flip requires the understanding of $`^kX`$ for larger $`k`$, it is not clear that the process will continue nicely beyond the second flip (at least for varieties of arbitrary dimension).
## 5. Construction of the Second Flip
Let $`X^{s_0}`$ be a smooth, irreducible variety that satisfies $`(K_2^1)`$. The diagram of varieties we study in this section is:
where $`_{(0,1,2)}`$ is the dominating variety of the pair of birational maps $`\phi _1:^{s_0}^{s_1}`$ and $`\phi _2:^{s_0}^{s_2}`$; and where we have yet to construct the two rightmost varieties. We write $`\mathrm{Pic}_{(0,1)}=\mathrm{Pic}_{(1,2)}=H+E`$ and $`\mathrm{Pic}_{(0,1,2)}=H+E_1+E_2`$ (recall all three spaces are smooth by Theorem 2.4).
###### Theorem 5.1.
Let $`X_{(0)}=^{s_0}`$ be a smooth, irreducible variety of dimension $`r`$ that satisfies $`(K_2^1)`$. Assume that $`X`$ is embedded by a complete linear system $`|L|`$ and that the following conditions are satisfied:
1. $`L`$ is $`(5+r)`$-very ample
2. If $`r2`$, then for every point $`pX`$, $`H^1(X,L_p^3)=0`$
3. $`Sec^2X=Sec_3^1(Sec^1X)`$ as schemes
4. The projection of $`X`$ into $`^m,m=s_01r`$, from any embedded tangent space is such that the image is projectively normal and satisfies $`(K_2)`$
Then the morphism $`\stackrel{~}{\phi _2}^+:_{(1,2)}_{(2)}`$ induced by $`𝒪_{_{(1,2)}}(3H2E)`$ is an embedding off the transform of $`Sec^2X`$, and the restriction of $`\stackrel{~}{\phi _2}^+`$ to the transform of $`Sec^2X`$ has fibers isomorphic to $`^2`$.
As the proof of Theorem 5.1 is somewhat involved, we break it into several pieces. We begin with a Lemma and a crucial observation, followed by the proof of the Theorem. The observation invokes a technical lemma whose proof is postponed until the end.
Remark On the Hypotheses 5.2 Note that if $`X`$ is a smooth curve embedded by a line bundle of degree at least $`2g+5`$, then conditions $`14`$ are automatically satisfied. Conjecture 3.10 would imply condition $`(K_2^1)`$ holds also. Furthermore, if $`r=2`$ and $`H^1(X,L)=0`$ then condition $`1`$ implies condition $`2`$.
If $`r2`$, then the image of the projection from the space tangent to $`X`$ at $`p`$ is $`\mathrm{Bl}_p(X)^m`$. Furthermore, by the discussion after Remark 3.8 any such projection of $`X`$ will be generated as a scheme by quadrics when $`SecX`$ is defined by cubics, hence condition $`4`$ is not unreasonable. $`\mathrm{}`$
###### Lemma 5.3.
With hypotheses as in Theorem 5.1, the image of the projection of $`X`$ into $`^m`$, $`m=s_01r`$, is $`\mathrm{Bl}_p(X)`$, hence is smooth. Furthermore, it contains no lines and it contains no plane quadrics except for the exceptional divisor, which is the quadratic Veronese embedding of $`^{r1}`$.
###### Proof.
If $`r=1`$ the statement is clear. Otherwise, let $`X^{}^m`$ denote the closure of the image of projection from the embedded tangent space to $`X`$ at $`p`$. As mentioned above, $`X^{}\mathrm{Bl}_p(X)`$, hence is smooth. Let $`E_pX^{}`$ denote the exceptional divisor. The existence of a line or plane quadric not contained in $`E_p`$ is immediately seen to be impossible by the $`(5+r)`$-very ampleness hypothesis.
As $`^m=\mathrm{\Gamma }(X^{},L𝒪(2E_p))`$ and as $`E_p^{r1}`$, we have $`L𝒪(2E_p)|_{E_p}𝒪_{^{r1}}(2)`$. Condition $`2`$ implies this restriction is surjective on global sections. ∎
Observation 5.4 Let $`_{(0,1,2)}_{(0)}`$ be the projection and let $`F_p`$ be the fiber over $`pX`$; hence $`F_p`$ is the blow up of $`^m`$ along a copy of $`\mathrm{Bl}_p(X)`$. We again denote this variety by $`X^{}^m`$, and the embedding of $`X^{}`$ into $`^m`$ satisfies $`(K_2)`$ by hypothesis. The restriction of $`𝒪_{_{(0,1,2)}}(3H2E_1E_2)`$ to $`F_p`$ can thus be identified with $`𝒪_{\mathrm{Bl}_X^{}(^m)}(2H^{}E^{})`$, and, noting Lemma 5.3, it seems that Theorem 2.2 could be applied. Unfortunately, it is not clear that this restriction should be surjective on global sections. However, by Lemma 5.7 below, the image of the morphism on $`F_p`$ induced by the restriction of global sections is isomorphic to the image of the morphism given by the complete linear system $`|𝒪_{\mathrm{Bl}_X^{}(^m)}(2H^{}E^{})|`$. Hence by the fourth hypothesis and Lemma 5.3, the only collapsing that occurs in $`F_p`$ under the morphism $`_{(0,1,2)}_{(2)}`$ is that of secant lines to $`X^{}^m`$.
Now, for some $`pX`$, suppose that a secant line $`S`$ in $`F_p`$ is collapsed to a point by the projection $`_{(0,1,2)}_{(2)}`$. Then $`S`$ is the proper transform of a secant line to $`X^{}^m`$, but every such secant line is the intersection of $`F_p`$ with a $`3`$-secant $`^2`$ through $`pX`$. For example, if $`SF_p`$ is the secant line through $`q,rX^{}`$, $`q,rE_p`$, then $`S`$ is the intersection of $`F_p`$ with the proper transform of the plane spanned by $`p,q,r`$. It should be noted that the two dimensional fiber associated to the collapsing of a plane spanned by a quadric in the exceptional divisor (Lemma 5.3) will take the place of a $`3`$-secant $`^2`$ spanned by a non-curvilinear scheme contained in the tangent space at $`p`$.
Therefore, all the collapsing in the exceptional locus over a point $`pX`$ is associated to the collapsing of $`3`$-secant $`2`$-planes. $`\mathrm{}`$
###### Proof.
(of Theorem 5.1) Let $`a_{(2)}`$ be a point in the image of $`\stackrel{~}{\phi _2}^+`$. The fiber over $`a`$ is mapped isomorphically into $`_{(1)}`$ by the projection $`_{(1,2)}_{(1)}`$. We are therefore able to study $`(\stackrel{~}{\phi _2}^+)^1(a)`$ by looking at the fiber of the projection $`_{(0,1,2)}_{(2)}`$, and projecting to $`_{(0,1)}`$ and to $`_{(1)}`$.
By applying Lemma 4.3 to the map $`_{(0,2)}_{(2)}`$, the projection to $`_{(0,1)}`$ is contained as a scheme in the total transform of one of the following (note the more refined division of possibilities):
1. a point in $`^{s_0}Sec^2X`$
2. a $`3`$-secant $`2`$-plane to $`X`$ not contained in $`SecX`$
3. a linear subspace of $`SecX`$ not tangent to $`X`$
4. a linear subspace of $`SecX`$ tangent to $`X`$
In the first case, there is nothing to show as the total transform of a point in $`^{s_0}Sec^2X`$ is simply a reduced point and the map $`\stackrel{~}{\phi _1}^+`$ to $`_{(1)}`$ is an embedding in a neighborhood of this point.
If the projection is a $`3`$-secant $`2`$-plane, then by Observation 5 the projection to $`_{(0,1)}`$ is a $`3`$-secant $`2`$-plane blown up at the three points of intersection, and so the image in $`_{(1)}`$ is a $`^2`$ that has undergone a Cremona transformation.
In the third case, Observation 5 shows that either the projection to $`_{(0,1)}`$ is the proper transform of a secant line to $`X`$, or that the projection to $`_{(0)}`$ is a linear subspace of $`SecX`$ that is not a secant line. In the first case, every such space is collapsed to a point by $`\stackrel{~}{\phi _1}^+`$. The second implies $`\stackrel{~}{\phi _2}^+`$ has a fiber of dimension $`d`$ that is contained in $`()_{(1,2)}`$. Because $`E_2()`$ is a $`^1`$-bundle, this implies the projection of the fiber to $`_{(0)}`$ is contained in a linear subspace $`M`$ of $`SecX`$ of dimension $`d+1`$. Furthermore, the proper transform of $`M`$ is collapsed to a $`d`$ dimensional subspace of $`_{(1)}`$, in particular the general point of $`M`$ lies on a secant line contained in $`M`$ by Theorem 2.3. Therefore $`Y=MX`$ has $`SecY=M`$, hence $`Sec^2Y=M`$ but this is impossible by Lemma 4.2 and the restriction that $`M`$ not be tangent to $`X`$.
In the final case, the proper transform in $`_{(0,1)}`$ of a linear space $`M^k`$ tangent to $`X`$ at a point $`p`$ is $`\mathrm{Bl}_p(^k)`$. Denote the exceptional $`^{k1}`$ by $`Q`$; Lemma 5.3 implies $`QE_p`$ is the quadratic Veronese embedding of $`^{k1}(\mathrm{\Gamma }(\mathrm{Bl}_p(X),L(2E_p)))`$. A simple dimension count shows that the restriction to $`Q`$ of the projective bundle $`E_2\stackrel{~}{SecX}`$ arising from the blow up of $`_{(0,1)}`$ along $`\stackrel{~}{SecX}`$ is precisely the restriction to $`Q`$ of the projective bundle arising from the induced blow up of $`(\mathrm{\Gamma }(\mathrm{Bl}_p(X),L(2E_p)))`$ along $`\mathrm{Bl}_p(X)`$; denote this variety $`_Q`$. Furthermore, the transform of $`\mathrm{Bl}_p(^k)`$ in $`_{(0,1,2)}`$ is a $`^1`$-bundle over $`_Q_{(1,2)}`$. Now by Lemma 5.7, every fiber of $`\stackrel{~}{\phi _2}^+`$ contained in $`_Q_{(1,2)}`$ is either a point or is isomorphic to a $`^2`$ spanned by a plane quadric in $`Q`$. ∎
Remark 5.5 For curves, parts $`3`$ and $`4`$ of the proof can also be concluded by showing that any line contained in $`SecX`$ must be a secant or tangent line (this is immediate from the $`6`$-very ample hypothesis). $`\mathrm{}`$
To complete the proof, we need Lemma 5.7 which itself requires a general result:
###### Lemma 5.6.
Let $`\pi :XY`$ be a flat morphism of smooth projective varieties. Let $`F=\pi ^1(p)`$ be a smooth fiber and let $`L`$ be a locally free sheaf on $`X`$. If $`R^i\pi _{}L=0`$ and $`H^i(F,L𝒪_F)=0`$ for all $`i>0`$, then $`R^i\pi _{}(_FL)=0`$ for all $`i>0`$.
###### Proof.
The hypotheses easily give the vanishing $`R^i\pi _{}(_FL)=0`$ for all $`i>1`$. For $`i=1`$, take the exact sequence on $`Y`$
$$0\pi _{}(_FL)\pi _{}L\pi _{}(𝒪_FL)R^1\pi _{}(_FL)0$$
Because $`\pi _{}(𝒪_FL)`$ is supported at the point $`p`$, it suffices to check that $`H^1(F,_F𝒪_FL)=H^1(F,N_{F/X}^{}L)=0`$. $`\pi `$ flat implies $`N_{F/X}^{}\pi ^{}(N_{p/Y}^{})`$, hence $`N_{F/X}^{}`$ is trivial. Now, $`H^1(F,L𝒪_F)=0`$ implies $`H^1(F,N_{F/X}^{}L)=0`$. ∎
###### Lemma 5.7.
Under the hypotheses of Theorem 5.1, The image of $`F_p`$ under the projection $`_{(0,1,2)}_{(2)}`$ is isomorphic to the image of $`F_p`$ under the morphism induced by the complete linear system associated to $`𝒪_{F_p}(2H^{}E^{})`$.
###### Proof.
Step 1: If $`a,bF_p`$ are mapped to the same point under the projection to $`_{(2)}`$, then $`a`$ and $`b`$ map to the same point under the projection to $`_{(0,2)}`$. This is clear from the construction of the maps in question as the projections $`^{s_0}\times ^{s_1}\times ^{s_2}^{s_2}`$ and $`^{s_0}\times ^{s_1}\times ^{s_2}^{s_0}\times ^{s_2}`$ respectively.
Step 2: Re-embed $`_{(0,2)}^N\times ^{s_2}`$ via the map associated to $`𝒪_{^{s_0}}(k)𝒪_{^{s_2}}(1)`$. This gives a map $`_{(0,1,2)}^N\times ^{s_2}`$ induced by a subspace of $`H^0(_{(0,1,2)},𝒪_{^{s_0}}(k)𝒪_{^{s_2}}(1)𝒪_{_{(0,1,2)}})`$ where $`𝒪_{^{s_0}}(k)𝒪_{^{s_2}}(1)𝒪_{_{(0,1,2)}}𝒪_{_{(0,1,2)}}((k+3)H2E_1E_2)`$. As $`_{(0,2)}^N\times ^{s_2}`$ is an embedding, the induced maps on $`F_p`$ have isomorphic images for all $`k1`$. We have, therefore, only to show $`H^0(^{s_0}\times ^{s_1}\times ^{s_2},𝒪_{^{s_0}}(k)𝒪_{^{s_2}}(1))`$ surjects onto $`H^0(F_p,𝒪_{F_p}(2H^{}E^{}))`$ for some $`k`$.
Step 3: The map
$$H^0(^{s_0}\times ^{s_1}\times ^{s_2},𝒪_{^{s_0}}(k)𝒪_{^{s_2}}(1))H^0(_{(0,1,2)},𝒪_{_{(0,1,2)}}((k+3)H2E_1E_2))$$
is surjective for all $`k0`$. This follows directly from the fact that $`SecX`$ is scheme theoretically defined by cubics and the construction of $`^{s_2}`$ as $`(\mathrm{\Gamma }(_{(0,1,2)},𝒪_{_{(0,1,2)}}(3H2E_1E_2)))`$.
Step 4: The map
$$H^0(_{(0,1,2)},𝒪_{_{(0,1,2)}}((k+3)H2E_1E_2))H^0(E_1,𝒪_{E_1}((k+3)H2E_1E_2))$$
is surjective for all $`k0`$.
We show $`H^1(_{(0,1,2)},𝒪_{_{(0,1,2)}}((k+3)H3E_1E_2))=0`$. Let $`\rho :_{(0,1,2)}_{(0)}`$ be the projection. By the projective normality assumption of Theorem 5.1, $`R^i\rho _{}𝒪_{E_1}((k+3)H\mathrm{}E_1E_2)=0`$ for all $`i,\mathrm{}>0`$ since $`E_1X`$ is flat. Ampleness of $`𝒪_{^{s_0}}(H)`$ implies $`H^1(E_1,𝒪_{E_1}(mH\mathrm{}E_1E_2))=0`$ for all $`mm_0`$, where $`m_0`$ may depend on $`\mathrm{}`$. From the exact sequence
$$0𝒪_{_{(0,1,2)}}(mH(\mathrm{}+1)E_1E_2)𝒪_{_{(0,1,2)}}(mH\mathrm{}E_1E_2)𝒪_{E_1}(mH\mathrm{}E_1E_2)0$$
a finite induction shows that if $`H^1(_{(0,1,2)},𝒪_{_{(0,1,2)}}(mH(\mathrm{}+1)E_1E_2))=0`$ for $`m0`$, some $`\mathrm{}>1`$ then $`H^1(_{(0,1,2)},𝒪_{_{(0,1,2)}}((k+3)H3E_1E_2))=0`$ for all $`k0`$.
As $`K_{_{(0,1,2)}}=𝒪_{_{(0,1,2)}}((s_01)H+(s_0r1)E_1+(s_02r2)E_2)`$, we have
$$𝒪_{_{(0,1,2)}}(mH(\mathrm{}+1)E_1E_2K)=𝒪_{_{(0,1,2)}}((m+s_0+1)H(\mathrm{}+s_0r)E_1(s_02r1)E_2)$$
As soon as $`\mathrm{}s_03r2`$, the right side is $`\rho `$-nef and, because $`\rho `$ is birational, the restriction of the right side to the general fiber of $`\rho `$ is big. Hence by \[Ko, 2.17.3\], $`R^i\rho _{}𝒪_{_{(0,1,2)}}(mH(\mathrm{}+1)E_1E_2)=0`$ for $`i1`$. Again by the ampleness of $`𝒪_{^{s_0}}(H)`$, we have $`H^1(_{(0,1,2)},𝒪_{_{(0,1,2)}}(mH(\mathrm{}+1)E_1E_2))=0`$ for $`m0`$, $`\mathrm{}`$ as above.
Step 5: The map $`H^0(E_1,𝒪_{E_1}((k+3)H2E_1E_2))H^0(F_p,𝒪_{F_p}(2H^{}E^{}))`$ is surjective for all $`k0`$. This is immediate by Lemma 5.6 and the projective normality assumption of Theorem 5.1. ∎
As in Theorem 2.3, we show that the restriction of $`\stackrel{~}{\phi _2}^+`$ to the transform of $`Sec^2X`$ is a projective bundle over $`^3X`$. By a slight abuse of notation, write $`\stackrel{~}{Sec^2X}_{(1,2)}`$ for the image of the proper transform of $`Sec^2X`$. Note the following:
###### Lemma 5.8.
Let $`S_Z=(\stackrel{~}{\phi _2}^+)^1(Z)^2`$ be a fiber over a point $`Z^3X`$. Then $`𝒪_{S_Z}(H)=𝒪_^2(2)`$ and $`𝒪_{S_Z}(E)=𝒪_^2(3)`$.
###### Proof.
This is immediate from the restrictions $`𝒪_{S_Z}(2HE)=𝒪_^2(1)`$ and $`𝒪_{S_Z}(3H2E)=𝒪_^2`$
###### Lemma 5.9.
There exists a morphism $`\stackrel{~}{Sec^2X}𝔾(2,s_0)`$ whose image is $`^3X`$.
###### Proof.
A point $`p\stackrel{~}{Sec^2X}`$ determines a unique $`2`$-plane $`S_Z`$ in $`\stackrel{~}{Sec^2X}`$ by Theorem 5.1. For every such $`p`$, the homomorphism $`H^0(_{(1,2)},𝒪_{_{(1,2)}}(H))H^0(_{(1,2)},𝒪_{S_Z}(H))`$ has rank $`3`$, hence gives a point in $`𝔾(2,s_0)`$. The image of the associated morphism clearly coincides with the natural embedding of $`^3X`$ into $`𝔾(2,s_0)`$ described in \[CG\]. ∎
As in \[V2, 3.5\], there is a morphism $`^3X_{(2)}`$ so that the composition factors $`\stackrel{~}{\phi _2}^+:\stackrel{~}{Sec^2X}_{(2)}`$. This is constructed by associating to every $`Z^3X`$ the rank $`1`$ homomorphism:
$$H^0(_{(1,2)},𝒪_{_{(1,2)}}(3H2E))H^0(_{(1,2)},𝒪_{S_Z}(3H2E))$$
where $`S_Z`$ is the $`^2`$ in $`_{(1,2)}`$ associated to $`Z`$.
Exactly as in Theorem 2.3, this allows the identification of $`\stackrel{~}{Sec^2X}`$ with a $`^2`$-bundle over $`^3X`$. Specifically, $`_2=(\stackrel{~}{\phi _2}^+)_{}(𝒪_{\stackrel{~}{Sec^2X}}(2HE))`$ is a rank $`3`$ vector bundle on $`^3X`$ and:
###### Proposition 5.10.
With notation as above, $`\stackrel{~}{\phi _2}^+:\stackrel{~}{Sec^2X}^3X`$ is the $`^2`$-bundle $`_{^3X}(_2)^3X`$. $`\mathrm{}`$
We wish to show further that blowing up $`Sec^2X`$ along $`X`$ and then along $`SecX`$ resolves the singularities of $`Sec^2X`$. By Theorem 2.4, $`h_1:_{(0,1,2)}_{(1,2)}`$ is the blow up of $`_{(1,2)}`$ along $`()`$, hence it suffices to show $`()(_2)`$ is a smooth subvariety of $`(_2)`$.
###### Proposition 5.11.
$`𝒟=()(_2)`$ is the nested Hilbert scheme $`Z_{2,3}(X)^2X\times ^3X`$, hence is smooth. Therefore $`\mathrm{Bl}_{\stackrel{~}{SecX}}(\mathrm{Bl}_X(Sec^2X))_{(0,1,2)}`$ is smooth and $`Sec^2X^{s_0}`$ is normal.
###### Proof.
Let $`𝒰_iX\times ^iX`$ denote the universal subscheme. We have morphisms $`\stackrel{~}{\phi _2}^+:𝒟^3X`$ and $`\stackrel{~}{\phi _1}^{}:𝒟^2X`$, and it is routine to check that $`(id_X\times \stackrel{~}{\phi _1}^{})^1(𝒰_2)(id_X\times \stackrel{~}{\phi _2}^+)^1(𝒰_3)`$. Hence (Cf. \[L, §1.2\]) $`\stackrel{~}{\phi _1}^{}\times \stackrel{~}{\phi _2}^+`$ maps $`𝒟`$ to the nested Hilbert scheme $`Z_{2,3}(X)^2X\times ^3X`$, where closed points of $`Z_{2,3}(X)`$ correspond to pairs of subschemes $`(\alpha ,\beta )`$ with $`\alpha \beta `$. Furthermore, via the description of the structure of the map $`\stackrel{~}{\phi _2}^+`$, it is clear that the morphism of $`^3X`$-schemes $`𝒟Z_{2,3}(X)`$ is finite and birational. It is shown in \[C, 0.2.1\] that $`Z_{2,3}(X)`$ is smooth, hence this is an isomorphism. ∎
Let $`_{(1,2,3)}`$ be the blow up of $`_{(1,2)}`$ along $`(_2)`$; note $`_{(1,2,3)}`$ is smooth. To construct $`_{(2,3)}`$, we first construct the exceptional locus as a projective bundle over $`^3X`$. Write $`\mathrm{Pic}_{(1,2,3)}=H+E_1+E_3`$.
###### Lemma 5.12.
Let $`p_3:E_3^3X`$ be the composition $`E_3(_2)^3X`$. Then $`_2=(p_3)_{}𝒪_{E_3}(4H3E_1E_3)`$ is locally free of rank $`s_03r2=codim(\stackrel{~}{Sec^2X},_{(1,2)})`$.
###### Proof.
Each fiber $`F_x`$ of $`p_3`$ is isomorphic to $`^2\times ^t`$, $`t+1=codim(\stackrel{~}{Sec^2X},_{(1,2)})`$. Furthermore $`H^0(F_x,𝒪_{F_x}(4H3E_1E_3))=H^0(^t,𝒪__t(1))`$ follows easily from Lemma 5.8. ∎
There is a map $`E_3(_2)`$ given by the surjection
$$p_3^{}_2𝒪_{E_3}(4H3E_1E_3)0$$
hence a diagram of exceptional loci:
It is important to note that
$$(_2)(p_3_{}𝒪_{E_3}(4H3E_1E_3+t(3H2E_1)))$$
for all $`t0`$ as the direct image on the right will differ from $`_2`$ by a line bundle. Hence for all $`t0`$ the same morphism $`E_3(_2)`$ is induced by the surjection
$$p_3^{}p_3_{}𝒪_{E_3}(4H3E_1E_3+t(3H2E_1))𝒪_{E_3}(4H3E_1E_3+t(3H2E_1))$$
One can now repeat almost verbatim \[V2, 4.7-4.10\] to construct the second flip; i.e. the space $`_{(2,3)}`$. Recall the following:
###### Proposition 5.13.
\[V2, 4.5\] Let $``$ be an invertible sheaf on a complete variety $`X`$, and let $``$ be any locally free sheaf. Assume that the map $`\lambda :XY`$ induced by $`||`$ is a birational morphism and that $`\lambda `$ is an isomorphism in a neighborhood of $`pX`$. Then for all $`n`$ sufficiently large, the map
$$H^0(X,^n)H^0(X,^n𝒪_p)$$
is surjective. $`\mathrm{}`$
Taking $`=𝒪_{_{(1,2,3)}}(4H3E_1E_3)`$ and $`=𝒪_{_{(1,2,3)}}(3H2E_1)`$, the map induced by the linear system associated to
$$𝒪_{_{(1,2,3)}}((4H3E_1E_3)+(k2)(3H2E_1))=𝒪_{_{(1,2,3)}}((3k2)H(2k1)E_1E_3)$$
is base point free off $`E_3`$ for $`k3`$. To show this gives a morphism, one shows the restriction of above linear system to the divisor $`E_3`$ induces a surjection on global sections, hence restricts to the map $`E_3(_2)`$ above. For this, define $`_\rho =𝒪((3\rho 2)H(2\rho 1)E_1E_3)`$ and write
$$𝒪_{_{(1,2,3)}}((3k2)H(2k1)E_12E_3)K_{_{(1,2,3)}}^1=_\alpha ^{s_03r1}𝒜$$
where $`\alpha =\frac{2k+s_0r1}{2s_06r2}`$ and $`𝒜=𝒪_{_{(1,2,3)}}\left(\left(\frac{3s_09r4}{2}\right)H(s_03r2)E_1\right)`$. By the above discussion, $`_\alpha ^{s_03r1}`$ is nef for $`k0`$ and it is routine to verify that $`𝒜`$ is a big and nef $``$-divisor; hence $`H^1(_{(1,2,3)},𝒪((3k2)H(2k1)E_12E_3))=0`$.
The variety $`_{(2,3)}`$ is defined to be the image of this morphism. This gives:
###### Proposition 5.14.
With hypotheses as in Theorem 5.1 and for $`k`$ sufficiently large, the morphism $`h_2:_{(1,2,3)}_{(2,3)}`$ induced by the linear system $`|_k|`$ is an embedding off of $`E_3`$ and the restriction of $`h_2`$ to $`E_3`$ is the morphism $`E_3(_2)`$ described above. $`\mathrm{}`$
Remark 5.15 The best (smallest) possible value for $`k`$ is $`k=3`$. This will be the case if $`Sec^3X^{s_0}`$ is scheme theoretically cut out by quartics. $`\mathrm{}`$
###### Lemma 5.16.
$`_{(2,3)}`$ is smooth.
###### Proof.
Because $`_{(2,3)}`$ is the image of a smooth variety with reduced, connected fibers it is normal (Cf. \[V1, 3.2.5\]). Let $`Z^2`$ be a fiber of $`h_2`$ over a point $`p(_2)`$. $`Z\times \{p\}`$ is a fiber of a $`^2\times ^t`$ bundle over $`^3X`$, hence the normal bundle sequence becomes:
$$0\underset{1}{\overset{s_03}{}}𝒪_^2N_{Z/_{(1,2,3)}}𝒪_^2(1)0$$
This sequence splits, and allowing the elementary calculations $`H^1(Z,S^rN_{Z/_{(1,2,3)}})=0`$ and $`H^0(Z,S^rN_{Z/_{(1,2,3)}})=S^rH^0(Z,N_{Z/_{(1,2,3)}})`$ for all $`r1`$, $`_{(2,3)}`$ is smooth by a natural extension of the smoothness portion of Castelnuovo’s contractibility criterion for surfaces given in \[AW, 2.4\]. ∎
Letting $`(_0)=_X(N_{X/^{s_0}}^{})=E_1`$ and $`(_0)=X`$, the analogue of Theorem 2.4 is:
###### Theorem 5.17.
Let $`X_{(0)}=^{s_0}`$ be a smooth, irreducible variety of dimension $`r`$ that satisfies $`(K_2^1)`$, with $`s_03r+4`$. Assume that $`X`$ is embedded by a complete linear system $`|L|`$ and that the following conditions are satisfied:
1. $`L`$ is $`(5+r)`$-very ample and $`Sec^2X=Sec_3^1(Sec^1X)`$ as schemes
2. The projection of $`X`$ into $`^m,m=s_01r`$, from any embedded tangent space is such that the image is projectively normal and satisfies $`(K_2)`$
3. If $`r2`$, then for every point $`pX`$, $`H^1(X,L_p^3)=0`$
Then there is a pair of flips as pictured below with:
1. $`_{(i,i+1)}`$ and $`_{(i,i+1,i+2)}`$ smooth
2. $`_{(i,i+1)}(_{i+1})_{(i+1,i+2)}(_{i+1})`$; as $`s_03r+4`$, $`\mathrm{Pic}_{(0,1)}\mathrm{Pic}_{(i+1,i+2)}`$
3. $`_i=\stackrel{~}{\phi _i_{}}^+𝒪_{\stackrel{~}{Sec^iX}}(iH(i1)E)`$ and $`_i=\stackrel{~}{\phi _i_{}}^{}𝒪_{\stackrel{~}{Sec^iX}}((i+2)H(i+1)E)`$
4. $`h_i`$ is the blow up of $`_{(i,i+1)}`$ along $`(_i)`$
5. $`_{(i,i+1,i+2)}_{(i,i+1)}`$ is the blow up along $`(_{i+1})`$
6. $`\stackrel{~}{\phi _i}^{}`$, induced by $`𝒪_{_{(i,i+1)}}((i+1)HiE)`$, is an embedding off of $`(_i)`$, and the restriction of $`\stackrel{~}{\phi _i}^{}`$ is the projection $`(_i)^{i+1}X`$
7. $`\stackrel{~}{\phi _i}^+`$, induced by $`𝒪_{_{(i1,i)}}((i+1)HiE)`$, is an embedding off of $`(_i)`$, and the restriction of $`\stackrel{~}{\phi _i}^+`$ is the projection $`(_i)^{i+1}X`$
8. $`(_i)(E_{i+1})_{(i,i+1)}`$ is isomorphic to the nested Hilbert scheme $`Z_{i+1,i+2}^iX\times ^{i+1}X`$, hence is smooth.
$`\mathrm{}`$ |
no-problem/9911/cond-mat9911374.html | ar5iv | text | # Monopole, half-quantum vortices and nexus in chiral superfluids and superconductors.
## Abstract
Two exotic objects are still not identified experimentally in chiral superfluids and superconductors. These are the half-quantum vortex, which plays the part of the Alice string in relativistic theories , and the hedgehog in the $`\widehat{𝐥}`$ field, which is the counterpart of the Dirac magnetic monopole. These two objects of different dimensionality are topologically connected. They form the combined object which is called nexus in relativistic theories . Such combination will allow us to observe half-quantum vortices and monopoles in several realistic geometries.
In relativistic quantum fields nexus is the monopole, in which $`N`$ vortices of the group $`Z_N`$ meet at a center (nexus) provided the total flux of vortices adds to zero (mod $`N`$) . In a chiral superfluid with the order parameter of the <sup>3</sup>He-A type, the analog of the nexus is the hedgehog in the $`\widehat{𝐥}`$ field, in which 4 vortices meet, each with the circulation quantum number $`N=1/2`$. The total topological charge of the four vortices is $`N=2`$ which is equivalent to $`N=0`$ because the homotopy group, which describes the <sup>3</sup>He-A vortices, is $`\pi _1=Z_4`$ , and thus $`N=0`$ (mod 2). Each $`N=1/2`$ vortex is the $`1/4`$ fraction of the ”Dirac string” in <sup>3</sup>He-A, the latter is the $`N=2`$ vortex terminating on the hedgehog . The hedgehog in the $`\widehat{𝐥}`$ field plays a part of the Dirac magnetic monopole: The distribution of the vector potential of the electromagnetic field $`𝐀`$ in the vicinity of the hedgehog in the electrically charged version of the <sup>3</sup>He-A (the chiral $`p`$-wave superconductor) is similar to that in the vicinity of magnetic monopole (see e.g. ).
The order parameter describing the vacuum manifold in chiral $`p`$-wave superfluid/superconductor (<sup>3</sup>He-A and also possibly the layered superconductor Sr<sub>2</sub>RuO<sub>4</sub> ) is
$$A_{\alpha i}=\mathrm{\Delta }\widehat{d}_\alpha (\widehat{e}_i^{(1)}+i\widehat{e}_i^{(2)}).$$
(1)
Here $`\widehat{𝐝}`$ is the unit vector of the spin-space anisotropy; $`\widehat{𝐞}^{(1)}`$ and $`\widehat{𝐞}^{(2)}`$ are unit mutually orthogonal vectors in the orbital space, they determine the superfluid velocity of the chiral condensate $`𝐯_s=\frac{\mathrm{}}{2m}\widehat{e}_i^{(1)}\widehat{e}_i^{(2)}`$, where $`2m`$ is the mass of the Cooper pair; the orbital momentum vector is $`\widehat{𝐥}=\widehat{𝐞}^{(1)}\times \widehat{𝐞}^{(2)}`$. The half-quantum vortex results from the identification of the points $`\widehat{𝐝}`$ , $`\widehat{𝐞}^{(1)}+i\widehat{𝐞}^{(2)}`$ and $`\widehat{𝐝}`$ , $`(\widehat{𝐞}^{(1)}+i\widehat{𝐞}^{(2)})`$, which correspond to the same order parameter Eq.(1). It is the combination of the $`\pi `$-vortex and $`\pi `$-disclination in the $`\widehat{𝐝}`$ field:
$$\widehat{𝐝}=\widehat{𝐱}\mathrm{cos}\frac{\varphi }{2}+\widehat{𝐲}\mathrm{sin}\frac{\varphi }{2},\widehat{𝐞}^{(1)}+i\widehat{𝐞}^{(2)}=e^{i\varphi /2}(\widehat{𝐱}+i\widehat{𝐲}),$$
(2)
where $`\varphi `$ is the azimuthal angle around the string.
The hedgehog in the orbital momentum field, $`\widehat{𝐥}=\widehat{𝐫}`$, produces the superfluid velocity field (or the vector potential in the corresponding superconductor):
$$𝐯_s=\frac{e}{mc}𝐀,𝐀=\underset{a}{}𝐀^a,$$
(3)
where $`𝐀^a`$ is the vector potential for the Dirac monopole with the $`a`$-th Dirac string, $`N_a`$ is the topological charge (number of circulation quanta) of the $`a`$-th string. Choosing the spherical coordinate system $`(r,\theta ,\varphi )`$ in such a way that the string $`a`$ occupies the lower half-axis $`z<0`$, the vector potential $`𝐀^a`$ of such string can be written as :
$$𝐀^a=\frac{\mathrm{}c}{4er}N_a\widehat{\varphi }\frac{1\mathrm{cos}\theta }{\mathrm{sin}\theta },$$
(4)
The superfluid vorticity and the corresponding magnetic field in superconductor are
$`\times 𝐯_s={\displaystyle \frac{\mathrm{}}{4m}}{\displaystyle \frac{𝐫}{r^3}}{\displaystyle \underset{a}{}}N_a+{\displaystyle \frac{h}{2m}}{\displaystyle \underset{a}{}}N_a{\displaystyle _0^R}𝑑r\delta (𝐫𝐫_a(r)),`$ (5)
$`𝐁={\displaystyle \frac{\mathrm{}c}{4e}}{\displaystyle \frac{𝐫}{r^3}}{\displaystyle \underset{a}{}}N_a+{\displaystyle \frac{hc}{2}}{\displaystyle \underset{a}{}}N_a{\displaystyle _0^R}𝑑r\delta (𝐫𝐫_a(r)),{\displaystyle \underset{a}{}}N_a=2.`$ (6)
Here $`𝐫_a(r)`$ is the position of the $`a`$-th line, assuming that the lines are emanating radially from the monopole, i.e. the coordinate along the line is the radial coordinate. The regular part of the magnetic field corresponds to the monopole with the magnetic charge $`g=\mathrm{}c/2e`$, the magnetic flux $`4\pi g`$ of the monopole is supplied by the Abrikosov vortices. The lowest energy of the monopole occurs when all the vortices emanating from the monopole have the lowest circulation number: this means that there must be four vortices with $`N_1=N_2=N_3=N_4=1/2`$.
The half-quantum vortices are accompanied by the spin disclinations. Assuming that the $`\widehat{𝐝}`$-field is confined in the plane the disclinations can be characterized by the winding numbers $`\nu _a`$ which have values $`\pm 1/2`$ in half-quantum vortices. The corresponding spin-superfluid velocity $`𝐯_{sp}`$ is
$$𝐯_{sp}=\frac{e}{mc}\underset{a=1}{\overset{4}{}}\nu _a𝐀^a,\underset{a=1}{\overset{4}{}}\nu _a=0,$$
(7)
where the last condition means the absence of the monopole in the spin sector of the order parameter. Thus we have $`\nu _1=\nu _2=\nu _3=\nu _4=1/2`$.
The spin-orbit coupling can be neglected if the size of the bubble is less than spin-orbit length (about $`10\mu m`$ in <sup>3</sup>He-A). Assuming that the superfluid velocity is everywhere perpendicular to $`\widehat{𝐥}`$ and has a form $`𝐯_s=\stackrel{~}{𝐯}_s(\theta ,\varphi )/r`$, the energy of the nexus in the spherical bubble of radius $`R`$ is
$`E={\displaystyle _0^R}r^2𝑑r{\displaystyle 𝑑\mathrm{\Omega }\left(\frac{1}{2}\rho _s𝐯_s^2+\frac{1}{2}\rho _{sp}𝐯_{sp}^2\right)}=R{\displaystyle 𝑑\mathrm{\Omega }\left(\frac{1}{2}\rho _s\stackrel{~}{𝐯}_s^2+\frac{1}{2}\rho _{sp}\stackrel{~}{𝐯}_{sp}^2\right)}=`$ (8)
$`{\displaystyle \frac{1}{2}}R{\displaystyle 𝑑\mathrm{\Omega }\left((\rho _s+\rho _{sp})\left[(\stackrel{~}{𝐀}^1+\stackrel{~}{𝐀}^2)^2+(\stackrel{~}{𝐀}^3+\stackrel{~}{𝐀}^4)^2\right]+2(\rho _s\rho _{sp})(\stackrel{~}{𝐀}^1+\stackrel{~}{𝐀}^2)(\stackrel{~}{𝐀}^3+\stackrel{~}{𝐀}^4)\right)},\stackrel{~}{𝐀}^a(\theta ,\varphi )={\displaystyle \frac{mcr}{e}}𝐀^a.`$ (9)
In the simplest case, which occurs in the ideal Fermi gas approximation when the Fermi liquid corrections are neglected, one has $`\rho _s=\rho _{sp}`$ . In this case the $`1/2`$-vortices with positive spin current circulation $`\nu `$ do not interact with $`1/2`$-vortices with negative $`\nu `$. The energy minimum occurs when the orientations of two positive-$`\nu `$ vortices are opposite, so that these two $`\frac{1}{4}`$ fractions of the Dirac strings form one line along the diameter (see Fig.1). The same happens for the other fractions with negative $`\nu `$. The mutual orientations of the two diameters is arbitrary in this limit. However, in real <sup>3</sup>He-A one has $`\rho _{sp}<\rho _s`$ . If $`\rho _{sp}`$ is slightly smaller than $`\rho _s`$, the positive-$`\nu `$ and negative-$`\nu `$ strings repel each other, so that the equilibrium angle between them is $`\pi /2`$. In the extreme case $`\rho _{sp}\rho _s`$, the ends of four half-quantum vortices form vertices of a regular tetrahedron.
Such monopole can be experimentally realized in the mixed <sup>4</sup>He/<sup>3</sup>He droplets obtained via the nozzle beam expansion of the He gases . The <sup>4</sup>He component of the mixture forms the cluster in a central region of the droplet . If the size of the cluster is comparable with the size of the droplet, the radial distribution of the $`\widehat{𝐥}`$ vector is stabilized by the boundary conditions on the the surface of the droplet and on the boundary of the cluster (see Fig.1). The <sup>4</sup>He cluster plays the part of the core of the nexus. The half-quantum vortices emanating from the nexus are well defined if the radius of the droplet exceeds the coherence length $`\xi 200500\AA `$.
In a $`p`$-wave superconductor such monopole will be formed in a thin spherical layer. In Sr<sub>2</sub>RuO<sub>4</sub> superconductor the spin-orbit coupling between the spin vector $`\widehat{𝐝}`$ and crystal lattice seems to align the $`\widehat{𝐝}`$ vector along $`\widehat{𝐥}`$. In this case the half quantum vortices are energetically unfavourable, and instead of 4 half-quantum vortices one would have 2 singly quantized vortices in the spherical shell.
Monopole of this kind can be formed also in the so called ferromagnetic Bose condensate in optical traps. Such condensate is described by vector or spinor chiral order parameter .
There are interesting properties of the system related to the fermionic spectrum of such objects. In particular, the number of fermion zero modes on $`N=1/2`$ vortex under discussion is twice less than that on the vortex with $`N=1`$. This is because such $`N=1/2`$ vortex can be represented as the $`N=1`$ vortex in one spin component with no vortices in another spin component. Thus, according to , in the core of the $`N=1/2`$ vortex there is one fermionic level (per 2D layer) with exactly zero energy. Since the zero-energy level can be either filled or empty, there is an entropy $`(1/2)\mathrm{ln}2`$ per layer related to the vortex. The factor $`(1/2)`$ appears because the particle excitation coincides with the antiparticle (hole) excitation in superconductors, i.e. the quasiparticle is a Majorana fermion, see also . Such fractional entropy also arises in the Kondo problem. According to , the $`N=1`$ vortex has spin $`S=1/4`$ per layer, this implies the spin $`S=1/8`$ per layer for $`N=1/2`$ vortex. Similarly the anomalous fractional charge of the $`N=1/2`$ vortex is 1/2 of that discussed for the $`N=1`$ vortex .
I thank M. Feigel’man and D. Ivanov for discussions. This work was supported in part by the Russian Foundations for Fundamental Research and by European Science Foundation. |
no-problem/9911/nucl-th9911065.html | ar5iv | text | # Scaling law for the electromagnetic form factors of the proton
## Abstract
The violation of the scaling law for the electric and magnetic form factors of the proton is examined within the cloudy bag model. We find that the suppression of the ratio of the electric and magnetic form factors is natural in the bag model. The pion cloud plays a moderate role in understanding the recent data from TJNAF.
The description of the electromagnetic structure of the nucleon requires two independent form factors. The usual Sachs form factors fully characterize the charge and current distributions inside the nucleon. Electromagnetic probes interact not only with valence quarks confined inside a quark core by nonlinear, gluon dynamics but also with the pion field required by chiral symmetry. A full understanding of the electromagnetic structure of the nucleon is of fundamental importance.
Historically, experimental determination of the electric ($`G_{Ep}`$) and magnetic ($`G_{Mp}`$) form factors of the proton was mainly based on the Rosenbluth separation of the unpolarized differential cross section data. The results of various analyses from the early experiments are summarized in a simple scaling law,
$$G_{Ep}(Q^2)=G_{Mp}(Q^2)/\mu _p=G_D(Q^2),$$
(1)
for momentum transfers, $`Q`$, up to several GeV. Here $`\mu _p`$ is proton’s magnetic moment and $`G_D(Q^2)`$ refers to the standard dipole form. However, from the Rosenbluth formula one sees that at large momentum transfer, the electric contribution to the cross section is kinematically suppressed relative to the magnetic contribution. Thus $`G_{Ep}`$ can not be determined as accurately as $`G_{Mp}`$ from such an analysis, especially at large $`Q^2`$. In the literature, the ratios, $`\mu _pG_{Ep}(Q^2)/G_{Mp}(Q^2)`$, obtained from different experiments are not consistent with each other, within the quoted errors.
With the advance of polarization technologies and the operation of high-duty electron machines, it is now possible to drastically reduce the systematic uncertainties in this ratio by direct measurement. That is, one can simultaneously measure the two components of the recoil proton polarization, $`P_T`$ and $`P_L`$, using a longitudinally polarized electron beam. As the transverse component behaves as $`P_TG_{Ep}G_{Mp}`$, and the longitudinal component as $`P_LG_{Mp}^2`$, the ratio, $`\mu _pG_{Ep}/G_{Mp}`$, can be determined directly from $`P_T/P_L`$.
Precise data for the nucleon electromagnetic form factors sets a strong constraint for various quark models of the nucleon. It helps us to develop an understanding of the composite nature of the nucleon as well as its long range chiral structure. In our previous work, we studied the nucleon electromagnetic form factors in an improved cloudy bag model (CBM) , where the center-of-mass motion correction and relativistic effects were treated explicitly. As a result, the region of validity of the calculation of the electric and magnetic form factors in the model was extended to larger momentum transfer than had previously been possible. Here we focus on the ratio, $`\mu _pG_{Ep}(Q^2)/G_{Mp}(Q^2)`$, and examine the mechanism for the violation of the scaling law, Eq. (1).
Let us start with the MIT bag model . Under the static cavity approximation, the bag surface is spherical and all valence quarks are in the lowest eigenmode. The electric and magnetic form factors for the proton can be written as
$`G_{Ep}(Q^2)`$ $`=`$ $`{\displaystyle _0^R}4\pi r^2𝑑rj_0(Qr)[g^2(r)+f^2(r)],`$ (2)
$`G_{Mp}(Q^2)`$ $`=`$ $`2m_N{\displaystyle _0^R}4\pi r^2𝑑r{\displaystyle \frac{j_1(Qr)}{Q}}[2g(r)f(r)],`$ (3)
where $`R`$ is the bag radius, $`Q^2=q^2=\stackrel{}{q}^{\mathrm{\hspace{0.17em}2}}`$ with $`\stackrel{}{q}`$ the three momentum of the photon in the Breit frame, and $`j_l(x)`$ refers to the spherical Bessel function. For a massless quark, the quark wave functions are given as $`g(r)=N_Sj_0(\omega _Sr/R)`$ and $`f(r)=N_Sj_1(\omega _Sr/R)`$, with $`\omega _S=2.04`$ and $`N_S^2=\omega _S/8\pi R^3j_0^2(\omega _S)(\omega _S1)`$.
At small $`Q^2`$, the ratio $`\mu _pG_{Ep}(Q^2)/G_{Mp}(Q^2)`$ is determined by the electromagnetic root-mean-squared (r.m.s.) radius, which is defined as followes: $`G_{E,M}(Q^2)1Q^2r^2_{E,M}/6`$, as $`Q0`$. A direct evaluation gives
$`r^2_{Ep}`$ $`=`$ $`4\pi N_S^2\left({\displaystyle \frac{R}{\omega _S}}\right)^5{\displaystyle \frac{\omega _S^2(2\omega _S^32\omega _S^2+4\omega _S3)}{3(2\omega _S^22\omega _S+1)}},`$ (4)
$`r^2_{Mp}`$ $`=`$ $`{\displaystyle \frac{8\pi N_S^2}{5\mu _p}}\left({\displaystyle \frac{R}{\omega _S}}\right)^6{\displaystyle \frac{\omega _S^2(8\omega _S^3+10\omega _S^220\omega _S+15)}{24(2\omega _S^22\omega _S+1)}}.`$ (5)
Note that $`\mu _p=R(4\omega _S3)/12\omega _S(\omega _S1)`$, thus the ratio $`r^2_{Ep}/r^2_{Mp}`$ is independent of $`R`$. It results in 1.36 using $`\omega _S=2.04`$. This means that the charge r.m.s. radius is considerably larger than the magnetic r.m.s. radius. In other words, $`G_{Ep}(Q^2)`$ decreases faster than $`G_{Mp}(Q^2)`$ at small momentum transfers, so the suppression of $`\mu _pG_{Ep}(Q^2)/G_{Mp}(Q^2)`$, as $`Q^2`$ increases, is a natural consequence of the MIT bag model.
The results of our numerical calculations are shown in Fig. 1, where $`R=0.8`$ fm is used. The long-dashed line is the ratio in the static MIT bag model. Two lowest curves are respectively electric and magnetic form factors of the proton within the MIT bag model. The experimental data are from the recent measurement at TJNAF . Note that the naive MIT bag model for the nucleon is only sensible in the region of very small momentum transfers. The ratio in a static bag calculation deviates from the data very quickly as $`Q^2`$ increases. The singularity in the long-dashed curve comes from a node in $`G_{Mp}(Q^2)`$, which is related to the sharp bag surface, and will be extremely sensitive to small admixtures in the nucleon ground state wave function. Such admixtures may be induced by residual one-gluon exchange or meson exchange interactions. More sophiscated models, such as the color dielectric and soliton bag models , would also help to remove the zero (or at least move it to larger $`Q^2`$), thus removing the singularity in the ratio.
As the bag model is inherently an independent particle model, the spurious center-of-mass motion must be removed in a realistic calculation. We have incorporated this correction by constructing momentum eigenstates using the Peierls-Thouless projection method. Such wave functions are consistent with Galilean translational invariance. In addition, since the static solution of the quark wave function is spherical, the calculated form factors are trustworthy only at very small momentum transfer. To account for relativistic effects we use the prescription proposed by Licht and Pagnementa . The technical details for implementing such corrections can be found in Ref. . The resulting ratio for the quark core of the bag model, after including the center-of-mass motion corrections and relativistic effects, is plotted as the solid curve in Fig. 1. The prominent zeros in the form factors are forced out much further and do not occur in the momentum region we are studying. Compared with the recent data, the improvement over the naive MIT bag model is significant, in particular, away from the region of small momentum transfers.
In the cloudy bag model the physical nucleon is a superposition of a quark core and a quark core plus a meson cloud . We keep only the dominant pion terms and treat them in the one loop approximation. Inside the pion loop, the intermediate baryon is either a $`N(939)`$ or a $`\mathrm{\Delta }(1232)`$. Thus the electromagnetic form factors receive contributions from the pion current in addition to the direct photon coupling to the confined quarks in the core. For more details and the explicit expressions for the form factors, we refer the readers to Eqs.(36-40) and Eqs.(A5,A6) in Ref. .
In our calculations the $`\pi NN`$ coupling constant is taken to be $`f_{\pi NN}^2=0.0791`$. Other coupling constants are fixed by SU(6) symmetry. The resulting ratios in the full calculations are presented in Fig. 2. Here the solid and long-dashed curves are respectively for massless and 10 MeV quarks in the nucleon bag, using a bag radius of $`R=0.8`$ fm. Clearly the ratio is insensitive to the quark mass variations. The addition of the pion cloud for $`R=0.8`$ fm increases the ratio by roughly 20% towards the recent data. In the same figure, the dot-dashed and dashed curves have similar meanings but with $`R=0.7`$ fm. This is about the smallest bag radius which allows a perturbative treatment of the pion cloud. The recent data indicate that a smaller bag radius is prefered.
One feature of our calculation is that the decline of the ratio of the Sachs form factors slows down as $`Q^2>2`$ Gev<sup>2</sup>. The corresponding ratio of the Pauli and Dirac form factors $`Q^2F_{2p}/F_{1p}`$ shows a rapid rise at small $`Q^2`$ (below 2 GeV<sup>2</sup>) and seemingly become flat at large $`Q^2`$. This is consistent with the prediction of pQCD, which claims $`F_{1p}1/Q^4`$ and $`F_{2p}1/Q^6`$, but is not obvious in other nucleon models .
In summary, we have studied the ratio, $`\mu _pG_{Ep}(Q^2)/G_{Mp}(Q^2)`$, in an improved quark model. We find that the sharp decline of the ratio as $`Q^2`$ increases is natural in a quark bag model, so the empirical scaling law for the nucleon electromagnetic form factors is unavoidably violated. The deviations of the ratio, $`\mu _pG_{Ep}(Q^2)/G_{Mp}(Q^2)`$ from unity can be understood semi-quantitatively within the cloudy bag model.
This work was supported in part by the National Science Council of ROC under grant No. NSC-89-2112-M002-038 and the Australian Research Council. |
no-problem/9911/chao-dyn9911025.html | ar5iv | text | # General properties of propagation in chaotic systems
## Abstract
We conjecture that in one-dimensional spatially extended systems the propagation velocity of correlations coincides with a zero of the convective Lyapunov spectrum. This conjecture is successfully tested in three different contexts: (i) a Hamiltonian system (a Fermi-Pasta-Ulam chain of oscillators); (ii) a general model for spatio-temporal chaos (the complex Ginzburg-Landau equation); (iii) experimental data taken from a $`CO_2`$ laser with delayed feedback. In the last case, the convective Lyapunov exponent is determined directly from the experimental data.
Since the recognition that deterministic chaos is an ubiquitous feature of nonlinear systems, much has been understood about their dynamical properties. An important example is the discovery of the relationships between Lyapunov exponents, measuring the divergence rate of nearby trajectories, and either geometrical or information-theoretic properties such as fractal dimensions and dynamical entropies . Nevertheless, little progress has been made to establish a link between chaotic indicators and directly observable properties. In this Letter we show the existence of a general and remarkable connection: the velocity of correlation propagation is determined by the vanishing of the convective Lyapunov exponent.
On the one hand, we know that chaotic systems are generally characterized by rapidly decaying correlations. This does not prevent the onset of sizable propagation phenomena, as apparent in Fig. 1 where, as an example, we report the space-time representation of the local heat-flux in a Fermi-Pasta-Ulam (FPU) chain (see later for a definition of both the observable and the model). A simple tool to pinpoint the existence of travelling processes is the autocorrelation function $`C(x,t)=w(t^{}+t,x^{}+x)w(t^{},x^{})`$ (or, equivalently, the structure function) of some observable $`w(x,t)`$, where $`x`$ and $`t`$ denote space and time variables, respectively, and $``$ denotes a time average (we shall always assume that ergodicity holds). The most effective propagation processes can then be identified by estimating the velocity $`v=\overline{v}`$ that minimizes the decay of $`C(vt,t)`$. In many chaotic systems the minimum is attained for a non-zero velocity.
On the other hand, Lyapunov exponents represent the right tool to investigate the evolution of an infinitesimal perturbation $`\delta (x,t)`$. If the initial perturbation $`\delta (x,0)`$ is localized around the origin, it has been shown that, in the limit $`t\mathrm{}`$,
$$\delta (x,t)\mathrm{exp}[\mathrm{\Lambda }(v=x/t)t],$$
(1)
where $`\mathrm{\Lambda }(v)`$ is the so-called convective or velocity-dependent Lyapunov exponent. In fact, it expresses the growth rate of a localized perturbation when observed in a frame moving with velocity $`v`$. The maximum value of $`\mathrm{\Lambda }(v)`$ coincides with the usual maximum Lyapunov exponent: in convectively unstable systems, this occurs for a suitable non-zero velocity, while it is located in the orgin ($`v=0`$) in spatially symmetric chaotic systems. An example of the typical behaviour of $`\mathrm{\Lambda }(v)`$ in this latter context can be seen in the inset of Fig. 1. There, one can notice a further general feature: moving away from the maximum, $`\mathrm{\Lambda }(v)`$ decreases, becoming negative above some critical velocity $`v^{}`$. Therefore, any perturbation “travelling” faster than the critical velocity is exponentially damped and thus becomes quickly negligible. This fast damping prevents any coherent propagation of information and, therefore, we expect that a meaningful long-term coherence cannot be maintained for such large velocities.
In the opposite limit of small velocities, it is well known that an exponential separation of trajectories leads to a fast decoherence, since nearby orbits rapidly enter different regions of the phase-space. Accordingly, strong correlations cannot again be maintained. The only moving frame in which neither of the two effects actively contributes to destroying correlations is precisely that one corresponding to the neutral point $`v^{}`$, so that the most effective propagation phenomena should occur exactly at the velocity $`v=v^{}`$.
The first system where we have tested this conjecture is a chain of FPU oscillators , an idealized microscopic Hamiltonian model for an insulating solid,
$$\ddot{q}_i=F(q_{i1}q_i)F(q_iq_{i+1})$$
(2)
where $`q_i`$ represents the displacement of the $`i`$th particle from its equilibrium position and $`F(x)=xx^3`$ is the force field. This is a simple nonlinear model, introduced to test the ergodic hypothesis in the first numerical experiment ever performed . Within the large number of features that have been found while investigating the dynamics of the above model, here we are interested in the propagation phenomena recently discovered in connection with the study of heat conductivity. In order to clarify the observed anomalous transport properties, the authors of Ref. have performed microcanonical simulations (with periodic boundary conditions), monitoring the local heat flux
$$j(t,i)=\frac{\dot{q}_i}{2}[F(q_{i1}q_i)+F(q_iq_{i+1})].$$
(3)
From the behaviour of the autocorrelation function $`j(t^{}+t,i^{}+i)j(t^{},i^{})`$, they clearly found a propagation velocity $`\overline{v}2.47`$, when the energy per particle is $`e=8.8`$. The space-time representation of $`j(t,i)`$ in Fig. 1 demonstrates directly the existence of symmetric propagation phenomena along the directions identified by the cross-correlation analysis and denoted by the tilted arrows in the figure.
For what concerns the spectrum of convective Lyapunov exponents, rather than letting an initially localized perturbation evolve, we have preferred to follow the procedure devised in Ref. , as it suffers much less problems of finite-size corrections . The readers interested in a thorough explanation of the procedure can consult Refs. ; here, for the sake of completeness, we summarize the key steps. Very briefly, the method consists in computing the maximum Lyapunov exponent $`\lambda (\mu )`$ of a special class of perturbations: those exhibiting an exponential profile with an imposed growth rate $`\mu `$. This can be done by linearizing Eqs. (2) to obtain the standard equations for a generic perturbation $`\delta q_i`$. By then introducing the Ansatz $`\delta q_i=\overline{\delta q_i}\mathrm{exp}(\mu i)`$, one obtains the differential equation for the “envelope” $`\overline{\delta q_i}`$. The corresponding growth rate is the generalized Lyapunov exponent $`\lambda (\mu )`$. The velocity-dependent Lyapunov spectrum $`\mathrm{\Lambda }(v)`$ is finally obtained by Legendre transforming $`\lambda (\mu )`$ ($`v=d\lambda /d\mu `$; $`\mathrm{\Lambda }=\lambda (\mu )\mu v`$). Mutatis mutandis, $`\lambda (\mu )/\mu `$ can be intepreted as the “phase velocity” of waves with wavelength “$`\mu `$”, while $`v`$ can be read as the corresponding group velocity.
The result for the FPU system is reported in the inset of Fig. 1: it reveals two symmetric zeros, whose value corresponds to the direction of the straight lines visible in the corresponding pattern. However, more importantly, the absolute value of the marginal velocity ($`2.46`$) is in full agreement with the velocity previously estimated from the behaviour of the correlation function. As a first check of the generality of this identity, we have repeated the same analysis for different energy densities, namely $`e=1.`$ and $`e=0.1`$. The data are altogether reported in Table 1, where one can see that $`\overline{v}`$ always agrees with $`v^{}`$ within the numerical error.
The second model we have considered is the complex Ginzburg-Landau (CGL) equation, a partial differential equation describing chaotic properties of generic systems close to oscillatory instabilities (it has been derived, e.g., in hydrodynamics and laser physics),
$$\frac{u}{t}=u(1ib)|u|^2u+(1+ic)\frac{^2u}{x^2}$$
(4)
where $`u`$ is a complex field which, in general, represents the slowly-varying amplitude of some relevant mode. For $`b=2.3`$ and $`c=0`$, propagation phenomena are clearly visible, as it can be noticed in the pattern reported in Fig. 2 (the equations have been integrated by using the algorithm described in Ref. and adopting again periodic boundary conditions). It is instructive to notice that, at variance with the previous Hamiltonian system, now the pattern is no longer invariant under time reversal. This is an obvious consequence of the dissipative nature of the CGL equation. From the computation of the correlation function $`|u|^2(t+t^{},x+x^{})|u|^2(t^{},x^{})`$, we can determine the optimal propagation velocities which are again symmetric (see the two arrows in Fig. 2) and practically coincide with the velocity of the dark structures visible in the pattern.
The comoving exponents can be again computed by linearizing Eq. (4) and assuming an exponential profile for the perturbation. The resulting spectrum is reported in the inset of Fig. 2. It is symmetric (because of the left-right symmetry in the model) and the zeros of the spectrum coincide with the propagation velocities of correlations (see Table 1) represented by the two tilted arrows in the same figure.
The last system where we have tested our conjecture, is an experimental one, namely, a $`CO_2`$ laser with delayed feedback. This is a physical system that has been used in the past to investigate several instabilities both in the regime of low and high-dimensional chaos . Although this is not, strictly speaking, a spatially extended system, it can be interpreted as such. The idea consists in decomposing the time variable as $`t=n+s\tau `$ , where the parameter $`\tau `$ is close to the delay time (fixed equal to $`400\mu s`$ in the experiment), $`n=\mathrm{Int}(t/\tau )`$ is the new, discrete, time variable, and $`s`$ ($`0<s<\tau `$) is a space-like variable. In a sense, this system is complementary to a chain of oscillators, since the discrete and continuous character of time and space axes are exchanged. The validity of this decomposition has been established in Ref. ; here, the reader can appreciate its meaningfulness by looking at the space-time representation of the intensity of the emitted radiation in Fig. 3. While looking at the pattern, it is important to realize that the slope of the various coherent structures depends on the value of $`\tau `$ adopted in the time decomposition. Here, for the sake of clarity, we have adjusted $`\tau `$ in such a way that the pseudo-rolls, responsible for the optimal correlations, are almost vertical (see arrow 1). A further important difference with the previous models is the absence of left-right symmetry. Even more, causality implies that no leftwards (negative) propagation is possible at all.
Since no theoretical model exists which reproduces the laser dynamics with a sufficient accuracy, one must exclusively rely on the experimental data. However, it has been recently developed an approach to reconstruct the dynamics of delayed systems, even when it is so high-dimensional that the standard embedding technique is bounded to fail . In short, given a time-series $`y_n`$, the method consists in constructing an embedding space composed of two-window vectors,
$$v_n(y_n,y_{n1},\mathrm{},y_{nm+1},y_{nT},\mathrm{},y_{nTm+1}),$$
(5)
where the distance $`T`$ between the two windows coincides with the delay, while the length $`m`$ is the effective number of variables involved in the dynamics. Once the proper values of both $`T`$ and $`m`$ have been identified (by minimizing the forecast error), a local linear model can be constructed, by fitting the behaviour in a suitable neighbourhood of each point in the embedding space. In the case of the $`CO_2`$ laser, $`m=5`$ guarantees an excellent reproduction of the original dynamics . From the knowledge of the empirical model, one can again compute the convective Lyapunov exponents by going through the same intermediate steps. The corresponding spectrum is reported in the inset of Fig. 3 (notice that, given the peculiarity of the space-time reconstruction, velocities are here expressed in time units - more precisely $`\mu s`$ per number of delay units), where we see that it is restricted to the positive-$`v`$ domain (see Ref. ). As a consequence of this asymmetry, both zeros are positive. Thus, in principle, one might expect two different propagation velocities. The correlation analysis of the experimental data has instead revealed a single velocity, which coincides with the smallest zero of the convective Lyapunov spectrum (see Table 1). We cannot rule out the possibility that a second, much less effective, propagation exists with a larger velocity in coincidence with the second zero. However, even if this is not the case, we can at least state that whenever propagation of correlations can be inferred from the evolution of some observable, it must coincide with a netrually stable velocity for the spectrum of convective Lyapunov exponents. What are the (possibly) additional ingredients ensuring the actual existence of propagation phenomena and determining their strength seems to be a much harder problem that we plan to attack in the future.
We conclude, by restating that we have found a clear evidence of a strict link between a “large”-scale feature like the optimal propagation of correlations and the evolution of infinitesimal perturbations, a problem which, because of its very nature, can be formulated in terms of Lyapunov exponents. Accordingly, albeit deterministic chaos contributes to a fast decay of correlations in low-dimensional systems, it is compatible with propagation phenomena as soon as spatial degrees of freedom come into play.
Financial support from the European Union (contract N. PSS 1043) is acknowleged. |
no-problem/9911/cond-mat9911264.html | ar5iv | text | # Bose condensation in a model microcavity
## Abstract
We study the equilibrium properties of a system of dipole-active excitons coupled to a single photon mode at fixed total excitation. Treating the presence or absence of a trapped exciton as a two-level system produces a model that is exactly soluble. It gives a simple description of the physics of polariton condensation in optical cavities beyond the low-density bosonic regime.
The coupled exciton-photon modes of a microcavity were first observed by Weisbuch et al.. Since at low densities excitons are bosons, at low densities coupled exciton-photon modes are also bosons. These bosonic excitations are known as cavity polaritons.
Considerable effort has been devoted to seeking experimental evidence for the bosonic nature of cavity polaritons. This evidence has been sought in the non-equilibrium behavior of microcavities. In this paper we will consider the simpler equilibrium problem, in particular the possibility of a Bose condensate of cavity polaritons.
Polaritons are not conserved particles, so there is ultimately no equilibrium condensate. We may, however, treat polaritons as conserved particles if their lifetime is much longer than the time required to achieve thermal equilibrium at a fixed polariton number. We will study such a quasi-equilibrium limit in a model microcavity.
While the theory of weakly-interacting bosons is well understood, it is far from obvious that this theory is appropriate to the cavity polariton condensate. The concept of a polariton is only valid in linear response; it is not valid for a substantial occupation of the excitons. Finite exciton densities introduce saturation of the electronic states, so the excitons cannot be treated as bosons. In general, finite densities of excitons also lead to exciton-exciton interactions, which can produce dephasing and ionization of the excitons. By considering a situation in which neither dephasing nor ionization are relevant, we will show how to generalize the concept of a polariton to include saturation of the exciton states. Saturation alone does not preclude Bose condensation.
We assume that the relevant electronic excitations in the microcavity are localized, physically separated, single excitons. Thus we can neglect the Coulomb interaction between excitons localized on different sites. We further assume that, because of the tiny effective mass of a cavity photon($`10^5m_e`$ for a $`1000`$ Å cavity), there is only a single relevant photon mode in the cavity.
These assumptions lead us to consider the well-known Dicke model. This consists of a single mode of the photon field, dipole coupled to a set of $`N`$ localized two-level oscillators. Each two-level oscillator represents one exciton state, localized on site $`n`$ with an energy $`E_g(n)`$. These exciton states are composed of conduction and valence electrons with fermionic annihilation operators $`b_n`$ and $`a_n`$ respectively; these fermions are subject to the local constraints $`b_n^{}b_n+a_n^{}a_n=1`$. For brevity, we suppress the site index $`n`$ on the fermion operators. Making the rotating wave approximation, we consider the Hamiltonian
$`H`$ $`=`$ $`{\displaystyle \frac{E_g(n)}{2}\left(b^{}ba^{}a\right)}+\omega _c\psi ^{}\psi `$ (2)
$`+{\displaystyle \frac{g}{\sqrt{N}}}{\displaystyle \left(b^{}a\psi +\psi ^{}a^{}b\right)}.`$
$`\psi `$ is the bosonic annihilation operator for the cavity mode and the summations are over the site index $`n`$.
The operator $`\sigma _z=\frac{1}{2}(b^{}ba^{}a)`$ measures the electronic excitation; such excitations are created by the operator $`\frac{1}{\sqrt{N}}b^{}a`$. They are approximately bosonic provided we remain near to the bare ground state, so that $`\sigma _zN/2`$. In this limit equation (2) becomes two coupled harmonic oscillators and we recover the usual bosonic polaritons.
Away from $`\sigma _z=N/2`$ saturation of the electronic states becomes relevant and the excitations are no longer bosonic. However, we can still define a polariton: it is the quantum of excitation of the non-linear coupled system. With this definition the polariton number operator is $`L=\psi ^{}\psi +\sigma _z`$, which is a conserved quantity for the Hamiltonian (2).
The thermal equilibrium of the Dicke model was originally solved by Hepp and Lieb. Here we are interested in the quasi-equilibrium problem posed by (2) at constant total excitation $`L`$, a generalization obtained by adding a chemical potential to constrain $`L`$. Thus at $`T=0`$ the relevant free energy is $`H\mu _{\mathrm{e}x}L`$, where $`\mu _{\mathrm{e}x}`$ is the chemical potential for excitations.
To keep the number of parameters to a minimum we restrict ourselves to the uniform case $`E_g(n)=E_g`$. There are then only three parameters in the problem: the temperature $`(k_B\beta )^1`$, the dimensionless detuning $`\mathrm{\Delta }=(\omega _cE_g)/g`$, and the excitation density $`\rho _{\mathrm{e}x}=L/N`$. Generalizing our solution to include a distribution of exciton energies is straightforward.
Following Kiry’anov and Yarunin’s work on the unconstrained problem, we solve the model using the large-$`N`$ expansion of a coherent-state path-integral. This expansion is possible because all the exciton states couple to a single mode of the photon field. More detailed derivations, as well as a variational approach based on the work of Comte and Nozières on the exciton condensate, will be presented in a future publication.
By integrating out the fermions and rescaling the photon field, the grand partition function for (2) can be represented as an imaginary-time path-integral over the photon field of the form $`Q𝒟\psi \mathrm{exp}(NS_{\mathrm{eff}})`$. In the thermodynamic limit $`N\mathrm{}`$ the partition function is dominated by those paths $`\psi _0(\tau )`$ which minimize the action $`S_{\mathrm{eff}}`$. The Euler-Lagrange equation for $`S_{\mathrm{e}ff}`$ is a self-consistency condition which relates $`\psi _0(\tau )`$ to the equilibrium polarization of a two-level system in the self-consistent field $`\psi _0(\tau )`$. It takes the form
$$(_\tau +\stackrel{~}{\omega }_c)\psi _0(\tau )+g\overline{a}(\tau )b(\tau )=0,$$
(3)
with $`\stackrel{~}{\omega _c}=\omega _c\mu _{\mathrm{e}x}`$.
The order parameter of the polariton condensate is the polarization of either the photon field or the exciton field. Although there are two physically distinct order parameters, we see from equation (3) that they are coupled by the dipole interaction. It is the dipole interaction which favors the formation of a condensate.
We assume that the extremal trajectories are independent of $`\tau `$, $`\psi _0(\tau )=\psi _0`$. We can then use the familiar eigenstates of a two-level system in a static external field to calculate the polarization term on the right of equation (3). A constrained thermal population of these renormalized eigenstates leads to a BCS-like equation for $`\psi _0`$,
$$\stackrel{~}{\omega }_c\psi _0=\frac{g^2\psi _0}{2E}\mathrm{tanh}(\beta E).$$
(4)
Here $`E`$ is the renormalized fermion energy $`E=\sqrt{\stackrel{~}{\epsilon }^2+g^2|\psi _0|^2}`$, with $`\stackrel{~}{\epsilon }=(E_g\mu _{\mathrm{e}x})/2`$.
The chemical potential $`\mu _{\mathrm{e}x}`$ is related to the partition function $`Q`$ in the usual manner. For the condensed solutions the leading term in the expansion of $`Q`$ around the extremal trajectories gives
$$\rho _{\mathrm{e}x}=|\psi _0|^2\frac{1}{g^2}\stackrel{~}{\epsilon }\stackrel{~}{\omega }_c,$$
(5)
while for the normal solution we have $`\rho _{\mathrm{e}x}=[\mathrm{tanh}(\beta \stackrel{~}{\epsilon })]/2`$.
At zero temperature (4) and (5) always have a condensed solution $`\psi _00`$. This solution is illustrated in Fig. 1. The upper part of Fig. 1 shows the chemical potential as a function of excitation density for detunings $`\mathrm{\Delta }=0,1`$ and $`3`$. In the low density limit, $`\rho _{\mathrm{e}x}=0.5`$, we are describing a condensate of conventional bosonic polaritons, so the chemical potential is given by the usual polariton energy $`\mu _{\mathrm{e}x}=\frac{1}{2}\left((\omega _c+E_g)g\sqrt{\mathrm{\Delta }^2+4}\right)`$. At high densities the electronic states are saturated and further excitation must be added as photons. Thus the chemical potential approaches the energy of the cavity mode. Between these two limits we find a discontinuity at $`\rho _{\mathrm{e}x}=0.5`$ if $`\mathrm{\Delta }>2`$. The lower part of Fig. 1 illustrates the composition of the condensate, again for detunings $`\mathrm{\Delta }=0,1`$ and $`3`$. Increasing detuning increases the electronic fraction of the condensate. In the high excitation limit the inversion approaches zero. This gives the maximum polarization of the electronic states and hence minimizes the dipole energy.
We study the stability of the solutions to (4) by considering the quadratic term $`S_2`$ in the functional Taylor series expansion of $`S_{\mathrm{eff}}`$ around the extremal trajectory $`\psi _0`$. The kernel of $`S_2`$, $`𝒢^1`$, is the inverse matrix Green’s function for fluctuations of the photon field. We find $`𝒢^1`$ by taking two functional derivatives of $`S_{\mathrm{e}ff}`$ and evaluating the result on the extremal trajectory $`\psi _0`$. This amounts to solving the Dyson-Beliaev equations for the photon Green’s functions. The self-energies are provided by the polarizability of a two-level system in the self-consistent field $`\psi _0`$.
The resulting expression for $`𝒢^1`$ in the condensed phase is somewhat lengthy, so we do not reproduce it here. We find that the eigenvalues of $`𝒢^1`$ in the condensed phase are, apart from the Goldstone mode, strictly positive provided $`\stackrel{~}{\omega _c}>0`$. This condition is automatically satisfied by the condensed solutions to (4).
To determine the phase diagram we assume a continuous transition between the normal and condensed states. The transition temperature $`(k_B\beta _c)^1`$ is determined by requiring (4) and (5) to have a repeated root $`\psi _0=0`$. This gives two transition temperatures
$$\beta _cg=\frac{4\mathrm{tanh}^1(2\rho _{\mathrm{e}x})}{\mathrm{\Delta }\pm \sqrt{\mathrm{\Delta }^28\rho _{\mathrm{e}x}}}.$$
(6)
We also obtain (6) as the temperature corresponding to the onset of a low frequency instability of the normal state. Thus our assumption of a continuous transition is correct.
The phase boundary (6) is illustrated in Fig. 2 for detunings $`\mathrm{\Delta }=0,1`$ and $`2`$. For $`\mathrm{\Delta }0`$ the phase diagram is straightforward. The transition temperature increases monotonically with density, reaching infinity at $`\rho _{\mathrm{e}x}=0`$. For $`\mathrm{\Delta }>0`$ there is a region in which the condensate exists on both the high and low temperature sides of the normal state. When $`\mathrm{\Delta }2`$ the condensate exists in two disconnected regions of the phase diagram.
This unusual phase diagram can be understood by considering the excitations of the normal state. The excitation energies follow in the usual manner from the locations of the poles of the Green’s function. We write the action for fluctuations of the photon field about the normal state as $`S_2=\beta _{\omega _n}\delta \overline{\psi }(\omega _n)𝒢_N^1(\omega _n)\delta \psi (\omega _n)`$, where $`\omega _n`$ is a bosonic Matsubara frequency. The normal-state Green’s function $`𝒢_N`$ takes the form
$$𝒢_N(\omega _n)=\frac{C_+}{i\omega _n+E_+}+\frac{C_{}}{i\omega _n+E_{}},$$
(7)
with $`E_\pm =[(\omega _c+E_g)\pm g\sqrt{\mathrm{\Delta }^28\rho _{\mathrm{e}x}}]/2\mu _{\mathrm{e}x}`$ and $`C_\pm =\pm (2\stackrel{~}{\epsilon }E_\pm )/(E_{}E_+)`$. The excitation energies, measured from the chemical potential $`\mu _{\mathrm{e}x}`$, are $`E_\pm `$.
The normal-state excitations are polaritons in the sense of Hopfield: coupled modes involving the linear response of the electronic system around its equilibrium state. The gap in the spectrum is increased over the bare detuning $`\mathrm{\Delta }`$ owing to the dipole coupling between the excitons and the cavity mode. The presence of excitation in the ground state, either driven by finite temperatures or by finite $`\mu _{\mathrm{e}x}`$, causes the two polariton branches to attract. This attraction can be understood in terms of an angular momentum representation for the collective states of the electronic system. The excitation of the electronic states $`\sigma _z`$ forms the z-component of an angular momentum and their polarization forms a raising operator $`\sigma _+`$. Thus the polarizability of the electronic states is a maximum at $`\sigma _z=N/2`$.
Figure 3 illustrates the excitation energies $`E_{\mathrm{e}x}=E_\pm +\mu _{\mathrm{e}x}`$ obtained from equation (7), for detunings $`\mathrm{\Delta }=0,1`$ and $`2`$. On the same axis we plot the normal state chemical potential, given by the expression immediately below equation (5), for $`\beta g=1`$. Figure 3 should be compared with the $`\beta g=1`$ line of the phase diagram in Fig. 2.
When $`\mathrm{\Delta }=0`$ and $`\rho _{\mathrm{e}x}=0.5`$ the system is in the normal state. Increasing $`\rho _{\mathrm{e}x}`$ populates the electronic excitations, increasing the chemical potential and decreasing the polariton splitting. Eventually the chemical potential crosses the lower polariton branch from below and the system condenses. In contrast, for $`\mathrm{\Delta }=1`$ and $`2`$ the chemical potential crosses the lower polariton branch at $`\rho _{\mathrm{e}x}=0`$ without the condensate appearing. It is not until the chemical potential crosses the upper polariton branch that the transition occurs. This can be understood by considering the signs of the quasiparticle weights $`C_\pm `$. A positive quasiparticle weight corresponds to absorption of an external field(particle-like excitations), whereas a negative quasiparticle weight corresponds to gain(hole-like excitations). For $`\rho _{\mathrm{e}x}>0`$, the lower polariton branch has a negative weight: it has become hole-like, and must be below the chemical potential for stability. Note that for large $`\mathrm{\Delta }`$ there is a particle-hole symmetry, so the region $`0.5<\rho _{ex}<0.5`$ of the phase diagram is symmetric about $`\rho _{\mathrm{e}x}=0`$.
In the condensed phase we find excitations with energies relative to the chemical potential $`E_\pm =\pm \sqrt{(\stackrel{~}{\omega _c}+2\stackrel{~}{ϵ})^2+g^2|\psi _0|^2}`$, along with a Goldstone mode. The polariton condensate will produce incoherent luminescence or absorption at the energies $`E_\pm +\mu _{\mathrm{e}x}`$ and coherent emission at $`\mu _{\mathrm{e}x}`$.
Since we expect polariton condensates to produce coherent light, a major issue experimentally will be distinguishing polariton condensates from semiconductor lasers. In semiconductor lasers excitons are conventionally assumed to be ionized. We list some aspects of the present work which should help to distinguish polariton condensation from lasing: (1)From Fig. 1, we see that the polariton condensate can exist without an inversion of the electronic system. (2)Approaching the condensation transition by increasing the excitation density we may observe a reduction of the normal-state excitation gap owing to the fermionic structure of the excitons. (3)The fermionic structure of the excitons shifts the coherent emission from the condensate away from the bosonic polariton energy. (4)The incoherent luminescence and absorption from the condensate exhibits a gap induced by the coherent photon field. In a conventional laser, this gap is destroyed by the very short relaxation time of the electronic polarization. Following , processes which destroy the electronic polarization could be incorporated into our solution. In superconductors, such processes produce a regime of gapless superconductivity where the order parameter survives without a gap in the excitation spectrum.
The applicability of our results to real microcavities is restricted by our neglect of exciton states with a significant wavefunction overlap. We have assumed that these states are at infinitely large energies. In reality, these states exist above an energy $`E_m`$. Thus our thermodynamic results are only valid when $`E_m\mu _{ex}`$ is large compared with $`\beta ^1`$ and $`g`$. By considering Fig. 3, we deduce that realizing the phase diagram of Fig. 2 requires an energy gap $`\mathrm{\Delta }E=E_mE_0g`$. This could occur in highly disordered materials with Frenkel-like excitons, such as organic semiconductors. An energy gap $`\mathrm{\Delta }E`$ could exist in inorganic quantum wells if the excitons move in a potential containing deep, well-separated traps, perhaps associated with interface islands. However, in both these cases it is likely that there will be several exciton states on each site, rather than the single state we have assumed.
If $`\mathrm{\Delta }E`$ is small compared with $`g`$ then the phase boundary (6) will only be realized when $`\mathrm{\Delta }0`$ and the temperature is low. In the normal state under these conditions both the exciton occupation and the effects of the dipole interaction are negligible. The transition temperature (6) then corresponds to non-interacting bosons with an unusual density of states.
To summarize, we have presented a theory of polariton condensation in the Dicke model. By studying the model at fixed excitation, we have generalized the concept of a polariton condensate from the low-density regime. We have found two states: a normal state of excitons and a condensate of polaritons. The polariton condensate is a superposition of a BCS-like state of the excitons and a coherent state of the photon field, and is favored over the normal state by the dipole coupling. We recover conventional polaritons both as the low-density limit of the condensate and as the linear-response excitations of the normal state.
This work was supported by funding from the Engineering and Physical Sciences Research Council, UK. |
no-problem/9911/cond-mat9911259.html | ar5iv | text | # Dynamic modulation of electron correlation by intramolecular modes in charge transfer compounds
\[
M. Meneghetti
Department of Physical Chemistry, University of Padova, 2, Via Loredan,I-35131 Padova, Italy
\] Electron-phonon and electron-electron interactions are in competition in determining the properties of molecular charge transfer conductors and superconductors. The direct influence of phonons on the electron-electron interaction was not before considered and in the present work the coupling of intramolecular modes to electron–electron interaction ($`U`$-vib interaction) is investigated. The effect of this coupling on the frequency of the normal modes of a dimer model is obtained and it is shown that frequency shifts of the Raman active modes are directly related to this coupling. The results are used to obtain the values of the $`U`$-vib coupling constants of intramolecular modes of a representative molecule of charge transfer conductors, like tetramethyltetratiafulvalene. Consequences of this coupling on the electron pairing are also suggested.
The interesting electronic and optical behaviors of organic $`\pi `$-conjugated systems derive from their low energy electronic properties. One recalls, for example, the superconducting properties of some charge transfer molecular crystals or the luminescent properties of some $`\pi `$-conjugated polymers used for producing light emitting diodes. The interaction of electrons with phonons has to be considered, in particular for these small band systems, as important as the Coulomb interaction between the $`\pi `$ electrons. In fact some properties, like a low symmetry ground state due to a Peierls or a spin-Peierls transition, derive from the electron-phonon interaction, but other properties like the charge transfer excitations are dominated by the Coulomb interaction. On the other hand it is not available a satisfactory model for the supercondunting properties of these systems based on a well defined interaction.
Hubbard or extended Hubbard models have been extensively used for describing the electronic properties of $`\pi `$-conjugated systems since they allow to describe the electron-electron interaction beyond the mean field approximation. Electron-phonon interaction has been included in such Hamiltonians on the basis of the Holstein model which considers the dependence of the site energies on the intra-molecular normal modes and of the SSH model, derived from the Peierls model, which takes into account the dependence of the transfer integrals on the inter-molecular modes. It is the aim of this paper to find evidences of the possible dependence of the intrasite Coulomb interaction of the Hubbard model on the intra-molecular vibrations ($`U`$-vib interaction). This interaction can be of direct importance for understanding the dynamic of electrons in low dimensional systems like organic charge transfer superconductors since it can favor the pairing of electrons if the modulation of the Coulomb interaction is comparable to that of the interaction itself.
The simplest model for studying the $`U`$-vib interaction is a half-filled charge transfer dimer with two identical molecules, namely two electrons on two sites. Since it is well known that one can easily obtain charge transfer dimers of radical molecules in solution, one has also the possibility of directly verifying the results of the model. We will consider in particular the cation radicals of tetramethyltetrathiafulvalene (TMTTF$`^{\dot{+}}`$), a molecule which is present in many charge transfer crystrals and in one superconducting compound.
The Hubbard Hamiltonian for the dimer is
$`H=`$ $`ϵ{\displaystyle \underset{i=1,2}{}}n_i+t{\displaystyle \underset{\sigma }{}}(a_{1,\sigma }^{}a_{2,\sigma }+a_{2,\sigma }^{}a_{1,\sigma })`$ (2)
$`+U{\displaystyle \underset{i=1,2}{}}{\displaystyle \underset{\sigma }{}}n_{i,\sigma }n_{i,\sigma }`$
where $`ϵ`$ is the site energy and can be taken as zero, $`t`$ is the hopping integrals and $`U`$ is the intramolecular Coulomb interaction.
The interaction with the intramolecular normal modes of vibration is introduced by expanding, to first order, $`ϵ`$ and $`U`$ on the basis of the normal modes. Because of the symmetry of the dimer, the following coordinates, for each $`Q_\alpha `$ intramolecular normal mode, are used $`Q_\alpha ^S=2^{1/2}(Q_{\alpha ,1}+Q_{\alpha ,2})`$ and $`Q_\alpha ^A=2^{1/2}(Q_{\alpha ,1}Q_{\alpha ,2})`$. On this basis the following two terms are added to Eq. 2:
$$\underset{\alpha }{}Q_\alpha ^A[\frac{g_\alpha }{\sqrt{2}}(n_2n_1)+\frac{g_\alpha ^U}{\sqrt{2}}(n_{1,}n_{1,}n_{2,}n_{2,})]$$
(3)
and
$$\underset{\alpha }{}Q_\alpha ^S\frac{g_\alpha ^U}{\sqrt{2}}(n_{1,}n_{1,}+n_{2,}n_{2,})$$
(4)
where $`g_\alpha =(ϵ/Q_\alpha )_{}`$ is the coupling constant based on the Holstein model and $`g_\alpha ^U=(U/Q_\alpha )_{}`$ is that for the $`U`$-vib interaction.
The contribution coming from the Holstein mechanism involves only the $`Q_\alpha ^A`$ coordinates which, by symmetry, can be observed in the infrared spectra, whereas the $`U`$-vib interaction affects also the $`Q_\alpha ^S`$ coordinates which are Raman active. Previously, attention was directed in particular to the infrared spectra where large and important effects have been observed. However, one can see that it is in the Raman spectra that one may find the sign of the $`U`$-vib interaction and it is in particular to these spectra of the charge transfer dimer that we direct our attention.
The evaluation of the electron-vibration interaction strength can be obtained from the frequency shifts of the vibrational modes observed in the infrared and Raman spectra. The case in which only the Holstein mechanism is present has already been considered. In this case the vibrational shift can be calculated for the infrared active $`Q_\alpha ^A`$ modes. The only difference one should consider in the present case for these coordinates is that the coupling constant for the $`\alpha `$-th mode change from $`g_\alpha `$ to $`(g_\alpha +g_\alpha ^U/2)`$. The result, can now be rewritten in a more convenient form as:
$$\mathrm{\Delta }_{IR}=\mathrm{}\omega _\alpha \mathrm{}\omega _{IR}\frac{4|\mathrm{\Delta }_{ST}|(g_\alpha +g_\alpha ^U/2)^2}{\mathrm{}\omega _{CT}(\mathrm{}\omega _{CT}+|\mathrm{\Delta }_{ST}|)}$$
(5)
where $`\mathrm{}\omega _\alpha `$ and $`\mathrm{}\omega _{IR}`$ are the unperturbed and perturbed energy, respectively, of the $`\alpha `$-th mode, $`\mathrm{}\omega _{CT}=U/2+(U^2/4+4t^2)^{1/2}`$ is the energy of the charge transfer excitation of the dimer and $`\mathrm{\Delta }_{ST}=U/2(U^2/4+4t^2)^{1/2}`$ is the energy separation between the singlet ground state and the triplet states. One finds that a coupled $`Q_\alpha ^A`$ mode shifts to lower frequencies.
The frequencies of the $`Q_\alpha ^S`$ modes, which can be observed in the Raman spectra, were previously considered as the frequencies of the unperturbed modes since these modes were not involved in the electron-vibration interaction. Now, if $`U`$-vib interaction is present, these modes are no more unperturbed and following the derivation of Eq. 5, one finds that their vibrational energy shift are:
$$\mathrm{\Delta }_R=\mathrm{}\omega _\alpha \mathrm{}\omega _R\frac{|\mathrm{\Delta }_{ST}|^2(g_\alpha ^U)^2}{(\mathrm{}\omega _{CT}+|\mathrm{\Delta }_{ST}|)^3}.$$
(6)
The reference, unperturbed state, which defines $`\mathrm{}\omega _\alpha `$, must now be the isolated radical molecule, namely the situation in which the charge transfer excitation is not present.
Before looking at the experimental spectra it is important to obtain some indications about the vibrational frequency shifts predicted by the model. Figures 1 and 2 show the calculated frequency shifts for the infrared ($`Q_\alpha ^A`$) and Raman ($`Q_\alpha ^S`$) modes, respectively, vs. the coupling constant $`g_\alpha ^U`$ for values of the other parameters characteristic of a charge transfer system: $`U`$=1.0 eV, $`t`$=0.2 eV and $`g_\alpha `$=0.1 eV.
One notes that the shifts for the infrared mode are more than one order of magnitude larger than those calculated for the Raman mode. The calculation shows that one can observe vibrational shift as large as 100 cm<sup>-1</sup> in infrared, as it was found many times, but that in the Raman spectrum one can predict vibrational shifts of the order of a few wavenumbers. This could be the reason why there have not been observations of frequency shifts of the Raman active modes for segregated systems like the dimer we are considering.
The small frequency shift one calculates for the Raman mode suggests care in obtaining the experimental data. In particular it is important that both the unperturbed and the perturbed frequency can be recorded using the same sample and in the same experiment so that to minimize possible influence of the experimental conditions. The opportunity of doing such an experiment is given by the equilibrium between the isolated molecules and the charge transfer dimers, which is present in solution of radical molecules like TMTTF$`^{\dot{+}}`$. From the absorptions in the visible and near-infrared spectral regions it is easy to identify the presence of the isolated molecules and of the charge transfer dimers.
Figure 3 shows the electronic spectra of a sulfolane solution of TMTTF$`^{\dot{+}}`$ radical cation in which the electronic transitions of the monomer and of the dimer are easily recognizable. The absorption at longer wavelength (805 nm) is the charge transfer excitation of the dimers. The electronic excitations at 670 nm and at 460 nm are those of the monomer and they shift to shorter wavelength (575 nm and 417 nm) when the dimer is formed. Since the absorption bands of the monomer and of the dimer are found at sufficiently different energies, one can obtain the Resonance Raman spectra of the two species using the appropriate excitation laser frequency. One finds that the Krypton laser lines at 568 and at 647 nm are in resonance with the electronic transitions of the dimer and of the monomer, respectively.
The same equilibrium between the monomer and the dimer can be found using dimethylsulfoxide (DMSO) as solvent. We have also obtained samples in which the dimers of TMTTF$`^{\dot{+}}`$ are embedded in a polymer like PMMA. In these cases, however, the equilibrium is shifted almost completely toward the dimers and it is not possible to obtain both the spectrum of the monomer and of the dimer using the same material.
The Raman spectrum of TMTTF$`^{\dot{+}}`$ shows two strong bands at 530 and at 1430 cm<sup>-1</sup>, and another one at 1057 cm<sup>-1</sup>. The bands have been assigned as $`a_g`$ $`\nu _{10}`$, $`a_g`$ $`\nu _4`$ and to an internal vibration of the methyl groups, respectively. $`a_g`$ $`\nu _{10}`$ was described as a displacement of the internal C-S bonds whereas $`a_g`$ $`\nu _4`$ was related to the stretching of the internal C=C bond. From the infrared data of dimerized TMTTF$`^{\dot{+}}`$ systems it was concluded that these two normal modes were strongly coupled to the charge transfer excitation since two strong new bands, which were not expected for the isolated molecule, can be observed in the infrared spectra and assigned to the infrared activation of these two normal modes.
The Raman spectra of the solutions have been recorded by using a Jobin-Yvon multichannel spectrometer (S3000) equipped with a CCD camera (Astromed) cooled at liquid nitrogen. The spectra of the solutions were obtained at room temperature with a rotating cell for liquids to avoid the decomposition of TMTTF$`^{\dot{+}}`$ due to a poor dissipation of the energy of the laser beam. The solutions were prepared by using (TMTTF)ClO<sub>4</sub> and were flushed with nitrogen to eliminate the presence of oxygen. We used about 40 mW and a beam waist estimated to be 200$`\mu `$m. It was verified that there was no decomposition of the solutions by recording the UV-Visible spectra after the Raman measurements. It was found that the reproducibility of the frequencies of the spectra was better than 1 cm<sup>-1</sup>.
The spectra of the sulfolane solutions showed two different behaviors for the $`a_g`$ $`\nu _{10}`$ at about 530 cm<sup>-1</sup> and for the $`a_g`$ $`\nu _4`$ at about 1430 cm<sup>-1</sup>. The first band at 530 cm<sup>-1</sup> showed a small positive shift of less then one wavenumbers when the exciting line was changed from 647 to 568 nm, namely from the resonance with the monomer to that of the dimer. On the contrary, the second band, reported in Figure 4, was found at 1435.1 cm<sup>-1</sup> when the resonance was with the monomer electronic transition and at 1428.6 cm<sup>-1</sup> when the resonance was with the dimer one. Therefore in this case one finds a frequency shifts of -6.5 cm<sup>-1</sup>. The fitting of the bands with gaussian functions showed that a single band was present in every case without a significant variation in the widths.
The same situation was found when the Resonance Raman spectra of the DMSO solutions were recorded. In this case the frequencies of the modes were different due to the different environment (for the monomer we observed 524.1 and 1430.3 cm<sup>-1</sup> for $`a_g`$ $`\nu _{10}`$ and $`a_g`$ $`\nu _4`$ respectively). However, we found that the frequency shifts of the two modes were similar to those of the previous situation with a small positive shift of 1.2 cm<sup>-1</sup> for $`a_g`$ $`\nu _{10}`$ and a negative shift of -5.0 cm<sup>-1</sup> for $`a_g`$ $`\nu _4`$.
The evidence was, therefore, that a significant shift is observed, when the dimer is formed, only for the mode related to the stretching of the central C=C bond of the molecule. The observed frequency shift is in agreement with the theory prediction of a small negative shift for a normal modes coupled with the charge transfer excitation.
However, to rule out the possibility that the formation of the dimer may induce some variation, not related to the charge transfer excitation, which could influence the vibrational normal modes, we have recorded the Raman spectra of the solution and of the solid of the neutral molecule. The crystal structure of the neutral molecule is not available but from these spectra one could reasonably obtain some indications about the difference between an almost free molecule in solution and one which is in close contact with other molecules. The $`a_g`$ $`\nu _4`$ mode has been observed, for the neutral molecule, at 1538 cm<sup>-1</sup> for the solid at about 20 K. One finds that, at room temperature, its frequency is 1532.5 cm<sup>-1</sup> for the crystalline material and less than one wavenumber higher for the solution in carbon sulfide in which the neutral molecule dissolves sufficiently. This can be taken as an indication that the aggregation alone has not a significant influence on the frequency of the normal mode related to the C=C central bond.
By using Eq. 6 one can estimate the $`g_\alpha ^U`$ coupling constant for $`a_g`$ $`\nu _4`$. The frequency of the charge transfer excitation is readly found in the near infrared spectrum where it can be located, below all the intramolecular electronic transitions, at 805 nm. $`\mathrm{\Delta }_{ST}`$, the energy separation between the ground state and the triplet states, can be measured in an EPR experiment in which the intensity of the lines associated to the Boltzman population of the triplet states, is measured against temperature. For the dimers of TMTTF$`^{\dot{+}}`$ these magnetic measurements are available. The dimers were embedded in PMMA where one can verify that almost only the dimeric species is present. It was found that $`\mathrm{\Delta }_{ST}=0.25`$ eV.
These data allow one to find $`|g_4^U|=256\pm 17`$ meV, considering the minimum and maximum observed Raman shift. With this value one can also find the value for the Holstein coupling $`g_4`$ by using Eq. 5 for the frequency shift of the infrared normal mode. In this case it is also important the relative sign of the two coupling constants $`g_4`$ and $`g_4^U`$ which, however, cannot be determined on the basis of the above equations. Therefore one takes into account that the sign of the two coupling constant may be different.
The frequency of the $`Q_4^A`$ infrared mode was already found to be 1361 cm<sup>-1</sup> for a solution of dimers of TMTTF$`^{\dot{+}}`$ in DMSO. Since the frequency of the unperturbed mode observed in the Raman spectrum of the DMSO solution is 1430.3 cm<sup>-1</sup>, one finds that the frequency shift for $`Q_4^A`$ is 69.3 cm<sup>-1</sup>. This shift is consistent with the predictions reported above on the basis of the dimer model and allow us to calculate $`g_4`$.
One finds that if the two coupling constant have the same sign $`|g_4|=32\pm 8`$ meV whereas in the case of different signs $`|g_4|=285\pm 8`$ meV. The reported errors are only indicative of the large and small values of $`g_4^U`$. Previously, the value of $`|g_4|`$ was reported to be 120 meV. Therefore, we find that the new estimate for $`g_4`$ are very different from the previous ones and this could be important for the dynamic of the electron transfer between the molecules.
In any case one finds a large $`g_\alpha ^U`$, namely an important dynamic modulation of the correlation term of the Hubbard Hamiltonian ($`U`$) which can be evaluated, from the value of $`\mathrm{}\omega _{CT}`$ and the value of $`\mathrm{\Delta }_{ST}`$ to be 1.29 eV. This situation is very interesting since it can be important for the formation of local pairs of electrons and this opens the way for an explanation of the superconducting properties of organic charge transfer salts following the condensation of bipolarons. One should recall, however, that such a superconducting ground state was never observed for a half-filled charge transfer compound and that this property was only found for quarter filled systems. In this case the relevant Coulomb interaction is not $`U`$ but $`V`$, namely the interaction between electrons on nearest neighbor sites, since there is one electron, or hole, on every second site . Therefore one could suggest that in this case it is the modulation of $`V`$ which is important for a pairing mechanism ($`V`$-vib interaction), a modulation that can derive both from the intra\- and the inter-molecular vibrations. Values of $`V`$, for charge transfer compounds, have been found to be of the order of the transfer integral and one can speculate that the modulation could be very large, due to the coupling to both type of vibrations, determining a strong dynamical pairing of electrons on neighboring sites. It would be therefore very important to obtain values of this type of electron-vibration interaction.
In conclusion, we have obtained, for the first time, experimental evidences that in organic charge transfer compounds there is a coupling between electronic correlation and intramolecular vibrations ($`U`$-vib interaction). In particular we have studied the Raman spectra of dimers of a representative molecule like TMTTF$`^{\dot{+}}`$ and found that its $`a_g`$ $`\nu _4`$ normal mode strongly modulates ($`|g_4^U|=256\pm 17`$ meV) the value of the intramolecular Coulomb interaction. This result shows that also the value of the coupling constants related to the Holstein mechanism ($`g_\alpha `$) have to be reconsidered. The value of $`g_4^U`$ seems to be too small to conclude that a pairing of electrons in organic superconductors follows from this interaction, but suggests that experimental data about the same mechanism but involving intersite Coulomb correlation ($`V`$-vib), have to be obtained.
I wish to thank Renato Bozio for valuable discussions and Gabriele Marcolongo and Gilda Deluca for technical help. Financial support by the Italian National Research Council also through its Progetto Finalizzato MSTA2 and the Ministry of the University and of Scientific and Technological Research are acknowledged. |
no-problem/9911/astro-ph9911333.html | ar5iv | text | # Theory of galaxy dynamics in clusters and groups
## 1. Introduction
Mergers of galaxies in slow collisions and tidal interactions in rapid collisions are two key dynamical processes that occur in groups and clusters of galaxies. Cosmological $`N`$-body simulations are beginning to approach the resolution necessary to study galaxy dynamics in groups and clusters (see Moore, in these proceedings). Moreover, mergers and tidal collisions leave significant observational signatures, in the form of tidal tails, asymmetries and generally disturbed morphologies and internal kinematics (see Amram, in these proceedings). Also, galaxy merging is an essential mechanism for driving elliptical galaxy morphologies given disk-like progenitors. As such, an understanding of galaxy merging is very important for semi-analytical modeling of galaxy formation.
In this review, I compute analytically the rates at which a galaxy of given mass and position within a cluster or group with a Navarro, Frenk, & White (1995, NFW) potential undergoes slow major mergers with lower mass galaxies and rapid tidal encounters. Since the collision cross-sections are strongly modulated by the tides from the group/cluster potential, I begin with a simple formalism for estimating the tides from the system potential. Note that I will not consider ram pressure stripping on galaxies.
## 2. Tides from the cluster/group potential
The potential of a cluster can exert a strong differential force on a galaxy orbiting within it, but these tides are strongly dependent on the galaxy orbit.
A galaxy on a nearly circular orbit is likely to be tidally locked, as the Moon is with respect to the Earth. In this case, the tidal force is simply (King, 1962)
$$F_{\mathrm{tide}}=\mathrm{\Delta }\left[\frac{GM(R)}{R^2}\mathrm{\Omega }^2(R)\right]$$
(1)
and equating $`F_{\mathrm{tide}}`$ in eq. (1) to the force, $`f=Gm(r)/r^2`$ that a galaxy exerts on one of its stars yields for $`rR`$ a galaxy tidal radius $`r_t`$ that satisfies a velocity modulated density criterion (see Mamon, 1995):
$$\overline{\rho }_g(r_t)=\overline{\rho }_{\mathrm{cl}}(R)\left[2\frac{3\rho _{\mathrm{cl}}(R)}{\overline{\rho }_{\mathrm{cl}}(R)}+\frac{V_p^2(R)}{V_{\mathrm{circ}}^2(R)}\right],$$
(2)
where $`V_p`$ and $`V_{\mathrm{circ}}`$ are respectively the galaxy’s velocity at pericenter and the cluster’s circular velocity. The term in brackets in eq. (2) is 2 for circular orbits. Moreover, for a singular isothermal law $`\rho r^2`$ for both galaxy and cluster, one finds (see White, 1983) that for circular orbits the galaxy size is proportional to its clustocentric radius: $`r_t=2^{1/2}(v_{\mathrm{circ}}/V_{\mathrm{circ}})R`$, where $`v_{\mathrm{circ}}`$ is the circular velocity of the galaxy. For general density profiles, writing $`\overline{\rho }(r)v_{\mathrm{circ}}^2(r)/r^2`$, one obtains
$$\frac{r_t/R}{v_{\mathrm{circ}}(r_t)/V_{\mathrm{circ}}(R)}=\left(23\frac{\rho }{\overline{\rho }}+V_p^2/V_{\mathrm{circ}}^2(R)\right)^{1/2}.$$
(3)
Merritt (1984) used a similar circular-tide criterion to argue that galaxies are strongly tidally limited by the cluster potential.
However, cosmological infall imposes elongated orbits. A galaxy on an elongated orbit experiences a strong tide during its rapid, hence short, passage at pericenter, prompting Ostriker, Spitzer, & Chevalier (1972) to introduce the term tidal shock. During this shock, a star in the galaxy experiences a velocity impulse
$$\mathrm{\Delta }vF_{\mathrm{tide}}\mathrm{\Delta }t\frac{GM(R_p)r}{R_p^3}\left(\frac{R_p}{V_p}\right)=\mathrm{cst}\frac{GM(R_p)r}{R_p^2V_p},$$
(4)
where we neglected the centrifugal term in $`F_{\mathrm{tide}}`$, because the galaxy falls in too fast to be phase locked. A more precise calculation by Spitzer (1958), who introduced the impulsive approximation where the point-mass perturber moves at constant $`𝐕`$, produces the same relation as in eq. (4) with a constant of order unity. The impulse approximation can also be applied to extended perturbers (Aguilar & White, 1995; Mamon, 1987). Recently, Gnedin, Hernquist, & Ostriker (1999) applied the impulsive approximation to the more realistic Hernquist (1990) potential and found a dependence of $`\mathrm{\Delta }v`$ matching that of eq. (4), with a very small dependence on $`R_p`$ and again the constant is found to be of order unity for elongated orbits (with order of unity changes when they performed orbit-integrated — instead of straight-line — tidal calculations).
The tidal radius can then be defined as that where the energy increment caused by the tidal perturbation is equal to the binding energy (White, 1983). With $`EGm(r_t)/r_t`$ and $`\mathrm{\Delta }E(\mathrm{\Delta }v)^2/2`$, one obtains another velocity modulated density criterion
$$\overline{\rho }_g(r_t)\overline{\rho }_{\mathrm{cl}}(R_p)\left(\frac{V_{\mathrm{circ}}(R_p)}{V_p}\right)^2,$$
(5)
which for any density profile yields
$$\frac{r_t/R}{v_{\mathrm{circ}}(r_t)/V_{\mathrm{circ}}(R)}=\frac{V_p}{V_{\mathrm{circ}}}.$$
(6)
Figure 1 shows the effects of velocity modulation on the tidal radii. The effective tidal radius should be taken as the largest of the circular and impulsive regimes (otherwise one would be left with discontinuities in the transition $`V_p`$ from near-circular to elongated orbits). Hence, with homogeneous cores, circular tidal theory produces increasingly smaller tidal radii for orbits of increasing but low elongation, as Merritt & White (1987) found in their $`N`$-body simulations. With cuspy cores as in the NFW profile, low-elongation non-circular orbits experience tidal shocks instead. The orbit elongation, $`R_p/R_a0.2`$, found in Ghigna et al.’s (1998) high-resolution cosmological simulations corresponds to $`V_p/V_{\mathrm{circ}}(R_p)1.52.7`$, roughly yielding
$$r_t(1.21.5)R_p\left(\frac{v_{\mathrm{circ}}}{V_{\mathrm{circ}}}\right)(1.21.5)R_p\left(\frac{v_g}{v_{\mathrm{cl}}}\right)$$
(7)
as for singular isothermal models, where $`v_g`$ and $`v_{\mathrm{cl}}`$ are the mean galaxy and cluster velocity dispersions, respectively. If $`V_pV_{\mathrm{circ}}(R_p)`$ and if galaxy and cluster density profiles are self-similar, then tides from the cluster potential would force the simple relation $`m(r_t)/m(r_{\mathrm{vir}})=M(R_p)/M(R_{\mathrm{vir}})`$. In fact, using $`V_p/V_{\mathrm{circ}}(R_p)`$ expected for the NFW profile, assuming $`R4R_p`$, and adopting the departures from self-similarity in the NFW profiles noted by Navarro, Frenk, & White (1997, lower mass NFW profiles are more centrally concentrated), one finds
$$\frac{m(r_t)}{m(r_{\mathrm{vir}})}\left[\frac{M(R)}{M(R_{\mathrm{vir}})}\right]^{b_m}a_r\left(\frac{R}{R_{\mathrm{vir}}}\right)^{b_r},$$
(8)
where for clusters ($`v_{\mathrm{cl}}=1000\mathrm{km}\mathrm{s}^1`$) and groups ($`v_{\mathrm{cl}}=300\mathrm{km}\mathrm{s}^1`$) we respectively have $`b_m=0.50`$ and 0.82, $`a_r=0.58`$ and 0.52, and $`b_r=0.57`$ and 0.78 (eq. is accurate to better than 5% for $`R>0.05R_{\mathrm{vir}}`$).
## 3. Galaxy merger rates in clusters and groups
### 3.1. Global merger rates for equal mass galaxies
The rate of mergers is obtained by integrating over velocities the merger cross-sections:
$$k\frac{1}{n^2}\frac{d^2N}{dtdV}=vs(v)=_0^{\mathrm{}}𝑑vf(v)vs(v),$$
(9)
where $`s(v)=\pi [p_{\mathrm{crit}}(v)]^2`$ is the merger cross-section and $`f(v)`$ is the distribution of relative velocities (with $`_0^{\mathrm{}}f(v)𝑑v=1`$). Hence, $`nkdN/dt`$ is the rate at which a galaxy suffers a merger. Within the virialized regions of clusters with 1D velocity dispersion $`v_{\mathrm{cl}}`$, the velocity distribution is a gaussian with standard deviation $`2^{1/2}v_{\mathrm{cl}}`$: $`f(v)=2^1\pi ^{1/2}v_{\mathrm{cl}}^3v^2\mathrm{exp}[v^2/(4v_{\mathrm{cl}}^2)]`$.
Roos & Norman (1979), Aarseth & Fall (1980) and Farouki & Shapiro (1982) have established merger cross-sections from very small $`N`$-body simulations of galaxy collisions, that were based upon the parameters at closest approach. The maximum distance of closest approach, $`r_p^{\mathrm{max}}`$, was 4 (Aarseth & Fall) or 11 (Farouki & Shapiro) times the mean galaxy half-mass radius, $`r_h`$. Mamon (1992) used the Roos & Norman cross-section with the Aarseth & Fall scaling to derive a merger rate.
However, the gaussian approximation for the relative velocity distribution implies that the cross-sections used in eq. (9) are based upon impact parameters (at infinity) and not at closest approach. Makino & Hut (1997) derived merger cross-sections using high resolution $`N`$-body simulations of colliding galaxies with more realistic density profiles. Their cross-sections are expressed in terms of impact parameters.
Krivitsky & Kontorovich (1997) used a simple gravitational focusing recipe to connect the cross-section at closest approach (which they assumed to be independent of pericentric velocity), in units of the velocity at infinity, to the cross-section at infinity (without making any assumption on the potential energy of interaction of the colliding pair). Note that Makino & Hut show that $`r_p^{\mathrm{max}}>10r_h`$, while Krivitsky & Kontorovich argue that it is approximately the sum of the galaxy *radii* (which are ill-defined).
Figure 2 shows the dimensionless merger rates $`k/(r_h^2v_g)`$, where $`v_g`$ is the mean galaxy internal velocity dispersion, derived from Mamon (1992), Makino & Hut (1997), and Krivitsky & Kontorovich (1997). I rescale the rates of Krivitsky & Kontorovich in terms of half-mass radii using $`R=9r_h`$ (to obtain merger rates similar to those of Makino & Hut), i.e., $`r_p^{\mathrm{max}}/r_h=18`$.
The agreement between the merger rates of Krivitsky & Kontorovich and Makino & Hut is remarkable, given the very simple analytical formulation of the former authors (but again this required a rescaling, or, in other words, a choice of $`r_p^{\mathrm{max}}/r_h=18`$). The very low merger rate of Mamon (1992) in the group regime ($`v_{\mathrm{cl}}v_g`$) is a consequence of the lack of gravitational focusing in that model.
The merger rates of Mamon (1992) and Makino & Hut agree to within 15% in the cluster regime ($`v_{\mathrm{cl}}>4v_g`$). This agreement is almost fortuitous since Mamon shows that the merger rate with the Roos & Norman cross-section scales as $`(r_p^{\mathrm{max}}/r_h)^2`$, while this ratio is very different in Makino & Hut’s cross-section. In any event, in the cluster regime the merger rate can then be written
$$k=b\frac{r_h^2v_g^4}{v_{\mathrm{cl}}^3}=a\frac{G^2m^2}{v_{\mathrm{cl}}^3},$$
(10)
where $`a8`$ (Mamon, 1992). With $`3v_g^20.4Gm/r_h`$ (Spitzer, 1969), appropriate for the Hernquist model, the Makino & Hut rate translates to $`a=12`$. Figure 2 shows that $`kv_{\mathrm{cl}}^3`$, whichever merger cross-section is used. In fact, it is easy to show that *for any merger cross-section rapidly decreasing with increasing velocity*, the merger rate should scale as $`v_{\mathrm{cl}}^3`$ for $`v_{\mathrm{cl}}v_g`$, as first found by Mamon (1992) for the Roos & Norman cross-section.
The important conclusion of Figure 2 is that *for given galaxy parameters, the merger rate is roughly 100 times lower in rich clusters than in poor groups of galaxies*.
### 3.2. Merger rates for different masses
If the critical merging velocity $`v_{\mathrm{crit}}`$ is a function of $`r_p/r_h`$ (Aarseth & Fall, 1980; Farouki & Shapiro, 1982), it is easy to show that $`kr_h^2`$, and if $`v_{\mathrm{crit}}`$ is a function of $`r_p/r_h`$, then $`kr_h^2`$ (see Mamon, 1992). Similarly, it is reasonable to expect that $`kv_g^2^2`$. Then, given eq. (10) and that $`mr_h^3`$, the rate of mergers of a galaxy of mass $`m`$ with a galaxy of mass $`\lambda m`$
$$k(m,\lambda m)=\frac{aG^2m^2}{v_{\mathrm{cl}}^3}\left(\frac{1+\lambda ^{1/3}}{2}\right)^2\left(\frac{1+\lambda ^{2/3}}{2}\right)^2.$$
(11)
A given galaxy undergoes mergers with other galaxies at a rate
$$n\overline{k}(m)=_{\lambda _{\mathrm{min}}}^{\lambda _{\mathrm{max}}}k(m,\lambda m)n(\lambda m)d(\lambda m),$$
(12)
where for major mergers with smaller galaxies (that transform disk galaxies into ellipticals), $`\lambda _{\mathrm{min}}1/3`$ and $`\lambda _{\mathrm{max}}=1`$, while for destruction by mergers with larger galaxies, $`\lambda _{\mathrm{min}}=1`$ and $`\lambda _{\mathrm{max}}\mathrm{}`$. Adopting a Schechter (1976) form for the mass function of galaxies, $`n(m)=(n_{}/m_{})x^\alpha \mathrm{exp}(x)`$, where $`x=m/m_{}`$, eqs. (11) and (12) yield
$$=n\overline{k}=\frac{aG^2n_{}m_{}^2}{16v_{\mathrm{cl}}^3}K(m/m_{}),$$
(13)
$$K_{\mathrm{major}}(x)=x^{3\alpha }\underset{j=0}{\overset{6}{}}\mathrm{Min}(j,7j)\left[\mathrm{\Gamma }(1+j/3\alpha ,x/3)\mathrm{\Gamma }(1+j/3\alpha ,x)\right],$$
(14)
$$K_{\mathrm{destr}}(x)=x^{3\alpha }\underset{j=0}{\overset{6}{}}\mathrm{Min}(j,7j)\mathrm{\Gamma }(1+j/3\alpha ,x).$$
(15)
Figure 3 shows the expected number of major mergers with smaller galaxies and destruction by mergers with larger galaxies that a galaxy of a given mass should expect in a Hubble time if it sits in a typical location of a rich cluster, assuming a constant rate in time. The rise in merger rates at low mass reflects the rise of merger cross-section with mass, while the decrease at high mass is caused by the sharp decrease in the galaxy mass function yielding few galaxies to merge with. Figure 3 clearly indicates that *the probability of merger for a given galaxy is always small.* Moreover, *low and intermediate mass galaxies ($`m<m_{}`$) are usually cannibalized before undergoing major mergers*.
### 3.3. Variation of merger rates with position in cluster
One can go one step further and predict the variation of the merger rate with position in the cluster. From eq. (13), the merger rate scales with radius as
$$(R,m)=\frac{aG^2m_{}\mu ^2(R)\rho _{\mathrm{cl}}(R)}{16\mathrm{\Gamma }(2\alpha ,x_m)v_{\mathrm{cl}}^3(R)}K(m/m_{}),$$
(16)
where $`\mu (R)=m(r_t)/m(r_{\mathrm{vir}})`$ (see eq. ), $`m_{}`$ is the mass at the break of the *field* galaxy mass function and $`x_m`$ is the minimum galaxy mass in units of $`m_{}`$.
Assuming a mass density profile $`\rho R^\beta `$ and arguing that cluster galaxies are severely tidally truncated by the cluster potential as $`r_{\mathrm{gal}}R`$ (i.e. with eq. and assuming constant $`R_p/R_a`$), Mamon (1992) showed that if galaxies also follow the law $`\rho r^\beta `$, then their masses obey $`mR^{3\beta }`$. Note that this sharp scaling of galaxy size with clustocentric distance is now confirmed in high resolution cosmological simulations of clusters (Ghigna et al., 1998). Hence, *the radial variation of merger rates are strongly modulated by potential tides*.
By writing $`n(R)\rho (R)/m(R)R^3`$, I derived $`nkR^{\beta /2}`$, hence a higher merger rate inside the cluster, with a slope agreeing perfectly with the observed elliptical fraction (Whitmore, Gilmore, & Jones, 1993), given $`\beta =9/4`$ as predicted in early models of cluster formation (Bertschinger, 1985). The derivation above has one flaw: although galaxy masses were correctly scaled to increase with $`R`$, I forgot to scale the fraction of cluster mass lying within galaxies in the same way. Therefore, one really expects $`n(R)\rho (R)R^\beta `$ and $`nkR^{33\beta /2}`$ yielding a slope $`d\mathrm{ln}(nk)/d\mathrm{ln}R=3/8`$ for $`\beta =9/4`$ and a null slope for $`\beta =2`$, both in disagreement with the logarithmic gradient of elliptical fraction found by Whitmore et al..
One can use the more realistic NFW density profiles to estimate the radial dependence of the merger rates. An essential parameter is $`R_p/R`$, which measures the effectiveness of the tides from the cluster potential. Because the dynamical friction time scales as $`M/m`$ times the orbital time (Mamon, 1995), orbit circularization, which to first order operates on a dynamical friction time scale, should be very slow for galaxies falling onto clusters, but fairly effective for galaxies falling into small groups.
Figure 4 shows the predicted number of major mergers in rich clusters and small groups, extrapolated over a Hubble time, using the non-self-similarity of the NFW profiles, an exact scaling of the typical galaxy mass $`m_{}`$ with radius (see eq. ), and partial orbit circularization in groups.
The merger rates for constant mass galaxies fall off sharply at small clustocentric radii, simply because tidal truncation of galaxies is so severe that there are no galaxies left that are massive enough to produce a major merger with our test galaxy. In general, the merger rates are maximum for intermediate radii for given test galaxy masses, and at low radii for fixed $`m/m_{}(R)`$ (recall though that low mass galaxies get cannibalized before they can undergo a major merger with a smaller galaxy). In any event, Figure 4 confirms that *mergers are ineffective in clusters, but very effective in small groups*. Note that without resorting to partial orbit circularization within groups, the expected number of mergers in groups is somewhat less than expected from the $`v_{\mathrm{cl}}^3`$ scaling, because, relative to clusters, the more concentrated NFW profiles in groups lead to stronger modulation of the merger rate by the potential tides.
## 4. Collisional tidal stripping in clusters and groups
Because non-merging galaxy collisions are by essence rapid, they can be treated as tidal shocks, and it is reasonable to assume that for tidal features to be visible, one requires $`\mathrm{\Delta }E\gamma |E|`$, where $`\gamma <1`$. Hence, $`\mathrm{\Delta }v(3\gamma )^{1/2}v_g`$. Denoting $`p`$ and $`V`$ the separation and relative velocity at pericenter, and $`v_{\mathrm{circ},\mathrm{p}}(p)`$ the circular velocity of the perturbing galaxy out to $`p`$, eq. (4) leads to a critical impact parameter
$$p_{\mathrm{crit}}=\frac{r}{(3\gamma )^{1/2}}\frac{v_{\mathrm{circ},\mathrm{p}}^2}{v_gV},$$
(17)
where we note that $`v_{\mathrm{circ},\mathrm{p}}`$, is almost independent of $`p`$ for realistic density profiles for the perturbing galaxy. Then, integrating the cross-sections derived from eq. (17), the rate of tidal interactions is
$$k=vs(v)=\frac{\pi ^{1/2}}{3\gamma }\left(\frac{r}{v_g}\right)^2\frac{v_{\mathrm{circ},\mathrm{p}}^4}{v_{\mathrm{cl}}},$$
(18)
and is virtually independent of the test galaxy parameters. Integrating eq. (18) over perturber mass and remarking that both the galaxy and the perturbers are tidally limited by the cluster potential, one obtains using eq. (5)
$$n\overline{k}=\frac{\mathrm{\Gamma }(7/3\alpha ,x_m)}{4\pi ^{1/2}\gamma G\mathrm{\Gamma }(2\alpha ,x_m)}\left(\frac{v_{\mathrm{circ}}}{v_g}\right)^2\frac{v_{\mathrm{circ},}^4}{m_{}}\frac{\rho _{\mathrm{cl}}(R)\mu ^{7/3}(R)}{\overline{\rho }_{\mathrm{cl}}(R_p)v_{\mathrm{cl}}(R)}\left[\frac{V_p}{V_{\mathrm{circ}}(R_p)}\right]^2,$$
(19)
where again $`\mu (R)=m(r_t)/m(r_{\mathrm{vir}})`$ (see eq. ), $`v_{\mathrm{circ},}`$ is the circular velocity at the virial radius for a field $`m_{}`$ galaxy. Again, *the rate of tidal encounters is independent of the galaxy mass*. Note that the $`\mu ^{7/3}(R)`$ dependence of the rate of tidal encounters illustrates the strong modulation of this rate by the tides from the cluster potential.
Figure 5 shows the expected (eq. ) number of strong tidal collisions for galaxies in clusters and groups, with $`R/R_p`$ as in Figure 4.
Although groups are preferential sites for strong tidal encounters, galaxies in the outskirts of clusters should also witness such interactions. However, the signature of tidal interactions lasts of order $`1h^1\mathrm{Gyr}`$, so that the fraction of galaxies in clusters and groups that are currently undergoing tidal interactions is roughly one-tenth of what is displayed in Figure 5.
## 5. Discussion
The strong radial dependence of galaxy masses, predicted by the tidal theory (eq. 8), is clear in the cosmological simulations of Ghigna et al. (1998). Should we then witness *inverse luminosity segregation* in clusters where, outside of the core, galaxies become more luminous towards the cluster periphery? Indeed, Adami, Biviano, & Mazure (1998) found a weak trend of mean galaxy magnitude versus radius for an ensemble of clusters, although they worry that this trend is caused by observational bias. It may be that incompleteness of the observational samples is washing out the trend rather than creating it.
The lack of mergers in present-day rich clusters has been noted in cosmological simulations of clusters (Ghigna et al., 1998). Figure 4 shows that in rich clusters, mergers are at best marginally probable for high mass galaxies lying in the cluster body. Given that high mass galaxies are rare, such merging will be difficult (but not impossible) to detect observationally or in simulations.
From their H$`\alpha `$ prism surveys of galaxies in clusters, Moss and co-workers (Moss & Whittle, 1993; Moss, Whittle, & Pesce, 1998; Bennett & Moss, 1998) note $`30\%`$ of spiral galaxies in rich clusters exhibit a *compact* H$`\alpha `$ morphology and roughly half of these tend to be morphologically disturbed *and* have nearby neighbors. Another half of these compact H$`\alpha `$ emission galaxies are in the cluster core, and presumably those have no tidal companions, suggesting that they are harassed by the cluster potential. But the first half, outside the core are probably *bona fide* cases of strong tidal interactions within clusters, leading to an overall frequency of 20% of all galaxies outside the cores of clusters. It remains to be seen if they are associated to substructures such as infalling groups. Correcting Figure 5 for the 1 Gyr duration of these tidal features produces an absolute frequency of tidally interacting galaxies in clusters in rough agreement with that found by Moss and co-workers. |
no-problem/9911/cond-mat9911356.html | ar5iv | text | # Treating Some Solid State Problems with the Dirac Equation
## Abstract
The ambiguity involved in the definition of effective-mass Hamiltonians for nonrelativistic models is resolved using the Dirac equation. The multistep approximation is extended for relativistic cases allowing the treatment of arbitrary potential and effective-mass profiles without ordering problems. On the other hand, if the Schrödinger equation is supposed to be used, our relativistic approach demonstrate that both results are coincidents if the BenDaniel and Duke prescription for the kinetic-energy operator is implemented. Applications for semiconductor heterostructures are discussed.
PACS numbers: 74.80.-g, 03.65.Pm, 73.40.Kp
The effective-mass theory has been successfully used in semiconductor heterostructures . An interesting aspect arises when we treat materials whose properties change from region to region. In particular, when the effective mass depends on position, the Schrödinger equation for an arbitrary potential profile is usually solved numerically by different methods. However, one of the problems of the effective-mass theory for semiconductors heterostructures, is to decide how to write out the Hamiltonian operator. This problem arise from canonical quantization of the classical Hamiltonian. For position-dependent carrier effective mass, we have an ordering problem with the kinetic energy operator (KEO). Some authors proposed different forms for the kinetic energy operator, all having the generic form proposed by von Roos
$$\widehat{T}=\frac{1}{4}\left(m^\alpha (x)\widehat{p}m^\beta \widehat{p}m^\gamma (x)+m^\gamma (x)\widehat{p}m^\beta \widehat{p}m^\alpha (x)\right),$$
(1)
where $`\alpha +\beta +\gamma =1`$, but the problem was not resolved because there is not a first principle to fix only one operator.
This ambiguity indicates that the Schrödinger equation is not rigorously suitable in the effective-mass aproximation with position-dependent effective mass. It is reasonable to try other equation that represents the same physics of the Schrödinger equation in the low energy limit.
The object of this Letter is to demonstrate that, in fact, the Dirac equation (at adequate limits) can successfully be used to describe quantum-mechanical systems where position-dependent effective mass is present. We recall that, in Dirac equation, the kinetic energy operator and the mass term appear separately, so there is no ordering problems in this context.
Obviously the considerations of this work only concern the mathematical issues related with the equations of motion. Indeed, a physically sensible application of the Dirac equation to semiconductor heterostructures would have to take in to account a relativistic extension of the Wannier-Slater theorem .
In this work we use a numerical method (multistep potential approximation ) which has been applied to solve Schrödinger equation for an arbitrary potential profile. Here we extend this algorithm to the relativistic case, in such a way that the ambiguity problem is overcome. In particular, we consider the Dirac equation with a one dimensional arbitrary potential well and find the energy levels for a particle. Also we apply the method for a particular type of heterostructure and compare the results to those obtained in the context of the Schrödinger equation with the form (1) for the KEO and several choices for the parameters $`\alpha ,\beta `$ and $`\gamma .`$ For a KEO in the Schrödinger equation of the form suggested by BenDaniel and Duke ($`\beta =1,\gamma =0`$) , we conclude that both equations lead to the same solutions in the energy range concerned.
Let us now introduce the numerical method that allows us to obtain the energy levels for a Dirac equation with space-dependent effective mass.
We will consider a particle with mass $`m(z)`$ that is submitted to an arbitrary one dimensional potential well $`V(z)`$. The time-independent Dirac equation is written as (in units with $`\mathrm{}=c=1`$)
$$\left(\alpha \widehat{p}+\beta m(z)\right)\mathrm{\Psi }=\left(EV(z)\right)\mathrm{\Psi },$$
(2)
where, $`\widehat{p}=i{\displaystyle \frac{d}{dz}}`$ is the momentum operator, $`E`$ is the electron energy, $`\alpha `$ and $`\beta `$ are $`4\times 4`$ matrices given by
$$\alpha =\left(\begin{array}{cc}0& \sigma ^3\\ \sigma ^3& 0\end{array}\right),\beta =\left(\begin{array}{cc}I\hfill & \hfill 0\\ 0\hfill & \hfill I\end{array}\right),$$
(3)
$`I`$ is the $`2\times 2`$ identity matrix and $`\sigma ^3`$ is a $`2\times 2`$ Pauli matrix defined as
$$\sigma ^3=\left(\begin{array}{cc}1\hfill & \hfill 0\\ 0\hfill & \hfill 1\end{array}\right)$$
(4)
Consider now an arbitrary well as sketched in fig.1. We split up the interval $`[a,b]`$ into $`N`$ infinitesimal intervals of length $`\mathrm{\Delta }z=(ba)/N`$. For the $`i`$-th interval, we approximate the potential and the mass by
$$V(z)=V(z_i)=V_iandm(z)=m(z_i)=m_i,forz_iz<z_{i+1}.$$
(5)
The wave function of the electron in Dirac’s equation with no spin flip in the $`i`$th interval is
$$\psi _i\left(z\right)=A_ie^{ip_iz}(\begin{array}{c}1\\ 0\\ \frac{p_i}{Ev_i+m_i}\\ 0\end{array})+B_ie^{ip_iz}(\begin{array}{c}1\\ 0\\ \frac{p_i}{Ev_i+m_i}\\ 0\end{array}),$$
(6)
where $`p_i=\sqrt{\left(EV_i\right)^2m_i^2}`$. The bound state conditions are given by $`|EV_0|<m_0`$ and $`|EV_{N+1}|<m_{N+1}`$. By imposing the continuity of the wave function at each $`z=z_i`$, we have a matrix $`M(E)`$ which relates the coefficients in the region where $`z<a`$ with the region where $`z>b`$.
$$\left(\begin{array}{c}A_{N+1}\\ B_{N+1}\end{array}\right)=M(E)\left(\begin{array}{c}A_0\\ B_0\end{array}\right)$$
(7)
The finiteness of the wave function requires that
$$M(E)_{22}=0$$
(8)
So, the solution of (8) give us the energy levels.
Note that this numerical method is specially convenient for treating wells and barriers with arbitrary profiles and it is nothing else than an extension of the transfer-matrix method for relativistic theories. To the best of our knowledge, this is the first numerical analysis for evaluating energy levels in an relativistic equation with mass position-dependent and arbitrary potential. However, our main point here is to demonstrate that the Dirac equation can be used to obtain unambiguously results in situations where the Schrödinger equation depends on ordering problems.
As an illustration, we applied the method described above for an electron in a one-dimensional GaAs/Al<sub>0.3</sub>Ga<sub>0.7</sub>As heterostructure. For the sake of comparison with previous results we take a square well, as sketched in fig. 2. The electron effective mass is $`0.67m_0`$ and $`0.86m_0`$ for GaAs and Al<sub>0.3</sub>Ga<sub>0.7</sub>As respectively ( $`m_0`$ is the free electron mass)(Fig. 2). In the Schrödinger equation we use the von Ross operator considering several values of the $`\alpha `$ parameter. Note that for abrupt heterojunctions only Hamiltonians with $`\alpha =\gamma `$ are viable, due to continuity conditions across the heterojunction . As we can see in fig. 3 there is an extraordinary coincidence between the results from the Dirac and from the Schrödinger equations for $`\alpha =0`$. As a matter of fact, this is expected since the maximum value of the energy involved is 3eV. Further, this result strongly supports the prescription of BenDaniel and Duke for the Schrödinger context.
As a second illustration consider the conduction-band structure of a GaAs/Al<sub>0.3</sub>Ga<sub>0.7</sub>As system with nonabrupt interface and assume that the effective mass changes linearly at the transition regions while the potential well varies quadratically in that regions (denoted by $`a`$) as it is shown in fig. 4. There the potential is given by
$$V(z)=C\left[ϵ_1\chi (z)+ϵ_2\chi (z)^2\right]\text{ ,}$$
(9)
where $`C=0.6`$ is the conduction band offset and $`ϵ_1=0.3`$, $`ϵ_2=0.7`$ are constants associated with the compositional dependence of the energy-gap difference between GaAs and AlGaAs (experimental parameters and details concerned can be seen in refs. ).
Again, we use our relativistic method and the Schrödinger equation with the BenDaniel and Duke prescription ($`\alpha =\gamma =0`$). Once more, as we can see in fig. 5, a complete coincidence between the relativistic and non-relativistic results is obtained.
In conclusion, we have shown that a relativistic method can be successfully used to overcome the ordering problem of the kinetic energy operator in non-relativistic models. Since the range of energy involved is extremely low (comparing to electron rest mass), the numerical results are perfectly coincident in both cases.
It is worthwhile to mention that, notwithstanding the Wannier-Slater theorem commented in the introduction, we claim attention for the coincidence shown above. Therefore, we believe that, after exhaustive and appropriate considerations about effective-mass approximation, band structures and the periodic potential, the relativistic approach constructed here it could be used to calculate physical parameters in the theory of abrupt and nonabrupt semiconductor heterostructures.
Moreover, the relativistic treatment developed here can be applied to all physical systems described by a Sturm-Liouville eigenvalue equation (within an appropriate range of energy), namely
$$\frac{d}{dz}\left[\frac{1}{p(z)}\frac{df}{dz}\right]+q(z)f(z)=\lambda w(z)f(z)\text{ }.$$
(10)
Here $`f`$ is an eigenfunction, $`\lambda `$ is an eigenvalue and $`p`$, $`q`$, and $`w`$ describe particular properties of the system. For example, this equation can describe the motion of electrons or phonons along the growth axis of a zinc-blende heterostructure . Thereby, the equation (10) can always be replaced by an correspondent Dirac equation for avoid ordering problems.
Acknowledgments
M. H. Pacheco was supported in part by Conselho Nacional de Desenvolvimento Científico e Tecnológico-CNPq.
FIGURE CAPTIONS
Figure 1: Generic profile of a one-dimensional quantum potential well. We split up the interval $`[a,b]`$ into $`N`$ infinitesimal intervals of length $`\mathrm{\Delta }z=(ba)/N`$
Figure 2: Potential well and effective mass, as a function of the position, for an abrupt heterostructure.
Figure 3: Eigenenergies in the conduction band of the GaAs/Al<sub>0.3</sub>Ga<sub>0.7</sub>As, with a conduction band-offset of $`0.6`$ vs the well width (abrupt heterojunction). The chosen value for $`a`$ is $`20\%`$ of the well width. Solid line shows the calculations performed using the Dirac equation. Dotted line shows the calculations for Schrödinger equation using the BenDaniel-Duke prescription ($`\alpha =0`$). The + + curve denotes calculations using the Zhu and Kroemer prescription ($`\alpha =0.5`$).
Figure 4: Potential well and effective mass, as a function of the position, for a nonabrupt heterostructure ( $`a`$ denotes the transition region).
Figure 5: Eigenenergies in the conduction band of the GaAs/Al<sub>0.3</sub>Ga<sub>0.7</sub>As, with a conduction band-offset of $`0.6`$ vs the well width (nonabrupt heterojunction). The chosen value for $`a`$ is $`20\%`$ of the well width. Solid line shows the calculations performed using the Dirac equation and dotted line shows the calculations for Schrödinger equation using the BenDaniel-Duke prescription ($`\alpha =\gamma =0`$). The + + curve denotes calculations using the Zhu and Kroemer prescription ($`\alpha =\gamma =0.5`$). |
no-problem/9911/quant-ph9911042.html | ar5iv | text | # Spin-Boson Hamiltonian and Optical Absorption of Molecular Dimers
## I Introduction
The spectral properties of the spin-boson Hamiltonian have frequently been adressed in the past and in the last years investigations on the spectral features of classical nonintegrability in the quantum case were performed (see e.g. and references therein). The interest in the spectral and dynamical properties of the spin-boson Hamiltonian is due to its numerous applications which include a large variety of phenomena in molecular and solid state physics with one realization being the optical and transfer properties of excitonic-vibronic coupled dimers. Thus the spectral and related optical properties of the corresponding dimer Hamiltonian have intensively been investigated over the years, as examples we point to the work exposed in . Most of these investigations were based on a Hamiltonian corresponding to a symmetric dimer configuration. However, from the point of view of a general asymmetric situation the symmetric case is singular, e.g. the quantum states possess parity properties in the symmetric case, which are not present in general. It should be noted that dimers often constitute subunits of more complicated molecular aggregates with an asymmetric structure, which breaks the symmetry of the dimer configuration (see e.g. ). Therefore the aim of this paper is to go beyond the symmetric case and to connect the properties of the eigenstates of a generalized and symmetry-broken spin-boson Hamiltonian with the line spectrum of absorption bands of asymmetric excitonic-vibronic coupled molecular dimers.
We combine a phase space analysis of the eigenstates, which is based on the method of Husimi projections , with the calculation of the matrix elements of optical transitions. Husimi projections have repeatedly been used in order to establish the features of nonintegrability of classical systems in the corresponding quantum case . Analyzing the Husimi projections of the eigenstates of the spin-boson Hamiltonian in relation to optical transitions we indicate a bridge between the eigenstate analysis of nonintegrable systems and observables in an optical experiment. In particular, we show how the intensity variations in the Husimi projections are related to the spectral randomness in the fine structure of absorption bands. Such spectral randomness in the eigenstates of the spin-boson Hamiltonian is also known as incipience of quantum chaos . In this case random features of the spectrum are just appearing while the system is still far from displaying universal spectral fluctuations described by random matrix theory and known for excited states of spectra of polyatomic molecules .
A central point in our eigenstate analysis will be to find out to what extent the adiabatic reference systems of the spin-boson Hamiltonian are present in its exact eigenstates and how the appearance of spectral randomness can be interpreted as a mixing of such reference systems. Following we show that Bloch projections are a useful quantitative characteristic to describe this mixing. Computing the Bloch projections of the eigenstates, it is possible to distinguish the spectral region where the adiabatic branches of the spectrum are still intact from the region with a substantial mixing of adiabatic reference systems. Furthermore, by performing projections of the numerically obtained eigenstates onto the ground state we obtain the details of the fine structure of the absorption bands in the spectral regions as indicated by the Franck-Condon principle. We note that the asymmetric case has some special features, which are not present in the symmetric situation: In the asymmetric case the final states of an optical transition can be located in the higher energetic regions of the corresponding adiabatic potentials. Then, as will be shown below for two representative cases, the fine structure of the spectrum and of the absorption bands can greatly vary from regular to irregular arrangements of absorption lines.
The paper is organized as follows: In Sec.II the model and the basic equations are presented. In Sec.III the eigenstate analysis for two representative cases of the asymmetric spin-boson Hamiltonian is performed and in Sec.IV this analysis is connected with the properties of the absorption bands in the corresponding spectral regions. Finally, in Sec.V our conclusions are summarized.
## II Model
We consider a dimer system described by the Hamiltonian
$$\widehat{H}=\underset{n}{}(ϵ_n+\gamma _nq_n)|nn|+\underset{\stackrel{n,m}{nm}}{}V_{nm}|nm|+\underset{n}{}\frac{1}{2}(p_n^2+\omega _n^2q_n^2),$$
(1)
where $`|n`$ ($`|m`$) are the excited states of the molecular monomers constituting the dimer ($`n,m=1,2`$) with the excitonic energies $`ϵ_n`$, the coupling constants to intramolecular vibrations $`\gamma _n`$ and the transfer matrix element $`V_{nm}`$. The variables $`q_n`$, $`p_n`$ and $`\omega _n`$ represent coordinate, momentum and frequency of the vibrations, respectively. The coupling of the excited dimer states to an incident light wave with electrical field vector $`\stackrel{}{E}_n(t)`$ is given by
$$\widehat{H}_{int}=\underset{n}{}\stackrel{}{\mu }_n\stackrel{}{E}_n(t)|n0|,$$
(2)
where $`\stackrel{}{\mu }_n`$ and $`\stackrel{}{E}_n(t)`$ are the optical transition matrix element and electric field at the nth monomer, respectively, and $`|0`$ is the excitonic vacuum. We consider an asymmetric configuration of the dimer system by assuming different monomer energies $`ϵ_n`$ and optical transition matrix elements $`\stackrel{}{\mu }_n`$. In order to keep the number of asymmetry parameters minimal we assume symmetric transfer rates, $`V_{nm}=V_{mn}=V`$, $`V>0`$, and equal coupling constants and frequencies, i.e. $`\gamma _1=\gamma _2=\gamma `$ and $`\omega _1=\omega _2=\omega `$. Then the coupling can be reduced to a single vibrational mode by introducing the relative displacement of the vibrations.
The Hamiltonian (1) can be represented as an operator in the space of two-dimensional vectors $`C=\left(\genfrac{}{}{0pt}{}{c_1}{c_2}\right)`$ using the standard Pauli spin matrices $`\widehat{\sigma }_i`$ ($`i=x,y,z`$). Passing to dimensionsless variables by measuring the energy in units of $`2V`$, $`H=H/2V`$, one obtains from (1) the spin-boson Hamiltonian
$$\widehat{H}=ϵ_+𝟙\frac{\mathrm{𝟙}}{\mathrm{𝟚}}\widehat{\sigma }_𝕩+\frac{\mathrm{𝟙}}{\mathrm{𝟚}}(\widehat{}^\mathrm{𝟚}+𝕣^\mathrm{𝟚}\widehat{}^\mathrm{𝟚})𝟙+(\sqrt{\frac{𝕡}{\mathrm{𝟚}}}𝕣\widehat{}+ϵ_{})\widehat{\sigma }_𝕫.$$
(3)
In (3), the dimensionless relative displacement and corresponding momentum, to which the exciton is coupled, are given by $`\widehat{Q}=\sqrt{2V}(\widehat{q}_1\widehat{q}_2)`$ and $`\widehat{P}=(\widehat{p}_1\widehat{p}_2)/\sqrt{2V}`$. The dimensionless parameters of the spin-boson Hamiltonian (3) are related to the dimer system (1) by
$`p`$ $`=`$ $`{\displaystyle \frac{\gamma ^2}{2V\omega ^2}},`$ (4)
$`r`$ $`=`$ $`{\displaystyle \frac{\omega }{2V}},`$ (5)
$`ϵ_\pm `$ $`=`$ $`{\displaystyle \frac{ϵ_1\pm ϵ_2}{4V}}.`$ (6)
The parameter $`p`$ expresses the coupling strength, and $`r`$ is the adiabatic parameter measuring the relative strength of quantum effects of the subsystems. The sum $`ϵ_+`$ represents the center of the excitation energy, whereas the difference $`ϵ_{}`$ is the asymmetry parameter. For $`ϵ_{}=0`$, one obtains the symmetric spin-boson Hamiltonian with conservation of total parity (given by the operation $`\widehat{Q}\widehat{Q}`$, $`\widehat{\sigma }_z\widehat{\sigma }_z`$). For $`ϵ_{}0`$, this symmetry is broken.
The adiabatic reference systems associated with (3) are introduced by considering the eigenvalue problem of the adiabatic part $`\widehat{h}_{\mathrm{ad}}`$,
$$\widehat{h}_{\mathrm{ad}}=\frac{1}{2}\widehat{\sigma }_x+(\sqrt{\frac{p}{2}}rQ+ϵ_{})\widehat{\sigma }_z+\frac{r^2}{2}Q^2𝟙.$$
(7)
The $`Q`$-dependent eigenvalues and eigenstates of $`\widehat{h}_{\mathrm{ad}}`$ are easily obtained. In particular, one finds the two adiabatic potentials $`U_{\mathrm{ad}}^\pm (Q)`$ for the upper (+) and lower (-) states
$$U_{\mathrm{ad}}^\pm (Q)=\frac{r^2}{2}Q^2\pm \sqrt{\frac{1}{4}+\left(ϵ_{}+\sqrt{\frac{p}{2}}rQ\right)^2},$$
(8)
from which the Hamiltonians of the adiabatic reference systems are obtained
$$H^\pm (Q)=\frac{1}{2}P^2+U_{\mathrm{ad}}^\pm (Q).$$
(9)
A key point in our analysis of the eigenstates and absorption bands will be the extent to which the exact eigenstates of the Hamiltonian (3) can be viewed as a mixing of adiabatic reference systems connected with the adiabatic Hamiltonians (9). As is shown below, the extent of this mixing can be controlled by analyzing Husimi projections and by computing Bloch projections from the numerically obtained eigenstates. Husimi projections constitute phase space representations of the eigenstates which can be compared with the phase space orbits of the adiabatic reference systems (9). Bloch projections constitute another independent indicator of the mixing of the adiabatic branches in the spectrum. For the adiabatic branches these projections are calculated from the eigenstates of the Hamiltonian (7)
$`|\phi _{\mathrm{ad}}^+(Q)`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(\sqrt{1+A(Q)}|\sqrt{1A(Q)}|\right)`$ (11)
$`|\phi _{\mathrm{ad}}^{}(Q)`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(\sqrt{1A(Q)}|+\sqrt{1+A(Q)}|\right)`$ (12)
where
$$A(Q)=\frac{2ϵ_{}+\sqrt{2p}rQ}{\sqrt{1+(2ϵ_{}+\sqrt{2p}rQ)^2}}.$$
(13)
Computing the expectation values $`x=\widehat{\sigma }_x`$ of the Pauli matrix $`\widehat{\sigma }_x`$ for the adiabatic eigenstates (II), one obtains the Bloch projections of the adiabatic branches
$$\phi _{\mathrm{ad}}^\pm (Q)|\widehat{\sigma }_x|\phi _{\mathrm{ad}}^\pm (Q)=\frac{1}{\sqrt{1+(2ϵ_{}+\sqrt{2p}rQ)^2}},$$
(14)
According to eq.(14) there is a distinction in the sign of the Bloch projections for the adiabatic eigenstates. Computing the Bloch projections from the numerically obtained eigenstates this distinction in the sign will be used as an indicator for the case in which the spectrum resolves into adiabatic branches.
## III Eigenstate analysis
Our analysis is based on a diagonalization of the spin-boson Hamiltonian using a basis of product states of spin states with harmonic oscillator eigenstates. In this basis the eigenstates are obtained in the form
$$|\lambda =\underset{m}{}\left(c_m^{(\lambda )}|+c_m^{(\lambda )}|\right)|m=\underset{m}{}|s_m^{(\lambda )}|m,$$
(15)
where $`|=\left(\genfrac{}{}{0pt}{}{1}{0}\right)`$, $`\left(|=\left(\genfrac{}{}{0pt}{}{0}{1}\right)\right)`$ denote the spin up (down) states, $`|m`$ the harmonic oscillator eigenstates and $`c_{zm}^{(\lambda )}`$ ($`z=,`$) the expansion coefficients of the $`\lambda `$ eigenstate. For a fixed $`m`$ we will also use the spin representation $`|s_m^{(\lambda )}`$ for the up and down coefficients $`c_{zm}^{(\lambda )}`$ ($`z=,`$) for a given oscillator state as indicated in the second part of (15). The matrix dimension in the numerical diagonalization was $`N=4000`$, for the eigenstate analysis the first $`1100`$ eigenstates were used.
For the phase space representation of the eigenstates Husimi projections are used. Husimi projections $`h_\lambda (\alpha (Q,P);s)`$ are defined by projecting an eigenstate $`|\lambda `$ onto a set of coherent oscillator states $`|\alpha (Q,P)`$, which scan the oscillator phase plane of the vibrational subsystem, while the spin projection $`|s=c_{}|+c_{}|`$ is kept fixed,
$$h_\lambda (\alpha (Q,P);s)=|\lambda |\alpha (Q,P),s|^2,$$
(16)
and $`|\alpha (Q,P),s=|\alpha (Q,P)|s`$. Using the explicit form of $`|\lambda `$ in (15) one finds
$$h_\lambda (\alpha (Q,P);s)=\left|\underset{m}{}(c_m^{(\lambda )}c_{}+c_m^{(\lambda )}c_{})m|\alpha (Q,P)\right|^2,$$
(17)
where $`m|\alpha (Q,P)=\frac{\alpha ^m}{\sqrt{m!}}\mathrm{e}^{\frac{|\alpha |^2}{2}}`$ and $`\alpha (Q,P)=\sqrt{\frac{r}{2}}\widehat{Q}+\frac{i}{\sqrt{2r}}\widehat{P}`$.
For the Bloch projection of an eigenstate,
$$x_\lambda =\lambda \left|\widehat{\sigma }_x\right|\lambda ,$$
(18)
one finds from the expansion (15)
$$x_\lambda =\underset{m}{}\left(c_m^{(\lambda )}c_m^{(\lambda )}+c_m^{(\lambda )}c_m^{(\lambda )}\right).$$
(19)
Before turning to the eigenstate analysis for definite parameter sets we note that the appearance of spectral irregularities can be expected in the high energetic region of the overlap of both adiabatic potentials only, where a sufficient number of states of both potentials mix. It is obvious that such a region can be probed by the final states of an optical transition in the asymmetric case only: In the symmetric case, as is evident from the adiabatic potentials (8), the optical transitions would necessarily terminate in the ground state region of the upper potential, where only a few states of the upper potential can mix with the lower potential. Therefore when investigating the structure of the spectrum and eigenstates of the spin-boson Hamiltonian we paid special attention to a systematic change of the asymmetry parameter $`ϵ_{}`$ combined with the coupling parameter $`p`$ in order to produce appropriate configurations of the adiabatic potentials. Such configurations were found for relatively high coupling $`p`$ and asymmetry values $`ϵ_{}`$. Below we present the eigenstate analysis for two typical asymmetric configurations in the adiabatic parameter region ($`r<1`$) with the parameter sets
A: $`p=4`$, $`r=0.1`$ and $`ϵ_{}=5`$ and
B: $`p=20`$, $`r=0.1`$ and $`ϵ_{}=10`$.
The parameter set B corresponds to a stronger coupling value $`p`$ and a larger asymmetry parameter $`ϵ_{}`$, as compared to set A. To indicate the location of the absorption bands, which will be analyzed in section IV, the adiabatic potentials of set B and the ground state of the system are shown in Fig. 1.
In Fig. 2 we start with the eigenstate analysis for the parameter set A by displaying the Bloch projections $`x(E_\lambda )`$ computed from the eigenstates according to eq.(19). In the low energy region one finds one adiabatic branch in the $`x(E_\lambda )`$ dependence associated with the lower adiabatic potential (for which $`x_\lambda >0`$ in accordance with the sign in (14)). The overlap of the two potentials is marked by the appearance of a second adiabatic branch in the $`x(E_\lambda )`$ dependence. The mixing of the eigenstates of the two adiabatic potentials is seen by the disappearance of both adiabatic branches and indicated by a broad band of Bloch projections in the high energy region. The presence of three characteristic regions in the Bloch projections with one adiabatic branch, two adiabatic branches and the mixing region is reflected in the Husimi projections of the eigenstates. In the regions where the adiabatic branches are intact one obtains regular Husimi projections concentrated around the phase space orbits of the integrable adiabatic reference Hamiltonians (9). Typical examples of such regular projections corresponding to equal spin projections $`c_{}=c_{}`$ for the energetic overlap region with two adiabatic branches present are shown in Fig. 3(a) and Fig. 3(b). In Fig. 3(a) the Husimi projection corresponding to an eigenstate in the upper adiabatic potential is shown, whereas in Fig. 3(b) a projection associated with an eigenstates of the lower adiabatic potential is displayed. Sequences of such projections clearly attributable to the coexistence of two independent adiabatic branches were observed until the mixing region is reached. Then Husimi projections incorporating the phase space features of both adiabatic potentials are observed, as shown in Fig. 3(c).
A closer inspection of these projections in the mixing region shows that although the distribution is concentrated on the regular phase space orbits of the adiabatic potentials, the intensity between these parts varies in an irregular and random way, when Husimi projections for a sequence of eigenstates are considered. The structure of the eigenstates which shows up in the random shift of intensity in the Husimi projections is a source of spectral randomness in the absorption bands. In the case of the parameter set A, however, this region of an irregular and random shift of the phase space distribution is not reached by the absorption processes. In this case the absorption band is located in the region, where the two adiabatic branches are still intact and no spectral mixing between their eigenstates occurs.
In the case of the parameter set B one finds a situation, where the spectral window, which is cut by the transition matrix elements of the absorption process, reaches the region of an irregular mixing of adiabatic states. The Husimi projections for selected eigenstates and equal spin projections which correspond to such a region and which are relevant as final states of an optical transition are shown in the set of Fig. 4(a)-(c). One observes a random variation of the intensity in the Husimi distributions of the eigenstates between the orbits of the two adiabatic potentials. As we show in the next section, as a consequence of this random variation intensities of lines for optical transitions terminating in these states vary randomly.
## IV Absorption
In the calculation of the optical absorption the dimer system is assumed to be small compared to the optical wavelength. Then the electric field vector is approximately equal at the monomer sites of the dimer configuration, i.e. $`\stackrel{}{E}_n(t)=\stackrel{}{E}(t)`$. Introducing the projections $`\mu _n`$ of the monomer transition matrix elements onto the direction of the field polarization, i.e. $`\stackrel{}{E}(t)\stackrel{}{\mu }_n=E(t)\mu _n`$, ($`n=1,2`$), the interaction of the dimer with the incident light wave is represented in the form
$$\widehat{H}_{int}=E(t)\widehat{M}_{int},$$
(20)
where $`\widehat{M}_{int}`$ is the interaction operator
$$\widehat{M}_{int}=\underset{n}{}|n\mu _n0|.$$
(21)
In the ground state there is no excitonic-vibronic interaction, i.e. the ground state of the system $`|g`$ is given by the direct product of the excitonic vacuum $`|0=|exc0`$ and the vibrational ground state. For the ground state of the vibrational subsystem we assume the zero temperature case, where it is in its lowest state $`|m=0`$, with $`|m=0`$ the $`m=0`$ Hermitian polynomial of the undisplaced $`Q`$ oscillator. Then $`|g=|exc0|m=0`$. For the optical transition matrix element $`M_{\lambda g}=\lambda |\widehat{M}_{int}|g`$ of the interaction operator (21), which corresponds to an absorption process from the ground state $`g`$ into an excited state $`\lambda `$, one then obtains using eq.(15) the expression
$$M_{\lambda g}=\mu _1c_0^{(\lambda )}+\mu _2c_0^{(\lambda )}.$$
(22)
The matrix element (22) is representable in the form of a particular projection of the spin vector associated with the $`m=0`$ vibrational state in (15), $`|s_0^{(\lambda )}`$. Introducing the angles
$$\mathrm{cos}\alpha =\frac{\mu _1}{\sqrt{\mu _1^2+\mu _2^2}},\mathrm{sin}\alpha =\frac{\mu _2}{\sqrt{\mu _1^2+\mu _2^2}}$$
(23)
and a spin vector $`|s_\mu `$ defined by
$$|s_\mu =\left(\genfrac{}{}{0pt}{}{\mathrm{cos}\alpha }{\mathrm{sin}\alpha }\right),$$
(24)
one represents (22) in the form
$$M_{\lambda g}=\sqrt{\mu _1^2+\mu _2^2}s_\mu |s_0^{(\lambda )}.$$
(25)
Measuring the square of (25), i.e. the absorption strength of an optical transition from $`|g`$ to $`|\lambda `$, in units of $`(\mu _1^2+\mu _2^2)`$ one finds for the dimensionsless absorption strength $`Q_{\lambda g}`$
$$Q_{\lambda g}=\frac{M_{\lambda g}^2}{\mu _1^2+\mu _2^2}=(s_\mu |s_0^{(\lambda )})^2.$$
(26)
Comparing eq.(26) with the expression for the Husimi projection (17) it follows that $`Q_{\lambda g}`$ is equal to a Husimi projection of the $`\lambda `$-eigenstate of (3) taken at the phase space point $`Q=P=0`$ with the spin projection being fixed at $`s_\mu `$ : For $`Q=P=0`$ one obtains $`\alpha (Q,P)=0`$, which selects the $`m=0`$ term in the sum of (17), inserting the value $`s_\mu `$ for the spin vector $`s`$, one finds that $`h_\lambda (0;s)`$ reduces to the r.h.s. of eq.(26), i.e.
$$Q_{\lambda g}=h_\lambda (0;s_\mu ).$$
(27)
Eq. (27) is in line with the Franck-Condon principle for optical transitions: The transition is vertical from the ground state region at $`Q=P=0`$ in the phase space representation and probes the intensity of the final state by its Husimi distribution in the same region. The optical transition matrix elements of the monomers define the spin projection $`s_\mu `$. For the special case of a symmetric dimer with $`\mu _1=\mu _2`$ the transition occurs into the symmetric combination of monomer states. Differences in the optical transition matrix elements, i.e. $`\mu _1\mu _2`$, introduce a second asymmetry parameter besides that of the site energy asymmetry $`ϵ_{}`$. These differences in the optical transition matrix elements express the optical asymmetry of the dimer and enter the optical dimer matrix element (22) through the spin projections (23).
In Fig. 5(a) the absorption bands $`Q_{\lambda g}`$ calculated from the numerically obtained eigenstates for the parameter set A are represented as stick spectra. Absorption bands are located in accordance with the Franck-Condon principle. This is seen by comparing their positions with the final states following from the adiabatic potentials in (8) by setting $`Q=0`$. One finds an energetically lower band for optical transitions terminating in the lower adiabatic potential and a higher band for transitions occuring into the region of the overlap of the two potentials. The shape of the lower band on the left side of Fig. 5(a) is completely regular. The lines of the higher band on the right side of Fig. 5(a) are also regularly arranged. A closer inspection of the higher band, however, reveals that between the lines shown in Fig. 5(a) lines with a much weaker intensity are embedded, which form another regular band. This is evident from the inset, Fig. 5(b), where parts of the lines of the weak and strong bands are shown together. In order to make the weak band visible in the inset a much smaller intensity scale is used. The two bands, into which the absorption spectrum of the high energy part resolves, are easily identified as to belong to the states of the two coexisting adiabatic branches analyzed in the preceding section for the parameter set A: The weak band with relatively small intensities is due to optical transitions terminating in the high energy states of the lower potential (the sign of the Bloch variables of the final states is $`x_\lambda >0`$ in accordance with the sign of the lower adiabatic branch (14)), whereas the strong band with much greater intensities is due to optical transitions into the low energy states of the upper potential (for these states $`x_\lambda <0`$, again in accordance with the sign of the adiabatic branch (14)).
In Fig. 6 the absorption bands are shown for the parameter set B. The bands are located in the spectral regions of the Franck-Condon energies indicated by the arrows in Fig. 1. The lower absorption band, shown on the left side of the Fig. 6, is regular like the lower band in the Fig. 5(a). The higher band, displayed on the right side of Fig. 6, however, is completely irregular and cannot be resolved into independent subbands as in the case of the parameter set A. This is a consequence of the mixing of the adiabatic reference systems and the irregular structure of the eigenstates analyzed by the Husimi projections in the preceding section. In particular, these projections (see Fig. 4) show a random variation of the intensity in the $`Q=P=0`$ region, which according to eq.(27) is relevant for the final states of the optical transitions. This random variation is probed by optical transition matrix elements and results in an irregular pattern of lines in the fine structure of the high energy absorption band.
The absorption bands in Fig. 5 and Fig. 6 were calculated for equal optical transition matrix elements at the monomers $`\mu _1=\mu _2`$, which correspond to spin projections of the eigenstates with equal components. The changes occuring in the lower and upper absorption bands in the limiting cases of an optical asymmetry with $`\mu _10,\mu _2=0`$ and $`\mu _1=0,\mu _20`$ for the case of the parameter set B are compared in the Figs. 7 (a) and (b), respectively. These limiting cases correspond to an optical asymmetry, when one of the monomers constituting the dimer is optically active only. Then according to eq.(22) the optical transition matrix element is determined by either the spin up $`c_0^{(\lambda )}`$ or the spin down $`c_0^{(\lambda )}`$ coefficients of the $`\lambda `$ eigenstate. As is seen from Fig. 7, which is representative for the case of a positive sign of the excitation energy asymmetry $`ϵ_{}`$ ($`ϵ_{}=10>0`$ in the case shown in Fig. 7), the lower and upper bands show a redistribution of their intensities with the optical asymmetry: For the given sign of the excitation energy asymmetry $`ϵ_{}>0`$ ($`ϵ_{}<0`$) the intensity of the upper band is increased (decreased) compared to the symmetric case, $`\mu _1=\mu _2`$, displayed in Fig. 6. The lower band behaves in the opposite way, the peak intensities of both bands differing by a factor of about $`10^3`$. In particular, maximum absorption is reached for the high energy band, when the monomer with the higher excitation energy is optically active only, i.e. $`ϵ_{}>0`$ with $`\mu _10`$ and $`\mu _2=0`$. Reversing the optical activity of the monomers but still considering the same asymmetry in the excitation energies, i.e. $`ϵ_{}>0`$ with $`\mu _1=0`$ and $`\mu _20`$, one finds maximum absorption for the low energy band (if $`ϵ_{}<0`$, the quantities $`\mu _1`$ and $`\mu _2`$ have to be interchanged in the above consideration). As in the case of equal transition matrix elements the lower bands shown on the left sides of Fig. 7(a) and Fig. 7(b) display a regular fine structure, whereas the fine structure of the upper bands, displayed on the right sides of Fig. 7(a) and Fig. 7(b) is irregular. We note that the appearance of the weak bands is due to the coupling of the excited states of both monomers: The monomer, which is optically not active, is still present in the absorption spectrum due to the coupling of the excited states in the dimer system. In the case displayed on the right side of Fig. 7(b), e.g. an optically nonactive monomer with a high excitation energy attached to an optically active monomer with a lower excitation energy produces a weak but irregular band in the dimer system. A closer inspection of the fine structure of the irregular bands for the limiting cases of an optical asymmetry, in Fig. 7(a) and Fig. 7(b), together with the irregular band for the symmetric case, shown in Fig. 6, reveals that the sequence of lines with a strong and small absorption strength is almost identical in all bands, there is only a small overall change in the line intensities in all three representations. The reason behind this behavior becomes evident after an inspection of the ratio of the spin projections
$$r^{(\lambda )}=c_0^{(\lambda )}/c_0^{(\lambda )},$$
(28)
which is presented in Fig. 8 as a function of the eigenstate energy $`E_\lambda `$ for the spectral regions corresponding to the lower and upper absorption bands, respectively. As is evident from Fig. 8 $`r^{(\lambda )}`$ is a smooth function of the energy $`E_\lambda `$ in the spectral regions of both bands. In particular, for the case of the irregular upper absorption bands, for which in a sequence of eigenstates the spin projection coefficients $`c_0^{(\lambda )}`$ and $`c_0^{(\lambda )}`$ vary in an irregular way, their ratio $`r^{(\lambda )}`$ considered between neighbouring eigenstates is practically identical. This correlation between the spin up and spin down ($`c_0^{(\lambda )}`$ and $`c_0^{(\lambda )}`$) coefficients results in an almost smooth interpolation of the fine structure of the absorption bands between the limiting cases displayed in the Fig. 7. The smoothness of the function $`r^{(\lambda )}`$ can be used to express the absorption strength for arbitrary optical asymmetry by one of the limiting cases. For example, representing the down coefficients $`c_0^{(\lambda )}`$ by the up coefficients $`c_0^{(\lambda )}`$ in (22) using eq. (28), one obtains
$$Q_{\lambda g}=\frac{(\mu _1+\mu _2r^{(\lambda )})^2}{\mu _1^2+\mu _2^2}Q_{\lambda g}^+,$$
(29)
where $`Q_{\lambda g}^+=[c_0^{(\lambda )}]^2`$ is the absorption strength for the case $`\mu _10`$ and $`\mu _2=0`$. The smoothness of the ratio of the different spin projections $`r^{(\lambda )}`$ together with eq. (29) allows an interpolation between absorption bands for different optical asymmetries. In particular, using this ratio in the prefactor in front of $`Q_{\lambda g}^+`$ the relative intensities of absorption bands corresponding to different optical asymmetries can be determined.
## V Conclusions
The transition from the coexistence to a mixing of the adiabatic branches in the eigenstates of a symmetry-broken spin-boson Hamiltonian can be controlled by Bloch and Husimi projections and it shows up in the structure of absorption bands in excitonic-vibronic coupled dimer systems with asymmetric adiabatic potential configurations. In the mixing region of the adiabatic branches phase space distributions of the eigenstates are concentrated on the phase space orbits of both adiabatic potentials and their intensity varies in a random and irregular way. This random variation in the intensitiy distribution of the eigenstates is probed by the optical transition matrix elements and it is a source of spectral randomness in the fine structure of the absorption bands. The asymmetry of the optical transition matrix elements of the dimer configuration is representable by suitably chosen spin projections of the spin part of the eigenstates. In the limiting cases of the optical asymmetry, in which one of the monomers constituting the dimer is optically active only, the intensities of the absorption lines are determined by either the spin up or spin down coefficients. For the states relevant for the optical absorption correlated spin down and spin up coefficients are obtained. As a result the shape of the fine structure of the irregular absorption bands remains almost the same for different optical asymmetries. Finally we point out that in the case of a symmetry-broken spin-boson Hamiltonian optical absorption processes can probe the high energy regions of the overlap of adiabatic potentials where it is possible to observe a transition from regular to irregular absorption spectra for asymmetric molecular dimer configurations. |
no-problem/9911/hep-th9911096.html | ar5iv | text | # Untitled Document
PUPT-1897 ITEP-TH-61/99 hep-th/9911096
Gravity Duals of Fractional Branes and Logarithmic RG Flow
Igor R. Klebanov<sup>1</sup> and Nikita A. Nekrasov<sup>1,2</sup>
<sup>1</sup> Joseph Henry Laboratories, Princeton University, Princeton, New Jersey 08544
<sup>2</sup> Institute for Theoretical and Experimental Physics, 117259 Moscow, Russia
Abstract
We study fractional branes in $`𝒩=2`$ orbifold and $`𝒩=1`$ conifold theories. Placing a large number $`N`$ of regular D3-branes at the singularity produces the dual $`\mathrm{𝐀𝐝𝐒}_5\times X^5`$ geometry, and we describe the fractional branes as small perturbations to this background. For the orbifolds, $`X^5=𝐒^5/\mathrm{\Gamma }`$ and fractional D3-branes excite complex scalars from the twisted sector which are localized on the fixed circle of $`X^5`$. The resulting solutions are given by holomorphic functions and the field-theoretic beta-function is simply reproduced. For $`N`$ regular and $`M`$ fractional D3-branes at the conifold singularity we find a non-conformal $`𝒩=1`$ supersymmetric $`SU(N+M)\times SU(N)`$ gauge theory. The dual Type $`\mathrm{II}`$B background is $`\mathrm{𝐀𝐝𝐒}_5\times 𝐓^{1,1}`$ with NS-NS and R-R 2-form fields turned on. This dual description reproduces the logarithmic flow of couplings found in the field theory.
11/99
1. Introduction
By now there exists an impressive body of evidence that Type $`\mathrm{II}`$B strings on $`\mathrm{𝐀𝐝𝐒}_5\times X^5`$ are dual to large $`N`$ strongly coupled 4-d conformal gauge theories, in the sense proposed in \[1,,2,,3\]. Here $`X^5`$ are positively curved 5-d Einstein spaces whose simplest example $`𝐒^5`$ corresponds to the $`𝒩=4`$ supersymmetric $`SU(N)`$ gauge theory. One may also quotient this duality by a discrete subgroup $`\mathrm{\Gamma }`$ of the $`SU(4)`$ R-symmetry \[4,,5\]. The resulting backgrounds with $`X^5=𝐒^5/\mathrm{\Gamma }`$ are dual to “quiver gauge theories” with gauge group $`S(U(N)^n)`$ and bifundamental matter , which describe D3-branes near orbifold singularities. In such orbifold theories, in addition to regular D-branes which can reside on or off the orbifold fixed plane there are also “fractional” D-branes pinned to the fixed plane \[7,,8\]. Our goal in this paper is to consider the effect of such fractional branes on the dual supergravity background. There is good motivation for studying this problem because, as we discuss below, introduction of fractional branes breaks the conformal invariance and introduces RG flow.
It has also been possible to construct dual gauge theories for $`X^5`$ which are not locally $`𝐒^5`$. The simplest example is $`X^5=𝐓^{1,1}=(SU(2)\times SU(2))/U(1)`$ which turns out to be dual to an $`𝒩=1`$ superconformal $`SU(N)\times SU(N)`$ gauge theory with a quartic superpotential for bifundamental fields \[9,,10\]. In this theory, which arises on D3-branes at the conifold singularity, it is also possible to introduce fractional D-branes \[11,,12\], and we study their effects in this paper.
Having constructed the gravity duals of fixed-point theories, the next logical step is to study the dual picture of the RG flow. A natural setup for this problem is provided by the supersymmetric flows connecting orbifold theories and the (generalized) conifold theories. It is clear, though, that in order to have a consistent picture one cannot restrict oneself to the $`\mathrm{\Gamma }`$ invariant supergravity fields only, and needs to add the massless fields coming from the twisted sectors. This program was initiated in \[9,,13\], and in this paper we focus on the role of the twisted sector fields in creating RG flows. Considerable progress on supersymmetric RG flows in other situations has also been made recently \[14\] (for a review see ). For important work done in studies of non-supersymmetric RG flows see \[16,,17\].
The basic picture common to all RG problems is that the radial coordinate $`r`$ of $`\mathrm{𝐀𝐝𝐒}_5`$ defines the RG scale of the field theory, hence the scale dependence of couplings may be read off from the radial dependence of corresponding supergravity fields.
Since the RG flows of couplings in physically relevant gauge theories are logarithmic, an important problem is to find gravity duals of logarithmic flows. Attempts in this direction have been made in the context of Type 0B string theory \[18,,19,,20\]. This is an NSR string with the non-chiral GSO projection $`(1)^{F+\stackrel{~}{F}}=1`$ which breaks all spacetime supersymmetry (it is also a $`(1)^{F_s}`$ orbifold of the Type $`\mathrm{II}`$B theory). Type 0B theory has two basic types of D3-branes, electric and magnetic, and we will see that it is appropriate to call them the fractional branes. If equal numbers of the electric and magnetic branes are stacked parallel to each other, then we find on their world volume a $`U(N)\times U(N)`$ gauge theory coupled to six adjoint scalars of the first $`U(N)`$, six adjoint scalars of the second $`U(N)`$, and Weyl fermions in bifundamental representations. This theory is a “regular” $`\text{ZZ}_2`$ orbifold of the $`𝒩=4`$ $`U(2N)`$ gauge theory and hence is conformal in the planar limit \[22,,23\]. <sup>1</sup> The $`\text{ZZ}_2`$ is generated by $`(1)^{F_s}`$, where $`F_s`$ is the fermion number, together with conjugation by $`\gamma =\left(\begin{array}{cc}I& 0\\ 0& I\end{array}\right)`$ where $`I`$ is the $`N\times N`$ identity matrix. This orbifold is called regular because $`\gamma `$ is traceless.
In general, in orbifold theories there are as many types of fractional branes as there are nodes of the quiver diagram (i.e. the number of gauge groups in the product), or the number of the irreducible representations of the orbifold group $`\mathrm{\Gamma }`$. If one takes a collection of the fractional branes of each type and the branes corresponding to the irreducible representation $`_i`$ are taken $`n_i=\mathrm{dim}_i`$ times then one gets a single regular D-brane which can depart to the bulk. More specifically, the charge of the fractional brane of $`_i`$ type is
$$\mathrm{q}_i=\frac{n_i}{|\mathrm{\Gamma }|}$$
where $`|\mathrm{\Gamma }|`$ is the order of the orbifold group. It is well known that $`_in_i\mathrm{q}_i=1`$.
The fractional branes act as sources for the twisted closed string states of the orbifold theory. In the Type 0B example discussed above the two $`U(N)`$ groups correspond to the electric and the magnetic D3-branes, hence these are the two types of fractional branes for this particular $`\text{ZZ}_2`$ orbifold. Indeed, such branes have tadpoles for the twisted RR 4-form and for the tachyon \[18,,24\]. If we stack $`N`$ parallel electric branes only, then we find an “irregular” orbifold theory where the $`\text{ZZ}_2`$ action does not act on the gauge indices. This gives $`SU(N)`$ gauge theory coupled to six adjoint scalars, which is not a CFT .
Equations satisfied by the gravity dual of this theory were derived in , and solved after some assumptions in \[19,,20\]. The RG flow of the dilaton, which is related to the gauge coupling, comes from the equation
$$^2\varphi =\frac{1}{4\alpha ^{}}T^2e^{\varphi /2}.$$
Since the tachyon field $`T`$ has a source $`F_5^2`$, it departs from zero and causes the dilaton to depend on $`r`$. Assuming that $`T`$ approaches a constant for large $`r`$ it was found that the RG flow is logarithmic in the UV \[19,,20\]. While this scenario has a number of uncertainties (due to the lack of detailed knowledge of the $`T`$-dependence in the effective action) it suggests a mechanism for RG flow of couplings in the dual gravity picture of fractional branes. In particular, the presence of “twisted” fields sourced by the fractional branes plays the crucial role.
The stack of electric D3-branes defines gauge theory coupled to six scalars fields, but it is obviously of more interest to remove the scalars and study the pure glue theory . One way of embedding it into string theory is to consider a $`\text{ZZ}_2`$ orbifold of Type 0B by reflection of six coordinates . The regular orbifold theory has gauge group $`U(N)^4`$ coupled to a chiral field content:
$``$four quadruples of bi-fundamental fermions transforming in $`(𝐍_𝐢,\overline{𝐍}_{𝐢+1})`$, $`𝐢=0,1,2,3;40`$,
$``$four sextets of bi-fundamental scalars in $`(𝐍_𝐢,\overline{𝐍}_{𝐢+2})`$.
Now there are 4 gauge groups in the product, hence there should be 4 different types of fractional branes. If we stack $`N`$ fractional D3-branes of the same type then we find pure glue $`U(N)`$ gauge theory on their world volume . This suggests that fractional branes may provide a link to string duals of realistic gauge theories.<sup>2</sup> It seems unclear, however, whether the fractional D3-branes in the $`\text{ZZ}_4`$ theory which are stuck to the fixed fourplane exist, for their flux has nowhere to escape to.
With this eventual goal in mind, in this paper we study gravity duals of fractional branes in supersymmetric conifold and orbifold theories: the SUSY removes some of the effective action uncertainties present in the Type 0B case. To simplify matters further we consider theories where $`\beta `$–functions for ‘t Hooft couplings $`g_{\mathrm{YM}}^2N`$ are of order $`1/N`$ rather than of order $`1`$. Such gauge theories occur on $`M`$ fractional D3-branes parallel to $`N`$ regular D3-branes, with $`M`$ held fixed in the large $`N`$ limit. In the simplest examples we consider, the gauge group is then $`SU(N+M)\times SU(N)`$.
The large number $`N`$ of regular D3-branes produces the dual $`\mathrm{𝐀𝐝𝐒}_5\times X^5`$ background, and for our purposes we may ignore the back-reaction of the $`M`$ fractional D3-branes on it. However, the fractional branes act as sources for an extra set of fields, namely the 2-form potentials $`B^{NSNS}`$ and $`B^{RR}`$. The flux of these 2-forms through a certain 2-cycle of $`X^5`$ (more precisely its deviation from the value at the orbifold point) defines the difference between $`g_{\mathrm{YM}}^2`$ for the gauge groups factors. In the $`𝒩=2`$ supersymmetric orbifold cases the 2-cycles are collapsed, so that the twisted sector fields corresponding to the 2-form fluxes are confined to $`\mathrm{𝐀𝐝𝐒}_5\times 𝐒^1`$ where $`𝐒^1`$ is the fixed circle of $`𝐒^5/\mathrm{\Gamma }`$. By studying the dependence of these twisted sector fields on the $`\mathrm{𝐀𝐝𝐒}_5`$ radial coordinate $`r`$ we find the logarithmic flow of the gauge couplings consistent with field theory expectations. In fact, the twisted sector fields are given by holomorphic functions of the complex variable $`z=re^{i\phi }`$ where $`\phi `$ is the $`𝐒^1`$ coordinate (previously known solutions of $`𝒩=2`$ gauge theories are also characterized by holomorphic functions \[27,,28,,29\]).<sup>3</sup> Supergravity duals of logarithmic RG flows in $`𝒩=2`$ gauge theories are being independently studied in . Remarkably, in the limit we are studying all string-scale corrections are small, so that the use of effective supergravity equations is justified.
2. Fractional Branes on the Conifold.
Let us start by reviewing what is known about regular D3-branes at conical singularities. If a large number $`N`$ of D3-branes is placed at the apex of a 6-d cone $`Y^6`$ with metric
$$ds_{\mathrm{cone}}^2=dr^2+r^2ds_5^2,$$
then the near-horizon region of the resulting 10-d geometry has the metric
$$ds^2=R^2\left[\frac{r^2}{R^4}(dt^2+dx_1^2+dx_2^2+dx_3^2)+\frac{dr^2}{r^2}+ds_5^2\right],R^4g_sN(\alpha ^{})^2.$$
This geometry is $`\mathrm{𝐀𝐝𝐒}_5\times X^5`$ where $`X^5`$ is the base of the cone (if $`Y_6`$ is Ricci-flat then $`X^5`$ is a positively curved Einstein space \[31,,9\]. Type $`\mathrm{II}`$B theory on this background is conjectured to be dual to the conformal limit of the gauge theory on $`N`$ D3-branes placed at the apex \[9,,10\].
An example where such a duality has been tested extensively is when $`Y^6`$ is the conifold, which is a singular Calabi-Yau manifold described in terms of complex variables $`w_1,\mathrm{},w_4`$ by the equation
$$\underset{a=1}{\overset{4}{}}w_a^2=0.$$
The base of this cone is $`𝐓^{1,1}=(SU(2)\times SU(2))/U(1)`$ whose Einstein metric may be written down explicitly ,
$$ds_{X_5}^2=\frac{1}{9}\left(d\psi +\mathrm{cos}\theta _1d\varphi _1+\mathrm{cos}\theta _2d\varphi _2\right)^2+\frac{1}{6}\underset{a=1}{\overset{2}{}}\left(d\theta _a^2+\mathrm{sin}^2\theta _ad\varphi _a^2\right).$$
The $`𝒩=1`$ superconformal field theory on $`N`$ regular D3-branes placed at the singularity of the conifold has gauge group $`SU(N)\times SU(N)`$ and global symmetry $`SU(2)\times SU(2)\times U(1)`$ . The chiral superfields $`A_1`$, $`A_2`$ transform as $`(𝐍,\overline{𝐍})`$ and are a doublet of the first $`SU(2)`$; the chiral superfields $`B_1`$, $`B_2`$ transform as $`(\overline{𝐍},𝐍)`$ and are a doublet of the second $`SU(2)`$. The R-charge of all four chiral superfields is $`1/2`$ and the theory has an exactly marginal superpotential $`W=ϵ^{ij}ϵ^{kl}\mathrm{Tr}A_iB_kA_jB_l`$.
$`\mathrm{II}`$B supergravity modes on $`\mathrm{𝐀𝐝𝐒}_5\times 𝐓^{1,1}`$ have been matched in some detail with operators in this gauge theory whose dimensions are of order $`1`$ in the large $`N`$ limit \[33,,34\]. In addition, string theory has heavy supersymmetric states obtained by wrapping D3-branes over 3-cycles of $`𝐓^{1,1}`$. They have been shown to correspond to “dibaryon” operators whose dimensions grow as $`3N/4`$ (schematically, these operators have the form $`\mathrm{Det}A`$ or $`\mathrm{Det}B`$). Further, one may consider a domain wall in $`\mathrm{𝐀𝐝𝐒}_5`$ obtained by wrapping a D5-brane over the 2-cycle of $`𝐓^{1,1}`$ (topologically, $`𝐓^{1,1}`$ is $`𝐒^2\times 𝐒^3`$). If this domain wall is located at $`r=r_{}`$ then, by studying the behavior of wrapped D3-branes upon crossing it, it was shown in that for $`r>r_{}`$ the gauge group changes to $`SU(N+1)\times SU(N)`$. Note that this is precisely the gauge theory expected on $`N`$ regular and one fractional D3-branes! Thus, a D5-brane wrapped over the 2-cycle is nothing but a fractional D3-brane placed at a definite $`r`$.<sup>4</sup> Similarly, a regular D3-brane serves as a domain wall between $`SU(N)\times SU(N)`$ and $`SU(N+1)\times SU(N+1)`$ gauge theory. This has a simple interpretation in terms of Higgsing the theory. The identification of a fractional D3-brane with a wrapped D5-brane is consistent with the results of \[35,,8,,36,,12\].
As shown in , this suggests a construction of the Type $`\mathrm{II}`$B dual for the $`𝒩=1`$ $`SU(N+M)\times SU(N)`$ gauge theory. In particular, the background has to contain $`M`$ units of R-R 3-form flux through the 3-cycle of $`𝐓^{1,1}`$:
$$_{C^3}H^{RR}=M.$$
If $`M`$ is fixed as $`N\mathrm{}`$ then the back-reaction of $`H^{RR}`$ on the metric and the $`F_5`$ background may be ignored to leading order in $`N`$. However, as we will show, the background must also include the NS-NS 2-form potential
$$B^{NSNS}=e^\varphi f(r)\omega _2,$$
where $`\omega _2`$ is the closed 2-form corresponding to the 2-cycle which is dual to $`C^3`$,
$$_{C^2}\omega _2=1.$$
The desired connection with the RG flow is due to the fact that \[5,,9,,11,,10\]
$$\frac{1}{g_1^2}\frac{1}{g_2^2}e^\varphi \left(_{C^2}B^{NSNS}\frac{1}{2}\right)$$
where $`g_1`$ and $`g_2`$ are the gauge couplings for $`SU(N+M)`$ and $`SU(N)`$ respectively. Therefore, the $`f(r)`$ in (2.1) gives the dual supergravity definition of the scale dependence of $`\frac{1}{g_1^2}\frac{1}{g_2^2}`$.
Before solving for $`f(r)`$ let us recall the $`\beta `$-function calculation in field theory. There we have
$$\begin{array}{cc}\hfill \frac{d}{d\mathrm{log}(\mathrm{\Lambda }/\mu )}\frac{1}{g_1^2}& 3(N+M)2N(1\gamma _A\gamma _B),\hfill \\ \hfill \frac{d}{d\mathrm{log}(\mathrm{\Lambda }/\mu )}\frac{1}{g_2^2}& 3N2(N+M)(1\gamma _A\gamma _B),\hfill \end{array}$$
where $`\gamma `$ are the anomalous dimensions of the fields $`A_i`$ and $`B_j`$. For $`M=0`$ we find a fixed point with $`\gamma _A=\gamma _B=1/4`$ which corresponds to R-charge $`1/2`$. This is the superconformal gauge theory dual to $`\mathrm{𝐀𝐝𝐒}_5\times 𝐓^{1,1}`$ with vanishing 2-form potentials. For $`M0`$ it is impossible to make both beta functions vanish (even if we allow the anomalous dimensions of $`A`$ and $`B`$ to be different) and the theory undergoes logarithmic RG flow:
$$\frac{1}{g_1^2}\frac{1}{g_2^2}M\mathrm{log}(\mathrm{\Lambda }/\mu )[3+2(1\gamma _A\gamma _B)].$$
Near the fixed point we expect $`\gamma _A+\gamma _B=1/2`$ plus small corrections, hence the RHS gives $`M\mathrm{log}(\mathrm{\Lambda }/\mu )`$.
Let us reproduce this result in supergravity. We need the Type $`\mathrm{II}`$B SUGRA equations of motion involving the 2-form gauge potentials. We will write these equations in the $`\mathrm{𝐀𝐝𝐒}_5\times 𝐓^{1,1}`$ background with constant $`\tau =C_0+ie^\varphi `$ (this is the $`\mathrm{SL}_2(\text{ZZ})`$ covariant combination of the dilaton and the R-R scalar of the Type $`\mathrm{II}`$B theory):<sup>5</sup> We will check later that this background has no corrections of order $`M/N`$ which is the order at which we are working.
$$d^{}G=iF_5G.$$
$`G`$ is the complex $`3`$-form field strength,
$$G=H^{RR}+\tau H^{NSNS},$$
which satisfies the Bianchi identity $`dG=0`$. Note that the RHS of (2.1) originates from the Chern-Simons term
$$C_4H^{RR}H^{NSNS}.$$
Since the fractional D3-brane (the wrapped D5-brane) creates R-R 3-form flux through $`𝐓^{1,1}`$, $`H^{RR}`$ should be proportional to the closed 3-form which was constructed in ,
$$H^{RR}Me^\psi (e^{\theta _1}e^{\varphi _1}e^{\theta _2}e^{\varphi _2}).$$
Here we are using the basis 1-forms
$$e^\psi =\frac{1}{3}(d\psi +\underset{i=1}{\overset{2}{}}\mathrm{cos}\theta _id\varphi _i),e^{\theta _i}=\frac{1}{\sqrt{6}}d\theta _i,e^{\varphi _i}=\frac{1}{\sqrt{6}}\mathrm{sin}\theta _id\varphi _i.$$
In these coordinates, the closed form $`\omega _2`$ which enters (2.1) is given by
$$\omega _2e^{\theta _1}e^{\varphi _1}e^{\theta _2}e^{\varphi _2},$$
so that
$$e^\varphi H^{NSNS}df(r)(e^{\theta _1}e^{\varphi _1}e^{\theta _2}e^{\varphi _2}).$$
Since $`F_5=\mathrm{vol}(\mathrm{𝐀𝐝𝐒}_5)+\mathrm{vol}(𝐓^{1,1})`$, $`F_5H^{NSNS}=0`$. Let us set the R-R scalar $`C_0=0`$. Then we see that the real part of (2.1) is satisfied for all $`f(r)`$. From the imaginary part we have
$$\frac{1}{r^3}\frac{d}{dr}\left(r^5\frac{d}{dr}f\left(r\right)\right)M,$$
which implies
$$f(r)M\mathrm{log}r.$$
Quite remarkably, our solution of Type $`\mathrm{II}`$B SUGRA equations has reproduced the field theoretic beta function for $`\frac{1}{g_1^2N}\frac{1}{g_2^2N}`$ to order $`M/N`$. This establishes the gravity dual of the logarithmic RG flow in the $`𝒩=1`$ supersymmetric $`SU(N+M)\times SU(N)`$ gauge theory on $`N`$ regular and $`M`$ fractional D3-branes placed at the conifold singularity.
It is not hard to see that there are no other effects of order $`M/N`$: the back-reaction of the 3-form field strengths on the metric, dilaton and the $`F_5`$ comes in at order $`(M/N)^2`$. Consider, for instance, the dilaton equation of motion:
$$^2\varphi =\frac{1}{12}(e^\varphi H_{RR}^2e^\varphi H_{NSNS}^2).$$
Even without checking the relative normalizations of $`H_{RR}^2`$ and $`H_{NSNS}^2`$, we can immediately see that the variation of $`e^\varphi /N`$ is at most of order $`(M/N)^2`$. In the field theory this quantity translates into $`\frac{1}{g_1^2N}+\frac{1}{g_2^2N}`$. The fact that there is no $`\beta `$-function of order $`M/N`$ for this quantity agrees with the field theory RG equations provided that the sum of the anomalous dimensions, $`\gamma _A+\gamma _B`$, has no corrections of order $`M/N`$. This is a simple gravitational prediction about the gauge theory. It is of further interest to study the order $`(M/N)^2`$ effects on the background and compare them with field theory, but we postpone these calculations for future work.
3. Fractional Branes in $`𝒩=2`$ Orbifold Theories.
In this section we will be concerned with the $`𝒩=2`$ supersymmetric orbifolds of Type $`\mathrm{II}`$B strings. Before introducing D-branes these are backgrounds of the form $`\mathrm{IR}^{5,1}\times \mathrm{IR}^4/\mathrm{\Gamma }`$. To write down gravity duals of fractional branes we follow the strategy used in the last section: we first stack a large number $`N`$ of regular D3-branes creating the dual $`\mathrm{𝐀𝐝𝐒}_5\times 𝐒^5/\mathrm{\Gamma }`$ background and then introduce the fractional branes as small perturbations on this background. We find an important difference from the conifold case, however, because $`𝐒^5/\mathrm{\Gamma }`$ does not have any finite volume 2-cycles. A blowup of this space produces 2-cycles but breaks the $`𝒩=2`$ supersymmetry. For this reason we work with the singular space where the 2-cycles are collapsed (we will see that this turns out to be simpler than the non-singular case discussed in the previous section).
Thus we consider the Type $`\mathrm{II}`$B superstring in the background $`\mathrm{𝐀𝐝𝐒}_5\times 𝐒^5/\mathrm{\Gamma }`$ where $`\mathrm{\Gamma }`$ is one of the A,D,E subgroups of $`SU(2)`$ whose action on $`𝐒^5`$ is induced from the standard action on $`𝐑^4\text{ }\mathrm{C}^2`$ times the trivial action on $`\mathrm{IR}^2`$. Therefore, we get the action of $`\mathrm{\Gamma }`$ on $`\mathrm{IR}^6`$ having a fixed two-plane $`0\times \mathrm{IR}^2`$. This action descends to $`𝐒^5`$ and the fixed plane becomes a fixed circle $`𝐒^1`$.
The metric on $`𝐒^5/\mathrm{\Gamma }`$ reads:
$$ds_5^2=d\theta ^2+\mathrm{cos}^2\theta d\phi ^2+\mathrm{sin}^2\theta ds_{𝐒^3/\mathrm{\Gamma }}^2,$$
and the fixed circle is at $`\theta =0`$.
String theory in this background has extra massless fields compared to the $`\mathrm{\Gamma }`$-invariant fields of the ten dimensional $`\mathrm{II}`$B supergravity. These fields are localized at the fixed circle $`𝐒^1`$. In the paper the multiplets of the five dimensional gauge supergravity these fields fall in were identified.
The simple meaning of these fields is the following: if one were to blow up the fixed circle to obtain a smooth Einstein metric then the topology of the resulting fivefold is such that the non-contractable two-spheres $`C_i`$ are supported. These spheres intersect each other according to the Dynkin diagram of the corresponding A,D,E Lie group:
$$C_iC_j=a_{ij}$$
Now, reducing the ten dimensional supergravity fields along these cycles leads to new fields in six dimensions spanned by $`𝐒^1\times \mathrm{𝐀𝐝𝐒}_5`$. Of particular importance for us are the NSNS and RR two-forms $`B_{NSNS}`$ and $`B_{RR}`$. They give rise to the scalars:
$$\beta _i^{NSNS}=_{C_i}B_{NSNS},\beta _i^{RR}=_{C_i}B_{RR}$$
These fields are present even if one did not perform the blowup: they come from the twisted sector of the string theory.
We now proceed with writing the effective action for these fields (here we have set the R-R scalar to zero):
$$_{𝐒^1\times \mathrm{𝐀𝐝𝐒}_5}\sqrt{g}g^{mn}a^{ij}\left(e^\varphi _m\beta _i^{NSNS}_n\beta _j^{NSNS}+e^\varphi _m\beta _i^{RR}_n\beta _j^{RR}\right)+a^{ij}F_5\beta _i^{NSNS}d\beta _j^{RR}$$
The last term comes from the 10-d Chern-Simons coupling of the form
$$F_5B_{NSNS}dB_{RR}.$$
Introduce the complex fields:
$$\gamma _i=\tau \beta _i^{NSNS}+\beta _i^{RR}.$$
These fields transform nicely under the $`\mathrm{SL}_2(\text{ZZ})`$ group:
$$\left(\begin{array}{cc}a& b\\ c& d\end{array}\right)\left(\begin{array}{c}\gamma _i\\ \tau \end{array}\right)=\frac{1}{c\tau +d}\left(\begin{array}{c}\gamma _i\\ a\tau +b\end{array}\right)$$
The metric on the $`𝐒^1\times \mathrm{𝐀𝐝𝐒}_5`$ space is:
$$ds^2=R^2\left(\frac{dr^2}{r^2}+d\phi ^2\right)+\frac{r^2}{R^2}(dt^2+dx_1^2+dx_2^2+dx_3^2),R^4=4\pi g_sN\mathrm{}_s^4,$$
where $`\phi `$ is the coordinate on the circle $`𝐒^1`$.
Let us introduce the complex coordinate on the space transverse to $`\mathrm{IR}^{1,3}`$: $`z=re^{i\phi }`$. We are interested in fields which have no $`\mathrm{IR}^{1,3}`$ dependence. For these fields the action (3.1) takes on the following form (assuming that $`\tau `$ is constant, otherwise we get covariant derivatives instead of ordinary ones):
$$S=r^5a^{ij}\frac{1}{\tau _2}\frac{\gamma _i}{\overline{z}}\frac{\overline{\gamma }_j}{z}𝑑r𝑑\phi $$
where $`\tau _2=e^\varphi =\mathrm{Im}\tau `$. This action is $`\mathrm{SL}_2(\text{ZZ})`$ invariant.
Gauge couplings. The theory on the boundary of the $`\mathrm{𝐀𝐝𝐒}_5`$ space is the superconformal quiver gauge theory with the gauge group $`SU(Nn_0)\times SU(Nn_1)\times \mathrm{}\times SU(Nn_r)`$ where $`n_i`$ are the Dynkin indices – the dimensions of the irreps of $`\mathrm{\Gamma }`$. $`n_0=1`$ is the dimension of the trivial representation $`_0`$.
The relation between the boundary values of $`\gamma _i`$ and the couplings of these gauge factors was shown in to be:
$$\begin{array}{cc}& \tau _i=\gamma _i,i=1,\mathrm{},r\hfill \\ & \tau _0=|\mathrm{\Gamma }|\tau \underset{i}{}n_i\tau _i\hfill \end{array}$$
where
$$\tau _j=\frac{\theta _j}{2\pi }+\frac{4\pi i}{g_j^2}$$
Running of the dilaton. Now let us discuss the validity of our assumption that $`\tau `$ is constant. The Lagrangian for the $`\stackrel{}{x}`$ independent dilaton reads as follows:
$$=r^3t^3𝑑r𝑑t𝑑\phi \frac{1}{\tau _2}\left(r^2_r\tau _r\overline{\tau }+(1t^2)_t\tau _t\overline{\tau }+\frac{1}{1t^2}_\phi \tau _\phi \overline{\tau }\right)$$
where we introduced the notation $`t=\mathrm{sin}\theta `$. The fixed circle is at $`t=0`$. The fields $`\gamma `$ acts as sources for the $`\tau `$ equations of motion. Irrespectively of the precise form of (3.1) the source term in the dilaton equation of motion is:
$$\delta (t)r^5e^\varphi a^{ij}(\gamma _i\overline{}\gamma _j+\overline{\gamma }_i\overline{}\overline{\gamma }_j),$$
(this expression is valid for $`C_0=0`$, $`\varphi =\mathrm{const}`$). Thus, the source term vanishes for holomorphic $`\gamma `$ and the dilaton is allowed to remain constant. This argument may suffer from some subtleties in case of badly singular $`\gamma _i`$.
Fractional branes. In string theory $`B_{NSNS},B_{RR}`$ do not have to be globally well-defined two-forms - they behave like gauge fields. The same applies to the scalars $`\beta _i`$ obtained by the reduction of the $`B`$-fields. In particular, if we add the $`k`$-th fractional threebrane (D5-brane wrapped over $`C_k`$) at some point $`z_{}`$ then the scalars $`\beta _j^{RR}`$ pick up a shift when circled around its location:
$$\beta _j^{RR}\beta _j^{RR}+a_{kj}$$
The brane being BPS does not spoil the holomorphicity of the $`\gamma `$’s, hence we conclude that the solution must have a logarithmic monodromy:
$$\gamma _j\frac{a^{kj}}{2\pi i}\mathrm{log}(zz_{})+\mathrm{}$$
One may be concerned about the appearance of the logarithm because $`\gamma _j`$ is defined on a torus with modular parameter $`\tau `$. Luckily, $`\tau _2=\frac{1}{g_s}`$ is of order $`N`$ in the ‘t Hooft limit and the periodicity in this direction may be ignored for large $`N`$ (this is because the $`\beta `$-functions for $`\gamma _j`$ are of order 1 for a finite number of fractional branes, so that the evolution of the couplings is relatively slow). In any event, the theory with a fractional brane may be regulated far in the UV by adding a fractional anti-brane at large $`r`$:
$$\gamma _j\frac{a^{kj}}{2\pi i}[\mathrm{log}(zz_{})\mathrm{log}(zz_{\mathrm{reg}})]+\mathrm{}$$
This way the theory becomes conformal again for $`r|z_{\mathrm{reg}}|`$. On the other hand, the singularities of $`\gamma _i`$ at the locations of fractional branes are presumably removed by the instanton corrections which will be discussed below.
Absence of the dilaton running from the field theory expectations. The fractional brane of the $`_i`$ type affects the gauge theory in a simple way: by changing the $`SU(Nn_i)`$ factor into $`SU(Nn_i+1)`$ and not changing the rest.
Clearly this induces a non-trivial beta functions for all couplings $`\tau _j`$, such that $`j=i`$ or $`a_{ij}0`$. In fact:
$$\frac{d}{d\mathrm{log}\left(\mathrm{\Lambda }/\mu \right)}\tau _j=2Nn_j\underset{ki}{}a_{jk}Nn_k+(Nn_i+1)a_{ij}=a_{ij}$$
$$\frac{d}{d\mathrm{log}\left(\mathrm{\Lambda }/\mu \right)}\tau _i=2(Nn_i+1)\underset{j}{}a_{ij}Nn_j=2$$
Clearly:
$$\frac{d}{d\mathrm{log}\left(\mathrm{\Lambda }/\mu \right)}\underset{k=0}{\overset{r}{}}n_k\tau _k=\left(\underset{j}{}a_{ij}n_j+2n_i\right)=0$$
On the other hand, from (3.1) we see that
$$\tau =\frac{1}{|\mathrm{\Gamma }|}\underset{k}{}n_k\tau _k$$
and we come to the complete agreement with the space-time picture of the dilaton being constant.
Instanton corrections. As the space-time dilaton is constant for the solutions we discuss we may hope that the solutions we find are valid, at least far away from the locations of the fractional branes, but not too far to make some of the gauge factors extremely strongly coupled. When we approach the branes, the $`\mathrm{log}(z)`$ behaviour $`\gamma `$ makes the fractional D-instantons favorable. The creation amplitude of the fractional D-instanton of the $`_j`$ type is clearly of the order of
$$\mathrm{exp}2\pi i\gamma _j\mathrm{}$$
near the fractional D3-brane. The fractional $`(p,q)`$ D-instantons are the euclidean $`(p,q)`$-string world sheets wrapped around the collapsed two cycles $`C_j`$. The field theory counterpart of the picture which we are advocating is the possibility of having instanton corrections coming from the instantons of each group factor. Also, the appearance of the $`(p,q)`$ types of the instantons is the consequence of the S-duality non-trivially realized in the field theories under consideration.
Acknowledgements
We are grateful to O. Bergman, C. Johnson, J. Minahan, A. Polyakov and S. Shatashvili for useful discussions. The work of I. R. K. was supported in part by the NSF grant PHY-9802484 and by the James S. McDonnell Foundation Grant No. 91-48; that of N. A. N by Robert H. Dicke Fellowship from Princeton University, partly by RFFI under grant 98-01-00327, partly by the grant 96-15-96455 for scientific schools.
References
relax J. Maldacena, “The Large N limit of superconformal field theories and supergravity,” Adv. Theor. Math. Phys.2 (1998) 231, hep-th/9711200 relax S.S. Gubser, I.R. Klebanov, and A.M. Polyakov, “Gauge theory correlators from noncritical string theory,” Phys. Lett. 428B (1998) 105, hep-th/9802109 relax E. Witten, “Anti-de Sitter space and holography,” Adv. Theor. Math. Phys.2 (1998) 253, hep-th/9802150 relax S. Kachru and E. Silverstein, “4d conformal field theories and strings on orbifolds,” Phys. Rev. Lett. 80 (1998) 4855, hep-th/9802183. relax A. Lawrence, N. Nekrasov and C. Vafa, “On conformal field theories in four dimensions,” Nucl. Phys. B533 (1998) 199, hep-th/9803015 relax M. Douglas and G. Moore, “D-branes, quivers, and ALE instantons,” hep-th/9603167 relax E.G. Gimon and J. Polchinski, “Consistency Conditions for Orientifolds and D Manifolds”, Phys. Rev. D54 (1996) 1667, hep-th/9601038 relax M.R. Douglas, “Enhanced Gauge Symmetry in M(atrix) theory”, JHEP 007(1997) 004, hep-th/9612126 relax I.R. Klebanov and E. Witten, “Superconformal Field Theory on Threebranes at a Calabi-Yau Singularity,” Nucl. Phys. B536 (1998) 199, hep-th/9807080. relax D.R. Morrison and M.R. Plesser, “Non-Spherical Horizons, I,” hep-th/9810201. relax S.S. Gubser and I.R. Klebanov, “Baryons and Domain Walls in an N=1 Superconformal Gauge Theory,” Phys. Rev. D58 (1998) 125025, hep-th/9808075. relax K. Dasgupta and S. Mukhi, “Brane Constructions, Fractional Branes and Anti-de Sitter Domain Walls,” hep-th/9904131. relax S. Gubser, N. Nekrasov, S. Shatashvili, “Generalized Conifolds and 4d N=1 SCFT,” hep-th/9811230. relax D.Z. Freedman, S.S. Gubser, K. Pilch, N.P. Warner, “Renormalization Group Flows from Holography–Supersymmetry and a c-Theorem,” hep-th/9904017. relax S.S. Gubser, “Non-conformal examples of AdS/CFT,” hep-th/9910117. relax L. Girardello, M. Petrini, M. Porrati, A. Zaffaroni, “Novel Local CFT and Exact Results on Perturbations of N=4 Super Yang Mills from AdS Dynamics,” hep-th/9810126. relax J. Distler and F. Zamora, “Nonsupersymmetric conformal field theories from stable anti-de Sitter spaces,” hep-th/9810206; “Chiral Symmetry Breaking in the AdS/CFT Correspondence,” hep-th/9911040. relax I. R. Klebanov and A. A. Tseytlin, “D-Branes and Dual Gauge Theories in Type 0 Strings,” Nucl. Phys. B546 (1999) 155, hep-th/9811035 relax J. Minahan, “Glueball Mass Spectra and Other Issues for Supergravity Duals of QCD Models,” hep-th/9811156 relax I.R. Klebanov and A.A. Tseytlin, “Asymptotic Freedom and Infrared Behavior in the Type 0 String Approach to Gauge Theory,” Nucl. Phys. B547 (1999) 143, hep-th/9812089 relax L. Dixon and J. Harvey, “String theories in ten dimensions without space-time supersymmetry”, Nucl. Phys. B274 (1986) 93 ; N. Seiberg and E. Witten, “Spin structures in string theory”, Nucl. Phys. B276 (1986) 272; C. Thorn, remarks at the workshop “Superstring Theories and the Mathematical Structure of Infinite-Dimensional Lie Algebras”, Santa Fe Institute, November 1985 relax I. R. Klebanov and A. A. Tseytlin, “Non-supersymmetric CFT from Type 0 String Theory,” JHEP 9903(1999) 015, hep-th/9901101 relax N. Nekrasov and S. Shatashvili, “On non-supersymmetric CFT in four dimensions,” hep-th/9902110., L. Okun Festschrift, North-Holland, in press relax O. Bergman and M. Gaberdiel, “A Non-supersymmetric Open String Theory and S-Duality,” Nucl. Phys. B499 (1997) 183, hep-th/9701137 relax A.M. Polyakov, “The Wall of the Cave,” hep-th/9809057 relax I.R. Klebanov, N. Nekrasov and S. Shatashvili, “An Orbifold of Type 0B Strings and Non-supersymmetric Gauge Theories,” hep-th/9909109 relax N. Seiberg and E. Witten, “Monopole Condensation and Confinement In N=2 Supersymmetric Yang-Mills Theory,” Nucl. Phys. B426 (1994) 19 relax E. Witten, “Solutions of Four-dimensional Field Theories via M-theory”, Nucl. Phys. B500 (1997) 3; hep-th/9703166 relax S. Katz, P. Mayr, C. Vafa, “Mirror symmetry and Exact Solution of 4D $`𝒩=2`$ gauge theories I”, hep-th/9706110, Adv.Theor.Math.Phys. 1 (1998) 53-114 relax C. Johnson, A. Peet and J. Polchinski, in preparation; reported by C. Johnson, talk at IAS. relax A. Kehagias, “New Type IIB Vacua and Their F-Theory Interpretation,” hep-th/9805131. relax P. Candelas and X. de la Ossa, “Comments on Conifolds,” Nucl. Phys. B342 (1990) 246. relax S.S. Gubser, “Einstein Manifolds and Conformal Field Theories,” Phys. Rev. D59 (1999) 025006, hep-th/9807164. relax A. Ceresole, G. Dall’Agata, R. D’Auria, and S. Ferrara, “Spectrum of Type IIB Supergravity on $`\mathrm{𝐀𝐝𝐒}_5\times 𝐓^{1,1}`$: Predictions On $`𝒩=1`$ SCFT’s,” hep-th/9905226. relax C.V. Johnson and R.C. Myers, “Aspects of Type $`\mathrm{II}`$B Theory on ALE Spaces”, Phys. Rev. D55 (1997) 6382, hep-th/9610140 relax D.-E. Diaconescu, M. Douglas and J. Gomis, “Fractional Branes and Wrapped Branes,” JHEP 02 (1998) 013. |
no-problem/9911/astro-ph9911050.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Several critical observational advances have taken place in the past few years to elucidate the formation and evolution of galaxies and the rates at which stars formed at different cosmological epochs. With the advent of magnitude limited samples of spectroscopic redshifts, it became possible to discuss general properties of galaxies through their comoving luminosity density or star formation rate (SFR) as a function of redshift (Lilly et al 1996, Madau et al 1996). Since then, we became familiar with the study of a new astronomical character, the ’$`1Mpc^3`$’ box of the universe. At the same time, the plot of the cosmic background was enriched by COBE (Puget et al 1996, Fixsen et al 1998, Lagache et al 1999, Hauser et al 1999, Dwek et al 1998) and the HST (Pozzetti et al 1998), on the far infrared to sub-millimeter part and on the optical-UV part respectively. These two global views of the past history of galaxy formation and evolution gave a new life to the debate that begun with the discovery by IRAS of Luminous InfraRed Galaxies (LIRGs, see the review by Sanders & Mirabel 1996). Galaxies brighter than $`10^{11}L_{}`$ radiate most of their light in the infrared (IR) range and exhibit the strongest star formation rates ever measured. No other observations, else than direct infrared observations, had prepared such a discovery. However, the infrared radiation of nearby galaxies ($`z<0.2`$) makes only $`30\%`$ of their optical light and only $`6\%`$ of it comes from LIRGs (based on IRAS results, Soifer & Neugebauer 1991), which consequently weight only $`2\%`$ by comparison with the optical light from ’normal’ galaxies. Prior to the COBE, SCUBA and ISO results, hence only three years ago, the matter of debate could have been summarized by the following question: should we consider LIRGs, as anything else than extreme cases, non representative of the general behavior of galaxies ?
When the previously quoted authors, after Puget et al (1996), found from the COBE FIRAS and afterwards DIRBE measurements, that the energetic content of the IR to sub-mm cosmic background was as high as that of the UV to near-IR background, the debate became more intense but still left opened the possibility to keep a scenario consistent with the one deduced from the ’$`1Mpc^3`$’ box of the universe story, i.e. that one could reproduce both the CIRB and the variation of the comoving UV luminosity density and SFR with redshift without the recourse of LIRGs (see Fall, Charlot & Pei 1996).
Several observations have now ruled out this picture. The redshift dependance of the comoving star formation rate of the universe needs not only the UV luminosity density to be known as a function of redshift but also the IR luminosity density, which is dominated by LIRGs at $`z1`$ and above (see below). The most directs of these observations, because they measure the IR light radiated by dust, come from three instruments: ISOCAM (Cesarsky et al 1996), in the mid-IR, and ISOPHOT (Lemke et al 1996), in the far-IR, on board ISO (Kessler et al 1996), and the SCUBA bolometer array at JCMT, in the sub-mm (see Lilly et al, in these proceedings). All three instruments present number counts much above expectations from no evolution models based on IRAS results on the local universe, indicating that the bulk of the cosmic IR background (CIRB) originates from LIRGs, or even brighter galaxies. Hence, answering NO to the previous question. So that we can say that we are leaving in the middle of an “Infrared Revolution”. The following discussion will illustrate the role that the ISOCAM and ISOPHOT extragalactic results are playing in it, being complementary to and consistent with the SCUBA results.
## 2 Description of the Surveys and Number Counts
ISOCAM is a mid-IR camera which allows to perform either wide band imagery or low resolution spectroscopy using Circular Variable Filters, in the range 5-18 $`\mu `$m. The two wide band filters that were used for the extragalactic surveys, LW2 & LW3, are centered at respectively 6.75 $`\mu `$m (5-8 $`\mu `$m) and 15 $`\mu `$m (12-18 $`\mu `$m) and were chosen because they cover the mid-IR emission from two different origins, the aromatic features and the hot dust continuum (see Section 3). Its pixel size is 6” in the most sensitive mode, for a FWHM PSF of 9” at 15 $`\mu `$m, but its astrometric accuracy can reach 2” in the micro-scanning mode of the deepest surveys.
ISOPHOT is an imaging photo-polarimeter covering the 2.5-240 $`\mu `$m range. The most performant filter for extragalactic surveys is centered at 175 $`\mu `$m (130-219 $`\mu `$m) and uses the four pixels camera, C200, with a 92” side length and a 1.9’ FWHM PSF at 175 $`\mu `$m.
The list of all ISOCAM extragalactic surveys performed during the ISO lifetime is given in the Table 1. We used 8 of them to produce the number counts in the Fig. 2 (from Elbaz et al 1999), for a total of 614 sources above an 80 $`\%`$ completeness limit. The ISOPHOT number counts from Dole et al (1999) superimposed in the Fig. 2 (left) use 208 sources brighter than 100 mJy in $``$3 square degrees from the fields quoted in the Table 2. The ISOCAM and ISOPHOT surveys were performed in regions selected for their low Galactic foreground emission, both in the northern and southern hemispheres to avoid strong contamination from large-scale structure.
## 3 Origin of the infrared emission
The origin of the mid-IR emission of galaxies has been a subject of debate prior to IRAS, when evolved stars such as OH/IR stars were expected to play a major role. However, based on the IRAS Point Source Catalog, source counts at 12 $`\mu `$m have shown that circumstellar dust shells accounted for only $``$10$`\%`$ of the 12 $`\mu `$m emission in the Galaxy at $`b10^{}`$ (see Review by Cox & Mezger 1989). It is now established that except for ellipticals (with no dust lanes), the mid-IR emission of galaxies does not originate primarily from individual stars but rather from the light of young stars (or of an AGN) absorbed by dust and re-radiated in the mid-IR mostly at the interface between HII regions and molecular clouds (the photo-dissociation regions, PDR, see Cesarsky & Sauvage 1999 and references therein). The mid-IR spectrum of a galaxy can be divided into three components: a set of broadband aromatic features, with their underlying continuum, narrow forbidden lines of ionized gas (negligible in ISOCAM wide band imagery), and a hot thermal continuum due to very small grains (VSGs) of dust heated by stellar radiation at temperatures greated than 150 K. The broadband aromatic features have been proposed to originate either from PAHs (for Polycyclic Aromatic Hydrocarbons, Léger $`\&`$ Puget 1984, Allamandola et al 1989) or from coal grains (Papoular 1999). In either case, they require the presence of aromatic structures, but their exact nature has not yet been definitively established. The low resolution spectra from ISOCAM have clearly established that in the wavelength range 5-8.5 $`\mu `$m of the 7 $`\mu `$m filter, the spectra of spiral and starburst galaxies were dominated by the aromatic broadband features located at 6.2, 7.7, 8.6, 11.3 and 12.7 $`\mu `$m. At redshifts of the order of z$``$0.8, the median redshift of the faint ISOCAM sources (see Section 4.2), the 15 $`\mu `$m band is shifted to this region of the spectrum dominated by the aromatic features, with a contribution of hot dust continuum.
At larger wavelengths, the spectral energy distribution (SED) of a galaxy is progressively dominated by the thermal emission of larger grains heated at a lower temperature (around 60 K), which produces the bulk of the overall IR emission originating from dust (around 60-100 $`\mu `$m). This ’cold’ dust dominates the IR emission measured by ISOPHOT. At 175 $`\mu `$m, the physical origin of the emission remains the same up to high redshifts and is well-known to be correlated with star formation and radio activity (Helou, Soifer & Rowan-Robinson 1985).
The link between mid-IR and star formation (SF) is less straightforward, but only LIRGs can be detected at $`z0.8`$ above 0.1 mJy at 15 $`\mu `$m. Hence mid-IR surveys can at least pinpoint IR active galaxies. Moreover, a correlation was found between mid-IR and UV flux densities (Boselli et al 1997, 1998) as well as with $`H_\alpha `$ (for the disks of spiral galaxies, Roussel et al, these proceedings). However, the complex link between mid-IR and SF is illustrated by the Fig. 1 (left), where we compare the SED of the ultra-LIRG Arp 220 and the nearby starburst M82, normalized by a factor 5 to reach the same energy density at $``$8 $`\mu `$m. At z$``$0.8, both galaxies would have the same 15 $`\mu `$m flux density but differ by a factor $``$20 in the far-IR. But Arp220 is an extreme case which can easily be rejected as a candidate SED for the ISOCAM galaxies (see Section 4.2), and M82’s SED, which is more typical of most LIRGs ($`L_{bol}L_{IR}>10^{11}L_{}`$) and ultra-LIRGs ($`L_{bol}>10^{12}L_{}`$) although it is only $`410^{10}L_{}`$, is probably a better candidate. Indeed, in the HDF-N seven galaxies, which were detected both at 15 $`\mu `$m and in the radio (Richards et al 1998), show consistent SF rates if we use the SED of M82 to convert the mid-IR flux into a SF rate (Aussel et al 1999, in preparation).
## 4 Discussion about the 15 $`\mu `$m results
### 4.1 The 15 $`\mu `$m counts
The first striking result of the 15 $`\mu `$m source counts is the consistency of the eight surveys (noted with a $``$ in the Table 1) over the full flux range (Fig. 2). Some scatter is nevertheless apparent; given the small size of the fields surveyed, we attribute it to clustering effects. The two main features of the observed counts are a significantly super-euclidean slope ($`\alpha =3.0`$) from 3 to 0.4 mJy and a fast convergence at flux densities fainter than 0.4 mJy. In particular, the combination of five independent surveys in the flux range 90-400 $`\mu `$Jy shows a turnover of the normalized differential counts around 400 $`\mu `$Jy and a decrease by a factor $`3`$ at 100 $`\mu `$Jy. We believe that this decrease, or the flattening of the integral counts, below $``$400 $`\mu `$Jy, is real. It cannot be due to incompleteness, which was quantified using the Monte-Carlo simulations. The differential counts can be fitted by two power laws by splitting the flux axis in two regions around 0.4 mJy. In units of mJy<sup>-1</sup> deg<sup>-2</sup>, we obtain, by taking into account the error bars ($`S`$ is in mJy):
$`{\displaystyle \frac{dN(S)}{dS}}=\{\begin{array}{cccc}(2000\pm 600)& S^{(1.6\pm 0.2)}& \mathrm{}& 0.1S0.4\\ & & & \\ (470\pm 30)& S^{(3.0\pm 0.1)}& \mathrm{}& 0.4S4\end{array}`$ (4)
The total number density of sources detected by ISOCAM at 15 $`\mu `$m down to 100 $`\mu `$Jy (no lensing) is ($`2.4\pm 0.4`$) arcmin<sup>-2</sup>. It extends up to ($`5.6\pm 1.5`$) arcmin<sup>-2</sup>, down to 50 $`\mu `$Jy, when including the lensed field of A2390 (Altieri et al 1999). The differential counts (Fig. 2 (right)), which are normalized to $`S^{2.5}`$ (the expected differential counts in a non expanding Euclidean universe with sources that shine with constant luminosity), present a turnover around $`S_{15\mu m}`$=0.4 mJy, above which the slope is very steep ($`\alpha =3.0\pm 0.1`$). No evolution predictions were derived assuming a pure k-correction in a flat universe ($`q_0=0.5`$), including the effect of the aromatic features in the galaxy spectra. The lower curve of the hatched area, which materializes the ’no evolution’ region in the Fig. 2, is based on the Fang et al (1998) IRAS 12 $`\mu `$m local luminosity function (LLF), using the spectral template of a relatively quiescent spiral galaxy (M51). The upper curve is based on the Rush, Malkan & Spinoglio (1993) IRAS-12 $`\mu `$m LLF, translated to 15 $`\mu `$m using as template the spectrum of M82, which is typical of most starburst galaxies in this band. More active and extincted galaxies, like Arp220, would lead to even lower number counts at low fluxes while flatter spectra like those of AGNs are less flat than the one of M51. In the absence of a well established LLF at 15 $`\mu `$m, we consider these two models as upper and lower bounds to the actual no-evolution expectations; note that the corresponding slope is $`2`$. The actual number counts are well above these predictions; in the 0.3 mJy to 0.6 mJy range, the excess is around a factor 10: clearly, strong evolution is required to explain this result (note the analogy with the radio source counts, Windhorst et al 1993).
For comparison, we have superimposed on Fig. 2 the optical B band (Metcalfe et al 1995) and near-IR K band (Gardner et al 1993) integral counts, normalized in terms of $`\nu S_\nu `$. For bright sources, with densities lower than 10 deg<sup>-2</sup>, these curves run parallel to an interpolation between the ISOCAM counts presented here and the IRAS counts; the bright K sources emit about ten times more energy in this band than a comparable number of bright ISOCAM sources at 15 $`\mu `$m. But the ISOCAM integral counts present a rapid change of slope around 1-2 mJy, and their numbers rise much faster than those of the K and B sources. The sources detected by ISOCAM are a sub-class of the K and B sources which harbor activity hidden by dust.
### 4.2 Nature of the 15 $`\mu `$m galaxies
We believe, according to the results obtained on the HDF (Aussel et al 1999a,b) and CFRS fields (Flores et al 1999a,b, see also these proceedings), that the sources responsible for the ’bump’ in the 15 $`\mu `$m counts are not the low mass faint blue galaxies which dominate optical counts and have a median redshift around z$``$0.6 (Pozzetti et al 1998). Instead, they most probably are bright and massive galaxies whose emission occurs essentially in the IR and could account for a considerable part of the star formation in the universe at $`z<1.5`$. Indeed, from the full sample of 15 $`\mu `$m galaxies with known redshift and optical-near IR magnitudes, we found that these galaxies are massive ($`10^{11}M_{}`$) and that their emission occurs essentially in the IR. Their median redshift is z$``$0.8 in a sample of 42 galaxies brighter than 0.1 mJy in the HDF+FF (Aussel et al 1999b) and z$``$0.7 in a sample of 41 galaxies brighter than 0.35 mJy in the CFRS-1415 field (Flores et al 1999b). The assessment of their bolometric luminosity requires the assumption of a spectral energy distribution, which is largely uncertain since the ratio of the far-IR over mid-IR flux densities is highly variable among galaxies. However, one can set limits to the bolometric luminosity using the following arguments. If all galaxies had extreme SEDs like the one of Arp220, they would produce a contribution to the SCUBA-850 $`\mu `$m number counts (see Blain et al 1999) much above observations (see Fig. 3). An Arp220-like SED can also be rejected for over-producing the 140 $`\mu `$m EBL with respect to the one measured by DIRBE on-board COBE (see Fig. 4). The light radiated in the far-IR by these galaxies however cannot be much lower than that emitted at 15 $`\mu `$m, since only galaxies luminous in the IR can be detected at z$`>`$0.5 with a 15 $`\mu `$m flux density larger than 0.1 mJy (the origin of the emission cannot be stellar without requiring excessive masses). We therefore estimate the mean bolometric luminosity of these galaxies to be of the order of a few $`10^{11}L_{}`$.
The galaxies from the HDF-N plus Flanking Fields whose 15 $`\mu `$m flux density is greater than 0.1 mJy (sensitivity limit of ISOCAM) harbour a I<sub>C</sub>-(H+K) color distribution very similar to that of the whole sample of galaxies for which we have access to both the I<sub>C</sub>-(H+K) colors and spectroscopic redshifts (see Fig. 1 (right)). Aussel et al (1999b and PhD thesis) showed that the sub-sample of galaxies with known spectroscopic redshift keeps the same color properties than the full sample of HDF+FF galaxies, hence we did not include a selection bias in the color distribution.
The median colors are I<sub>C</sub>-(H+K)= 2.3 for the 15 $`\mu `$m galaxies and I<sub>C</sub>-(H+K)= 2 for all galaxies, hence only differ by 0.3 dex. However, the linear fit to the I<sub>C</sub>-(H+K) versus redshift plot of the two samples of galaxies in Fig. 1 (right) shows that the ’dusty’ galaxies tend to redden slightly faster with increasing redshift, than the natural reddening of the whole population of field galaxies which is due to k-correction: the color difference increases from 0.2 dex below z=0.7 to 0.4 dex above z=0.7. Nevertheless, this difference is not strong enough to allow one to separate the infrared galaxies from the whole sample like for Lyman-break galaxies (even accounting for the other set of optical colors existing for the HDF galaxies). The origin of this reddening with redshift is not clear but it is probably due to a selection of the galaxies suffering from more extinction, hence emitting more in the infrared, at larger redshifts.
Hence the population of galaxies producing the 15 $`\mu `$m number counts excess is very distinct from the one which dominates the deep optical counts, known to be made of low mass galaxies with blue luminosities, although at similar redshifts. In other words, the star formation activity responsible for the light emitted by the 15 $`\mu `$m galaxies is not the other face of the same star formation activity already quantified from the optical surveys, but instead it should be considered as a second component, which was previously missed.
## 5 Discussion about the 175 $`\mu `$m results
### 5.1 The 175 $`\mu `$m counts
As for the 15 $`\mu `$m ISOCAM counts, the 175 $`\mu `$m ISOPHOT counts present a strong excess with respect to IRAS of about one order of magnitude (Puget et al 1999, Dole et al 1999). The slope of the integral counts is $`\alpha =2.2`$, hence much above the Euclidean value of -1.5, and far from flattening, contrary to the ISOCAM counts which go deep enough to detect the beginning of the flattening, with a slope decreasing from -2.4 around 0.5 mJy to -1.2 around 0.1 mJy. It is therefore natural that the detected sources account only for 10 $`\%`$ of the CIRB measured by COBE (see Fig. 4).
### 5.2 Nature of the 175 $`\mu `$m galaxies
The search for optical counterparts to the ISOPHOT-175 $`\mu `$m galaxies is not an easy task because of the large uncertainty on the position of the ISOPHOT sources and of the absence of a clear optical signature. Hence the use of other wavelengths with better spatial resolution and sensitivity to the IR activity is mandatory in this study. Because of the good radio/far-IR correlation, the radio domain is an obvious candidate and VLA surveys of the northern fields (ELAIS N1 & N2) have been carried out by Ciliegi et al (1999) in the framework of the ELAIS project (Rowan-Robinson et al 1999). A sample of ten 21 cm radio counterparts of the ISOPHOT galaxies, i.e. with $`S_{175\mu m}>100mJy`$, have been selected for their high $`175\mu m/21cm`$ ratio by Scott et al (1999) for a follow-up at 450 and 850 $`\mu `$m with SCUBA. Assuming a dust temperature of $`T_d=40K`$, they find a range of low to moderate redshifts of $`z=01.5`$ (no spectroscopic redshift are available yet), when accounting for the SCUBA measurements or upper limits together with the ISOPHOT ones.
In these two fields, the ISOCAM surveys are not sensitive enough to strongly constrain the SED of the ISOPHOT galaxies and miss the majority of them ($`S_{15\mu m}>3mJy`$). In the southern field however (called MFB, for ’Marano FIRBACK’ since it is shifted with respect to the original Marano Field, Marano, Zamorani & Zitelli, 1988, in order to minimize the foreground Galactic cirrus emission), deep ISOCAM surveys have been performed over 0.5 square degree down to 0.3 mJy (and a completeness limit of $`0.4mJy`$). Half of the 24 FIRBACK galaxies in the field have a counterpart, sometimes several, at 15 $`\mu `$m above $`S_{15\mu m}=0.3mJy`$ and closer than 30” (except one at 50”). One can compare their $`S_{175\mu m}/S_{15\mu m}`$ ratio to the one of any template SED at any redshift and determine a possible redshift range for this SED. The striking result of this study is that none of the 12 galaxies present any 15 $`\mu `$m counterpart compatible with the SED of M82 at any redshift (unless they are brighter than $`10^{13}L_{}`$), while all but one of them are consistent with an Arp220-like galaxy of similar luminosity ($`10^{12}L_{}`$) and at a redshift between z=0.1-0.4. For the remaining 12 galaxies, they could be compatible with an M82-like SED for $`z>1.2`$ and $`L_{bol}>410^{12}L_{}`$ or with an Arp220-like SED (though fainter), if only their redshift is larger than $`z=0.3`$. In any case, the FIRBACK galaxies are most probably LIRGs or ULIRGs and their large number density implied by the number counts implies a strong evolution of this population of galaxies with respect to IRAS, consistently with ISOCAM, although the 175 $`\mu `$m galaxies are most probably brighter than the 15 $`\mu `$m ones.
## 6 Conclusion
The ISO extragalactic surveys have demonstrated the fundamental role of IR in the understanding of galaxy evolution. However, only ground-based follow-ups will allow us to understand the nature of the population of galaxies responsible for the strong evolution exhibited by the IR number counts. Are they particularly rich in metals ? Do they harbor active galactic nuclei ? We have a large enough sample of galaxies to be able to quantify the evolution of the luminosity function in the IR as a function of redshift. However, we are still far from understanding the origin of the huge cosmic IR background revealed by COBE, but SIRTF and FIRST are still to come… |
no-problem/9911/math9911237.html | ar5iv | text | # Untitled Document
|
no-problem/9911/quant-ph9911052.html | ar5iv | text | # Coherent states, Yang-Mills theory, and reduction
## 1. Introduction
The quantization of Yang-Mills theory is an important example of the quantization of reduced Hamiltonian systems. This paper concerns the simplest non-trivial case of quantized Yang-Mills theory, namely, pure Yang-Mills on a spacetime cylinder. Most of this paper is an exposition of a joint work \[DH1\] with Bruce Driver, with an emphasis on the concepts rather than the mathematical technicalities.
Driver and I use as our main tool the Segal–Bargmann transform, or equivalently, coherent states. We reach two main conclusions. First, upon reduction the ordinary coherent states on the space of connections become the generalized coherent states in the sense of \[H1\] on the finite-dimensional compact structure group. Second, coherent states provide a way to make rigorous the generally accepted idea that upon reduction the Laplacian for the infinite-dimensional space of connections becomes the Laplacian on the structure group. In the rest of the introduction I give a schematic description of the the paper. More details are found in the body of the paper and in \[DH1\].
Driver and I use the canonical quantization approach rather than the path-integral approach, and we work in the temporal gauge. As stated, we assume that spacetime is a cylinder, namely, $`S^1\times .`$ We fix a compact connected structure group $`K,`$ which I will assume here is simple connected, with Lie algebra $`𝔨.`$ The configuration space for the classical theory is the space $`𝒜`$ of $`𝔨`$-valued connection 1-forms over the spatial circle. The gauge group $`𝒢`$, consisting of maps of the spatial circle into $`K,`$ acts naturally on $`𝒜.`$ The based gauge group $`𝒢_0,`$ consisting of gauge transformations that equal the identity at one fixed point in the spatial circle, acts freely on $`𝒜,`$ and the quotient $`𝒜/𝒢_0`$ is simply the compact structure group $`K.`$ This reflects that in this simple case the only gauge-invariant quantity is the holonomy of a connection around the spatial circle.
Meanwhile we have the complexification of $`𝒜,`$ namely, $`𝒜_{}=𝒜+i𝒜,`$ which is identifiable with the cotangent bundle of $`𝒜`$ and is the phase space for the unreduced system. We have also the complexification $`K_{}`$ of the structure group $`K,`$ which is identifiable with the cotangent bundle of $`K.`$ Here $`K_{}`$ is the unique simply connected complex Lie group whose Lie algebra is $`𝔨+i𝔨.`$ One defines in the obvious way the (based) complexified gauge group $`𝒢_{0,}`$, which acts holomorphically on $`𝒜_{}.`$ The quotient $`𝒜_{}/𝒢_{0,}`$ is $`K_{}.`$ This is the reduced phase space for the theory.
Now we have the ordinary Segal–Bargmann tranform for $`𝒜`$, which maps from an $`L^2`$ space of functions on $`𝒜`$ to an $`L^2`$ space of holomorphic functions on $`𝒜_{}.`$ Much more recently there is a generalized Segal–Bargmann transform for $`K`$ \[H1\], which maps from an $`L^2`$ space of functions on $`K`$ to an $`L^2`$ space of holomorphic functions on $`K_{}.`$ The gist of \[DH1\] is that the ordinary Segal–Bargmann transform for $`𝒜,`$ when restricted to the gauge-invariant subspace is precisely the generalized Segal-Bargann transform for $`𝒜/𝒢_0=K.`$ To say the same thing in the language of coherent states, taking the ordinary coherent states for $`𝒜`$ and projecting them onto the gauge-invariant subspace gives the generalized coherent states for $`K`$ (in the sense of \[H1\]). So \[DH1\] gives a new way of understanding the generalized Segal–Bargmann transform (or generalized coherent states) for a compact Lie group $`K.`$
There is another purpose to the paper \[DH1\]. The Segal–Bargmann transform for $`𝒜`$ may be expressed in terms of the Laplacian $`\mathrm{\Delta }_𝒜`$ for $`𝒜.`$ The generalized Segal–Bargmann transform for $`K`$ is expressed in a precisely parallel way in terms of the Laplacian for $`K.`$ Thus the result that the Segal–Bargmann transform for $`𝒜`$ becomes the generalized Segal–Bargmann transform for $`K`$ when restricted to the gauge-invariant subspace, this result gives a mathematically precise meaning to the following generally accepted principle.
(1)
$$\text{On the gauge-invariant subspace, }\mathrm{\Delta }_𝒜\text{ reduces to}\mathrm{\Delta }_K\text{.}$$
Driver and I wish to interpret Theorem 5.2 of \[DH1\] as a mathematically rigorous version of this principle, which, as explained below, does not really make mathematical sense as written. Since there is no spatial curvature when space is one-dimensional, the Hamiltonian for our theory is just a multiple of $`\mathrm{\Delta }_𝒜.`$ So we may say that upon restriction to the gauge-invariant subspace the Hamiltonian becomes a multiple of $`\mathrm{\Delta }_K.`$
I discuss two additional points. First, I discuss why, even at a formal level, $`\mathrm{\Delta }_𝒜`$ should go to $`\mathrm{\Delta }_K`$ on the invariant subspace. For the general situation of a manifold modulo an isometric group action, even in the finite-dimensional case, the Laplacians before and after reduction do not agree. So beyond the technicalities associated to the infinite-dimensionality, something special is happening in this case. Second, I consider the possibility of doing things in the opposite order, namely, first passing to the reduced phase space $`K_{},`$ and then constructing coherent states by means of geometric quantization. It turns out that the two procedures give the same answer, provided that on includes as part of the geometric quantization the mysterious “half-form correction.”
Acknowledgments. The idea of deriving the generalized Segal–Bargmann transform from the infinite-dimensional ordinary Segal–Bargmann transform is due to L. Gross and P. Malliavin \[GM\]. However, \[GM\] was not intended to be about Yang-Mills theory. What I am here calling the gauge group $`𝒢_0`$ they call the loop group, and its action in \[GM\] is not unitary. To get the Yang-Mills interpretation that we were striving for, Driver and I modified the approach of Gross and Malliavin so as to make the action of $`𝒢_0`$ unitary. (More precisely, we take a certain limit under which the action of $`𝒢_0`$ becomes formally unitary.)
The idea that the generalized coherent states for $`K`$ could be obtained from the ordinary coherent states for $`𝒜`$ by reduction is due to K. Wren \[W\]. Wren uses the “Rieffel induction” approach proposed by Landsman \[L1\] and carried out in the abelian case by Landsman and Wren \[LW\]. See also the exposition in the book of Landsman \[L2, IV.3.7\]. I describe in Section 5 the relationship of our results to those of Wren.
I am endebted to Bruce Driver for clarifying to me many aspects of what is discussed here. I also acknowledge valuable discussions with Andrew Dancer, and I thank Dan Freed for a valuable suggestion regarding the half-form correction.
## 2. Classical Yang-Mills theory on a spacetime cylinder
Yang-Mills theory on a spacetime cylinder is an exactly solvable model \[R\]. Nevertheless, I believe that there are things to learn here, both classically and quantum mechanically, by comparing what happens before gauge symmetry is imposed to what happens afterward. I begin with the classical theory, borrowing heavily from the treatment of Landsman \[L2\].
We work on the spacetime manifold $`S^1\times ,`$ with $`S^1`$ being space and $``$ time. Fix a connected compact Lie group $`K`$, the structure group, which for simplicity I take to be simply connected, and fix an Ad-invariant inner product on the Lie algebra $`𝔨`$ of $`K`$. We work in the temporal gauge, which has the advantage of allowing the classical Yang-Mills equations to be put into Hamiltonian form. The temporal gauge is only a partial gauge-fixing, leaving still a large gauge group $`𝒢,`$ namely the group of mappings of the space manifold $`S^1`$ into the structure group $`K.`$ Note that the gauge group is just a loop group in this case. I will concern myself only with the based gauge group $`𝒢_0,`$ consisting of maps of $`S^1`$ into $`K`$ which equal the identity at one fixed point in $`S^1.`$ The remaining gauge symmetry can easily be added later.
In the temporal gauge, the Yang-Mills equations have a configuration space $`𝒜`$ consisting of connections on the space manifold. The connections are 1-forms with values in the Lie algebra $`𝔨.`$ Since our space manifold is one-dimensional, we may think of the connections as $`𝔨`$-valued functions. There is a natural norm on $`𝒜`$ given by
$$A^2=_0^1\left|A\left(\tau \right)\right|^2𝑑\tau ,A𝒜.$$
Here $`S^1`$ is the interval $`[0,1]`$ with ends identified, and $`\left|A\left(\tau \right)\right|^2`$ is computed using the inner product on $`𝔨.`$ The norm allows us to define a distance function
$$d(A,B):=AB.$$
The gauge group $`𝒢_0`$ acts on $`𝒜`$ by
(2)
$$\left(gA\right)_\tau =g_\tau A_\tau g_\tau ^1g_\tau ^1\frac{dg}{d\tau }.$$
The map $`AgAg^1`$ is linear, invertible, and norm-preserving, hence a “rotation” of $`𝒜.`$ So the action of each $`g𝒢_0`$ is distance-preserving, a combination of a rotation and a translation in $`𝒜.`$
The phase space of the theory is the cotangent bundle of $`𝒜,`$ $`T^{}\left(𝒜\right)𝒜+𝒜.`$ The action of $`𝒢_0`$ on $`𝒜`$ extends in a natural way to an action on $`𝒜+𝒜`$ given by
$$g(A,P)=(gA,gPg^1).$$
So the translation part of (2) affects only the “position” $`A`$ and not the “momentum” $`P.`$
The Yang-Mills equations take place in the phase space $`𝒜+𝒜`$ and have three parts. First we have a dynamical part. The equations of motion are just Hamilton’s equations, for the Hamiltonian function
$$H\left(A+iP\right)=\frac{1}{2}P^2.$$
Normally there would be another term involving the curvature of $`A,`$ but that term is necessarily zero in this case, since $`S^1`$ is one-dimensional. Thus the solutions of Hamilton’s equations are embarassingly easy to write down: the general solution is
$$(A\left(t\right),P\left(t\right))=(A_0+tP_0,P_0).$$
This is just free motion in $`𝒜.`$ Observe that the Hamiltonian $`H`$ is invariant under the action of $`𝒢_0`$ on $`𝒜+𝒜.`$
Second we have a constraint part. This says that the solutions (trajectories in $`𝒜+𝒜`$) have to lie in a certain set, which I will denote $`J^1\left(0\right),`$ which is “the zero set of the moment mapping for the action of $`𝒢_0.`$” I will not repeat here the formulas, which may be found for example in \[DH1, Sect. 2\]. This constraint is of a simple sort, in that $`J^1\left(0\right)`$ is invariant under the dynamics and under the action of $`𝒢_0`$ on $`𝒜+𝒜.`$ So the constraint does not alter the dynamics, it merely restricts us to certain special solutions of the original equations of motion.
Third we have a philosophical part. This says that the only functions on phase space that are physically observable are ones that are gauge-invariant.
The last two points together say that we may as well think of the dynamics as taking place in $`J^1\left(0\right)/𝒢_0,`$ which is the same as $`T^{}\left(𝒜/𝒢_0\right).`$ This is Marsden–Weinstein or symplectic reduction. Since the Hamiltonian function $`H`$ is $`𝒢_0`$-invariant, it makes sense as a function on $`J^1\left(0\right)/𝒢_0.`$
Now, we are in a very simple situation, with the space manifold being just a circle. In this case two connections are gauge-equivalent if and only if they have the same holonomy around the spatial circle. So the orbits of $`𝒢_0`$ are labeled by the holonomy $`h\left(A\right)`$ of a connection $`A`$ around the circle, where for $`A𝒜,`$ $`h\left(A\right)`$ is an element of the structure group $`K.`$ It is easily seen that any $`xK`$ can be the holonomy of some $`A,`$ and so we have
$$𝒜/𝒢_0K.$$
Thus
$$J^1\left(0\right)/𝒢_0T^{}\left(𝒜/𝒢_0\right)T^{}\left(K\right).$$
After this reduction, the dynamics become geodesic motion in $`K.`$ The geodesics may be written explicitly as $`xe^{tX}`$with $`xK`$ and $`X𝔨.`$
We require one last discussion before turning to the quantum theory. We may think of $`𝒜+𝒜`$ as the complex vector space $`𝒜_{}=𝒜+i𝒜,`$ in the same way that we think of $`T^{}\left(\right)+`$ as $`.`$ We may then think of elements of $`𝒜_{}`$ as functions (or 1-forms) with values in the complexified Lie algebra $`𝔨_{}=𝔨+i𝔨.`$ The action of $`𝒢_0`$ extends to an action on $`𝒜_{}`$ by
$$\left(gZ\right)_\tau =g_\tau Z_\tau g_\tau ^1g_\tau ^1\frac{dg}{d\tau },$$
where $`Z:[0,1]𝔨_{}.`$ Note that the translation part in in the real direction; that is, $`g_\tau ^1\frac{dg}{d\tau }`$ is in $`𝒜.`$ One can think of elements of $`𝒜_{}`$ as complex connections and thus define their holonomy. But the holonomy now takes values in the complexified group $`K_{},`$ where $`K_{}`$ is the unique simply connected complex Lie group with Lie algebra $`𝔨+i𝔨.`$ For example, if $`K=SU(n)`$ then $`K_{}=SL(n;).`$ The complexified (based) gauge group $`𝒢_{0,}`$ is then the group of based loops with values in $`K_{}.`$ The same reasoning as on $`𝒜`$ shows that the only $`𝒢_{0,}`$-invariant quantity on $`𝒜_{}`$ is the holonomy; so $`𝒜_{}/𝒢_{0,}=K_{}.`$
It turns out that restricting to the zero set of the moment mapping and then dividing out by the action of $`𝒢_0`$ gives the same result as working on the whole phase space and then dividing out by the action of $`𝒢_{0,}.`$ Thus
$$J^1\left(0\right)/𝒢_0=𝒜_{}/𝒢_{0,}=K_{}.$$
On the other hand, we have already said that $`J^1\left(0\right)/𝒢_0`$ is identifiable with $`T^{}\left(K\right).`$ So we have a natural identification
$$K_{}T^{}\left(K\right).$$
This is explained in detail in Section 6 and the resulting identification is given there explicitly.
## 3. Formal and semiformal quantization
In this section we will see what is involved in trying to quantize this system. This discussion will set the stage for the entrance of the Segal–Bargmann transform and the coherent states in the next two sections.
Let us first try to quantize our Yang-Mills example at a purely formal level, that is, without worrying too much whether our formulas make sense. I want to do the quantization before the reduction by $`𝒢_0.`$ If we did the reduction before the quantization, then we would have a finite-dimensional system, which is easily quantized. So it is of interest to do the quantization first and see if this gives the same result. See \[R\], where quantization is done after the reduction, and \[Di\], where quantization is done before the reduction.
Since our system has a configuration space $`𝒜,`$ we may formally take our unreduced quantum Hilbert space to be
$$L^2(𝒜,𝒟A),$$
where $`𝒟A`$ is the fictitious Lebesgue measure on $`𝒜.`$ The quantization of the constraint equation (see \[DH1, Sect. 2\]) then imposes the condition that our wave functions be $`𝒢_0`$-invariant. Note that the quantization of the second part of the classical theory (the constraint) automatically incorporates the third part as well (the $`𝒢_0`$-invariance). So we want the reduced (physical) quantum Hilbert space to be
$$L^2(𝒜,𝒟A)^{𝒢_0}:=\left\{fL^2(𝒜,𝒟A)\right|f\left(gA\right)=f\left(A\right),g,A\}.$$
Recall that in our example, in which space is a circle, two connections are $`𝒢_0`$-equivalent if and only if they have the same holonomy around the spatial circle. That means that a $`𝒢_0`$-invariant function must be of the form
(3)
$$f\left(A\right)=\varphi \left(h\left(A\right)\right),$$
where $`h\left(A\right)`$ $`K`$ is the holonomy of $`A`$ and where $`\varphi `$ is a function on the structure group $`K.`$ Furthermore, as we shall see more clearly in the next section, it is reasonable to think that for a function of the form (3), integrating $`\left|f\left(A\right)\right|^2`$ with respect to $`𝒟A`$ is the same as integrating $`\left|\varphi \left(g\right)\right|^2`$ with respect to a multiple of the Haar measure on $`K.`$ Thus
(4)
$$L^2(𝒜,𝒟A)^{𝒢_0}L^2(K,Cdg),$$
for some (probably infinite) constant $`C.`$ This is our physical Hilbert space.
Next we consider the Hamiltonian. Formally quantizing the function $`\frac{1}{2}P^2`$ in the usual way gives
(5)
$$\widehat{H}=\frac{\mathrm{}^2}{2}\mathrm{\Delta }_𝒜=\frac{\mathrm{}^2}{2}\underset{k=1}{\overset{\mathrm{}}{}}\frac{^2}{x_k^2},$$
where the $`x_k`$’s are coordinates with respect to an orthonormal basis of $`𝒜.`$ We must now try to determine how $`\widehat{H}`$ acts on the $`𝒢_0`$-invariant subspace. In light of what happens when performing the reduction before the quantization, it is reasonable to guess that on the invariant subspace $`\mathrm{\Delta }_𝒜`$ reduces to $`\mathrm{\Delta }_K,`$ that is,
(6)
$$\mathrm{\Delta }_𝒜\left[\varphi \left(h\left(A\right)\right)\right]=\left(\mathrm{\Delta }_K\varphi \right)\left(h\left(A\right)\right).$$
(See also \[Di, W\].) If we accept this and if we ignore the infinite constant $`C`$ in (4) then we conclude that our quantum Hilbert space is
$$L^2(K,dx)$$
and our Hamiltonian is
$$\widehat{H}=\frac{\mathrm{}^2}{2}\mathrm{\Delta }_K.$$
This concludes the formal quantization of our system.
We now begin to consider how to make this mathematically precise. One approach is to forget about the measure theory (i.e. the Hilbert space) and to try to prove (6). As it turns out, the answer is basis-dependent–choosing different bases in (5) will give different answers. Another way of saying this is that the matrix of second derivatives of a function $`f`$ of the form (3) is in general non-trace-class. However, if one uses the most obvious sort of basis, then indeed it turns out that (5) is true. See the appendix of \[DH1\].
Even without the problem of basis-dependence, the above approach is unsatisfying because we would like to define $`\widehat{H}`$ as an operator in some Hilbert space. Since Lebesgue measure $`𝒟A`$ does not actually exist, one reasonable procedure is to “approximate” $`𝒟A`$ by a Gaussian measure $`dP_s\left(A\right)`$ with large variance $`s.`$ This means that $`P_s`$ is formally given by the expression
$$dP_s\left(A\right)=c_se^{A^2/2s}𝒟A,$$
where $`c_s`$ is supposed to be a normalization constant that makes the total integral one. Formally as $`s\mathrm{}`$ we get back a multiple of Lebesgue measure $`𝒟A.`$ The measure $`P_s`$ does exist rigorously, provided that one allows sufficiently non-smooth connections.
There is good news and bad news with this approach. First the good news. 1) Even though the connections in the support of $`P_s`$ are not smooth, the holonomy of such a connection makes sense, as the solution to a stochastic differential equation. 2) If we define the gauge-invariant subspace to be
$$L^2(𝒜,P_s)^{𝒢_0}=\left\{f\text{ }\right|\text{for all }g𝒢_0,\text{ }f\left(gA\right)=f\left(A\right)\text{ for }P_s\text{-almost every }A\},$$
then the Gross ergodicity theorem \[G2\] asserts that $`L^2(𝒜,P_s)^{𝒢_0}`$ is precisely what we expect, namely, the space of functions of the form $`f\left(A\right)=\varphi \left(h\left(A\right)\right),`$ with $`\varphi `$ a function on $`K.`$ 3) There is a natural dense subspace of $`L^2(𝒜,P_s)`$ on which $`\mathrm{\Delta }_𝒜`$ is unambiguously defined, consisting of smooth cylinder functions. Here a cylinder function is one which depends on only finitely many of the infinitely many variables in $`𝒜.`$ See \[DH1, Defn. 4.2\].
Note that the map which takes $`f\left(A\right)`$ to $`f\left(gA\right)`$ is not unitary, because the measure $`P_s`$ is not invariant under the action of $`𝒢_0.`$ Driver and I wish to avoid “unitarizing” the action of $`𝒢_0,`$ because if we did unitarize then there would be no gauge-invariant subspace. See \[DH2\]. Instead of unitarizing the action for a fixed value of $`s,`$ we will eventually let $`s`$ tend to infinity, at which point unitarity is formally recovered.
The bad news about this approach is that $`\mathrm{\Delta }_𝒜`$ is not a closable operator, and that functions of the holonomy are not cylinder functions. This means that if we approximate $`\varphi \left(h\left(A\right)\right)`$ by cylinder functions, then the value of $`\mathrm{\Delta }_𝒜\varphi \left(h\left(A\right)\right)`$ depends on the choice of approximating sequence. So we still have a major problem in making mathematical sense out of the quantization.
## 4. The Segal–Bargmann transform to the rescue
In this section I will explain how the Segal–Bargmann transform can be used to make sense out of the quantization. At the same time, we will see how the generalized Segal–Bargmann transform for the structure group $`K`$ arises from the restriction of the ordinary Segal–Bargmann transform for the gauge-invariant subspace. Although it is technically easier to describe the quantization in terms of the Segal–Bargmann transform, there is a formally equivalent description in terms of coherent states, as I will explain in the next section. See \[B, S1, S2, S3\] and also \[BSZ, H6\] for results on the ordinary Segal–Bargmann transform.
Let me explain the normalization of the Segal–Bargmann transform that I wish to use, first for the finite-dimensional space $`^d.`$ Let $`\left(^d\right)`$ denote the space of holomorphic (complex analytic) functions on $`^d.`$ For any positive constant $`\mathrm{},`$ define
$$C_{\mathrm{}}:L^2(^d,dx)\left(^d\right)$$
by the formula
(7)
$$C_{\mathrm{}}f\left(z\right)=\left(2\pi \mathrm{}\right)^{d/2}_^de^{\left(zq\right)^2/2\mathrm{}}f\left(q\right)𝑑x,z^d.$$
Here $`\left(zq\right)^2`$ means $`\mathrm{\Sigma }\left(z_kq_k\right)^2.`$ If we restrict attention to $`z^d,`$ then this is the standard expression for the solution of the heat equation $`u/\mathrm{}=\frac{1}{2}\mathrm{\Delta }u,`$ at time $`\mathrm{}`$ with initial condition $`f.`$ Thus we may write
$$C_{\mathrm{}}f=\text{ analytic continuation of }e^{\mathrm{}\mathrm{\Delta }/2}f.$$
Here $`e^{\mathrm{}\mathrm{\Delta }/2}f`$ is just a mnemonic for the solution of the heat equation with initial condition $`f,`$ and the analytic continuation is in the space variable (analytic continuation from $`^d`$ to $`^d`$). Because $`\mathrm{}`$ is playing the role of time in the heat equation, it is tempting call this parameter $`t`$ instead of $`\mathrm{};`$ this is what we do in \[DH1\].
Now let $`\nu _{\mathrm{}}`$ be the measure on $`^d`$ given by
$$d\nu _{\mathrm{}}\left(z\right)=\left(\pi \mathrm{}\right)^{d/2}e^{\left(\mathrm{Im}z\right)^2/\mathrm{}}dz,$$
where $`dz`$ refers to the $`2d`$-dimensional Lebesgue measure on $`^d.`$
###### Theorem 1 (Segal-Bargmann transform).
For each positive value of $`\mathrm{},`$ $`C_{\mathrm{}}`$ is a unitary map of $`L^2(^d,dq)`$ onto $`L^2(^d,\nu _{\mathrm{}}),`$ where $`L^2`$ denotes the space of entire holomorphic functions on $`^d`$ which are square-integrable with respect to $`\nu _{\mathrm{}}.`$
This is not quite the form of the transform given by either Segal or Bargmann. Comparing to Bargmann’s map $`A`$ (and taking $`\mathrm{}=1`$ since that is what Bargmann does) we have
$$C_1f\left(z\right)=\left(4\pi \right)^{d/4}e^{z^2/4}Af\left(\frac{z}{\sqrt{2}}\right).$$
The factor in front of $`Af`$ converts from the measure in \[B\] to the measure $`\nu _{\mathrm{}}`$ that I am using, and also emphasizes the role of the heat equation. The factor of $`\sqrt{2}`$ accounts for the difference between Bargmann’s convention that $`z=\left(q+ip\right)/\sqrt{2}`$ and my convention that $`z=q+ip,`$ which is preferable for me because on a more general manifold, the map $`zz/\sqrt{2}`$ does not make sense.
The $`C_{\mathrm{}}`$ form of the Segal–Bargmann transform has the advantage of making explicit the symmetries of position-space. The measure $`dq`$ on $`^d`$ and the measure $`\nu _{\mathrm{}}`$ on $`^d`$ are both invariant under rotations and translations of $`q`$-space, and the transform commutes with rotations and translations of $`q`$-space. Since a gauge transformation is just a combination of a rotation and a translation, this property of $`C_{\mathrm{}}`$ will be useful.
On the other hand, as it stands this form of the Segal–Bargmann transform does not permit taking the infinite-dimensional limit, as we must do if we want to quantize $`𝒜,`$ since neither $`dq`$ nor $`\nu _{\mathrm{}}`$ makes sense when $`d`$ tends to infinity. Fortunately, it is not too hard to fix this problem by adding a little bit of Gaussian-ness to our measures in the $`q`$-directions. It turns out that if we do this correctly, then we can keep the same formula for the Segal–Bargmann transform while making a small change in the measures, and still have a unitary map.
###### Theorem 2.
For all $`s>`$ $`\mathrm{}/2`$, let $`P_s`$ denote the measure on $`^d`$ given by
$$dP_s\left(q\right)=\left(2\pi s\right)^{d/2}e^{q^2/2s}dq$$
and let $`M_{s,\mathrm{}}`$ denote the measure on $`^d`$ given by
$$dM_{s,\mathrm{}}\left(q+ip\right)=\left(\pi \mathrm{}\right)^{d/2}\left(\pi r\right)^{d/2}e^{q^2/r}e^{p^2/\mathrm{}},$$
where $`r=2s`$ $`\mathrm{}.`$ Then the map $`S_{s,\mathrm{}}:L^2(^d,P_s)\left(^d\right)`$ given by
$$S_{s,\mathrm{}}f=\text{ analytic continuation of }e^{\mathrm{}\mathrm{\Delta }/2}f$$
is a unitary map of $`L^2(^d,P_s)`$ onto $`L^2(^d,M_{s,\mathrm{}}).`$
If we multiply the measures on both sides by $`\left(2\pi s\right)^{d/2}`$ and then let $`s\mathrm{}`$ we recover the $`C_{\mathrm{}}`$ version of the transform. On the other hand, for any finite value of $`s`$ it is possible to let $`d\mathrm{}`$ to get a transform that is applicable to our gauge-theory example. So we consider $`L^2(𝒜,P_s),`$ where $`P_s`$ is the Gaussian measure on $`𝒜`$ described in Section 3, which is just the infinite-dimensional limit of the measures $`P_s`$ on $`^d.`$ We consider also the Gaussian measure $`M_{s,\mathrm{}}`$ on $`𝒜_{}`$ that is the infinite-dimensional limit of the corresponding measures on $`^d.`$
We then work with cylinder functions in $`L^2(𝒜,P_s),`$ that is, functions that depend on only finitely many of the infinitely many variables in $`𝒜.`$ (See \[DH1, Defn. 4.2\].) On such functions the Segal–Bargmann transform makes sense, since on such functions $`\mathrm{\Delta }_𝒜`$ reduces to the Laplacian for some finite-dimensional space. It then follows from Theorem 2 that the Segal–Bargmann transform $`S_{s,\mathrm{}}`$ is an isometric map of the space of cylinder functions in $`L^2(𝒜,P_s)`$ into $`L^2(𝒜_{},M_{s,\mathrm{}}).`$ This transform extends by continuity to a unitary map of $`L^2(𝒜,P_s)`$ onto $`L^2(𝒜_{},M_{s,\mathrm{}}).`$ Recall that $`\mathrm{\Delta }_𝒜`$ by itself is a non-closable operator as a map of $`L^2(𝒜,P_s)`$ to itself. Considering $`e^{\mathrm{}\mathrm{\Delta }_𝒜/2}`$ as a map from $`L^2(𝒜,P_s)`$ to itself will not help matters. But by considering $`e^{\mathrm{}\mathrm{\Delta }_𝒜/2}`$ followed by analytic continuation, as a map from $`L^2(𝒜,P_s)`$ to $`L^2(𝒜_{},M_{s,\mathrm{}}),`$ we get a map which is not only closable but continuous (even isometric). It then makes perfect sense to apply this operator (the Segal–Bargmann transform) to functions of the holonomy.
The following theorem summarizes the above discussion.
###### Theorem 3.
For all $`s>`$ $`\mathrm{}/2`$ the map $`S_{s,\mathrm{}}`$ given by
$$S_{s,\mathrm{}}f=\text{ analytic continuation of }e^{\mathrm{}\mathrm{\Delta }_𝒜/2}f$$
makes sense and is isometric on cylinder functions, and extends by continuity to a unitary map of $`L^2(𝒜,P_s)`$ onto $`L^2(𝒜_{},M_{s,\mathrm{}}).`$
We are now ready to state the main result (Theorem 5.2) of \[DH1\].
###### Theorem 4.
Suppose $`fL^2(𝒜,\stackrel{~}{P}_s)`$ is of the form
$$f\left(A\right)=\varphi \left(h\left(A\right)\right)$$
where $`\varphi `$ is a function on $`K.`$ Then there exists a unique holomorphic function $`\mathrm{\Phi }`$ on $`K_{}`$ such that
$$S_{s,\mathrm{}}f\left(C\right)=\mathrm{\Phi }\left(h_{}\left(C\right)\right).$$
The function $`\mathrm{\Phi }`$ is given by
$$\mathrm{\Phi }=\text{analytic continuation }e^{\mathrm{}\mathrm{\Delta }_K/2}\varphi .$$
Note that in light of the definition of $`S_{s,\mathrm{}}`$ , this result says that on the gauge-invariant subspace, $`e^{\mathrm{}\mathrm{\Delta }_𝒜/2}`$ (followed by analytic continuation) reduces to $`e^{\mathrm{}\mathrm{\Delta }_K/2}`$ (followed by analytic continuation). Thus Theorem 4 is a formally equivalent to the principle (1) with which we started.
Now, the gauge-invariant subspace $`L^2(𝒜,P_s)^{𝒢_0}`$ consists of functions of the form $`f\left(A\right)=\varphi \left(h\left(A\right)\right),`$ with $`\varphi `$ a function on $`K.`$ It may be shown that
$$_𝒜\left|\varphi \left(h\left(A\right)\right)\right|^2𝑑P_s\left(A\right)=_K\left|\varphi \left(x\right)\right|^2\rho _s\left(x\right)𝑑x,$$
where $`\rho _s`$ is the heat kernel at the identity on $`K`$ at time $`s.`$ Similarly,
$$_𝒜_{}\left|\mathrm{\Phi }\left(h_{}\left(Z\right)\right)\right|^2𝑑M_{s,\mathrm{}}\left(Z\right)=_K_{}\left|\mathrm{\Phi }\left(g\right)\right|^2\mu _{s,\mathrm{}}\left(g\right)𝑑g,$$
where $`\mu _{s,\mathrm{}}`$ is a suitable heat kernel on $`K_{}`$ and $`dg`$ is Haar measure on $`K_{}.`$ So the gauge-invariant subspace on the real side is identifiable with $`L^2(K,\rho _s\left(x\right)dx)`$ and on the complex side with $`L^2(K_{},\mu _{s,\mathrm{}}\left(g\right)dg).`$ So we have the following commutative diagram in which all maps are unitary.
(8)
$$\begin{array}{ccc}L^2(𝒜,P_s)^{𝒢_0}& \underset{}{e^{\mathrm{}\mathrm{\Delta }_𝒜/2}}& L^2(𝒜_{},M_{s,\mathrm{}})^{𝒢_0}\\ & & \\ L^2(K,\rho _s\left(x\right)dx)& \underset{}{e^{\mathrm{}\mathrm{\Delta }_K/2}}& L^2(K_{},\mu _{s,\mathrm{}}\left(g\right)dg)\end{array}$$
The horizontal maps contain an implicit analytic continuation.
This result embodies a rigorous version of the principle (1) and also shows that a form of the Segal–Bargmann transform for $`𝒜`$ can descend to a Segal–Bargmann transform for $`𝒜/𝒢_0=K.`$ But so far we still have the regularization parameter $`s,`$ which we are supposed to remove by letting it tend to infinity. On the full space $`L^2(𝒜,P_s)`$ or $`L^2(𝒜_{},M_{s,\mathrm{}})`$ the limit $`s\mathrm{}`$ does not exist; this was the point of putting in the $`s`$ in the first place. But on the gauge-invariant subspaces, identified with functions on $`K`$ or $`K_{},`$ the limit does exist. As $`s\mathrm{},`$ the heat kernel measure $`\rho _s`$ on $`K`$ converges to normalized Haar measure on $`K.`$ This confirms our earlier conjecture that the fictitious Lebesgue measure on $`𝒜`$ (formally the $`s\mathrm{}`$ limit of $`P_s`$) pushes forward to the Haar measure on $`K.`$ Meanwhile, the measure $`\mu _{s,\mathrm{}}`$ converges as $`s\mathrm{}`$ to a certain measure I call $`\nu _{\mathrm{}},`$ which coincides with the “$`K`$-averaged heat kernel measure” of \[H1\]. So taking the limit in the bottom line of (8) gives
(9)
$$\begin{array}{ccc}L^2(K,dx)& \underset{}{e^{\mathrm{}\mathrm{\Delta }_K/2}}& L^2(K_{},\mu _{s,\mathrm{}})\end{array}$$
This supports the expected conclusion that our reduced quantum Hilbert space is $`L^2(K,dx)`$ and that the quantum Hamiltonian is $`\left(\mathrm{}^2/2\right)\mathrm{\Delta }_K.`$ It further shows that the generalized Segal–Bargmann transform for $`K,`$ as given in (9), arises naturally from the ordinary Segal–Bargmann transform for the space of connections, upon restriction to the gauge-invariant subspace. The transform in (9) is precisely the $`K`$-invariant form of the transform which was previously constructed in \[H1\] from a purely finite-dimensional point of view.
## 5. Coherent states: from $`𝒜_{}`$ to $`K_{}`$
Let us now reformulate the results of the last section in terms of coherent states. Klauder and Skagerstam \[KS\] think of coherent states as a collection of states $`\psi _\alpha `$ in some Hilbert space, labeled by points some parameter space $`X`$ such that there is a resolution of the identity
(10)
$$I=_X|\psi _\alpha \psi _\alpha |𝑑\nu \left(\alpha \right)$$
for some measure $`\nu `$ on $`X.`$ One may then define a “coherent state transform,” that is, a linear map $`C:HL^2(X,\nu )`$ given by taking the inner product of a vector in $`H`$ with each of the coherent states:
$$C\left(v\right)\left(\alpha \right)=\psi _\alpha |v.$$
The resolution of the identity implies that
$`{\displaystyle _X}\left|\psi _\alpha |v\right|^2𝑑\nu \left(\alpha \right)`$ $`={\displaystyle _X}v|\psi _\alpha \psi _\alpha |v𝑑\nu \left(\alpha \right)`$
$`=v\left|{\displaystyle _X}\right|\psi _\alpha \psi _\alpha \left|d\nu \left(\alpha \right)\right|v`$
$`=v|v.`$
Thus the resolution of the identity (10) is at least formally equivalent to the statement that $`C`$ is an isometric linear map. Note that $`C`$ is only isometric but not unitary; in all the interesting cases the image of $`C`$ is a proper subspace of $`L^2(X,\nu ),`$ which may be characterized by a certain reproducing kernel condition. Although the resolution of the identity looks on the surface like an orthonormal basis expansion, it is in fact quite different. The coherent states are typically non-orthogonal and “overcomplete.” The overcompleteness is reflected in the fact that $`C`$ does not map onto $`L^2(X,\nu ).`$
As an example, consider the finite-dimensional Segal–Bargmann transform, in my normalization. The coherent states are then the states $`\psi _zL^2(^n,dx)`$ given by
$$\psi _z\left(x\right)=\left(2\pi \mathrm{}\right)^{n/2}e^{\left(\overline{z}x\right)^2/2\mathrm{}},z^n.$$
This means that the coherent state transform is given by
$$\left(C_{\mathrm{}}f\right)\left(z\right)=\psi _z|f_{L^2(^n,dx)}=_^n\left(2\pi \mathrm{}\right)^{n/2}e^{\left(zx\right)^2/2\mathrm{}}f\left(x\right)𝑑x,$$
as above. In this case the parameter space $`X`$ is $`^n`$ and the measure on $`X`$ is the measure $`\nu _{\mathrm{}}`$ of the last section. The overcompleteness of the coherent states means here that the image of $`C_{\mathrm{}}`$ is not all of $`L^2(^n,\nu ),`$ but only the holomorphic subspace.
Next consider what happens to a set of coherent states under reduction. Suppose we have a set of coherent states in a Hilbert space $`H,`$ satisfying a resolution of the identity (10). Then suppose that $`V`$ is a closed subspace of $`H`$ and that $`P`$ is the orthogonal projection onto $`V.`$ Since $`P^2=P^{}=P,`$ (10) gives
$$P=PIP=_X|P\psi _\alpha P\psi _\alpha |𝑑\nu \left(\alpha \right).$$
Thus by projecting each coherent state into $`V`$ we get a resolution of the identity (and hence a coherent state transform) for the subspace $`V.`$ Note that at the moment the parameter space for the coherent states, and the measure on it, are unchanged by the projection. However, it may happen that certain sets of distinct coherent states become the same after the projection is applied. In that case we may reduce (or “collapse”) the parameter space $`X`$ by identifying any two parameters $`\alpha `$ and $`\beta `$ for which $`P\psi _\alpha =P\psi _\beta .`$ The measure $`\nu `$ then pushes forward to a measure $`\stackrel{~}{\nu }`$ on the reduced parameter space $`\stackrel{~}{X}.`$
This is what happens in our Yang-Mills case. We have states $`\psi _Z^{(s)}L^2(𝒜,P_s)`$ defined by the condition that
(11)
$$S_{s,\mathrm{}}f\left(z\right)=\psi _Z^{(s)}|f_{L^2(𝒜,P_s)},$$
where now the parameter space, that is, the set of $`Z`$’s, is the space $`𝒜_{}`$ of complexified connections. I suppress the dependence of $`\psi _Z^{(s)}`$ on $`\mathrm{}.`$ We want to project the $`\psi _Z`$’s onto the gauge-invariant subspace, that is, onto the space of functions of the form $`\varphi \left(h\left(A\right)\right).`$ The projection amounts to the same thing as restricting attention in (11) to $`f`$’s of the form $`f\left(A\right)=\varphi \left(h\left(A\right)\right).`$ For such $`f`$’s, Theorem 4 tells us that
$`\psi _Z^{(s)}|f`$ $`=\left[S_{s,\mathrm{}}\left(\varphi h\right)\right]\left(Z\right)`$
$`=\mathrm{\Phi }\left(h_{}\left(Z\right)\right),`$
where $`\mathrm{\Phi }`$ is the analytic continuation to $`K_{}`$ of $`e^{\mathrm{}\mathrm{\Delta }_K/2}\varphi .`$ We see then that for $`f`$ in the invariant subspace, the right side of (11) depends only on the holonomy of $`Z.`$ Thus upon projection into the gauge-invariant subspace the parameter space for the coherent states collapses from $`𝒜_{}`$ to $`𝒜_{}/𝒢_{0,}=K_{}.`$
If we identify the gauge-invariant subspace with $`L^2(K,\rho _s)`$ as in the previous section, then the reduced coherent states are the vectors $`\stackrel{~}{\psi }_g^{(s)}L^2(K,\rho _s),`$ with $`gK_{},`$ given by
$$\stackrel{~}{\psi }_g^{(s)}\left(x\right)=\frac{\overline{\rho _{\mathrm{}}\left(gx^1\right)}}{\rho _s\left(x\right)},gK_{},$$
so that, as required, we have
$`\stackrel{~}{\psi }_g^{(s)}|\varphi _{L^2(K,\rho _s)}`$ $`={\displaystyle _K}{\displaystyle \frac{\rho _{\mathrm{}}\left(gx^1\right)}{\rho _s\left(x\right)}}\varphi \left(x\right)\rho _s\left(x\right)𝑑x`$
$`=\mathrm{\Phi }\left(g\right).`$
Here $`\rho _{\mathrm{}}\left(gx^1\right)`$ refers to the analytic continuation of the heat kernel from $`K`$ to $`K_{},`$ and for $`gK,`$ the convolution $`_K\rho _{\mathrm{}}\left(gx^1\right)\varphi \left(x\right)𝑑x`$ is nothing but $`\left(e^{\mathrm{}\mathrm{\Delta }_K/2}\varphi \right)\left(g\right).`$
The $`\stackrel{~}{\psi }_g^{(s)}`$ satisfy a resolution of the identity with respect to the measure $`\mu _{s,\mathrm{}}`$ on $`K_{}.`$ This measure is the one which is naturally induced from the measure $`M_{s,\mathrm{}}`$ on $`𝒜_{},`$ upon reduction from $`𝒜_{}`$ to $`K_{}.`$ That is, $`\mu _{s,\mathrm{}}`$ is the “push-forward” of $`M_{s,\mathrm{}}`$ from $`𝒜_{}`$ to $`K_{},`$ under the map $`h_{}.`$ Now, as $`s\mathrm{},`$ $`\rho _s\left(x\right)`$ converges to the constant function $`\mathrm{𝟏}`$. Thus we obtain in the limit coherent states $`\stackrel{~}{\psi }_gL^2(K,dx)`$ given by
(12)
$$\stackrel{~}{\psi }_g\left(x\right):=\underset{s\mathrm{}}{lim}\stackrel{~}{\psi }_g^{(s)}\left(x\right)=\overline{\rho _{\mathrm{}}\left(gx^1\right)},gK_{}.$$
These satisfy the following resolution of the identity:
$$I=_K_{}|\stackrel{~}{\psi }_g\stackrel{~}{\psi }_g|𝑑\nu _{\mathrm{}}\left(g\right),$$
where $`\nu _{\mathrm{}}=lim_s\mathrm{}\mu _{s,\mathrm{}}.`$ The measure $`\nu _{\mathrm{}}`$ coincides with the “$`K`$-averaged heat kernel measure” of \[H1\].
Although we are “supposed to” let $`s\mathrm{},`$ we get a well-defined coherent state theory for any $`s>`$ $`\mathrm{}/2.`$ The case $`s=`$ $`\mathrm{}`$, as well as the limiting case $`s\mathrm{},`$ had previously been described in \[H1\]. For other values of $`s`$ we get something new, which I investigate from a finite-dimensional point of view in \[H5\].
Let me compare the above results to those in the paper of Wren \[W\], which motivated Driver and me to develop our paper \[DH1\]. Wren uses the “Rieffel induction” method proposed by Landsman \[L1\], applied to this same problem of Yang-Mills theory on a spacetime cylinder. The commutative case was considered previously by Landsman and Wren in \[LW\]. Wren uses a fixed Gaussian measure and a “unitarized” action of the gauge group. In this approach there is no gauge-invariant subspace (see \[DH2\]) and so an integration over the gauge group is used to define a reduced Hilbert space, which substitutes for the gauge-invariant subspace. Wren shows that the reduced Hilbert space can be identified with $`L^2(K,dx)`$ and further shows that under the reduction map the ordinary coherent states map precisely to the coherent states $`\stackrel{~}{\psi }_g`$ in (12). So the appearance of these coherent states in \[DH1\] was expected on the basis of Wren’s results.
The paper \[DH1\] set out to understand better two issues raised by \[W\]. First, because in \[W\] there is no true gauge-invariant subspace to project onto, the resolution of the identity for the classical coherent states does not survive the reduction. That is, Rieffel induction does not tell you what the right measure is to get a resolution of the identity. Of course, the relevant measure had already been described in \[H1\], but it would be nice not to have to know this ahead of time. By contrast, in our approach the measure $`\nu _{\mathrm{}}`$ arises naturally by pushing forward the Gaussian measure $`M_{s,\mathrm{}}`$ to $`K_{}`$ and then letting $`s`$ tend to infinity. Second, the calculation in \[W\] concerning the reduction of the Hamiltonian is non-rigorous, mainly because the unconstrained Hamiltonian is not well-defined. Driver and I used the Segal–Bargmann transform in order to get some form of the Hamiltonian $`\mathrm{\Delta }_𝒜`$ to make rigorous sense.
Finally, let me mention that the generalized coherent states on $`K`$ are do not fall into the framework of Perelomov \[P\], because there does not seem to be in the compact group case anything analogous to the irreducible unitary representation of the Heisenberg group on $`L^2\left(^n\right)`$.
## 6. Identification of $`T^{}\left(K\right)`$ with $`K_{}`$
I am thinking of $`K`$ as the configuration space for the reduced classical Yang-Mills theory, and of $`K_{}`$ as the corresponding phase space. For this to be sensible, there should be an identification of $`K_{}`$ with the standard phase space over $`K,`$ namely the cotangent bundle $`T^{}\left(K\right).`$ In this section I will explain how such an identification comes out of the reduction process. The resulting identification coincides with the one described in \[H3, H4\] from an intrinsic point of view.
So let us see what comes out of the reduction process. From the symplectic point of view we have the Marsden–Weinstein symplectic quotient $`J^1\left(0\right)/𝒢_0.`$ Since the action of $`𝒢_0`$ on $`𝒜_{}=T^{}\left(𝒜\right)`$ arises from an action of $`𝒢_0`$ on the configuration space $`𝒜,`$ general principles tell us that $`J^1\left(0\right)/𝒢_0`$ coincides with $`T^{}\left(𝒜/𝒢_0\right)=T^{}\left(K\right).`$ On the other hand, from the complex point of view we may analytically continue the action of $`𝒢_0`$ on $`𝒜_{}`$ to get an action of $`𝒢_{0,}`$ on $`𝒜_{}.`$ Dividing out by this action gives $`𝒜_{}/𝒢_{0,}=K_{}.`$ But in this case there is a natural identification of $`J^1\left(0\right)/𝒢_0`$ with $`𝒜_{}/𝒢_{0,}`$: each orbit of $`𝒢_{0,}`$ intersects $`J^1\left(0\right)`$ in precisely one $`𝒢_0`$-orbit. This may be seen from \[L2\].
This result is not a coincidence. In general, given a Kähler manifold $`M`$ (in our example $`𝒜_{}`$) and an action of a group $`G`$ that preserves both the complex and symplectic structure of $`M,`$ we may analytically continue to get an action of $`G_{}`$ on $`M,`$ an action which preserves the complex but not the symplectic structure of $`M.`$ Then if $`J^1\left(0\right)`$ is the moment mapping for the action of $`G,`$ one expects that
(13)
$$J^1\left(0\right)/G=M/G_{}.$$
This would mean that for each orbit $`O`$ of $`G_{}`$ in $`M`$ the intersection of $`O`$ with $`J^1\left(0\right)`$ is precisely a single $`G`$-orbit. Now, (13) is not actually true in general, but only with various provisos and qualifications. (See \[Ki, MFK\].) Still, this is an important idea and in our case it works out exactly.
So putting everything together we have the following identifications.
$$\begin{array}{ccccc}𝒜_{}/𝒢_{0,}& =& J^1\left(0\right)/𝒢_0& =& T^{}\left(𝒜/𝒢_0\right)\\ & & & & \\ K_{}& & & & T^{}\left(K\right)\end{array}$$
If one does the calculations, one obtains the following explicit identification of $`T^{}\left(K\right)`$ with $`K_{}.`$ First, use left-translation to trivialize the cotangent bundle, so that $`T^{}\left(K\right)K\times 𝔨^{}.`$ Then use the inner product on $`𝔨`$ to identify $`K\times 𝔨^{}`$ with $`K\times 𝔨.`$ Finally map from $`K\times 𝔨`$ to $`K_{}`$ by the map
(14)
$$\mathrm{\Phi }(x,Y)=xe^{iY},xK,Y𝔨.$$
The map $`\mathrm{\Phi }`$ is a diffeomorphism of $`K\times 𝔨`$ onto $`K_{},`$ and $`\mathrm{\Phi }^1`$ is called the polar decomposition of $`K_{}.`$
For example, suppose $`K=SU(n)`$ so that $`K_{}=SL(n;).`$ Then given $`gSL(n;)`$ we may use the standard polar decomposition for matrices to write
$$g=xp$$
with $`x`$ unitary and $`p`$ positive. Since $`detg=1`$ it follows that $`detx=detp=1.`$ Then $`p`$ has a unique self-adjoint logarithm $`\xi `$, which will have trace zero. Letting $`Y=\xi /i`$ we have
$$g=xe^{iY}$$
with $`Y`$ skew and trace zero, so $`Ysu\left(n\right).`$
Now in \[H3\] (see also \[H4\]) I argued from an intrinsic, finite-dimensional point of view that the above identification of $`T^{}\left(K\right)`$ with $`K_{}`$ was natural. The argument was based on the notion of “adapted complex structures.” There is a good reason that the reduction argument gives the same identification as the adapted complex structures do. Suppose $`X`$ is a finite-dimensional compact Riemannian manifold such that $`T^{}\left(X\right)`$ has a global adapted complex structure, and suppose $`G`$ is a compact Lie group which acts freely and isometrically on $`X.`$ Then a result of R. Aguilar \[A\] says that $`T^{}\left(X/G\right)`$ has a global adapted complex structure and that this complex structure coincides with the one inherited from $`T^{}\left(X\right)`$ by means of reduction. We have the same sort of situation here, with $`X=𝒜`$ and $`G=𝒢_0.`$ Of course, $`𝒢_0`$ is not compact and $`𝒜`$ is neither comact nor finite-dimensional, but nevertheless what happens is reasonable in light of Aguilar’s result.
## 7. Reduction of the Laplacian
Why should $`\mathrm{\Delta }_𝒜`$ correspond to $`\mathrm{\Delta }_K`$ on gauge-invariant functions? Let us strip away the infinite-dimensional technicalities and consider the analogous question in finitely many dimensions. Suppose $`X`$ is a finite-dimensional connected Riemannian manifold and suppose $`G`$ is a Lie group that acts by isometries on $`X.`$ For simplicity I will assume that $`G`$ is compact and that $`G`$ acts freely on $`X.`$ Then $`X/G`$ is again a manifold, which has a unique Riemannian metric such that the quotient map $`q:XX/G`$ is a Riemannian submersion. This means that at each point $`xX`$, the differential of $`q`$ is an isometry when restricted to the orthogonal complement of the tangent space to the $`G`$-orbit through $`x.`$
Given this metric on $`X/G`$ we may consider the Laplace-Beltrami operator $`\mathrm{\Delta }_{X/G}`$. For a smooth function $`f`$ on $`X/G`$ we may ask whether $`\left(\mathrm{\Delta }_{X/G}f\right)q`$ coincides with $`\mathrm{\Delta }_X\left(fq\right).`$ This amounts to asking whether $`\mathrm{\Delta }_X`$ and $`\mathrm{\Delta }_{X/G}`$ agree on the $`G`$-invariant subspace of $`C^{\mathrm{}}\left(X\right).`$ Since $`\mathrm{\Delta }_X`$ commutes with isometries, it will at least preserve the $`G`$-invariant subspace.
The answer in general is no, $`\mathrm{\Delta }_X`$ and $`\mathrm{\Delta }_{X/G}`$ do not agree on $`C^{\mathrm{}}\left(X\right)^G.`$ For example, consider $`SO\left(2\right)`$ acting on $`^2\left\{0\right\}`$ by rotations. The quotient manifold is diffeomorphic to $`(0,\mathrm{}),`$ with the point $`r(0,\mathrm{})`$ corresponding to the orbit $`x^2+y^2=r^2`$ in $`^2\left\{0\right\}.`$ The induced metric on $`(0,\mathrm{})`$ is the usual metric on $`(0,\mathrm{})`$ as a subset of $`.`$ So the Laplace-Beltrami operator on $`(0,\mathrm{})`$ is just $`d^2/dr^2.`$ On the other hand, the formula for the two-dimensional Laplacian on radial functions is
(15)
$$\left(\frac{^2}{x^2}+\frac{^2}{y^2}\right)f\left(\sqrt{x^2+y^2}\right)=\left[\frac{d^2f\left(r\right)}{dr^2}+\frac{1}{r}\frac{df\left(r\right)}{dr}\right]|_{r=\sqrt{x^2+y^2}}.$$
The source of the trouble is the discrepancy between the intrinsic volume measure $`dr`$ on $`(0,\mathrm{})`$ and the push-forward of the volume measure from $`^2\left\{0\right\},`$ which is $`2\pi rdr.`$
In general, each $`G`$-orbit in $`X`$ inherits a natural Riemannian metric from $`X`$, and we may compute the total volume of this orbit with respect to the associated Riemannian volume measure. The function $`\mathrm{Vol}\left(Gx\right)`$ on $`X/G`$ measures the discrepancy between the intrinsic volume measure on $`X/G`$ and the push-forward of the volume measure on $`X.`$ The two Laplacians on $`C^{\mathrm{}}\left(X\right)^G`$ will be related by the formula
(16)
$$\mathrm{\Delta }_X=\mathrm{\Delta }_{X/G}+\left(\mathrm{log}\mathrm{Vol}\left(Gx\right)\right).$$
(The gradient may be thought of as that for $`X/G`$, although this coincides in a natural sense with that for $`X,`$ on $`G`$-invariant functions.) Formula (15) is a special case of (16) with volume factor $`2\pi r.`$ So the two Laplacians agree if and only if the $`G`$-orbits all have the same volume.
Let us return, then, to the case of $`𝒜/𝒢_0.`$ By considering the appendix of \[DH1\] it is easily seen that the metric on $`K`$ that makes the map $`h:𝒜K`$ a Riemannian submersion is simply the bi-invariant metric on $`K`$ induced by the inner product on $`𝔨.`$ (We use on $`𝒜`$ the metric coming from the $`L^2`$ norm as in Section 2.) So in light of (16) the statement that $`\mathrm{\Delta }_𝒜`$ and $`\mathrm{\Delta }_K`$ agree on the $`𝒢_0`$-invariant subspace is formally equivalent to the statement that all the $`𝒢_0`$-orbits have the same volume. But again from \[DH1\] it may be seen that there exist isometries of $`𝒜`$ that map any $`𝒢_0`$-orbit to any other, so formally all should have the same volume.
To look at it another way, we need to see that the (fictitious) volume measure $`𝒟A`$ on $`𝒜`$ pushes forward to a multiple of the Haar measure on $`K.`$ Accepting $`P_s`$ as an approximation to $`𝒟A,`$ this pushes forward to the measure $`\rho _s\left(x\right)dx,`$ which indeed converges to $`dx`$ as $`s`$ tends to infinity. It should be noted, however, that nothing so simple is likely to happen in higher-dimensional Yang-Mills theory. See for example \[Ga\].
## 8. Does quantization commute with reduction?
When quantizing a reduced Hamiltonian system such as Yang-Mills theory, one may ask whether the quantization should be done before or after the reduction. If we were very optimistic, we might hope that it doesn’t matter, that one gets the same answer either way. If this were so, we could say that quantization commutes with reduction. Of course the question of whether quantization commutes with reduction may well depend on the system being quantized and on how one interprets the question. I want to consider this question from the point of view of geometric quantization and I want specifically to compare the Segal–Bargmann space obtained by first quantizing $`𝒜_{}`$ and then reducing by $`𝒢_0`$ to the one obtained by directly quantizing $`K_{}.`$
In geometric quantization \[Wo\] one begins with a symplectic manifold $`(M,\omega )`$ and constructs over $`M`$ a Hermitian complex line bundle $`L`$ with connection, whose curvature form is $`i\omega /\mathrm{}.`$ If $`M`$ is a cotangent bundle then such a bundle exists and may be taken to be topologically and Hermitianly trivial (though the connection is necessarily non-trivial). The “prequantum Hilbert space” is then the space of sections of $`L`$ which are square-integrable with respect to the symplectic volume measure on $`M.`$ To obtain the “quantum Hilbert space” one picks a “polarization” and restricts to the space of square-integrable polarized sections of $`L.`$ If $`M`$ is a Kähler manifold, i.e. it has a complex structure which is compatible in a natural sense with $`\omega ,`$ then there is a natural Kähler polarization. In that case $`L`$ may be given the structure of a holomorphic line bundle and the quantum Hilbert space becomes the space of square-integrable holomorphic sections of $`L.`$
In the case $`M=^n`$ the resulting bundle is holomorphically trivial. So by choosing a nowhere vanishing holomorphic section, the space of holomorphic sections of $`L`$ may be identified with the space of holomorphic functions on $`^n.`$ This nowhere vanishing section will not, however, have constant norm. This means that the inner product on the space of holomorphic functions will be an $`L^2`$ inner product with respect to a measure which is Lebesgue measure times the norm-squared of the trivializing section. Working this out we get simply the Segal–Bargmann space, with different normalizations of the space coming from different possible choices of the trivializing section. The construction depends on Planck’s constant $`\mathrm{}.`$ In summary: applying geometric quantization to $`^n,`$ using a Kähler polarization, yields the Segal–Bargmann space.
To apply geometric quantization to the infinite-dimensional space $`𝒜_{}`$ we may try to quantize $`^n`$ and then let $`n`$ tend to infinity. For this to make sense with my normalization, we need to add the additional parameter $`s.`$ So we obtain the Segal–Bargmann space $`L^2(𝒜_{},M_{s,\mathrm{}}).`$ We then want to reduce by the action of $`𝒢_0,`$ which amounts to restricting to the space of functions in $`L^2(𝒜_{},M_{s,\mathrm{}})`$ which are $`𝒢_0`$-invariant, and thus by analyticity, $`𝒢_{0,}`$-invariant. The resulting space is identifiable with $`L^2(K_{},\mu _{s,\mathrm{}}).`$ Finally, letting $`s`$ tend to infinity we obtain $`L^2(K_{},\nu _{\mathrm{}}).`$ It is therefore reasonable to say that $`L^2(K_{},\nu _{\mathrm{}})`$ is the space obtained by quantizing $`𝒜_{}`$ and then reducing by $`𝒢_0.`$
Meanwhile, we may apply geometric quantization directly to $`K_{}.`$ I do this calculation in \[H4\] and find that geometric quantization yields the space $`L^2(K_{},\gamma _{\mathrm{}}),`$ where $`\gamma _{\mathrm{}}`$ and $`\nu _{\mathrm{}}`$ are related by the formula
(17)
$$d\nu _{\mathrm{}}\left(g\right)=a_{\mathrm{}}u\left(g\right)d\gamma _{\mathrm{}}\left(g\right).$$
Here $`a_{\mathrm{}}`$ is an irrelevant constant and $`u`$ is a function which is non-constant except when $`K`$ is commutative. So it seems that quantizing $`K_{}`$ directly does not yield the same answer. However, this is not the end of the story. One can quantize $`K_{}`$ taking into account the “half-form correction” (also known as the “metaplectic correction”). This “corrected” quantization yields an extra factor in the measure, a factor that coincides precisely with the factor $`u\left(g\right)`$ in (17)! On the other hand, in the $`^n`$ case the half-form correction does not affect the answer, so even with the half-form correction we would get $`L^2(𝒜_{},M_{s,\mathrm{}})`$ and then ultimately $`L^2(K_{},\nu _{\mathrm{}}).`$ So our conclusion is the following: In this example, if we use geometric quantization with a Kähler polarization and the half-form correction, quantization does in fact commute with reduction.
Let me conclude by mentioning a related setting in which one can ask whether quantization commutes with reduction. In an influential paper \[GS\], Guillemin and Sternberg consider the geometric quantization of a compact Kähler manifold $`M.`$ They assume then that there is an action of a compact group $`G`$ on $`M`$ that preserves the complex structure and the symplectic structure of $`M`$ and they consider as well the Marsden–Weinstein quotient $`M^G:=J^1\left(0\right)/G,`$ where $`J`$ is the moment mapping for the action of $`G.`$ Under certain conditions they show that there is a natural invertible linear map between on the one hand the $`G`$-invariant subspace of the quantum Hilbert space over $`M`$ and on the other hand the quantum Hilbert space over $`M^G.`$ They interpret this result as a form of quantization commuting with reduction.
However, Guillemin and Sternberg do not show that this invertible linear map is unitary, and indeed there seems to be no reason that it should be in general. So in their setting we may say that quantization fails to commute unitarily with reduction. Dan Freed \[F\] has suggested to me that inclusion of the half-form correction in the quantization might the map unitary, and indeed our Yang-Mills example seems to confirm this. (It was Freed’s suggestion that led me to work out that $`u`$ is just the half-form correction.) After all, upon inclusion of the half-form correction we get the same measure (except for an irrelevant overall constant) and therefore the same inner product whether quantizing before or after the reduction. Nevertheless, I do not believe that one will get a unitary correspondence in general. So we are left with the following open question.
> Given a Kähler manifold $`M`$ with an action of a group $`G,`$ under what conditions on $`M`$ and $`G`$ will quantization commute unitarily with reduction?
Although the question may be considered with or without the half-form correction, what little evidence there is so far suggests that the answer is more likely to be yes if the half-form correction is included.
## 9. Notes
Section 2. One should say something about the degree of smoothness assumed on the connections and gauge transformations. Although it does not matter so much at the classical level, it seems natural to take the space of connections to be the Hilbert space of square-integrable connections. This amounts to completing $`𝒜`$ with respect to the natural norm, the one which appears in the formula for the classical Hamiltonian. We may then take the gauge group to be the largest group whose action on $`𝒜`$ makes sense. This is the group of “finite energy” gauge transformations, namely, the ones for which $`g^1dg`$ is finite. It is easily shown that in our example of a spatial circle, two square-integrable connections are related by a finite energy gauge (based) gauge transformation if and only if they have the same holonomy. In the quantized theory we will be forced to consider a larger space of connections.
Section 3. The measure $`P_s`$ is a Gaussian measure, about which there is an extensive theory. For example, see \[G1, K, GJ\]. The distinctive feature of Gaussian measures on infinite-dimensional spaces is the presence of two different spaces, a Hilbert space $`H`$ whose norm enters the formal expression for the measure, and a larger topological vector space $`B`$ on which the measure lives. Although one should think of the Gaussian measure as being canonically associated to $`H,`$ the measure lives on $`B,`$ and $`H`$ is a measure-zero subspace. In our example $`H`$ is the space of square-integrable connections and $`B`$ is a suitable space of distributional connections. Since the elements of $`B`$ are highly non-smooth, the holonomy must be defined as the solution of a stochastic differential equation.
If one glosses over questions of smoothness, the Gross ergodicity theorem \[G2\] sounds as if it ought to be trivial. But we have just said that we must enlarge the space of connections in order for the measure $`P_s`$ to exist. Unfortunately, we may not correspondingly enlarge the gauge group without losing the quasi-invariance of the measure $`P_s`$ under the action of $`𝒢_0,`$ without which the definition of $`L^2(𝒜,P_s)^{𝒢_0}`$ does not make sense. So we end up unavoidably in a situation in which two connections with the same holonomy are not necessarily $`𝒢_0`$-equivalent, because the would-be gauge transformation is not smooth enough to be in $`𝒢_0.`$ It was the “J-perp” theorem, which arose as a corollary of Gross’s proof of the ergodicity theorem, which led him to suggest to me to look for an analog of the Segal–Bargmann transform on $`K.`$
Section 4. Driver and I define the holomorphic subspace of $`L^2(𝒜_{},M_{s,\mathrm{}})`$ to be the $`L^2`$ closure of the space of holomorphic cylinder functions. An important question then is whether a function of the form $`F\left(Z\right)=\mathrm{\Phi }\left(h_{}\left(Z\right)\right),`$ with $`\mathrm{\Phi }L^2(K_{},\mu _{s,\mathrm{}}),`$ is in this holomorphic subspace. The answer is yes, but the proof that we give is indirect.
I am defining $`L^2(𝒜_{},M_{s,\mathrm{}})^{𝒢_0}`$ to be the image of $`L^2(𝒜,P_s)^{𝒢_0}`$ under the Segal–Bargmann transform. Certainly every element of $`L^2(𝒜_{},M_{s,\mathrm{}})^{𝒢_0}`$ is actual invariant under the action of $`𝒢_0`$ on $`𝒜_{}.`$ The converse is probably true as well, namely that every element of $`L^2(𝒜_{},M_{s,\mathrm{}})`$ which is $`𝒢_0`$-invariant is in the image of $`L^2(𝒜,P_s)^{𝒢_0},`$ but we have not proved this.
Section 5. Except when $`s=`$ $`\mathrm{}`$ the coherent states $`\psi _Z^{(s)}`$ in $`L^2(𝒜,P_s)`$ are non-normalizable states. When $`s=`$ $`\mathrm{},`$ the coherent states $`\psi _Z^{(s)}`$ are normalizable states provided that $`Z`$ is a square-integrable (complex) connection \[HS, Sect. 2.3\]. But even then the measure $`M_\mathrm{},\mathrm{}`$ does not live on the space of square-integrable connections, and so it is a bit delicate to formulate the resolution of the identity. This shows that it is technically easier to formulate things in terms of the Segal–Bargmann transform instead of the coherent states. Nevertheless, we may think continue to think of unitarity for the Segal–Bargmann transform as formally equivalent to a resolution of the identity for the coherent states.
Section 6. There are several obstructions to (13) holding in general. One needs some condition to guarantee that the analytic continuation of the $`G`$-action exists globally. Even when it does, one needs to worry about the possibility of “unstable points,” that is points whose $`G_{}`$-orbit does not intersect the zero set of the moment mapping, and about the possibility that the $`G_{}`$-orbits may not be closed. In the case of a cotangent bundle of a compact Riemannian manifold whose cotangent bundle admits a global adapted complex structure, none of these problems actually arises. See \[A, Sect. 7\].
Section 8. I jumping to conclusions about the correct action of the gauge group $`𝒢_0`$ on the Segal–Bargmann space $`L^2(𝒜_{},M_{s,\mathrm{}}).`$ One should properly use geometric quantization to determine this action. To do this, we restrict first to the finite-dimensional space $`L^2(^n,\nu _{\mathrm{}})`$ and then consider the action of the group of rotations and translations of $`^n`$ on this space. Going through the calculations, on finds that with my normalization these rotations and translations act in the obvious way, namely, by composing a function in $`L^2(^n,\nu _{\mathrm{}})`$ with the rotation or translation. Note that this holds only for rotations and translations in the $`x`$-directions. Now we have said that the action of $`𝒢_0`$ on $`𝒜`$ consists just of a rotation and a translation. So, taking $`L^2(𝒜_{},M_{s,\mathrm{}})`$ as the best approximation of $`L^2(^n,\nu _{\mathrm{}})`$ when $`n=\mathrm{},`$ it is reasonable to say that the action of $`𝒢_0`$ on $`L^2(𝒜_{},M_{s,\mathrm{}})`$ should be just $`F\left(Z\right)F\left(g^1Z\right).`$
There is a large body of work extending the results of \[GS\]; see for example the survey article of Sjamaar \[Sj\]. |
no-problem/9911/cond-mat9911139.html | ar5iv | text | # Ground-state energy of a high-density electron gas in a strong magnetic field
## 1 INTRODUCTION
The interacting electron gas is a widely used model in condensed matter, chemistry, and astrophysics and has been studied in great detail. Many investigations have focussed on the calculation of the ground-state energy. Gell-Mann and Brueckner derived an exact formula for the ground-state energy of a high-density electron gas. Their method will also be employed in this work.
In contrast to the field-free case, the electron gas in an externally applied magnetic field has received much less attention. So far most of the works have concentrated on studying the ground-state properties in random-phase approximation .
Clearly, it is of general interest to find analytical expressions for at least some limiting cases. In doing so, we concentrate in our calculation on the strong-field limit in which only the lowest Landau level is assumed to be occupied and all electrons are aligned antiparallel to the magnetic field. In this case the energy associated with a Landau level exceeds the Fermi energy. This regime has been investigated by Horing et al. , Isihara and Tsai , Keldysh and Onishchenko , and recently by Skudlarski and Vignale , and by Steinberg and Ortner . The latter have obtained an asymptotically exact result by deriving an analogous expansion of the ground state energy as it was established by Gell-Mann Brueckner. Within this calculation the result derived by Horing et al. was found to be the leading term in the expansion. Keldysh et al. considered the limiting case of a small filling factor of the lowest Landau level. Additionally, they performed an expansion of the RPA correlation energy in inverse powers of the Wigner-Seitz radius $`r_s`$ and derived an result that becomes exact in the limit of small densities, i.e., $`r_s>1`$. Strictly speaking, the RPA is not valid in this regime. However, at metallic densities the RPA is considered to be a reasonable approximation of the correlation energy.
Physically, the strong-field limit may be achieved in laboratory plasmas, such as semiconductors, but one also expects these conditions on the surface of neutron stars. Strong magnetic fields are predicted to be observable in laser induced plasmas . The results for the ground-state energy in the high density approximation may also be used as a starting point for the construction of the local-density approximation for the density-functional theory.
Unlike in the field-free electron gas, the Fermi energy is no longer a function of the Wigner-Seitz radius $`r_s`$ alone, but also of the magnetic field. Therefore, it is convenient to introduce the new parameters
$$r_B=\frac{1}{\pi a_Bk_F}=\frac{2}{3\pi ^2}\alpha ^2r_s^3,t=\frac{ϵ_F}{\mathrm{}\omega _c}=\frac{9\pi ^2}{8}\frac{1}{\alpha ^6r_s^6},$$
(1)
where $`\omega _c=eB/m`$ is the cyclotron frequency, $`\alpha =a_B/l_B`$ is the ratio of the Bohr radius and the magnetic length $`l_B=\sqrt{\mathrm{}/(eB)}`$. At strong magnetic fields $`r_B`$ takes the role of the expansion parameter. The second parameter $`t`$ may be regarded as a filling parameter. In the quantum strong-field limit it satisfies the condition $`t1`$. Now the ground-state energy (per Rydberg and electron) of a high-density electron gas in the quantum strong-field limit is of the following form
$$ϵ_g=\frac{1}{3\pi ^2}\frac{1}{r_B^2}+\frac{A(t)}{r_B}+B(t)\mathrm{ln}(r_B)+C(t)+\mathrm{terms}\mathrm{that}\mathrm{vanish}\mathrm{as}\mathrm{r}_\mathrm{B}0.$$
(2)
This gives the asymptotically exact ground-state energy in the limit $`r_B0`$. The coefficients $`A(t)`$, $`B(t)`$, and $`C(t)`$ in this expansion depend on the filling factor $`t`$. The leading term in (2) is the kinetic energy, while the second term describes the first-order exchange effects. The sum of all remaining contributions is known as the correlation energy. Only the RPA correlation energy contributes to $`B(t)`$ in the high-density limit, i.e. $`r_B0`$. To obtain $`C(t)`$, we must calculate the second-order exchange diagram in addition to the sum of all ring diagramms.
In an earlier work the authors have given analytic expressions for the constants $`A(t)`$, $`B(t)`$, and $`C(t)`$ in the limit of small filling factors $`t0`$. In this regime all electrons occupy the lowest Landau level and transitions to higher levels are negelected. We will now generalize these results and consider the case of arbitrary filling factors by taking into account transition to higher levels. With that, one can improve the quality of the convergence of the series. The validity domain of this expansion in the $`r_s`$-$`t`$ plane is schematically shown in Figure 1.
## 2 HARTREE-FOCK AND CORRELATION ENERGY
The dominant interaction contribution to the internal energy of a high-density electron gas comes from the first-order exchange term. It was first calculated by Danz and Glasser in the quantum strong-field limit. They derived an integral representation for $`A(t)`$, which may be expressed in terms of Meijer’s $`G`$-function. By extending their calculation we have found the following series
$$A(t)=\frac{1}{\pi ^2}\left(3𝐂\mathrm{ln}(4t)\right)\frac{1}{\pi ^2}\underset{s=1}{\overset{\mathrm{}}{}}\frac{(4t)^s}{s!(2s+1)(s+1)}\left(\psi (s+1)\mathrm{ln}(4t)+\frac{4s+3}{(2s+1)(s+1)}\right),$$
(3)
where $`𝐂`$ is Euler’s constant $`𝐂0.5772`$. The first term determines the asymptotic behavior for small filling factors.
The starting point for the calculation of the correlation energy of a high-density electron gas is the determination of the polarization function in the random-phase approximation. An explicit expression for the polarisation function in the quantum strong-field limit was derived by Horing et al.. Using these results and following the original work of Gell-Mann and Brueckner, we expand the polarization function in powers of the momentum transfer. In order to avoid divergencies, we introduce a cut-off in the following momentum integration. However, the final result will be independent of this cut-off procedure. After performing all momentum integrations, we find the exact result for the logarithmic contribution
$$B(t)=\frac{1}{16\pi ^2t}.$$
(4)
Furthermore, we can extend this procedure to obtain the next term in this series, which will be independent of $`r_B`$. The constant of this expansion may be split into contributions coming from the ring approximation and from the exchange term according to $`C(t)=C^{RPA}(t)+C^{ex}(t)`$. In the following, we only consider the RPA contribtuion to $`C(t)`$, since even the numerical evaluation of $`C^{ex}(t)`$ is very complicated. In Ref., we have analytically found an expression for $`C^{RPA}(t)`$ at small values of $`t`$. Additionally, we have derived an integral representation for $`C^{RPA}(t)`$ at any filling factor, which we have now numerically integrated. The final expression for $`C^{RPA}(t)`$ may be obtained from a numerical fit to the results and is given by
$$C^{RPA}(t)=0.00633\frac{\mathrm{ln}(t)}{t}\frac{0.02706}{t}0.16666\mathrm{ln}(t)0.59661+at^b+ct^{1/2},$$
(5)
with a=-2.23106, b=0.5677 and c=2.77534. The first three terms were analytically established in and give the exact asymptotics at $`t0`$. The remaining contributions are fitted in such a way that the total correlation energy, including the logarithmic term, is accurate within $`1\%`$.
In Figure 2 we have plotted the ground-state energy in the Hartree-Fock and the high-density approximation (2). For comparison we have included the numerical results for the full RPA ground-state energy. At lower densities ($`r_s0.52`$ at $`\alpha =4.673`$) the high-density ground-state energy deviates from the full RPA ground-state energy indicating the breakdown of the high-density expansion. At these densities the correlation energy should be calculated from an improved polarisation function including local-field corrections .
## 3 Acknowledgments
This work was supported by the Deutsche Forschungsgemeinschaft under grant#Eb 126/5-1. We thank Werner Ebeling for useful discussions. |
no-problem/9911/cond-mat9911110.html | ar5iv | text | # Why Effective Medium Theory Fails in Granular Materials
## Abstract
Experimentally it is known that the bulk modulus, $`K`$, and shear modulus, $`\mu `$, of a granular assembly of elastic spheres increase with pressure, $`p`$, faster than the $`p^{1/3}`$ law predicted by effective medium theory (EMT) based on Hertz-Mindlin contact forces. Further, the ratio $`K/\mu `$ is found to be roughly pressure independent but the measured values are considerably larger than the EMT predictions. To understand the origin of these discrepancies, we have undertaken numerical simulations of a granular assembly of spherical elastic grains. Our results for $`K(p)`$ and $`\mu (p)`$ are in good agreement with the existing experimental data. We show, also, that EMT can describe their pressure dependence if one takes into account the fact that the number of grain-grain contacts increases with $`p`$. Most important, the affine assumption (which underlies EMT), is found to be valid for $`K(p)`$ but to breakdown seriously for $`\mu (p)`$. This explains why the experimental and numerical values of $`\mu (p)`$ are much smaller than the EMT predictions.
PACS: 81.06.Rm, 81.40.Jj
The study of sound propagation and nonlinear elasticity in unconsolidated granular matter is a topic of great current interest . In the simplest experiments, a packing of glass beads is confined under hydrostatic conditions and the compressional and shear sound speeds, $`v_p`$ and $`v_s`$, are measured as functions of pressure, $`p`$ . In the long-wavelength limit, the sound speeds are related to the elastic constants of the aggregate: $`v_p=\sqrt{(K+4/3\mu )/\rho ^{}}`$ and $`v_s=\sqrt{\mu /\rho ^{}}`$, where $`\rho ^{}`$ is the system’s density. In a recent Letter acoustic measurements were made on bead packs under uniaxial stress and it was suggested that long wavelength compresional waves can be described in terms of an effective medium. Thus, it would of great value to have a reliable EMT to describe sound propagation as a function of applied stress. However, our analysis, together with the work of others, raises serious question about the validity of the generally accepted theoretical formulation. The EMT predicts that $`K`$ and $`\mu `$ both vary as $`p^{1/3}`$, and that the ratio $`K/\mu `$ is a constant (independent of pressure and coordination number) dependent only on the Poisson’s ratio of the material from which the individual grains are made.
Experimentally (see Fig. 1), the bulk and shear moduli increase more rapidly than $`p^{1/3}`$ and the values of $`K/\mu `$ are considerably larger than the EMT prediction. These discrepancies between theory and experiment could be due to the breakdown of the Hertz-Mindlin force law at each grain contact as proposed in for the case of metallic beads with an oxide layer, and in for grains with sharp angularities. Alternatively, they could be associated with the breakdown of some of the assumptions underlying the EMT, for example, that the number of contacts per grain is pressure independent, which may not be the case as several authors have suggested .
In this Letter we report calculations of $`K(p)`$ and $`\mu (p)`$ based on granular dynamics (GD) simulations using the Discrete Element Method developed by Cundall and Strack . Here it is assumed at the outset that one has an assembly of spherical soft grains which interact via the Hertz-Mindlin force laws. We find good agreement with the existing experimental data, thus confirming the validity of Hertz-Mindlin contact theory to glass bead aggregates. Further, we can explain the two problems with EMT described above. First, if the calculated increase of the average coordination number with $`p`$ is taken into account, the modified EMT gives an accurate description of the bulk modulus found in the simulations, $`K(p)`$; for $`\mu (p)`$ we obtain a curve whose shape is in good agreement with the simulation data but whose values are seriously offset therefrom. Second, the EMT makes the affine assumption in which the motion of each grain is specified simply in terms of the externally applied strain (see below). We show that while the affine assumption is approximately valid for the bulk modulus, it is seriously in error for the shear modulus; this is why the EMT prediction of $`K/\mu `$ differs significantly from the experimental value.
Numerical Simulations. At the microscopic level the grains interact with one another via (1) non-linear Hertz normal forces and (2) friction generated transverse forces. The normal force, $`f_n`$, has the typical 3/2 power law dependence on the overlap between two spheres in contact, while the transverse force, $`f_t`$ depends on both the shear and normal displacements between the spheres . For two spherical grains with radii $`R_1`$ and $`R_2`$:
$$f_n=\frac{2}{3}C_nR^{1/2}w^{3/2},$$
(2)
$$\mathrm{\Delta }f_t=C_t(Rw)^{1/2}\mathrm{\Delta }s.$$
(3)
Here $`R=2R_1R_2/(R_1+R_2)`$, the normal overlap is $`w=(1/2)[(R_1+R_2)|\stackrel{}{x}_1\stackrel{}{x}_2|]>0`$, and $`\stackrel{}{x}_1`$, $`\stackrel{}{x}_2`$ are the positions of the grain centers. The normal force acts only in compression, $`f_n=0`$ when $`w<0`$. The variable $`s`$ is defined such that the relative shear displacement between the two grain centers is $`2s`$. The prefactors $`C_n=4G/(1\nu )`$ and $`C_t=8G/(2\nu )`$ are defined in terms of the shear modulus $`G`$ and the Poisson’s ratio $`\nu `$ of the material from which the grains are made. In our simulations we set $`G=29`$ GPa and $`\nu =0.2`$. We assume a distribution of grain radii in which $`R_1=0.105`$ mm for half the grains and $`R_2=0.095`$ mm for the other half. Our results are, in fact, quite insensitive to the choice of distribution. We also include a viscous damping term to allow the system to relax toward static equilibrium .
Our calculations begin with a numerical protocol designed to mimic the experimental procedure used to prepare dense packed granular materials. In the experiments the initial bead pack is subjected to mechanical tapping and ultrasonic vibration in order to increase the solid phase volume fraction, $`\varphi _s`$. The simulations begin with a gas of 10000 spherical particles located at random positions in a periodically repeated cubic unit cell approximately 4 mm on a side. At the outset, the transverse force between the grains is turned off ($`C_t=0`$). The system is them compressed slowly until a specified value of $`\varphi _s`$ is attained (see dashed lines in Fig. 2). The compression is then stopped and the grains are allowed to relax. If the compression is stopped before reaching the critical volume fraction, $`\varphi _s0.64`$, corresponding to random close packing (RCP), the system will relax to zero pressure and zero coordination number, since the system cannot equilibrate below RCP. The compression is then continued to a point above the critical packing fraction and the target pressure is maintained with a “servo” mechanism which constantly adjusts the applied strain until the system reaches equilibrium. Because there are no transverse forces, the grains slide without resistance during the relaxation process and the system reaches the high volume fractions found experimentally.
The simulated granular aggregate relaxes to equilibrium states in which the average coordination number, $`<Z(p)>`$, increases with pressure as seen in Fig. 2. For low pressures compared with the shear modulus of the beads $`<Z>6`$, while in two dimensions the same preparation protocol gives $`<Z>4`$. Such low coordination numbers can be understood in terms of a constraint argument for frictionless rigid balls , which gives $`<Z>=2D`$, where D is the dimension. These values should be valid in the limit of low pressure when the beads are minimally connected near RCP (or in the rigid ball limit $`G\mathrm{}`$). For large values of the confining pressure more grains are brought into contact, and the coordination number increases . Empirically, we find that
$$<Z(p)>=6+\left(\frac{p}{0.06\text{MPa}}\right)^{1/3}.$$
(4)
Comparison with Experiment. Consider, now, the calculation of the elastic moduli of the system as function of pressure. Beginning with the equilibrium state describe above, we first restore the transverse component of the contact force interaction. We then apply a small perturbation to the system and measure the resulting response. The shear modulus is calculated in two ways, from a pure shear test, $`\mu =(1/2)\mathrm{\Delta }\sigma _{12}/\mathrm{\Delta }ϵ_{12}`$, and also from a biaxial test, $`\mu =(\mathrm{\Delta }\sigma _{22}\mathrm{\Delta }\sigma _{11})/2(\mathrm{\Delta }ϵ_{22}\mathrm{\Delta }ϵ_{11})`$. The bulk modulus is obtain from a uniaxial compression test, $`K+4/3\mu =\mathrm{\Delta }\sigma _{11}/\mathrm{\Delta }ϵ_{11}`$. Here the stress, $`\sigma _{ij}`$, is determined from the measured forces on the grains , and the strain, $`ϵ_{ij}`$, is determined from the imposed dimensions of the unit cell.
In Fig. 1 our calculated values of the elastic moduli as a function of pressure are compared with EMT and with experimental data. Because there is a considerable degree of scatter in the experimental results we performed our own experiments with standard sound propagation techniques. A set of high quality glass beads of diameter $`45\mu `$m are deposited in a flexible container of 3 cm height and 2.5 cm radius. Transducers and a pair of linear variable differential transformers (for measurement of displacement) were placed at the top and bottom of the flexible membrane, and the entire system was put into a pressure vessel filled with oil. Before starting the measurements, a series of tapping and ultrasonic vibrations were applied to the container in order to let the grains settle into the best possible packing. We then applied confining pressures ranging from 5 MPa to 100 MPa. The pressure was cycled up and down several times until the system exhibited minimal hysteresis. At this point shear and compressional waves were propagated by applying pulses with center frequencies of 500 KHz. The sound speeds and corresponding moduli were obtained by measuring the arrival time for the two sound waves.
From Fig. 1 we see that our experimental and numerical results are in reasonably good agreement. Also shown are measured data from Domenico and Yin . Clearly, the experimental data are somewhat scattered. This scatter reflects the difficulty of the measurements, especially at the lowest pressures where there is significant signal loss. Nevertheless, our calculated results pass through the collection of available data. Also shown in Fig. 1 are the EMT predictions
$$K=\frac{C_n}{12\pi }(\varphi _sZ)^{2/3}\left(\frac{6\pi p}{C_n}\right)^{1/3},$$
(6)
$$\mu =\frac{C_n+(3/2)C_t}{20\pi }(\varphi _sZ)^{2/3}\left(\frac{6\pi p}{C_n}\right)^{1/3}.$$
(7)
The EMT curves are obtained using the same parameters as in the simulations; we also set $`Z=6`$ and $`\varphi _s=0.64`$, independent of pressure. At low pressures we see that $`K`$ is well described by EMT. At larger pressures, however, the experimental and numerical values of $`K`$ grow faster than the $`p^{1/3}`$ law predicted by EMT. The situation with the shear modulus is even less satisfactory. EMT overestimates $`\mu (p)`$ at low pressures but, again, underestimates the increase in $`\mu (p)`$ with pressure.
To investigate the failure of EMT in predicting the correct pressure dependence of the moduli, we plot in Fig. 3 the shear modulus divided by $`p^{1/3}`$. For such a plot, EMT predicts a horizontal straight line but we see that the numerical and experimental results are clearly increasing with $`p`$. We can explain this behavior by modifying Eq. (3b) to take account of the pressure dependence of the coordination number $`<Z(p)>`$ from Fig. 2 and the pressure dependence of $`\varphi _s(p)`$ (which is a much smaller effect). The modified EMT is also plotted in Fig. 3 and we see that it predicts the same trend with pressure as the simulations. The experimental data also seem to be following this trend but more data over a larger pressure range are clearly needed. Not shown in Fig. 3 is a similar analysis of $`K(p)`$ but the result is that the modified EMT is in essentially exact agreement with our numerical simulations. It is for this reason that we focus on $`\mu (p)`$.
Another way of seeing the breakdown of EMT is to focus on the ratio $`K/\mu `$. According to Eqs. (7), $`K/\mu =5/3(2\nu )/(54\nu )`$, independent of pressure, a value which depends only on the Poisson’s ratio of the bead material. The experiments give $`K/\mu 1.11.3`$. Our simulations give $`K/\mu 1.05\pm 0.1`$ in good agreement with experiments. EMT predicts $`K/\mu =0.71`$, if we take $`\nu =0.2`$ for the Poisson’s ratio of glass. \[The EMT prediction is quite insensitive to variations of $`\nu `$; $`K/\mu =0.71\pm 0.04`$ for $`\nu =0.2\pm 0.1`$.\]
To understand why $`\mu `$ is over predicted by EMT we must examine the role of transverse forces and rotations in the relaxation of the grains. \[These effects do not play any role in the calculation of the bulk modulus.\] Suppose we re-define the transverse force by introducing a multiplicative coefficient $`\alpha `$, viz: $`\mathrm{\Delta }f_t=\alpha C_t(Rw)^{1/2}\mathrm{\Delta }s`$; with $`\alpha =1`$ we recover our previous results. To quantify the role of the transverse force on the elastic moduli, we calculate $`K(\alpha )`$ and $`\mu (\alpha )`$ at a given pressure $`p=100`$ KPa \[Fig. 4a\]. This pressure is low enough that the changing number of contacts is not an issue. Surprisingly, the shear modulus becomes negligibly small as $`\alpha 0`$. As expected, $`K`$ is independent of the strength of the transverse force. To compare with the theory we also plot the prediction of the EMT, Eq. (7) in which $`C_t`$ is rescaled by $`\alpha C_t`$. We see that the EMT fails in taking into account the vanishing of the shear modulus as $`\alpha 0`$. However it accurately predicts the value of the bulk modulus, which is independent of $`\alpha `$.
There are two main approximations in the EMT: (1) All the spheres are statistically the same, and it is assumed that there is an isotropic distribution of contacts around a given sphere. (2) An affine approximation is used, i.e., the spheres at position $`X_j`$ are moved a distance $`\delta u_i`$ in a time interval $`\delta t`$ according to the macroscopic strain rate $`\dot{ϵ}_{ij}`$ by $`\delta u_i=\dot{ϵ}_{ij}X_j\delta t`$. The grains are always at equilibrium due to the assumption of isotropic distribution of contacts and further relaxation is not required.
In the GD calculation of the shear modulus an affine perturbation is first applied to the system. The shear stress increases (from A to B in Fig. 4b) and the grains are far from equilibrium since the system is disordered. The grains then relax towards equilibrium (from B to C), and we measure the resulting change in stress from which the modulus is calculated. To better understand the approximations involved in the EMT, suppose we repeat the GD calculations taking into account only the affine motion of the grains and ignoring the subsequent relaxation. The resulting values of the moduli are plotted in Fig. 4a as open symbols and we see that the moduli calculated this way are very close to the EMT predictions. Thus, the difference between the GD and EMT results for the shear modulus lies in the non-affine relaxation of the grains; this difference being largest when there is no transverse force. By contrast, grain relaxation after an applied isotropic affine perturbation is not particularly significant and the EMT predictions for the bulk modulus are quite accurate.
The surprisingly small values of $`\mu `$ found as $`\alpha 0`$ can be understood as a melting of the system that occurs when the system is close to the RCP fraction. This fluid like behavior (when $`C_t=0`$) is closely related to the melting transition seen in compressed emulsions and foams . At the RCP fraction the system behaves like a fluid with no resistance to shear. By contrast, molecular dynamics simulations of glasses, in which the atoms interact by purely longitudinal forces, predict non-vanishing shear speeds . The crucial difference between these two systems is the local coordination of the particles. In the granular system, the coordination number near RCP (where the balls are only weakly deformed) is $`<Z>=6`$; the system is at its minimal coordination number. In glasses, however, the number of neighbors is closer to 10 and the motion of the grain is highly constrained.
Conclusions. Our GD simulations are in good agreement with the available experimental data on the pressure dependence of the elastic moduli of granular packings. They also serve to clarify the deficiencies of EMT. Grain relaxation after an infinitesimal affine strain transformation is an essential component of the shear (but not the bulk) modulus. This relaxation is not taken into account in the EMT. In the limit $`\alpha 0`$ a packing of nearly rigid particles responds to an external isotropic load with an elastic deformation and a finite $`K`$. By contrast, such a system cannot support a shear load ($`\mu 0`$) without severe particle rearrangements. This may indicate a “fragile” state of the system where inter particle forces are organized along “force chains” (stress paths carrying most of the forces in the system) oriented along the principal stress axes. Such fragile networks support, elastically, only perturbations compatible with the structure of force chains and deform plastically otherwise. Clearly, there is a need for an improved EMT; recent work on stress transmission in minimally connected networks may provide an alternative formulation and allow to describe properly the response of granular materials to perturbations.
ACKNOWLEDGMENTS. We would like to thank J. Dvorkin, J. St. Germain, B. Halperin, J. Jenkins, and D. Pissarenko for many stimulating discussions. |
no-problem/9911/cond-mat9911456.html | ar5iv | text | # S-wave Scattering Length from Effective Positronium-Positronium Interaction for Bose-Einstein Condensates
## I Introduction
Recent realizations of Bose-Einstein condensates for trapped Alkali-atom gas have made a great development in quantum physics of many-body system where the condensation phenomenon is one of the most important concepts and used in a vast area of physics such as atomic, laser, condensed-matter, nuclear, elementary-particle physics and also in cosmology. After the experimental success of the condensation, a lot of physical phenomena has been observed and studied on Alkali-atom BEC that include collective excitations, sound propagations, quantum interference, etc.
The success of the Alkali-atom BEC and the developments of the laser cooling and trapping techniques encouraged the study for the BEC of bose/fermi particles other than Alkali atoms: recently, the BEC of Rb isotope with the sympathetic cooling technique and the potassium fermi condensates have been observed. The bose/fermi mixed condensates are also studied for the potassium isotopes.
The positronium (Ps), the bound state of the electron and the positron, has been also an interesting candidate for the BEC:
* The Ps is the lightest atom with the mass $`m_{\text{Ps}}2m_e`$ where $`m_e`$ is the electron mass: $`10^3`$ times lighter than the hydrogen atom. It leads to the large value of the critical temperature $`T_C`$ of the Bose-Einstein phase transition: in free-gas and continuous approximation, $`T_C`$ becomes 10 times higher than that of hydrogen BEC for the trapping potential of same size ($`T_Cm^{1/3}`$). The effect of the Ps-Ps interaction for $`T_C`$ will be discussed in this paper.
* The Ps has finite life-time and self-annihilates into photons: $`Ps2\gamma `$ with $`\tau _{\text{pPs}}0.125\mathrm{ns}`$ for para-Ps and $`Ps3\gamma `$ with $`\tau _{\text{oPs}}142\mathrm{ns}`$ for ortho-Ps. Thus, the Ps BEC should be an unstable condensate and decays with gamma-ray radiations, the coincident observation of which will give the precise measurements for position/momentum distributions of the BEC. Those radiated gamma-ray photons are also suggested to show new physical phenomena such as gamma-ray laser by spontaneous amplifications and the supperradiance.
However, it should be noted that these very short life-times will make technical difficulties in experimental Ps BEC formations: $`\tau _{\text{Ps}}`$ are in the comparable order with the relaxation time $`\tau _{rel}\mathrm{ns}`$ in atomic phenomena so that the cooling methods that depend on the natural relaxation (like the evaporative cooling) might be less effective for the Ps BEC formation. At present, several faster cooling methods are planned by several experimental groups for Ps-BEC formations, and the preliminary experiments are progressing.
Originally, the Ps BEC was considered as $`e^+`$-$`e^{}`$ condensates as a source of gamma-ray laser, the $`e^+`$-$`e^{}`$ plasma in the strong magnetic field at the neutron star surface and the storage ring, and also as the source of the gamma-ray burst in astro-physics. The Ps BEC itself has been suggested as a source of the gamma-ray laser, but detailed studies for the properties of Ps BEC including the Ps-Ps interaction effects have not been done.
In the binding structure, the Ps is similar with the hydrogen atom (H), where the main differences are mass of positively charged particle ($`m_p/m_{e^+}2000`$) (and their stability). The BEC of poralized H has been reported with $`N10^9`$ atoms . More similar object are excitons in semiconductors: the bound states of electron and hole in valence band with $`m_h/m_e0.7`$. The exciton has also finite life-time $`\tau 30\mathrm{ns}`$ and decays into phonon, so that it is quite similar with the Ps. The BEC of excitons in CuO has also been reported . The experimental success of those BEC encourages strongly studies of Ps BEC.
In this paper, we derive the Ps-Ps interaction potential with the long-range van der Waals interaction and the short-range cut-off that represents the hard core repulsion of positronium. The s-wave scattering length is estimated from the solution of the Schrödinger equation using this potential. With the given scattering length, the Gross-Pitaevski equation for the ground-state wave function (order parameter) of Ps BEC is given, and, solving this equation with the trapping potential, we study the ground-state properties of Ps BEC. We also estimate the Ps-Ps interaction effects for the critical temperature $`T_C`$ in the mean-field approximation, which are shown to be negligible except large particle number BEC.
## II Effective Ps-Ps Interaction Potential
i) van der Waals Potential for Ps-Ps Interaction
We consider the system of two positronium Ps(1) and Ps(2). The spacial coordinates of the electron and positron composing Ps(1) are denoted by $`𝐱_1`$ and $`𝐱_a`$ and those of Ps(2) are by $`𝐱_2`$ and $`𝐱_b`$, with which the relative coordinates are defined by $`𝐫_{ij}=𝐱_j𝐱_i`$. Especially, for relative coordinates between electron and positron, we denote
$$𝐫_1𝐫_{a1},𝐫_2𝐫_{b2}.$$
(1)
The hamiltonian of the system is
$$H=H_{CM}+H_1+H_2+H_{int},$$
(2)
where $`H_{CM}`$ is the center-of-mass part that can be neglected in the derivation of the Ps-Ps interaction because of the translational invariance. The $`H_{1,2}`$ in (2) are the internal parts of Ps(1,2) with the Coulomb potential:
$$H_{1,2}=\frac{𝐩_{1,2}^2}{m_e}\frac{\alpha \mathrm{}c}{r_{1,2}},$$
(3)
where $`r_{1,2}=|𝐫_{1,2}|`$ and $`𝐩_{1,2}`$ are the momenta conjugate with $`𝐫_{1,2}`$. The $`\alpha `$ in (3) is a fine structure constant and $`\alpha =e^2/(4\pi ϵ_0\mathrm{}c)1/137`$ in the SI unit system. The interaction part of Ps(1,2) is represented by
$$H_{int}=\alpha \mathrm{}c\left(\frac{1}{r}+\frac{1}{r_{12}}\frac{1}{r_{a2}}\frac{1}{r_{b1}}\right),$$
(4)
where $`r_{ij}=|𝐫_{ij}|`$ and $`rr_{ab}=|𝐫_{ab}|`$ is the distance between two positrons, Ps(1) and Ps(2).
Let us consider the case where the distance between Ps(1,2) are very large: $`rr_{ab}r_1,r_2`$. In that case, the Ps-Ps interaction potential can be determined by the van der Waals potential.
Eq. (4) can be expanded by $`1/r`$, and we get the dipole-dipole interaction:
$$H_{int}\frac{\alpha \mathrm{}c}{r^3}\left\{3\frac{(𝐫_1𝐫)(𝐫_2𝐫)}{r^2}𝐫_1𝐫_2\right\}.$$
(5)
When $`r0`$, the interaction term $`H_{int}1/r^3`$ in (5) can be evaluated with perturbation. The unperturbed part is just $`H_0=H_1+H_2`$ defined by (3) and the zeroth-order unperturbed wave functions that diagonalize the $`H_0`$ are $`\psi _{nlm}(1)\psi _{n^{}l^{}m^{}}(2)`$ where $`\psi _{nlm}(i)`$ is the eigenstate of Ps(i), $`H_i\psi _{nlm}(i)=E_n^{(0)}\psi _{nlm}(i)`$, with the binding energy:
$$E_n^{(0)}=m_ec^2\alpha ^2/\{4(n+1)^2\}.$$
(6)
The $`n`$, $`l`$, $`m`$ are principal, azimuthal and magnetic quantum numbers. In the Ps state function, we dropped the spin angular momentum part and spin quantum numbers $`S=0,1`$ and $`S_z`$, because the spin-spin/spin-orbital interactions are weak, and give a very small modifications that can be neglected in the present semi-qualitative estimations.
From the second-order perturbation for $`H_{int}`$, we get the van der Waals potential:
$$E^{(2)}=\alpha ^2(\mathrm{}c)^2\frac{6}{r^6}\underset{n0,l,m,n^{}0,l^{},m^{}}{}\frac{|0|d_1|\psi _{nlm}(1)|^2|0|d_2|\psi _{n^{}l^{}m^{}}(2)|^2}{E_n^{(0)}+E_n^{}^{(0)}E_0^{(0)}E_0^{(0)}},$$
(7)
where $`d_{1,2}=z_{1,2}`$ are dipole-moment operators of Ps(1,2) and $`|0=|\psi _{000}`$. Using eq. (6), the denominator of eq. (7) becomes
$$E_n^{(0)}+E_n^{}^{(0)}E_0^{(0)}E_0^{(0)}=\frac{\alpha \mathrm{}c}{4a_0}[2(n+1)^2(n^{}+1)^2],$$
(8)
where $`a_0=\mathrm{}c/(\alpha m_ec^2)0.0529\mathrm{nm}`$ is the Bohr radius. The term $`(n+1)^2`$ in the denominator of eq. (8) is very small for large value of $`n`$, so that it can be neglected in the first approximation
$$E_n^{(0)}+E_n^{}^{(0)}E_0^{(0)}E_0^{(0)}\frac{\alpha \mathrm{}c}{2a_0},$$
(9)
Using eq. (9) and the completeness condition: $`_{n,l,m}0|d_1|\psi _{nlm}\psi _{nlm}|d_1|0=0|z^2|0`$, eq. (7) becomes
$$E^{(2)}=\frac{6\alpha \mathrm{}c(2a_0)}{r^6}|0|z^2|0|^2.$$
(10)
Using $`0|z^2|0=(2a_0)^2`$ with the Coulomb wave function $`\psi _{000}`$ for the ground state, we get van der Waals potential:
$$E^{(2)}=6\alpha \frac{\mathrm{}c(2a_0)^5}{r^6}.$$
(11)
The error for the above approximation (11) has been evaluated by numerical summation for (7) for the H-H interaction . When their results are rescaled for Ps potential, they obtained 6.47 instead of 6 in (11). It suggests that the present results (11) should be a reasonable estimation.
ii) Effective Ps-Ps interaction potential with the short-range hard core
In the short range part, the real part of the Ps-Ps interaction should be characterized by repulsive hard core at the distance of atomic diameter where two positronium begin to overlap. In the present paper, we treat this hard core potential phenomenologically in two manners: A) the cut-off potential at $`r=r_C`$, $`V_{\text{Ps-Ps}}^{CO}`$ (Fig. 2) and B) the Lenard-Jones potential with $`r^{12}`$-hard core (Fig. 3):
$$V_{\text{Ps-Ps}}^{LJ}(r)=\alpha \mathrm{}c\left[6\frac{(2a_0)^5}{r^6}+C\frac{(2a_0)^{11}}{r^{12}}\right],$$
(12)
The phenomenological parameters $`r_C`$ and $`C`$ in $`V_{\text{Ps-Ps}}^{CO}`$ and $`V_{\text{Ps-Ps}}^{LJ}`$ are fixed to reproduce the binding energy $`B_{\text{Ps}_2}`$ of the molecular state Ps<sub>2</sub>, which has not been confirmed experimentally but predicted theoretically/numerically with the hamiltonian (2,3,4) . Recent results for $`B_{\text{Ps}_2}`$ by elaborate numerical calculations are around $`0.435\mathrm{eV}`$ .
To reproduce $`B_{\text{Ps}_2}=0.435\mathrm{eV}`$ with solving the Schrödinger equation for $`V_{\text{Ps-Ps}}`$ numerically, we obtain A) $`r_C=1.78a_0`$ (with the potential depth $`v_0=6.13\alpha \mathrm{}c/a_0`$) for $`V_{\text{Ps-Ps}}^{CO}`$, and B) $`C=2.25`$ ($`v_0=2.0\alpha \mathrm{}c/a_0`$ and $`r_C^{}=1.90a_0`$: the distance where $`V_{\text{Ps-Ps}}(r_C^{})=0`$) for $`V_{\text{Ps-Ps}}^{LJ}`$. It should be noticed that two phenomenological potentials have almost the same hard-core distances $`r_C,r_C^{}1.8a_0`$, which are consistent with the diameter of Ps, $`2a_0`$.
We should comment about the effect of Ps spin state. Because of the Pauli principle, the Ps<sub>2</sub> bound state exists in singlet state, so that the effective Ps-Ps potential using the binding energy of Ps<sub>2</sub> as an input parameter should be considered for singlet channel. However, because the $`r_C`$ of $`V_{\text{Ps-Ps}}^{CO,LJ}`$ is very close to $`2a_0`$ that should be expected to be the hard core length for nonsinglet channel , the present potential should be applicable also for these channels.
## III s-wave Scattering Length by Effective Ps-Ps Interaction
For the Ps-Ps s-wave scattering, the relative wave function of two positronium becomes $`\mathrm{\Psi }(𝐫)=r\chi (r)`$ that satisfies the Schödinger equation:
$$\frac{\mathrm{}^2}{2\mu }\frac{d^2}{dr^2}\chi (r)+V_{\text{Ps-Ps}}(r)\chi (r)=E\chi (r),$$
(13)
where $`\mu `$ is a reduced mass $`\mu =m_{\text{Ps}}/2m_e`$.
The s-wave scattering length $`a`$ is obtained from the asymptotic form of the $`E=0`$ solution for eq. (13) as
$$\chi (r\mathrm{})1\frac{r}{a}.$$
(14)
In fig. 4, we show the numerical results for the $`E=0`$ wave function $`\chi (r)`$ of eq. (13) with $`V_{\text{Ps-Ps}}^{CO,LJ}`$ defined in the previous section. The one peak at $`r3.5a_0`$ shows that (only) one weak bound state exists for this potential and it corresponds to the Ps<sub>2</sub> molecular state whose binding energy has been used as an input for $`V_{\text{Ps-Ps}}`$. As seen in fig. 4, for $`r5a_0`$, $`\chi (r)`$ in fig. 4 are in the asymptotic linear region, so that the scattering length $`a`$ can be read off by $`\chi (a)=0`$:
$$\text{A) }a=0.440\mathrm{nm}\text{for }V_{\text{Ps-Ps}}^{CO},\text{B) }a=0.437\mathrm{nm}\text{for }V_{\text{Ps-Ps}}^{LJ}.$$
(15)
Both results are almost consistent, and it means that the s-wave scattering length (the low-energy limit) do not depend on the detailed form of the short-range potential. Existence of only one weakly bound state in Ps-Ps system is also essential for the rigidness of $`a`$ given with the effective potential
For the illustration of the effects by bound states for the scattering length, the $`s`$-wave scattering state function $`\chi (r)`$ is shown in fig. 5 for the effective interaction potential fitted with H-H binding energy. That potential has the larger $`v_0`$ than the Ps-Ps potential and produces many excited molecular states that can be seen as the oscillating behavior of $`\chi (r)`$ for $`r5a_0`$. Its behavior depends on the short-range part of the potential given only in phenomenological manner in the effective potential, so that it leads to the impreciseness of the connected long-range behavior of $`\chi (r)`$ from which the scattering length $`a`$ is given. For the Ps-Ps case, it should also be too rough approximation to treat the molecular bound state Ps<sub>2</sub> itself as that of the effective potential given in this paper, but the long-range part of the scattering state should be more reliable because its short-range behavior is very simple and the potential has only one bound state that was used for the input parameter. The consistency between the scattering lengths given with $`V_{\text{Ps-Ps}}^{CO}`$ and $`V_{\text{Ps-Ps}}^{LJ}`$ also supports it.
Currently-used $`s`$-wave scattering lengths for Alkali atoms are, $`5.77\mathrm{nm}`$ for <sup>87</sup>Rb , $`2.75\mathrm{nm}`$ for <sup>23</sup>Na and $`1.45\mathrm{nm}`$ for <sup>7</sup>Li . The resultant Ps-Ps $`s`$-wave scattering lengths in (15) are $`a0.44\mathrm{nm}`$, which is smaller than those for Alkali atoms in one order but larger than that for polarized H atom (triplet), $`a=0.065\mathrm{nm}`$ . It shows that the Ps-gas behaves more like perfect gas than the Alkali atom gas. The rough estimation for the Ps-Ps scattering length can also be made directly from the binding energy of Ps<sub>2</sub> : $`a(2m_eB_{\text{Ps}_2})^{1/2}0.3\mathrm{nm}`$ , which is also consistent with the present result.
## IV Ps BEC with Ps-Ps Interaction
The dynamical behavior of BEC is represented by the ground-state wave function (order parameter) $`\psi (x)`$ and, for weak-interacting bose gas in the harmonic oscillator trapping potential at $`T=0`$, it satisfies the Gross-Pitaevski (GP) equation :
$$i\mathrm{}\frac{\psi }{t}=\frac{\mathrm{}^2}{2m_{\text{Ps}}}^2\psi +\frac{1}{2}m_{\text{Ps}}\omega _{HO}^2r^2\psi +gN|\psi |^2\psi ,$$
(16)
where $`\omega _{HO}`$ is the angular frequency of the harmonic oscillator potential, and $`N`$ is the condensed bose particle number: the normalization condition for $`\psi `$ is taken as
$$d^3x|\psi (x)|^2=1.$$
(17)
For low-density bose gas, the boson-boson interaction is dominated by the low-energy s-wave scattering, and the interaction constant $`g`$ in (16) is given with the s-wave scattering length $`a`$ :
$$g=\frac{4\pi \mathrm{}^2}{m_{\text{Ps}}}a.$$
(18)
Using the dimensionless variables $`\stackrel{~}{𝐫}=𝐫/a_{HO}`$, $`\stackrel{~}{t}=\omega _{HO}t`$ and $`\stackrel{~}{\psi }=a_{HO}^{3/2}\psi `$ with the harmonic oscillator length $`a_{HO}=\sqrt{\mathrm{}/(m_{\text{Ps}}\omega _{HO})}`$, the GP equation (16) becomes
$$i\frac{\stackrel{~}{\psi }}{\stackrel{~}{t}}=\frac{1}{2}\stackrel{~}{}^2\stackrel{~}{\psi }+\frac{\stackrel{~}{r}^2}{2}\stackrel{~}{\psi }+\stackrel{~}{g}N|\stackrel{~}{\psi }|^2\stackrel{~}{\psi },$$
(19)
where the dimensionless constant $`\stackrel{~}{g}N`$ is
$$\stackrel{~}{g}N=\frac{4\pi a}{a_{HO}}N.$$
(20)
Thus, the solutions of GP equation are similar for the same value of $`\stackrel{~}{g}N`$. It should be noticed that the interaction effects come into eq. (19) through $`\stackrel{~}{g}N`$-term, so that they are more effective for large value of $`\stackrel{~}{g}N`$.
In fig. 6, the solutions of eq. (16) are shown with $`a=0.437\mathrm{nm}`$ in (15) and $`\omega =3.1\mathrm{sec}^1`$, for several boson numbers: $`N=2.6\times 10^7`$ (curve 2), $`2.0\times 10^8`$ (curve 3), $`7.9\times 10^8`$ (curve 4), $`2.5\times 10^9`$ (curve 5). With $`a(Ps)=0.44\mathrm{nm}`$ for Ps, we obtain $`a(Ps)/a_{HO}=2.16\mathrm{eV}^{1/2}\sqrt{\mathrm{}\omega _{HO}}`$, which is much smaller than $`a(Rb)/a_{HO}=7962\mathrm{eV}^{1/2}\sqrt{\mathrm{}\omega _{HO}}`$ for Rb atom: this large difference comes not only from weak Ps-Ps interaction $`a(Ps)/a(Rb)1/130`$ but from the mass difference $`\sqrt{m_{\text{Ps}}/m_{Rb}}1/281`$. Because of the small $`a/a_{HO}`$ for the Ps-Ps interaction, for large particle number $`N=2.6\times 10^7`$ (curve 2 in fig. 6), the calculated $`|\psi (r)|`$ changes only $`16\%`$ from that for the ideal gas $`a=0`$ (curve 1 in fig. 6). To realize $`70\%`$ change for $`|\psi (0)|`$, the Ps of $`N=2.5\times 10^9`$ should condensate (curve 5 in fig. 6). For Rb BEC, because of large scattering length $`a=5.77\mathrm{nm}`$, $`N=6.7\times 10^5`$ is enough for realization of the profile 5 in fig. 6.
For the ideal gas with no interactions, the critical temperature $`T_C^0`$ can be calculated in thermodynamical continuum limit :
$$k_BT_C^0=0.94\mathrm{}\omega _{HO}N^{1/3},$$
(21)
where $`k_B`$ is a Boltzmann constant. The particle interaction has an effect to shift the critical temperature: $`T_C=T_C^0+\delta T_C`$. The $`\delta T_C`$ can be evaluated with the mean-field and semiclassical approximation :
$$\frac{\delta T_C}{T_C^0}=1.33\frac{a}{a_{HO}}N^{1/6}.$$
(22)
For Alkali atom BEC, eq. (22) is consistent with experimentally observed temperature shift ($`2\%`$ shift for Rb BEC of $`N10^4`$). In the case of Ps BEC, eq. (22) becomes
$$\frac{\delta T_C}{T_C^0}=2.87\mathrm{eV}^{1/2}\sqrt{\mathrm{}\omega _{HO}}N^{1/6}.$$
(23)
It gives very small shift for $`T_C`$ for Ps BEC, and can be neglected in the first approximation except BEC of large $`N`$.
It should be noted that the concept of $`T_C`$ is exact for equilibrium (thermodynamical) system. For real experimental situations, the Ps BEC might be performed as nonequilibrium system due to the small Ps life-time that is almost the same order with relaxation time. In that case, the $`T_C`$ evaluated here should be considered as estimation for the energy scale.
## V Summary and Discussions
In the present paper, we studied the effective Ps-Ps interaction potentials with interpolating the long-range van der Waals interaction and the short-range repulsive hard core. The latter was taken phenomenologically in two kinds of parametrizations: A) sudden cut-off and B) Lenard-Jones (and cut-off) types. The Ps-Ps s-wave scattering length was calculated from the solutions of the Scrödinger equation with those effective potentials, and we obtained $`a0.44\mathrm{nm}`$. It is ($`210`$)-times smaller than that of the Alkali atoms, and larger than that of the polarized hydrogen atom. The stability of $`a`$ for the different parametrizations of the short-range interaction was also checked. Using that value of the scattering length, we studied the ground-state wave functions for Ps BEC with solving the Gross-Pitaevski equation. The Ps BEC is shown to be more perfect-gas-like than Alkali-atom BEC because of weaker interaction. The critical temperature shift $`\delta T_C`$ due to the Ps-Ps interaction was also estimated and shown to be very small because of the weak interaction and the small mass of Ps atom.
As commented in the last section, the inclusion of the dissipative effects to the Gross-Pitaevski equation is essentially important for further studies of Ps BEC physics discussed in the introduction. It will be discussed in a future publication. |
no-problem/9911/cond-mat9911049.html | ar5iv | text | # Modulation structure of a frustrated spin ladder
## Abstract
We study a two-leg spin-1/2 ladder with isotropic exchanges and biquadratic interactions in the basic plaquettes. It is shown that for the extremely frustrated case, the system exhibits a self-organized phase separation. In some parameter regions, the singlet rungs form a Wigner-like lattice in the triplet-rung host. There are three types of elementary excitations in this modulation phase, i.e., the spinons in a triplet domain, the broken singlet rungs and the deformation of the Wigner-like lattice. The flux phase induced by an external magnetic field in the rung-dimerized phase is also discussed.
After the discovery of the ladder superconductors, the current interest in the coupled spin chains has been greatly renewed. It is well established that the regular spin-1/2 ladders with even number of legs have a gaped spin-liquid ground state, while the odd-legged ladders have a gapless spin-liquid ground state. On the other hand, the generalized spin ladders including other couplings beyond the nearest neighbor exchanges, which can interpolate among a variety of systems, have been demonstrated exhibiting remarkably rich behavior. The hamiltonian of the generalized spin ladder reads:
$`H={\displaystyle \frac{1}{4}}J_1{\displaystyle \underset{j=1}{\overset{N}{}}}(\stackrel{}{\sigma }_j\stackrel{}{\sigma }_{j+1}+\stackrel{}{\tau }_j\stackrel{}{\tau }_{j+1})+{\displaystyle \frac{1}{4}}J_2{\displaystyle \underset{j=1}{\overset{N}{}}}(\stackrel{}{\sigma }_j\stackrel{}{\tau }_{j+1}+\stackrel{}{\tau }_j\stackrel{}{\sigma }_{j+1})`$ (1)
$`+{\displaystyle \frac{1}{2}}J_3{\displaystyle \underset{j=1}{\overset{N}{}}}\stackrel{}{\sigma }_j\stackrel{}{\tau }_j+{\displaystyle \frac{1}{4}}U_1{\displaystyle \underset{j=1}{\overset{N}{}}}(\stackrel{}{\sigma }_j\stackrel{}{\sigma }_{j+1})(\stackrel{}{\tau }_j\stackrel{}{\tau }_{j+1})`$ (2)
$`+{\displaystyle \frac{1}{4}}U_2{\displaystyle \underset{j=1}{\overset{N}{}}}(\stackrel{}{\sigma }_j\stackrel{}{\tau }_{j+1})(\stackrel{}{\sigma }_{j+1}\stackrel{}{\tau }_j)+{\displaystyle \frac{1}{4}}U_3{\displaystyle \underset{j=1}{\overset{N}{}}}(\stackrel{}{\sigma }_j\stackrel{}{\tau }_j)(\stackrel{}{\sigma }_{j+1}\stackrel{}{\tau }_{j+1}),`$ (3)
where $`\stackrel{}{\sigma }_j`$ ($`\stackrel{}{\tau }_j`$) are Pauli matrices on site $`j`$ of the upper (lower) leg, $`J_\alpha `$ ($`\alpha =1,2,3`$) are the exchange coupling constants and $`U_\alpha `$ are the biquadratic coupling constants. The hamiltonian (1) contains all possible $`SU(2)`$-invariant interactions in a basic plaquette formed by two nearest rungs. We note the biquadratic interactions may be effectively mediated by phonons and their importance for some properties of $`CuO_2`$ plaquette has been pointed out. In addition, current experiments have revealed that such multi-spin interactions are realized in the two-dimensional (2D) solid $`{}_{}{}^{3}He`$, 2D Wigner solid of electrons formed in a $`Si`$ inversion layer, and the $`bcc`$ solid $`{}_{}{}^{3}He`$.
The model (1) has been studied by many authors in some parameter regions. The possible phases including the dimerized phase, the Haldane phase and the rung-dimerized phase have been reported. In addition, the model (1) has a variety of integrable cases and some gapless phases have been found. All these results indicate that the generalized spin ladder has a very rich ground state phase diagram. In this letter, we study the extremely frustrated, i.e., $`J_1=J_2`$, $`U_1=U_2`$ case. We show that there are three possible phases in this case, i.e., a rung-dimerized (RD) phase, a triplet-rung (TR) phase interpolating between the gapless spin liquid and the Haldane spin liquid or the $`VBS`$ state, and a new gapful phase with modulation structure. The latter phase is a mixed state (MS) consisting of both singlet rungs and triplet rungs and its structure is very similar to a Wigner lattice (crystallization of the singlet rungs) or the flux phase of the type II superconductors (with the singlet rungs as the effective fluxes).
In the extremely frustrated case ($`J_1=J_2`$, $`U_1=U_2`$), the hamiltonian (1) can be rewritten as
$`H=\stackrel{~}{J}{\displaystyle \underset{j=1}{\overset{N}{}}}[\stackrel{}{S}_j\stackrel{}{S}_{j+1}\beta (\stackrel{}{S}_j\stackrel{}{S}_{j+1})^2]`$ (4)
$`+U{\displaystyle \underset{j=1}{\overset{N}{}}}(2\stackrel{}{S}_j^2)(2\stackrel{}{S}_{j+1}^2)J{\displaystyle \underset{j=1}{\overset{N}{}}}(2\stackrel{}{S}_j^2)+C,`$ (5)
where $`\stackrel{}{S}_j=(\stackrel{}{\sigma }_j+\stackrel{}{\tau }_j)/2`$ is the total spin of the $`j`$-th rung, $`\stackrel{~}{J}=2J_1+U_1`$, $`\beta =2U_1/(2J_1+U_1)`$, $`U=U_3`$, $`J=J_3+U_32U_1`$ and $`C=J_3/2+U_3/45U_1/2`$ is an irrelevant constant. For convenience, we put $`\stackrel{~}{J}=1`$ and omit the irrelevant constant $`C`$ in the following text. The total spin $`S_j`$ takes two possible values $`0,1`$, corresponding to the singlet rungs and triplet rungs, respectively. Based on Eq.(2), the phase diagram of the system can be conjectured roughly. There are three possible types of ground state configurations. For a large enough $`J`$, the singlet rungs are more stable than the triplet rungs. The ground state is a simple product of $`N`$ singlet rungs, i.e., a rung-dimerized state. For a small $`J`$ and large $`U`$, the triplet rungs are dominant over the singlet rungs and the ground state is exactly the same of a generalized spin-1 chain. In some intermediate $`J,U`$ regions, the singlet rungs and the triplet rungs may coexist in the ground state. Since there is no genuine interactions between the singlet rungs and the triplet rungs as we can read off from Eq.(2), a self-organized phase separation occurs in this case. The singlet rungs simply cut the system into disconnected domains of triplet rungs and a positive $`U`$ will stablize a Wigner-like lattice of the singlet rungs in the triplet liquid. Each triplet domain in the mixed state behaves as an open spin-1 chain. Such a mixed state can only be accompanied by a negative boundary energy, as the flux phase in a type II superconductor. We note that $`\beta =\pm 1`$ represent two integrable points of the model (2) and when $`\beta =1/3`$, a $`VBS`$ ground state can be constructed. Very interestingly, the translational invariance is broken in the MS phase and the ground state is no longer a spin liquid but a periodic array of spin-liquid domains.
To give a clear picture, we study one of the integrable cases $`\beta =1`$. In this case, each triplet domain behaves as an $`SU(3)`$-invariant spin-1 chain with open boundaries and can be solved exactly via Bethe ansatz. Let us consider a triplet domain with $`M`$ rungs. The effective hamiltonian of this domain reads
$`H_M={\displaystyle \underset{j=1}{\overset{M1}{}}}[\stackrel{}{S}_j\stackrel{}{S}_{j+1}+(\stackrel{}{S}_j\stackrel{}{S}_{j+1})^2]`$ (6)
with $`S_j=1`$. The solution of Eq.(3) is exactly the same of the $`SU(3)`$-invariant $`tJ`$ model with pure open boundaries. Its spectrum is determined by the following Bethe ansatz equations (BAE’s)
$`\left({\displaystyle \frac{\lambda _j\frac{i}{2}}{\lambda _j+\frac{i}{2}}}\right)^{2M}={\displaystyle \underset{r=\pm 1}{}}\left[{\displaystyle \underset{lj}{\overset{M_1}{}}}{\displaystyle \frac{\lambda _jr\lambda _li}{\lambda _jr\lambda _l+i}}{\displaystyle \underset{\alpha =1}{\overset{M_2}{}}}{\displaystyle \frac{\lambda _jr\mu _\alpha +\frac{i}{2}}{\lambda _jr\mu _\alpha \frac{i}{2}}}\right],`$ (7)
$`{\displaystyle \underset{r=\pm 1}{}}{\displaystyle \underset{\beta \alpha }{\overset{M_2}{}}}{\displaystyle \frac{\mu _\alpha r\mu _\beta i}{\mu _\alpha r\mu _\beta +i}}={\displaystyle \underset{r=\pm 1}{}}{\displaystyle \underset{j=1}{\overset{M_1}{}}}{\displaystyle \frac{\mu _\alpha r\lambda _j\frac{i}{2}}{\mu _\alpha r\lambda _j+\frac{i}{2}}},`$ (8)
where $`\lambda _j`$ and $`\mu _\alpha `$ are the rapidities of the spinons and $`M_2<M_1<M`$. The eigen energy of the triplet domain reads:
$`E_M={\displaystyle \underset{j=1}{\overset{M_1}{}}}{\displaystyle \frac{1}{\lambda _j^2+1/4}}+2(M1).`$ (9)
Notice that the second term in Eq.(5) is also relevant since the number of triplet rungs is not fixed in the whole system. It contributes an amount of -2 to the boundary energy, which is crucial to stablize an MS. For $`M>>1`$, the ground state energy of the triplet domain can be expressed as $`E_M=Mϵ_0+ϵ_b(M)`$, where $`ϵ_0`$ is the energy density of an infinite $`SU(3)`$ spin-1 chain, $`ϵ_b(M)`$ is the boundary energy including the $`O(M^1)`$ finite size correction. Denote the ground state distributions of $`\lambda `$ and $`\mu `$ for $`M\mathrm{}`$ as $`\rho _1(\lambda )`$ and $`\rho _2(\mu )`$ respectively. From the BAE’s (4) we get the following integral equations
$`\rho _1(\lambda )+[2]\rho _1(\lambda )[1]\rho _2(\lambda )={\displaystyle \frac{1}{2M}}a_2(\lambda )+a_1(\lambda ),`$ (10)
$`\rho _2(\mu )+[2]\rho _2(\mu )[1]\rho _1(\mu )={\displaystyle \frac{1}{2M}}a_2(\mu ),`$ (11)
where $`a_n(\lambda )=n/2\pi [\lambda ^2+(n/2)^2]`$ and $`[n]`$ is the integrable operator with kernel $`a_n(\lambda )`$. By solving the above integral equations with Fourier transformation, we obtain
$`\rho _1(\lambda )=\rho _1^0(\lambda )+{\displaystyle \frac{1}{M}}\delta \rho _1(\lambda ),`$ (12)
$`\rho _1^0(\lambda )={\displaystyle \frac{1}{2\pi }}{\displaystyle e^{i\omega \lambda }\frac{2\mathrm{cosh}\frac{\omega }{2}}{4\mathrm{cosh}^2\frac{\omega }{2}1}𝑑\omega },`$ (13)
$`\delta \rho _1(\lambda )={\displaystyle \frac{1}{4\pi }}{\displaystyle e^{i\omega \lambda }\frac{e^{\frac{|\omega |}{2}}}{2\mathrm{cosh}\frac{\omega }{2}1}𝑑\omega }{\displaystyle \frac{1}{2}}\delta (\lambda ),`$ (14)
where $`\rho _1^0(\lambda )`$ is the density of $`\lambda `$ of the ground state for $`M\mathrm{}`$ with periodic boundary conditions. The $`\delta (\lambda )`$ term in the third equation of Eq.(7) comes from the $`\lambda =0`$ mode, which is forbidden in an open boundary system. $`ϵ_0`$ and $`ϵ_b(M)`$ can be derived as
$`ϵ_0={\displaystyle \frac{\rho _1^0(\lambda )}{\lambda ^2+1/4}𝑑\lambda }+2=2\mathrm{ln}3{\displaystyle \frac{\pi }{3\sqrt{3}}},`$ (15)
$`ϵ_b(M)={\displaystyle \frac{\delta \rho _1(\lambda )}{\lambda ^2+1/4}}2+O(M^1)={\displaystyle \frac{4}{9}}\sqrt{3}\pi 4+O(M^1).`$ (16)
Based on Eq.(8), the phase boundaries can be determined exactly. As we discussed above, there are three possible phases, i.e., the RD phase, the TR liquid and the MS phase containing both singlet rungs and triplet rungs. We note that the density of the ground state energy of the rung-dimerized state reads $`ϵ_r=4U2J`$, as we can easily derive from Eq.(2). Therefore, the phase boundary between the RD phase and the TR liquid phase is given by $`4U2J=ϵ_0`$. On the other hand, generating a singlet rung in the TR liquid implys a broken triplet rung and an open boundary. The excitation energy of this process is $`ϵ_b(\mathrm{})ϵ_02J`$. Therefore, the phase boundary between the TR phase and the mixed state is given by $`2J=ϵ_b(\mathrm{})ϵ_0`$. We note the singlet rungs are very similar to the fluxes in a type II superconductor. Here $`ϵ_b(\mathrm{})<0`$ corresponds to the surface energy, while $`2Jϵ_0`$ corresponds to the self energy of the fluxes. When $`2J>ϵ_b(\mathrm{})ϵ_0`$, some singlet rungs appear in the triplet liquid. Here $`J_{c1}=(ϵ_b(\mathrm{})ϵ_0)/2`$ serves as the lower critical field. Due to the finite size correction of $`ϵ_b(M)`$, the singlet rungs are unfavorable to close each other (note $`ϵ_b(M)`$ is an increasing function of $`1/M`$). The lengths of the triplet domains are mainly controlled by the finite size correction. Suppose we have $`N_M=N/(M+1)`$ triplet domains with lengths $`M+\delta _m`$ respectively. The small shifts $`\delta _m`$ satisfy the condition $`_m\delta _m=0`$. The total correction energy reads
$`E_c={\displaystyle \underset{m=1}{\overset{N_M}{}}}[ϵ_b(M+\delta _m)ϵ_b(\mathrm{})]{\displaystyle \underset{m=1}{\overset{N_M}{}}}{\displaystyle \frac{1}{M+\delta _m}}.`$ (17)
By minimizing Eq.(9), we get $`\delta _m=0`$, which indicates a periodic array of the singlet rungs in the whole system. This novel modulation structure is very similar to a Wigner lattice or an Abrikosov lattice but with a very different physical interpretation. As in the type II superconductors, there is also an upper bound of $`J`$ corresponding to the upper critical field, which determines the phase boundary between the RD phase and the MS phase. Suppose an MS is stable at $`M=M(J)`$. The phase boundary is given by
$`4U2J=ϵ_{M(J)}+{\displaystyle \frac{1}{M(J)+1}}[ϵ_{M(J)}2J].`$ (18)
When $`M(J)1`$, $`J_{c2}4U`$.
Now we turn to the elementary excitations in the MS phase. There are three types of elementary excitations in this case, i.e., the spinons in the triplet domains, broken singlet rungs and the deformation of the Wigner-like lattice. Suppose the lattice is stablized with a period $`M+1`$ ($`M`$ triplet rungs and one singlet rung). The spinons in the present case have a finite energy gap in the order of $`M^1`$ due to the finite length of the domains. For $`M>>1`$, the energy gap can be derived with the well known finite-size-correction techniques. As in the usual integrable models, the excitation energies of the spinons are given by the so-called dressed energies which in our case read (for $`M\mathrm{}`$)
$`ϵ_1(\lambda )=2\pi a_1(\lambda )+[1]ϵ_2(\lambda )[2]ϵ_1(\lambda ),`$ (19)
$`ϵ_2(\mu )=[1]ϵ_1(\mu )[2]ϵ_2(\mu ).`$ (20)
The velocities of the two branches of low-energy spinons are equivalent, which read
$`v=\underset{\lambda \mathrm{}}{lim}{\displaystyle \frac{ϵ_1^{}(\lambda )}{2\pi \rho _1^0(\lambda )}}=\underset{\mu \mathrm{}}{lim}{\displaystyle \frac{ϵ_2^{}(\mu )}{2\pi \rho _2^0(\mu )}}`$ (21)
The energy gap associated with the spinon excitations is thus $`\mathrm{\Delta }\pi v/M`$. There are three different pictures of the spinon excitations: (i)For $`M=3n`$, there is no holes in the $`\lambda `$\- and $`\mu `$-seas in the ground state. The spinons are generated by spin flips. The simplest excitation is a one $`\lambda `$-hole and two $`\mu `$-hole state. (ii)For $`M3n`$, there are some $`\lambda `$\- and $`\mu `$\- holes in the ground state. The mobility of the holes gives the simplest excitation. (iii)For arbitrary $`M`$, there are string excitations. We note all these excitations have the same energy gap.
A broken singlet rung in the modulation phase indicate that two neighboring triplet domains combined to a single domain with length $`2M+1`$. In a stable MS phase, $`M`$ is determined by minimizing the per site energy $`ϵ_M=[ϵ_b(M)ϵ_02J]/(M+1)+ϵ_0`$. For large $`M`$, from $`ϵ_M/M=0`$ we readily obtain $`J[ϵ_b(\mathrm{})ϵ_0]/2M^1`$. Therefore, the excitation energy of a broken singlet rung is $`ϵ_b(2M+1)ϵ_b(M)+ϵ_0+2J`$, which is still in the order of $`M^1`$.
The deformations of the Wigner-like lattice represent another type of excitations in the modulation phase. This type of excitations are static rather than dynamic due to the complete phase separation. Suppose one singlet rung is moved from its equilibrium position by $`\delta M`$. The two neighboring triplet domains connected by this singlet rung are thus enlarged and compressed by $`\delta M`$ respectively. The energy of this process comes mainly from the finite size corrections $`ϵ_b(M+\delta M)+ϵ_b(M\delta M)2ϵ_b(M)`$ and is in the order of $`M^3`$ (notice that $`ϵ_b(M)ϵ_b(\mathrm{})M^1`$), implying the larger the $`M`$, the more unstable the lattice.
A ladder system may exhibit interesting behavior in an external magnetic field. In the TR phase, the system behaves as a two-component Luttinger liquid. The response of the system to the magnetic field is rather usual, i.e., the zero field susceptibility shows a simple Pauli law. In the MS phase, a magnetic field will enlarge the period of the Wigner-like lattice and at a critical field $`H_c`$, the singlet rungs are no longer stable even in the ground state, implying a phase transition between TR phase and MS phase. Two different situations may appear when a magnetic field is applied on the RD phase. Roughly speaking, the magnetic field depresses the effect of $`J`$. Therefore, with the increase of the field, the system flows either toward the TR phase or toward the MS phase. If the system flows to the MS under a magnetic field, some flux phase may appear. Here the triplet rungs with $`S_z=1`$ serve as the fluxes in the singlet-rung host. Since the “repulsive interaction” $`U`$ occurs only between the nearest neighbor singlet rungs, the only possible flux phase has the structure $`|0>_1|1>_2|0>_3\mathrm{}|0>_{N1}|1>_N`$, where $`|0>_i`$ and $`|1>_j`$ indicate the singlet rungs and the triplet rungs, respectively and $`N`$ (even) represents the length of the ladder. With the increase of the external field, the fluxes will clusterize and finally form a triplet-rung liquid or a completely polarized state.
In conclusion, a generalized spin ladder is studied. It is found that in the extremely frustrated case, a modulation structure which represents a Wigner-like lattice of the singlet rungs in the triplet-rung host can exist in some parameter regions. Though only the $`\beta =1`$ case is studied in detail, similar phenomena may exist for arbitrary $`\beta `$. The only difference is that the TR domains in $`\beta =1`$ case may be replaced by Haldane domains or VBS domains.
The author is indebted to the hospitality of Institut für Physik, Universität Augsburg, where this work was initiated. This work was partially supported by AvH-Stiftung, NSFC, FCAS and NOYSFC. |
no-problem/9911/nucl-th9911044.html | ar5iv | text | # Reaction mechanism and characteristics of 𝑇₂₀ in d + 3He backward elastic scattering at intermediate energies
## I Introduction
For the last few decades, elastic scattering of deuterons by protons at the backward angle at intermediate energies has intensively been studied as an important source of information of nuclear interactions and reaction dynamics, by including non-nucleonic degrees of freedom in the consideration . Since <sup>3</sup>He has the same spin as the proton’s, the spin structure of the scattering amplitude of the d+<sup>3</sup>He (d<sup>3</sup>He) system is similar to that of the d+p (dp) one when <sup>3</sup>He is considered as a single body . Then the backward elastic scattering of the deuteron by <sup>3</sup>He attracts our attention, for investigations of probable differences as well as similarities of the information when compared to the dp scattering.
In the dp backward scattering, the assumption of mechanism by neutron transfer from the deuteron to the proton (Fig. 1(a)) has been fundamentally successful in explaining energy dependence of observables. For example, the neutron-transfer mechanism produces the overall agreement to the experimental cross section $`\sigma `$ of the scattering up to GeV-energy region when the empirical momentum distribution of the deuteron is employed in the relativistic framework and the simple PWIA calculation by the mechanism describes qualitative features of the measured tensor analyzing power $`T_{20}`$ and polarization transfer coefficient $`K_y^y`$(d$``$p) at few hundreds MeV . Further, the measured $`\sigma `$ and $`T_{20}`$ of the scattering have broad resonance-like structures around $`E_d`$ = 1 GeV in the plot against the incident energy , which have been interpreted as the signal of excitations of the transferred neutron to $`\mathrm{\Delta }`$-states (Fig. 1(b)) . The PWIA calculation by the transfer mechanism has reproduced the observed structure of $`T_{20}`$ when effects of other reaction channels like the $`\mathrm{\Delta }`$ excitation have been phenomenologically included . For the transfer mechanism, the spin observables are described by the $`w(k)`$ to $`u(k)`$ ratio, where $`u(k)`$ and $`w(k)`$ are the deuteron $`S`$\- and $`D`$\- wave functions in the momentum space, and thus the information of the nuclear interaction obtained by the spin observables is generally of different nature from the one by the cross section .
In the case of the <sup>3</sup>He target, the possible reaction mechanism of the backward scattering of the deuteron will be proton transfer from <sup>3</sup>He to the deuteron (Fig. 1(c)) as follows. The backward cross section of direct scattering, for example by potentials, is small compared to the forward one by several order of magnitude at intermediate energies. On the other hand, the cross section of the proton transfer reaction is large for the forward emission of <sup>3</sup>He, i.e. the backward for the deuteron, as in usual pick-up reactions. When the transfer mechanism is adopted, the analogy to the dp scattering suggests that the spin observables of the d<sup>3</sup>He scattering in the PWIA are described by the $`w(k)`$ to $`u(k)`$ ratio, where $`u(k)`$ and $`w(k)`$ are now the wave functions of the p-d relative motion in <sup>3</sup>He. The wave functions, which are supplied by the Faddeev calculation, have ambiguities due to the choice of interactions as the input . The information of the $`w(k)`$ to $`u(k)`$ ratio obtained from the spin observables will be useful in solving such ambiguities.
Further, the spin structure of the scattering amplitude for the d<sup>3</sup>He system is similar to that for the dp one even in the nucleon transfer mechanism because in both scattering the spin of the projectile is transformed in the same way, from one to one half, and similarly that of the target from one half to one. Thus, the spin observables can be described in similar forms in these scattering and therefore some of their characteristics in the dp scattering will also be observed in the d<sup>3</sup>He one. These stimulate us to investigate the spin observables of the d<sup>3</sup>He scattering at the backward angle by theories as well as by experiments, which have not been attempted so far. In particular, $`T_{20}`$ will be investigated in detail because of its sensitivity to the wave functions. The theoretical prediction is performed up to GeV-energy region similarly to the dp case although the present measurements are limited to a few hundreds MeV. As the first step of investigations, we will try to clarify global features of $`T_{20}`$ in the scattering theoretically and to provide experimental evidence for the reaction mechanism. The interests of the investigation will be focused on (i) the validity of the assumption of the proton-transfer mechanism, (ii) effects due to the difference between the <sup>3</sup>He wave function and the deuteron one and (iii) the overview of effects of coupling to other reaction channels. In the following, experimental details are described in Sec. II, the PWIA analyses are given in Sec. III, where in addition to the cross section and the analyzing power the polarization transfer coefficient and the spin correlation one are briefly discussed, and in Sec. IV the effects of other reaction channels are investigated. Section V is devoted to a summary, discussion and perspective.
## II Experimental details and results
The d<sup>3</sup>He experiment was carried out at RIKEN Accelerator Research Facility for $`\sigma `$ and $`T_{20}`$. Polarized deuteron beam was provided by the high-intensity polarized ion source . Three polarization modes (unpolarized and two tensor-polarized modes) were cycled every 5 seconds. The beam polarization was measured with a polarimeter based on the $`d+p`$ elastic scattering after acceleration to 140, 200 and 270 MeV. It was monitored continuously during a run and obtained to be typically 60–80% of the ideal value. At $`E_d=`$200 and 270 MeV, detectors were placed at the angle where $`T_{20}`$ vanishes so that the polarimeter could work as a beam intensity monitor independent of beam polarization.
A cryogenic <sup>3</sup>He gas target was bombarded by the polarized deuteron beams. The size of the target was 13 mm (20 mm) wide, 15 mm (20 mm) high and 20 mm (10 mm) thick in the case of 270 MeV (140, 200 MeV) measurement. Entrance and exit windows were 6 $`\mu `$m thick Havar foils. The gas pressure ($`1`$ atm) was measured with a Baratron gauge. Temperature of the target was monitored by using a diode thermo-sensor and found to be about 11 K throughout the experiment. The density of the $`{}_{}{}^{3}\mathrm{He}`$ gas was $`6.6\times 10^{20}`$ cm<sup>-3</sup>.
Scattered <sup>3</sup>He particles were momentum-analyzed in the magnetic spectrograph SMART and detected by a multi-wire drift chamber and plastic scintillators placed at the focal plane. The total beam charge was measured with a Faraday cup placed in the first dipole magnet and the beam intensity was monitored by the polarimeter described above.
In the off-line analysis, data for $`\theta _{\mathrm{lab}}1.4^{}`$ were used. The background spectrum from the $`(d,{}_{}{}^{3}\mathrm{He})`$ reactions on Havar foils was subtracted to obtain the net yield for the d$`{}_{}{}^{3}\mathrm{He}`$ events. Signal-to-noise ratio at the peak region was 4–20% depending on the beam energy. The tensor analyzing power $`T_{20}`$ was deduced from the ratios of yields for three polarization modes.
The measured cross sections and analyzing powers with statistical errors are given in Table I. The systematic errors of $`\sigma `$ and $`T_{20}`$ are $`\pm 10\%`$ and $`\pm 2\%`$, respectively. The former is mainly due to the uncertainties of both the beam intensity and the target thickness, whereas the latter is due to the uncertainty of the beam polarization.
## III PWIA analyses
The theoretical analyses are performed first by the PWIA. The approximation is rather crude but is still useful to obtain the qualitative feature of the reaction at the intermediate energies . Analyses by extended formulas which include the effect of virtual excitations of other reaction channels will be presented in the next section.
The <sup>3</sup>He wave function in the pd cluster configuration is given by
$`\mathrm{\Psi }_{He}`$ $`=`$ $`{\displaystyle \underset{\nu _d}{}}\left({\displaystyle \frac{1}{2}}1\nu _p\nu _d|{\displaystyle \frac{1}{2}}\nu _{He}\right)\chi _{\nu _p}\chi _{\nu _d}{\displaystyle \frac{u(k)}{\sqrt{4\pi }}}`$ (2)
$`+{\displaystyle \underset{\nu _d,m}{}}\left({\displaystyle \frac{1}{2}}1\nu _p\nu _d|{\displaystyle \frac{3}{2}}\nu \right)\left({\displaystyle \frac{3}{2}}2\nu m|{\displaystyle \frac{1}{2}}\nu _{He}\right)\chi _{\nu _p}\chi _{\nu _d}Y_{2m}(\widehat{𝒌})w(k),`$
where $`𝒌`$ is the proton-deuteron relative momentum in <sup>3</sup>He, $`\chi _{\nu _p}`$ and $`\chi _{\nu _d}`$ are the wave functions of the proton and the deuteron, and $`\nu ^,`$s are the z-components of their spins. In the PWIA for the d<sup>3</sup>He scattering, the proton transfer is induced by proton-deuteron interactions as the neutron transfer by neutron-proton interactions in the dp scattering . We denote the proton-deuteron scattering amplitude at the momentum $`k`$ by $`t(k)`$, neglecting the spin dependence similarly to the dp case. Referring to the calculation of the dp scattering, we get non-vanishing independent T matrix elements $`\nu _{He}^{},\nu _d^{}|𝑴|\nu _{He},\nu _d`$ for the proton transfer in the d<sup>3</sup>He backward elastic scattering as
$$\frac{1}{2},1|𝑴|\frac{1}{2},1=\frac{t(k)}{4\pi }\left\{\frac{2}{3}u(k)^2+\frac{2\sqrt{2}}{3}u(k)w(k)+\frac{1}{3}w(k)^2\right\},$$
(3)
$$\frac{1}{2},0|𝑴|\frac{1}{2},1=\frac{1}{2},1|𝑴|\frac{1}{2},0=\frac{t(k)}{4\pi }\left\{\frac{\sqrt{2}}{3}u(k)^2\frac{1}{3}u(k)w(k)\frac{\sqrt{2}}{3}w(k)^2\right\},$$
(4)
$$\frac{1}{2},0|𝑴|\frac{1}{2},0=\frac{t(k)}{4\pi }\left\{\frac{1}{3}u(k)^2\frac{2\sqrt{2}}{3}u(k)w(k)+\frac{2}{3}w(k)^2\right\},$$
(5)
where $`k`$ is related to the incident deuteron momentum $`k_d`$ as $`k=\frac{1}{3}k_d`$ . The matrix elements (3)-(5) are equivalent to those in the dp backward scattering except for the factor $`\frac{2}{3}`$, when $`\nu _{He}`$ is replaced by $`\nu _p`$ and $`t(k)`$ by the n-p scattering amplitude. Then the cross section $`\sigma `$ and the tensor analyzing power $`T_{20}`$ are given as
$$\sigma =(\frac{\mu }{2\pi })^2\frac{1}{3}(\frac{|t(k)|}{4\pi })^2(u(k)^2+w(k)^2)^2$$
(6)
and
$$T_{20}=\frac{2\sqrt{2}rr^2}{\sqrt{2}(1+r^2)}\text{with}r=\frac{w(k)}{u(k)}.$$
(7)
Here $`\mu `$ is the reduced mass of the d<sup>3</sup>He system. Using Eqs. (3)-(5), one can also calculate other polarization observables, which are equivalent to those in the dp scattering. For example, the polarization transfer coefficient $`\kappa _0=\frac{3}{2}K_y^y`$(d$``$<sup>3</sup>He) and the spin correlation coefficient $`C_{yy}`$ are given as
$$\kappa _0=\frac{1}{1+r^2}\left(1\frac{1}{\sqrt{2}}rr^2\right)$$
(8)
and
$$C_{yy}=\frac{2}{9(1+r^2)^2}\left(1\frac{5}{\sqrt{2}}r+3r^2+\sqrt{2}r^32r^4\right).$$
(9)
The quantities $`T_{20}`$ and $`\kappa _0`$ satisfy the equation of a circle as in dp scattering,
$$(T_{20}+\frac{1}{2\sqrt{2}})^2+\kappa _{0}^{}{}_{}{}^{2}=\frac{9}{8}.$$
(10)
At the low-energy limit where $`r=0`$, $`T_{20}=0`$ and $`\kappa _0=1`$. With the increase of the incident energy, the magnitude of $`r`$ is increased and the point defined by a set of $`\kappa _0`$ and $`T_{20}`$ in the $`\kappa _0T_{20}`$ plane moves along the above circle clockwise for $`r<0`$ and counterclockwise for $`r>0`$. As will be seen later, the former is the case of the dp scattering and the latter that of the d<sup>3</sup>He one. More details are referred to Refs. .
For the Faddeev calculation of <sup>3</sup>He, we will specify the nucleon-nucleon (NN) force to the AV14 potential and the three-nucleon (3N) force to the 2$`\pi `$-exchange Brazil model . The Coulomb interaction is discarded and the <sup>3</sup>He nucleus is approximated by the triton. These potentials give the binding energy of the three-nucleon system as 8.34 (7.68) MeV with (without) the 3N force . In Fig. 2(a), the calculated $`u(k)`$ and $`w(k)`$ are shown as the functions of $`k`$ and are compared to those for the deuteron by the same NN force. The wave functions of <sup>3</sup>He have similar $`k`$-dependence to those of the deuteron in a global view but with the opposite sign of $`w(k)`$ for most $`k`$. Due to the different sign of $`w(k)`$, the calculated $`r`$ of <sup>3</sup>He has the opposite sign to that of the deuteron except at large $`k`$ as is shown in Fig. 2(b). This characteristic of $`r`$ is consistent with the result of Ref. that the asymptotic D-state to S-state ratio in <sup>3</sup>He has the opposite sign to that in the deuteron. The quantity $`r`$ is infinite at the zero point of $`u(k)`$, $`k=k_{0u}`$, which is located at almost the same $`k`$ for <sup>3</sup>He and the deuteron, $`k_{0u}=0.40`$ GeV/c. The zero point of $`w(k)`$ for <sup>3</sup>He, $`k=k_{0w}`$, is located at about $`k_{0w}`$=0.79 GeV/c (0.96 GeV/c) with (without) the 3N force.
The analysis of the measured cross section will not fully been performed because the cross section depends on the proton-deuteron scattering amplitude $`t(k)`$ as is seen in Eq. (6), and reliable information of $`t(k)`$ is not available at the present. In Fig. 3, however, a crude estimation which assumes the $`k`$-dependence of $`t(k)`$ to be negligible for the relevant energies is presented. The calculation describes the global feature of the energy dependence of the measured cross section. This will encourage further investigations of the scattering by the assumption of the transfer mechanism.
The characteristics of $`r`$ in Fig. 2 are reflected to $`T_{20}`$ of the scattering as shown in Fig. 4(a), where comparisons will be made in two ways; one is the comparison of $`T_{20}`$ between the d<sup>3</sup>He scattering and the dp one and the other is that of $`T_{20}`$ between the calculated and the measured for the same scattering. Due to the opposite sign of $`r`$, in a small-$`k`$ region, the calculated $`T_{20}`$ for the d<sup>3</sup>He scattering has the opposite sign to that for the dp one. In the figure, the measured $`T_{20}`$ for the d<sup>3</sup>He scattering by the present experiment is shown together with that for the dp scattering in Ref. and their signs are opposite to each other. Since the dp and d<sup>3</sup>He scattering in the transfer model are essentially (d,p) stripping and (d,<sup>3</sup>He) pickup reactions, although the incident energies are higher than the conventional stripping and pickup reactions, the above feature is compared to that of measured $`T_{20}`$ for (d,p) and (d,t) reactions at low energies, where $`T_{20}`$ of the former reactions has the opposite sign to that of the latters in a wide angular range . This nature of the sign of the measured $`T_{20}`$ indicates that the proton-pickup mechanism is a reasonable model for the d<sup>3</sup>He backward scattering. In both of the d<sup>3</sup>He and dp scattering, the calculated $`T_{20}`$ at the small $`k`$ reproduces the qualitative feature of the $`k`$ dependence of the measured one, although the magnitudes of the calculated are larger than those of the measured. This qualitative agreement of the calculation to the experiment supports the assumption of the proton-transfer mechanism for the d<sup>3</sup>He scattering. In Fig. 4(b), as an example of calculations by other potentials, $`T_{20}`$ by the recently proposed NijmII potential is compared to that by the AV14 one. The calculated $`T_{20}`$ are very similar to each other up to $`k0.6`$ GeV/c and the difference due to the potential employed is seen at larger $`k`$. In the figure, the effect of the 3N force, which arises through high-momentum components of the <sup>3</sup>He wave function, is shown for the AV14 potential. In large $`k`$ region, the contribution of the 3N force to $`T_{20}`$ becomes large and the second zero point of $`T_{20}`$ is shifted from $`k`$=0.96 GeV/c to $`k`$=0.79 GeV/c due to the 3N force.
In Figs. 5(a) and 5(b), the calculated $`\kappa _0`$ and $`C_{yy}`$ are shown for the AV14 and NijmII potentials. The dependence of $`\kappa _0`$ and $`C_{yy}`$ on the choice of the potential is weak, up to about $`k0.6`$ GeV/c. The 3N force effect is shown in the figure but is less remarkable compared to that in $`T_{20}`$.
## IV Virtual excitation of other reaction channels
Now we will extend the theoretical framework to a more general one so as to include effects of other reaction channels . Let us first transform the original $`T`$-matrix elements $`\nu _{He}^{},\nu _d^{}|𝑴|\nu _{He},\nu _d`$ into $`U`$, $`T`$ and $`T^{}`$ as
$$U=\frac{9}{2\sqrt{2}}\frac{1}{2},1|𝑴|\frac{1}{2},1+3\frac{1}{2},0|𝑴|\frac{1}{2},1+\frac{3}{\sqrt{2}}\frac{1}{2},0|𝑴|\frac{1}{2},0,$$
(11)
$$T=2\frac{1}{2},1|𝑴|\frac{1}{2},1\sqrt{2}\frac{1}{2},0|𝑴|\frac{1}{2},12\frac{1}{2},0|𝑴|\frac{1}{2},0,$$
(12)
$$T^{}=\frac{1}{4}\frac{1}{2},1|𝑴|\frac{1}{2},1\frac{1}{\sqrt{2}}\frac{1}{2},0|𝑴|\frac{1}{2},1+\frac{1}{2}\frac{1}{2},0|𝑴|\frac{1}{2},0.$$
(13)
Here, $`U`$, $`T`$ and $`T^{}`$ are free from the PWIA in principle. It is shown by the invariant-amplitude method that $`U`$ is the scalar amplitude in the spin space and describes the scattering by central interactions and $`T`$ and $`T^{}`$, the second-rank tensor ones, describe the scattering by tensor interactions. In the PWIA limit, as will be seen later, contribution of the S-state of <sup>3</sup>He ($`u(k)^2`$) is included in $`U`$, that of the interference between the S-state and the D-state ($`u(k)w(k)`$) in $`T`$ and that of the D-state ($`w(k)^2`$) is divided into two amplitudes, $`U`$ and $`T^{}`$, according to their tensorial character. Secondly, we introduce the relative magnitudes and phases between $`U`$, $`T`$ and $`T^{}`$ as
$$\frac{|T|}{|U|}R,\frac{|T^{}|}{|U|}R^{},\theta _T\theta _U\mathrm{\Theta },\text{and}\theta _T^{}\theta _U\mathrm{\Theta }^{}.$$
(14)
Then we get the general form of $`T_{20}`$ in terms of $`R,R^{},\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ as
$$T_{20}=\{2\sqrt{2}R\mathrm{cos}\mathrm{\Theta }R^232R_{}^{}{}_{}{}^{2}+12RR^{}\mathrm{cos}(\mathrm{\Theta }^{}\mathrm{\Theta })\}/N_R,$$
(15)
$$N_R=\sqrt{2}+2\sqrt{2}R^2+34\sqrt{2}R_{}^{}{}_{}{}^{2}4R^{}\mathrm{cos}\mathrm{\Theta }^{},$$
(16)
which is exact except for the proton-transfer assumption. Since these formulas are the same as those in the dp case , we will follow their analyses.
We will calculate $`R`$ and $`R^{}`$ by the PWIA and treat $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ as the adjustable parameters. This treatment of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ takes into account effects of virtual excitations of other reaction channels phenomenologically since the virtual excitations of the channel induces imaginary parts in the transition amplitudes to vary $`\mathrm{\Theta }`$ and/or $`\mathrm{\Theta }^{}`$. These effects can be neglected in $`R`$ and $`R^{}`$ since the neglect has not induced significant errors in the calculation of $`T_{20}`$ of the dp scattering except at very large $`k`$ . The range of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ will generally be $`\pi \mathrm{\Theta }(\mathrm{\Theta }^{})\pi `$. In the following, $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ are treated to be $`k`$-independent for simplicity, although $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ vary with $`k`$ in principle. The amplitudes $`U`$, $`T`$ and $`T^{}`$ in the PWIA limit are calculated as
$$U=3\sqrt{2}(u(k)^2+\frac{1}{4}w(k)^2)t(k),$$
(17)
$$T=3\sqrt{2}u(k)w(k)t(k)$$
(18)
and
$$T^{}=\frac{3}{4}w(k)^2t(k).$$
(19)
Then $`R`$ and $`R^{}`$ are obtained as
$$R=\frac{4|r|}{4+r^2}\text{and}R^{}=\frac{r^2}{\sqrt{2}(4+r^2)}.$$
(20)
In addition, when we choose the PWIA limit for $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$, i.e. $`\mathrm{\Theta }=0`$ for $`k<k_{0u}`$ and $`k>k_{0w}`$ and $`\pi `$ for $`k_{0u}<k<k_{0w}`$, and $`\mathrm{\Theta }^{}=0`$, Eqs. (15), (16) and (20) give the previous result, Eq. (7). Since $`\mathrm{\Theta }`$ changes at $`k=k_{0u}`$ and $`k=k_{0w}`$, in the following we will identify $`\mathrm{\Theta }`$ by the value for $`k<k_{0u}`$, for simplicity.
By the numerical calculation, we will study the effects on $`T_{20}`$ by varying $`\mathrm{\Theta }`$ and/or $`\mathrm{\Theta }^{}`$, specifying the potential to the AV14 one. First we will examine the calculation without the 3N force. In Fig. 6(a), effects of variations of $`\mathrm{\Theta }`$ are described for typical $`\mathrm{\Theta }`$ by fixing $`\mathrm{\Theta }^{}`$ to the PWIA limit, where the result is shown for positive $`\mathrm{\Theta }`$ since $`T_{20}`$ in this case is independent of the sign of $`\mathrm{\Theta }`$ as seen in Eqs. (15) and (16). The $`k`$ dependence of $`T_{20}`$ varies remarkably with $`\mathrm{\Theta }`$. At $`k=k_{0u}(r=\mathrm{})`$ and $`k=k_{0w}(r=0)`$, $`T_{20}`$ are independent of $`\mathrm{\Theta }`$ and are $`\frac{1}{\sqrt{2}}`$ and zero, respectively. The calculation by $`\mathrm{\Theta }=60^{}`$ describes the present data very well and indicates the importance of the virtual excitation of other channels for the quantitative description of the data. In Fig. 6(b), we show the $`\mathrm{\Theta }^{}`$ dependence of $`T_{20}`$ with typical $`\mathrm{\Theta }^{}`$ by fixing $`\mathrm{\Theta }=0`$, where the result is shown for positive $`\mathrm{\Theta }^{}`$ because of the independence on the sign of $`\mathrm{\Theta }^{}`$ (see Eqs. (15) and (16)). None of the calculations with variations of $`\mathrm{\Theta }^{}`$ reproduces our data of $`T_{20}`$ in the small-$`k`$ region. However, the calculations for $`\mathrm{\Theta }^{}=120^{}`$ and $`180^{}`$ produce resonance-like structures around $`k=k_{0u}`$, which are similar to the structure found in the dp scattering observed at $`k=0.30.45`$ GeV/c in Fig. 4(a). There the calculation by Eqs. (15) and (16) with $`\mathrm{\Theta }=180^{}`$ (PWIA limit) and $`\mathrm{\Theta }^{}=120^{}`$ for the dp scattering is shown to exhibit that the virtual-excitation effect reproduces the structure.
As examples of combined effects of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$, we vary $`\mathrm{\Theta }^{}`$ fixing $`\mathrm{\Theta }=60^{}`$ in Fig. 6(c). The calculated $`T_{20}`$ in the small-$`k`$ region is little affected by the variation of $`\mathrm{\Theta }^{}`$, reproducing our data. On the other hand, the calculations by $`\mathrm{\Theta }^{}=60^{}`$, and $`120^{}`$ produce structures similar to those in Fig. 6(b). Therefore $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ play different roles in the calculation of $`T_{20}`$; i.e. $`T_{20}`$ in the small-$`k`$ region is mainly governed by the magnitude of $`\mathrm{\Theta }`$, while the resonance-like structure in the medium $`k`$ is produced by proper choices of $`\mathrm{\Theta }^{}`$. In Fig. 6(c), the calculated $`T_{20}`$ at $`k=k_{0u}`$ is concentrated in a narrow range of the magnitude of $`T_{20}`$. This is due to the weak dependence of $`T_{20}`$ on $`\mathrm{\Theta }^{}`$ at $`k=k_{0u}`$ as seen in Eqs. (15) and (16), and thus the measurement at $`k=k_{0u}`$ will provide less ambiguous examinations of the reaction mechanism. These features of $`T_{20}`$ are similar to those in the dp scattering .
In Figs. 6(a) and 6(b), we show the effect of the 3N force, as examples, for $`\mathrm{\Theta }=120^{}`$ in Fig. 6(a), and $`\mathrm{\Theta }^{}=120^{}`$ in Fig. 6(b). The qualitative nature of the effect is similar to that in the PWIA limit in Fig. 4(b); $`T_{20}`$ for large $`k`$ is remarkably affected by the 3N force, reflecting that the 3N force dominantly affects the p-d wave function at small distance. The second zero point of $`T_{20}`$, which is independent of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$, is moved to $`k=0.79`$ GeV/c from $`0.96`$ GeV/c by the force.
## V Summary, discussion and perspective
In the present paper, we have investigated the cross section and the spin observables in the d<sup>3</sup>He backward elastic scattering by the theory and the experiment. The theory which assumes the proton-transfer mechanism has predicted $`T_{20}`$, $`\kappa _0`$ and $`C_{yy}`$ in a wide range of the intermediate energies. In the low-energy region of the range, $`\sigma `$ and $`T_{20}`$ have been measured. The global feature of the energy dependence of $`\sigma `$ is reproduced by the PWIA calculation with the energy-independent p-d scattering amplitude and the measured $`T_{20}`$ is described qualitatively by the PWIA prediction. The <sup>3</sup>He wave function is obtained by solving the Faddeev equation. At small $`k`$, the sign of $`w(k)/u(k)`$ of <sup>3</sup>He is different from that of the deuteron and this explains the sign of the measured $`T_{20}`$ in the d<sup>3</sup>He scattering to be opposite to that in the dp scattering. These give strong support to the assumption that the dominant reaction mechanism is the proton transfer from <sup>3</sup>He to the deuteron. The quantitative difference between the measured $`T_{20}`$ and the PWIA one has been explained by including the effect of the coupling to other channels. More details will be discussed later. The formula of $`T_{20}`$ in the d<sup>3</sup>He scattering looks similar to the one in the dp scattering. However, the difference of the internal wave functions between <sup>3</sup>He and the deuteron produces the different features of $`T_{20}`$ for most $`k`$. The calculated $`T_{20}`$ depends weakly on the NN interaction employed, up to $`k0.6`$ GeV/c. The 3N force contributes to $`T_{20}`$ through the high-momentum components of the p-d wave functions in <sup>3</sup>He. The 3N forces, which have a long history since the pioneer works in Ref. , have recently been investigated in detail by analyzing cross sections, analyzing powers and some spin-correlation effects in the n-d scattering and the cross sections in the n-t one . The measurement of $`T_{20}`$ at large $`k`$ in the d<sup>3</sup>He backward elastic scattering will provide additional information of the 3N force.
The effect of the virtual excitation of other reaction channels is considered through the imaginary part of the scattering amplitudes, which are parameterized by $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$. The calculations with $`\mathrm{\Theta }=60^{}`$ are successful in reproducing the small-$`k`$ data. This effect will be interpreted as the virtual-breakup effects, mainly of the deuteron, by the analogy to the dp case. In the case of dp scattering the calculations by $`\mathrm{\Theta }=120^{}135^{}`$ which differ from the PWIA limit ($`\mathrm{\Theta }=180^{}`$) by $`60^{}45^{}`$ have described the measured $`T_{20}`$ at the small $`k`$ and this $`\mathrm{\Theta }`$ effect has been interpreted as the deuteron-breakup contribution, because below the pion threshold the breakup is the only one open channel strongly coupled to the dp one. In the present scattering, the large magnitude of $`\mathrm{\Theta }^{}`$ produces the resonance-like structure, which is similar to that observed in the dp scattering. If the virtual excitation of the transferred nucleon to the $`\mathrm{\Delta }`$-state (Fig. 1(b)) is responsible for producing the structure in the dp scattering, similar effects will also be expected in the d<sup>3</sup>He scattering as shown in Fig. 1(d). Thus it will be interesting to examine by experiments if such structures are observed. It should be noted that the effects of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ on $`T_{20}`$ have been studied by using the AV14 NN potential and the Brazil 3N force. However, the essential features of the effects will be unchanged by other choices of the nuclear interactions except in the large-$`k`$ region.
In the present investigation, the $`k`$ dependence of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ is not considered. Since they depend on $`k`$ in principle, a particular set of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ might be valid in a limited range of $`k`$. At present, $`\mathrm{\Theta }=60^{}`$ with any $`\mathrm{\Theta }^{}`$ reproduces the data between $`k=0.1`$ and $`0.2`$ GeV/c, indicating that such sets of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ are valid in this range of $`k`$. In the dp scattering, as shown in Fig. 4(a), a set of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ reproduces the global feature of the data in a wide range of $`k`$. This suggests a set of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$ to be applicable in a wide range of $`k`$ in the d<sup>3</sup>He scattering as well.
Finally, although numerical results are not displayed, the effects of virtual excitation of other channels on $`\kappa _0`$ and $`C_{yy}`$ are small at small $`k`$. However, their contributions have appreciable magnitudes at $`k0.2`$ GeV/c. Therefore, the measurements of these quantities in this $`k`$ region will give information of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$, such as the validity of the phenomenological value, $`\mathrm{\Theta }60^{}`$. Because the predicted $`T_{20}`$ has interesting features, experiments at higher energies are desirable, which will provide valuable information of the nuclear interaction and the reaction mechanism. In particular, the measurement of $`T_{20}`$ at $`k0.4`$ GeV/c will be one candidate to obtain the convincing evidence of the reaction mechanism because $`T_{20}`$ there is almost independent of $`\mathrm{\Theta }`$ and $`\mathrm{\Theta }^{}`$. Further refinements of the theory, which include relativistic effects, will be made when higher-energy data become available. |
no-problem/9911/nucl-ex9911011.html | ar5iv | text | # Entropy in hot 161,162Dy and 171,172Yb nuclei
## I Introduction
Already in 1936 H. Bethe introduced the Fermi gas to describe nuclear properties at high temperature . This simple independent particle model has been modified by including residual interactions between the nucleons. In the low excitation region long-range pair correlations play an important role and are roughly described within the so-called back-shifted Fermi gas model .
There is evidence for the existence of paired nucleons (Cooper pairs) at low temperature. In high spin physics, the backbending phenomenon is a beautiful manifestation of the breaking of pairs. The mechanism is induced by Coriolis forces tending to align single particle angular momenta along the nuclear rotational axis . Theoretical models also predict reduction in the pair correlations at higher temperatures .
The breaking of pairs is difficult to observe as a function of intrinsic excitation energy. Recent theoretical and experimental works indicate that the process of breaking pairs takes place over several MeV of excitation energy. The corresponding critical temperature is measured to be $`T_c`$ 0.5 MeV/$`k_B`$ , where $`k_B`$ is Boltzmann’s constant.
The aim here is to extract the entropy of the <sup>161,162</sup>Dy and <sup>171,172</sup>Yb isotopes, and deduce the number of excited quasi-particles as function of excitation energy. In Sect. 2 we describe the experimental techniques and analyzing tools. Sect. 3 presents results for the entropy using a simple pairing Hamiltonian with an even and odd number of fermions distributed over $`L`$ single-particle levels with double degeneracy. This is a model, which for small numbers of fermions, typically less than 20, can be solved numerically yielding all possible eigenstates. It results in the full level density and can in turn be used to extract thermodynamical quantities. Since we expect pairing correlations to be important in nuclei, such a simple model should mimic to a certain extent the entropy extracted from the experimental level density. In Sect. 4 we present the experimental findings and relate them to the simple pairing model of Sect. 3. Concluding remarks are given in Sect. 5.
## II Experimental methods
The Oslo cyclotron group has developed a method to extract nuclear level densities at low spin from measured $`\gamma `$-ray spectra . The main advantage of utilizing $`\gamma `$-rays as a probe for level density is that the nuclear system is likely thermalized prior to the $`\gamma `$-emission. In addition, the method allows for the simultaneous extraction of level density and $`\gamma `$-strength function over a wide energy region.
The experiments were carried out with 45 MeV <sup>3</sup>He-projectiles accelerated by the MC-35 cyclotron at the University of Oslo. The experimental data were recorded with the CACTUS multidetector array using the (<sup>3</sup>He,$`\alpha \gamma `$) reaction on <sup>162,163</sup>Dy and <sup>172,173</sup>Yb self-supporting targets. The charged ejectiles were detected with eight particle telescopes placed at an angle of 45 relative to the beam direction. Each telescope comprises one Si $`\mathrm{\Delta }E`$ front and one Si(Li) $`E`$ back detector with thicknesses of 140 and 3000 $`\mu `$m, respectively.
An array of 28 NaI $`\gamma `$-ray detectors with a total efficiency of $``$15% surrounded the target and particle detectors. In addition, two Ge detectors were used to monitor the spin distribution and selectivity of the reactions.
From the reaction kinematics the measured $`\alpha `$-particle energy can be transformed to excitation energy $`E`$. Thus, each coincident $`\gamma `$-ray can be assigned to a $`\gamma `$-cascade originating from a specific excitation energy. The data are sorted into a matrix of $`(E,E_\gamma )`$ energy pairs. At each excitation energy $`E`$ the NaI $`\gamma `$-ray spectra are unfolded , and this matrix is used to extract the first-generation (or primary) $`\gamma `$-ray matrix with the well-established subtraction technique of Ref. .
The resulting matrix $`P(E,E_\gamma )`$, which describes the primary $`\gamma `$-spectra obtained at initial excitation energy $`E`$, is factorized according to the Brink-Axel hypothesis by
$$P(E,E_\gamma )\rho (EE_\gamma )\sigma (E_\gamma ).$$
(1)
The assumptions and methods behind the factorization of this expression are described in Refs. , and only a short outline is given here.
Both the level density $`\rho `$ and the $`\gamma `$-energy dependent function $`\sigma `$ are unknown. In the new iteration procedure the first $`\rho ^0`$ function is simply taken as a straight line, and the first $`\sigma ^0`$ is calculated from Eq. (1). Then new $`\rho `$ and $`\sigma `$ functions are analytically calculated by minimizing the least square fit $`\chi ^2`$ to the data set $`P`$. This procedure is repeated until a global minimum is obtained with respect to the values at all ($`E`$, $`E_\gamma `$) pairs. About 50 iterations are necessary for fitting the $``$ 150 free parameters to the $``$ 1500 data points of $`P`$. Due to methodical problems in the first-generation procedure, we only use data with $`\gamma `$-energies $`E_\gamma >1`$ MeV and excitation energies $`E>`$ 2.5 and 4.0 MeV in the odd-even and even-even isotopes, respectively .
It has been shown that if one solution for $`\rho `$ and $`\sigma `$ is known, it is possible to construct infinitely many solutions with the same $`\chi ^2`$ using the substitution
$$\rho (EE_\gamma )A\mathrm{exp}[\alpha (EE_\gamma )]\rho (EE_\gamma )$$
(2)
and
$$\sigma (E_\gamma )B\mathrm{exp}(\alpha E_\gamma )\sigma (E_\gamma ),$$
(3)
where $`A`$, $`B`$ and $`\alpha `$ are arbitrary parameters. In the new product of these two functions, a factor $`AB\mathrm{exp}(\alpha E)`$ is left over, which is absorbed in $`P`$, since the sum $`_{E_\gamma }P(E,E_\gamma )`$ is undetermined.
In the case of <sup>162</sup>Dy, Fig. 1 demonstrates how the parameters $`A`$ and $`\alpha `$ are determined to obtain a level density function (data points) with correct number of levels around the ground state (histogram). In addition, the parameters reproduce the level density calculated from the spacing of neutron resonances at the neutron binding energy $`B_n`$, see insert of Fig. 1.
In the following, we concentrate only on the information given by the level density, which is assumed to be independent of particular $`\gamma `$-ray decay routes.
## III Entropy from a simple pairing model
The level density<sup>*</sup><sup>*</sup>*Hereafter we use $`\rho `$ for the level density in the microcanonical ensemble. Furthermore, since we are dealing with a system with discrete energies from a quantal system, the microcanonical partition function is defined by the number of states at a given energy $`E`$ . $`\rho `$ defines the partition function for the microcanonical ensemble, the latter being the appropriate one for statistical descriptions of isolated systems such as finite nuclei. The partition function for the canonical ensemble is related to that of the microcanonical ensemble through a Laplace transform
$$Z(\beta )=_0^{\mathrm{}}𝑑E\rho (E)\mathrm{exp}(\beta E).$$
(4)
Here we have defined $`\beta =1/k_BT`$, where $`T`$ is the temperature and $`k_B`$ is Boltzmann’s constant. Since we will deal with discrete energies, the Laplace transform of Eq. (4) takes the form
$$Z(\beta )=\underset{E}{}\mathrm{\Delta }E\rho (E)\mathrm{exp}(\beta E),$$
(5)
where $`\mathrm{\Delta }E`$ is the energy bin used.
In nuclear and solid state physics, thermal properties have mainly been studied in the canonical and grand-canonical ensemble. In order to obtain the level density, the inverse transformation
$$\rho (E)=\frac{1}{2\pi i}_i\mathrm{}^i\mathrm{}𝑑\beta Z(\beta )\mathrm{exp}(\beta E),$$
(6)
is normally used. Compared with Eq. (4), this transformation is rather tricky to perform since the integrand $`\mathrm{exp}\left(\beta E+\mathrm{ln}Z(\beta )\right)`$ is a rapidly varying function of the integration parameter. In order to obtain the density of states, approximations like the saddle-point method, viz., an expansion of the exponent in the integrand to second order around the equilibrium point and subsequent integration, have been used widely. For the ideal Fermi gas, this gives the following density of states
$$\rho _{\mathrm{ideal}}(E)=\frac{\mathrm{exp}(2\sqrt{aE})}{E\sqrt{48}},$$
(7)
where $`a`$ in nuclear physics is a factor typically of the order $`a=A/8`$ with dimension MeV<sup>-1</sup>, $`A`$ being the mass number of a given nucleus.
Ideally, the experiment should provide the level density as function of excitation energy and thereby the ’full’ partition function for the microcanonical ensemble. In the microcanonical ensemble we could then extract expectation values for thermodynamical quantities like temperature $`T`$, or the heat capacity $`C`$. The temperature in the microcanonical ensemble is defined as
$$T=\left(\frac{dS(E)}{dE}\right)^1.$$
(8)
It is a function of the excitation energy, which is the relevant variable of interest in the microcanonical ensemble. However, since the extracted level density is given only at discrete energies, the calculation of expectation values like $`T`$, involving derivatives of the partition function, is not reliable unless a strong smoothing over energies is performed. This case is discussed at large in Ref. and below. Another possibilityThe transformation to the canonical ensemble represents also a smoothing. is to employ the Laplace transformation of Eq. (5) in order to evaluate various thermodynamical quantities in the canonical ensemble. As an example, we can evaluate the entropy in the canonical ensemble using the definition of Helmholtz free energy
$$F(T)=k_BT\mathrm{ln}Z(T)=E(T)TS(T).$$
(9)
Note that the temperature $`T`$ is now the variable of interest and the energy $`E`$ is given by the expectation value $`E`$ as function of $`T`$. Similarly, the entropy $`S`$ is also a function of $`T`$.
In this section, we extract the exact level density from a simple theoretical modelA similar analysis within the framework of several BCS ansatz based approaches was done by Døssing et al. in Ref. . Whereas our approach includes all possible eigenvalues in order to determine the level density, Døssing et al. perform their diagonalization within a space spanned by number-projected states. The qualitative behavior of their results is however similar to that presented here.. The Hamiltonian we use to obtain the eigenvalues and the level density $`\rho (E)`$ is the simple pairing Hamiltonian
$$H=\underset{i}{}\epsilon _ia_i^{}a_i\frac{1}{2}G\underset{ij>0}{}a_i^{}a_{\overline{ı}}^{}a_{\overline{ȷ}}a_j,$$
(10)
where $`a^{}`$ and $`a`$ are fermion creation and annihilation operators, respectively. The indices $`i`$ and $`j`$ run over the number of levels $`L`$, and the label $`\overline{ı}`$ stands for a time-reversed state. The parameter $`G`$ is the strength of the pairing force while $`\epsilon _i`$ is the single-particle energy of level $`i`$.
We assume that the single-particle levels are equidistant with a fixed spacing $`d`$ such that Eq. (10) becomes
$$H=d\underset{i}{}ia_i^{}a_i\frac{1}{2}G\underset{ij>0}{}a_i^{}a_{\overline{ı}}^{}a_{\overline{ȷ}}a_j.$$
(11)
Moreover, in our simple model, the degeneracy of the single-particle levels is set to $`2J+1=2`$, with $`J=1/2`$ being the spin of the particle.
Introducing the pair-creation operator
$$S_i^+=a_{im}^{}a_{im}^{},$$
(12)
and
$$S_i^{}=a_{im}a_{im},$$
(13)
one can rewrite the Hamiltonian in Eq. (11) as
$$H=d\underset{i}{}iN_i\frac{1}{2}G\underset{ij>0}{}S_i^+S_j^{},$$
(14)
where
$$N_i=a_i^{}a_i$$
(15)
is the number operator. The latter commutes with the Hamiltonian $`H`$. In this model quantum numbers like seniority $`𝒮`$ are good quantum numbers, and the eigenvalue problem can be rewritten in terms of blocks with good seniority. Loosely speaking, the seniority quantum number $`𝒮`$ is equal to the number of unpaired particles, see Ref. for further details.
The reason why we focus on such a simple model is twofold. Firstly, we expect the ground state of nuclei to be largely dominated by pairing correlations. This is mainly due to the strong singlet $`{}_{}{}^{1}S_{0}^{}`$ state in the nucleon-nucleon interaction, see e.g., Refs. . For even-even systems this is typically reflected in an energy gap between the ground state and the first excited state, a gap that is larger than that seen in odd nuclei. This is taken as an evidence of strong pairing correlations in the ground state. More energy is needed in order to excite the system when all fermions are paired, i.e. when we have a system with an even number of particles. Since pairing correlations are important in nuclear systems, we expect the above model to exhibit at least some of the properties seen in finite nuclei.
Secondly, for particle numbers up to $`N20`$, the above model can be solved exactly through numerical diagonalization, since seniority is a good quantum number. This means that we can subdivide the full eigenvalue problem into minor blocks with given seniority and diagonalize these separately. In our case we use for the even system $`N=12`$ particles which are distributed over $`L=12`$ two-fold degenerate levels giving a total of
$$\left(\begin{array}{c}2L\\ N\end{array}\right)=\left(\begin{array}{c}24\\ 12\end{array}\right)=\mathrm{2.704.156}$$
(16)
states. Of this total, for $`𝒮`$ $`=0`$, i.e. no broken pairs, we have
$$\left(\begin{array}{c}L\\ N/2\end{array}\right)=\left(\begin{array}{c}12\\ 6\end{array}\right)=924,$$
(17)
states. Since the Hamiltonian does not connect states with different seniority $`𝒮`$, we can diagonalize a $`924\times 924`$ matrix and obtain all eigenvalues with $`𝒮`$ $`=0`$. Similarly, we can subdivide the Hamiltonian matrix into $`𝒮`$ $`=2`$, $`𝒮`$ $`=4`$,… and $`𝒮`$ $`=12`$ (all pairs broken) blocks and obtain all $`\mathrm{2.704.156}`$ eigenvalues for a system with $`L=12`$ levels and $`N=12`$ particles. As such, we have the exact density of levels and can compute observables like the entropy, heat capacity etc. This numerically solvable model enables us to compute exactly the entropy in the microcanonical and the canonical ensembles for systems with odd and even numbers of particles. In addition, varying the level spacing $`d`$ and the pairing strength $`G`$, may reveal features of e.g., the entropy which are similar to those of the experimentally extracted entropy. Recall that the nuclei studied represent both even-even and even-odd nucleon systems.
Here we study two systems in order to extract differences between odd and even systems, namely by fixing the number of doubly degenerated single-particle levels to $`L=12`$, whereas the number of particles is set to $`N=11`$ and $`N=12`$. Fig. 2 pictures the ground state of these systems and possible excited states.
These two systems result in a total of $`3\times 10^6`$ eigenstates. In the calculations, we choose a single-particle level spacing of $`d=0.1`$, which is close to what is expected for rare earth nuclei. In this sense, if we are to assign energies with dimension MeV, our results may show some similarity with experiment. We select three values of the pairing strength, namely $`G=1,0.2`$ and $`0.01`$, resulting in the ratio $`\delta =d/G=0.1`$, $`\delta =d/G=0.5`$ and $`\delta =d/G=10`$, respectively. The first case represents a strong pairing case, with almost degenerate single-particle levels. The second is an intermediate case where the level spacing is of the order of the pairing strength, while the last case results in a weak pairing case. As shown below, the results for the latter resemble to a certain extent those for an ideal gas.
The calculational procedure is rather straightforward. First we diagonalize the large Hamiltonian matrix (which is subdivided into seniority blocks) and obtain all eigenvalues $`E`$ for the odd and even particle case. This defines also the density of levels $`\rho (E)`$, the partition function and the entropy in the microcanonical ensemble. Thereafter we can obtain the partition function $`Z(T)`$ in the canonical ensemble through Eq. (5). The partition function $`Z(T)`$ enables us in turn to compute the entropy $`S(T)`$ using
$$S(T)=k_B\mathrm{ln}Z(T)+E(T)/T.$$
(18)
Since this is a model with a finite number of levels and particles, unless a certain smoothing is done, the microcanonical entropy may vary strongly from energy to energy. This is seen in Fig. 3, where we plot the entropy for the odd (upper part) and even (lower part) system using $`\delta =d/G=0.5`$. The entropy is given by discrete points, since we do not have eigenvalues at all energies. However, we can also perform a moderate smoothing which conserves the basic features of the model, namely an increase in entropy when pairs are broken. This was performed with a Gaussian smoothing
$$\stackrel{~}{S}_i=\frac{_kS_k\mathrm{exp}((E_iE_k)^2/2\sigma ^2)}{_k\mathrm{exp}((E_iE_k)^2/2\sigma ^2)}$$
(19)
where $`S_k`$ and $`E_{i,k}`$ are the entropies and energies from the diagonalization of the pairing Hamiltonian. $`\stackrel{~}{S}`$ is the smoothed entropy. With a smoothing parameter of $`\sigma =0.2`$ we see that the smoothed entropy still keeps track of the points where the entropy experiences an increase due to breaking of pairs.
Figure 3 clearly reveals the energies where two, three, four and so forth quasi-particles are created, i.e., where sudden increases in entropy take place. For the even system with the ground state at $`E_{GS}=2.44`$, the first seniority $`𝒮`$ $`=2`$ (formation of two quasi-particles) state appears at an excitation energy of $`E=2.2`$, the first $`𝒮`$ $`=4`$ state appears at $`E=4.06`$ and the first $`𝒮`$ $`=6`$ state is at $`E=5.41`$. Note well that in the figures of calculations we do not show the absolute energies. If we wish to employ dimensions in MeV, the first excited state for the even system would be close to what is expected experimentally.
For the odd system, the first excited states are just one-quasi-particle states, i.e., excitations of the last and least bound single-particle. Since the level spacing is much smaller around the ground state energy for the odd case (with energy $`E_{GS}=1.65`$), these states appear rather close to the ground state. When a pair is broken, we create a three-quasi-particle state (one broken pair plus a quasi-particle), or seniority $`𝒮`$ $`=3`$ state. This appears at an excitation energy<sup>§</sup><sup>§</sup>§Note that the first state with a broken pair appears at a lower excitation energy for the odd system, as expected. of $`E=2.01`$, whilst the seniority $`𝒮`$ $`=5`$ state (two broken pairs plus one quasi-particle) appears at $`E=3.58`$. We note from Fig. 3 that at an energy of $`E89`$, the entropy starts decreasing (population inversion), reflecting thereby the limited size of our model.
For $`\delta =0.5`$, where the single-particle spacing is only half the pairing strength, the energy eigenvalues are fairly well distributed over the given energy range. If we decrease $`\delta `$ however, we approach the degenerate limit, and the eigenvalues and the entropy are sharply concentrated around those eigenvalues where pairs are broken. This is seen in Fig. 4 for $`\delta =0.1`$ for the even case with $`N=12`$. The odd case with $`N=11`$ exhibits a similar behavior. Clearly, if we wish to evaluate the temperature according to Eq. (8) for $`\delta =0.1`$, even with a strong smoothing, we cannot obtain reliable values for e.g., $`T`$. Thus, rather than performing a certain smoothing, we will choose to present further results for the entropy in the canonical ensemble, using the Laplace transform of Eq. (5).
The results for the entropy in the canonical ensemble as functions of $`T`$ for the above three sets of $`\delta =d/G`$ are shown in Fig. 5. For the two cases with strong pairing, we see a clear difference in entropy between the odd and the even system. The difference in entropy between the odd and even systems can be easily understood from the fact that the lowest-lying states in the odd system involve simply the excitation of one single-particle to the first unoccupied single-particle state, and is interpreted as a single-quasi-particle state. These states are rather close in energy to the ground state and explain why the entropy for the odd system has a finite value already at low temperatures (recall also the discussion in connection with Fig. 3). Higher lying excited states include also breaking of pairs and can be described as three-, five- and more-quasi-particle states. For $`\delta =10`$, the odd and even systems merge together already at low temperatures, indicating that pairing correlations play a negligible role. For small single-particle spacing, also the difference in energy between the first excited state and the ground state for the odd system is rather small.
For our choice of $`d`$ we observe that the maximum entropy is of the order of $`S14k_B`$ in the canonical ensemble, whereas in the microcanonical ensemble, see Figs. 3 and 4, the maximum value is $`S1012k_B`$. Obviously, when performing the transformation to the canonical ensemble, since we have a small system, there may be larger fluctuations in expectation values like the entropy. In the limit $`N\mathrm{}`$, the two ensembles should result in equal values for $`T`$, $`E`$ and $`S`$, see Ref. for an in depth discussion.
For $`\delta =0.5`$ we note that at a temperature of $`k_BT0.50.6`$, the even and odd system approach each otherIf we wish to make contact with experiment, we could again assign units of MeV to $`k_BT`$ and $`E`$.. The temperature where this occurs corresponds to an excitation energy $`E`$ in the canonical ensemble of $`E4.75.0`$. Recalling Fig. 3, this corresponds to excitation energies where we have $`46`$ quasi-particles, seniority $`𝒮`$ $`=46`$, in the even system and $`57`$ quasi-particles, seniority $`𝒮`$ $`=57`$, in the odd system. The almost merging together of the even and odd systems at these temperatures, can be retraced to the features seen in Fig. 3. For higher excitation energies in Fig. 3, we saw that higher seniority values show less marked bumps in the entropy, indicating that the level density at higher excitation energies contains many more states and that we are getting closer to a phase where pairing plays a less significant role.
For small systems like finite nuclei, where the size of the system is not large compared to the range of the strong interaction, the entropy is not an extensive quantity, i.e., it does not scale with the size of the system . However, if we assume that the entropy is an extensive quantity, then $`S=nS_1`$, with $`n`$ the number of particles and $`S_1`$ the single-particle entropy in the canonical ensemble. In our case $`S_1`$ should correspond to the single-quasi-particle entropy. If we label the entropy excess $`\mathrm{\Delta }S`$ as the difference between the odd and even entropies, namely $`\mathrm{\Delta }S=S_{\mathrm{odd}}S_{\mathrm{even}}`$ with $`S_{\mathrm{odd}}=n_{\mathrm{odd}}S_1`$ and $`S_{\mathrm{even}}=n_{\mathrm{even}}S_1`$, we can in turn define the number of quasi-particles in the odd and even systems as
$$n_{\mathrm{odd}}(E)=\frac{S_{\mathrm{odd}}}{\mathrm{\Delta }S}\mathrm{and}n_{\mathrm{even}}(E)=\frac{S_{\mathrm{even}}}{\mathrm{\Delta }S},$$
(20)
respectively. The odd system has one more quasi-particle than the even system, i.e., $`n_{\mathrm{odd}}=n_{\mathrm{even}}+1`$.
Fig. 6 shows the number of quasi-particles in the odd and even systems for the three values of $`\delta `$ using the definition in Eq. (20). We note that for all cases the differences between the odd and even systems remain equal and close to one, demonstrating that the entropy is an extensive quantity as function of the number of quasi-particles. Furthermore, for $`\delta =0.5`$ (central panel), we see that the excitation energies where $`1,2,3,\mathrm{}`$ quasi-particles appear, agree with the results discussed in Fig. 3 in the microcanonical ensemble. To give an example, for the odd system, three quasi-particle appear at an energy of $`E=1.8`$, which should be compared with the exactly calculated one in the microcanonical ensemble of $`E=2.01`$. Five quasi-particles show up at $`E=3.4`$, which again should be compared with the result obtained in the microcanonical ensemble of $`E=3.58`$. The agreement for the even case is slightly worse. For $`\delta =0.1`$, the strong pairing case, we note that more energy is needed in order to create $`2,4,\mathrm{}`$ and $`3,5,\mathrm{}`$ quasi-particles in the even and odd systems, respectively. This agrees also with the microcanonical result of Fig. 4. For the weak pairing case $`\delta =10`$, higher seniority states appear already at low excitations energies, indicating that pairing plays a minor role, as expected.
Fig. 6 carries also an interesting message. If one can extract the number of quasi-particles as function of excitation energies, the steepness of the curve provides useful information about the relation between the single-particle spacing and the pairing strength.
In summary, varying $`\delta `$, allows us to extract qualitative informations about thermodynamical properties such as the entropy and the number of quasi-particles in even and odd systems. Especially, two properties are worth paying attention to concerning the discussion in the next section. Firstly, for the two cases with strong pairing ($`\delta =0.1`$ and $`\delta =0.5`$), Fig. 5 tells us that at temperatures where we have $`46`$ quasi-particles in the even system and $`57`$ quasi-particles in the odd system, the odd and even system tend to merge together. This reflects the fact that pairing correlations tend to be less important and we approach the non-interacting case. For the weak pairing case, $`\delta =10`$, the odd and even systems yield similar results at much lower temperatures. In a simple model with just pairing interactions, it is thus easy to see where, at given temperatures and excitation energies, certain degrees of freedom prevail. For the experimental results in the next section, this may not be the case since the interaction between nucleons is much more complicated. The hope however is that pairing may dominate at low excitation energies and that the features seen in e.g., Fig. 5 are qualitatively similar to the experimental ones.
Secondly, we can read from Fig. 6 the excitation energy where different numbers of quasi-particles appear. With a realistic value for the level spacing, a comparison with experiment may tell us something about the strength of the pairing force.
## IV Experimental results and discussion
The experimental level density $`\rho (E)`$ at excitation energy $`E`$ is proportional to the number of levels accessible in $`\gamma `$-decay. For the present reactions the spin distribution is centered around $`J`$ 4.4 $`\mathrm{}`$ with a standard deviation of $`\sigma _J`$ 2.4 $`\mathrm{}`$ . Hence, the entropyThe experiment reveals the level density and not the state density. Thus, also the observed entropy reveals the number of levels. The state density can be estimated by $`\rho _{\mathrm{state}}(2J+1)\rho _{\mathrm{level}}`$ 9.8 $`\rho _{\mathrm{level}}`$. can be deduced within the microcanonical ensemble, using
$$S(E)=k_B\mathrm{ln}N(E)=k_B\mathrm{ln}\frac{\rho (E)}{\rho _0},$$
(21)
where $`N`$ is the number of levels in the energy bin at energy $`E`$. The normalization factor $`\rho _0`$ can be determined from the ground state band in the even-even nuclei, where we have $`N(E)`$ 1 within a typical experimental energy bin of $``$ 0.1 MeV.
The extracted entropies for the <sup>161,162</sup>Dy and <sup>171,172</sup>Yb nuclei are shown in Figs. 7 and 8. In the transformation from level density to entropy we use Eq. (21) with $`\rho _0`$ 3 MeV<sup>-1</sup>. The entropy curves are rather linear, but with small oscillations or bumps superimposed. The curves terminate around 1 MeV below their respective neutron binding energies due to the experimental cut excluding $`\gamma `$-rays with $`E_\gamma <`$ 1 MeV. All four curves reach $`S`$ 13 $`k_B`$, which by extrapolation correspond to $`S`$ 15 $`k_B`$ at the neutron binding energy $`B_n`$. This is the maximum entropy that a nucleus in this mass region can reach before neutron emission.
The calculations for odd and even systems (see Fig. 3) show clear increases in the entropy at the excitation energies where Cooper pairs are broken. This behavior is not very pronounced in the experimental data, probably due to residual couplings in real nuclei. In particular, our pairing model excludes collective excitations, which are known to contribute strongly at low excitation energy. For <sup>172</sup>Yb in Fig. 8 one can identify bumps at 1.5 MeV and 2.8 MeV of excitation energy, that could be interpreted as increased entropy due to the breaking of two and four quasi-particles, respectively.
For the odd system the valence particle (or hole) is expected to perform blocking, and indeed the calculations of Fig. 3 reveal effects of smearing out the entropy structures as function of excitation energy. The smoother experimental entropy curves for <sup>161</sup>Dy and <sup>171</sup>Yb (see Figs. 7 and 8) seems also evident, in particular for the <sup>161</sup>Dy case.
The experimental entropy of the even-odd system follows closely the entropy for the even-even system, but the even-odd system has an entropy excess. The difference of entropy in the even-odd system compared to the even-even system is evaluated in Fig. 9 for <sup>161,162</sup>Dy and <sup>171,172</sup>Yb. The observed entropy excesses in the 1.5 MeV $`<E<`$ 5.5 MeV excitation region are $`\mathrm{\Delta }S`$ 1.8(1) $`k_B`$ and $``$ 1.6(1) $`k_B`$ for dysprosium and ytterbium, respectively. The calculations of Fig. 5 show that the entropy excesses are due to the additional degrees of freedom imposed by a valence particle (or hole). In the center panel of Fig. 5, using $`\delta =0.5`$, the entropy excess is $`\mathrm{\Delta }S3k_B`$ at $`k_BT00.3`$. However, the model is based on the density of states formed by 1/2 spin particles. The average total spin is $`J1`$ giving $`\rho _{\mathrm{level}}\rho _{\mathrm{state}}/3`$. Hence, the calculated entropy excess based on level density should be on the average $`\mathrm{\Delta }S(3\mathrm{ln}3)k_B2k_B`$, which is close to experiment. The temperature region up to 0.3 MeV corresponds roughly to the experimental excitation energy region discussed, and the model thus supports that excitation can be described by some sort of quasi-particles.
The experimental level density can be used to determine the canonical partition function $`Z(T)`$. However, in the evaluation of Eq. (5), we have to extrapolate the experimental $`\rho `$ curve to $``$ 40 MeV. Here, we use the back-shifted level density formula of Refs. with
$$\rho =f\frac{\mathrm{exp}[2\sqrt{aU}]}{12\sqrt{2}a^{1/4}U^{5/4}\sigma },$$
(22)
where the back-shifted energy is $`U=EE_1`$ and the spin cut-off parameter $`\sigma `$ is defined through $`\sigma ^2=0.0888A^{2/3}\sqrt{aU}`$. The level density parameter $`a`$ and the back-shift parameter $`E_1`$ are defined by $`a=0.21A^{0.87}`$ MeV<sup>-1</sup> and $`E_1=C_1+\mathrm{\Delta }`$, respectively, where the correction factor is given by $`C_1=6.6A^{0.32}`$ according to Ref. . The factor $`f`$ is introduced by us to adjust the theoretical level density to experiment at $`EB_n1`$ MeV. The parameters employed are listed in Table 1. From our semi-experimental partition function, the entropy can be determined from Eq. (18). The results are shown in Fig. 10. The entropy curves show a splitting at temperatures below $`k_BT=0.50.6`$ MeV which reflects the experimental splitting shown in the microcanonical plots of Figs. 7 and 8. However, the strong averaging produced by the summing in Eq. (5), modifies the entropy excess due to components from the theoretical extrapolation of $`\rho `$. Even so, the curves agree qualitatively with the calculations in Fig. 5 using $`\delta =0.5`$. The effect of pairing seems in both cases to vanish above $`0.50.6`$ MeV. This agrees with our previous work , giving a critical temperature of $`k_BT_c=`$ 0.5 MeV for the existence of pair correlations.
The observation that one quasi-particle carries 1.7 $`k_B`$ of entropy, can be utilized to estimate the number of quasi-particles as function of excitation energy. Analogously to Eq. (20), we estimate from the experimental entropies $`S_{\mathrm{eo}}`$ and $`S_{\mathrm{ee}}`$ in neighboring even-odd and even-even isotopes the entropy excess $`\mathrm{\Delta }S=S_{\mathrm{eo}}S_{\mathrm{ee}}`$. The number of quasi-particles in the even-odd and even-even systems is given by Eq. (20), except that the odd system is replaced by an odd-even nucleus and the even system by an even-even nucleus.
The extracted number of quasi-particles $`n(E)`$ in <sup>162</sup>Dy and <sup>172</sup>Yb is shown in Fig. 11. The number of quasi-particles raises to a level of $`n`$ 2 around $`E=1.52`$ MeV, which could be a signal for the formation of two quasi-particle states. However, the creation of four and six quasi-particles shows no clear step-like function. The breaking of additional pairs is spread out in excitation energy giving a rather smooth increase in the number of quasi-particles as function of excitation energy. In the excitation region $`0.55`$ MeV the $`n(E)`$ curve gives on the average 1.6 MeV of excitation energy to create a quasi-particle pair. This value is consistent with pairing gap parameters of this mass region, see Table 1. The theoretical calculation<sup>\**</sup><sup>\**</sup>\**The reader should keep in mind that the number of particles in the theoretical calculation and experiment are rather different. In experiment, if one assumes <sup>132</sup>Sn as closed shell core, the number of valence protons and neutrons is of the order of $`3040`$. However, performing the above theoretical calculations with say 10 or 14 particles results in qualitatively similar results as those presented here. with $`\delta =`$ 0.5 gives an energy of 1.7 MeV per broken pair, which is close to the experimental finding of 1.6 MeV. Hence, with a single-particle spacing of $`d=0.10.2`$ MeV, the pairing strength is determined to $`G=0.20.4`$ MeV.
## V Conclusions
The entropy as function of excitation energy has been extracted for the <sup>161,162</sup>Dy and <sup>171,172</sup>Yb nuclei. The observed entropy excess in the even-odd nuclei compared to the even-even nuclei is interpreted as the entropy for a single quasi-particle (particle or hole) outside an even-even core. The entropy excess remains at a level of $`\mathrm{\Delta }S`$ 1.7 $`k_B`$ as function of excitation energy. A simple pairing model with an equidistant level spacing of $`d`$ and a pairing strength of $`G`$, gives a qualitatively similar description of these features.
The number of excited quasi-particles has been extracted from data. The onset of two quasi-particle excitations seems evident; however, the breaking of additional pairs is smeared out in excitation energy and is difficult to observe. The maximum number of excited quasi-particles is measured to $`n`$ 6 at an excitation energy of 5.5 MeV in the <sup>162</sup>Dy and <sup>172</sup>Yb isotopes.
The quasi-particle picture has been a success in describing rotational bands in cold nuclei. The present results indicate that quasi-particles also can describe certain thermodynamical properties of hot nuclei. This gives hope for realistic modeling of nuclei up to high intrinsic energy with several quasi-particles excited.
The authors are grateful to E.A. Olsen and J. Wikne for providing the excellent experimental conditions. We wish to acknowledge the support from the Norwegian Research Council (NFR). |
no-problem/9911/math-ph9911005.html | ar5iv | text | # 1. Introduction
## 1. Introduction
Kramer has introduced the icosahedrally symmetric tiling of the (3dimensional) space by seven (proto)tiles . Sadoc and Mosseri have rebuilt these prototiles and reduced their number to four: $`z`$, $`h`$, $`s`$ and $`a`$. But the inflation class (inflation specie ) of the tilings of the space by the four Mosseri–Sadoc prototiles has lost the icosahedral symmetry . In Ref. a projection class (projection specie ) of the tilings $`𝒯^{(MS)}`$ by the Mosseri–Sadoc prototiles has been locally derived from the canonical icosahedrally projected (from the lattice $`D_6`$) local isomorphism class of the tilings $`𝒯^{(2F)}`$.
In this paper we derive the inflation rules for the decorated prototiles $`z`$, $`r`$, $`m`$, $`s`$ and $`a`$ of the projection class $`𝒯^{(MS)}`$. These rules enforce the inflation rules for the decorated Mosseri–Sadoc tiles $`z`$, $`h`$, $`s`$ and $`a`$ in agreement with those suggested by Sadoc and Mosseri , up to the decoration. Using the Dehn invariants for the prototiles (apart from the volumes of the prototiles) we reproduce algebraically the inflation matrix for the prototiles.
## 2. Inflation of Decorated Mosseri–Sadoc Tiles
Mosseri and Sadoc have given the inflation rules for their $`z`$, $`h`$, $`s`$ and $`a`$ tiles . These rules were for the Steininflation of the tiles. By definition, an inflation is a Steininflation if the inflated tiles are composed of the whole original tiles.
The inflation factor for the Mosseri–Sadoc tiles is $`\tau =\frac{1+\sqrt{5}}{2}`$. The inflation matrix of the Steininflation of the tiles is clearly the matrix with integer coefficients. It has been given by Sadoc and Mosseri
$`M=\left(\begin{array}{cccc}1& 1& 1& 1\\ 2& 1& 2& 2\\ 1& 1& 1& 2\\ 0& 0& 1& 2\end{array}\right),`$ (5)
in the following ordering of the tiles: $`z`$, $`h`$, $`s`$ and $`a`$.
In the case of the Mosseri–Sadoc tiles, the Steininflation is breaking the symmetry of the tiles. The authors of Ref. haven’t given a decoration of the tiles which would take care about the symmetry breaking and uniquely define the inflation–deflation procedure at every step.
In it has been shown that the projection class of the locally isomorphic tilings $`𝒯^{(2F)}`$ (the tilings of the 3dimensional space by the six tetrahedra with all edges parallel to the 2fold symmetry axis of an icosahedron, of two lengths, the standard one denoted by $``$2 , the “short” edge, and $`\tau `$ $``$2 , the “long” edge, ) by the “golden” tetrahedra can be locally transformed into the tilings $`𝒯^{(MS)}`$, $`𝒯^{(2F)}`$ $``$ $`𝒯^{(MS)}`$. The class $`𝒯^{(MS)}`$ of locally isomorphic tilings by Mosseri–Sadoc tiles has been defined by the projection . The important property is that the minimal packages of the six golden tetrahedra in $`𝒯^{(2F)}`$, such that their equilateral faces (orthogonal to the directions of the 3fold symmetry axis of an icosahedron) are covered, led to five tiles $`a`$, $`s`$, $`z`$, $`r`$ and $`m`$ .
In Fig. 1 we show how the decorated prototiles $`A^{}`$, $`B^{}`$, $`C^{}`$, $`D^{}`$, $`F^b`$, $`F^r`$, $`G^b`$ and $`G^r`$ of the projection class of tilings $`𝒯^{(2F)}`$ are locally transformed into the decorated composite tiles denoted by $`a`$, $`m`$, $`r`$, $`z`$ and $`s`$.
Moreover, the tiles $`r`$ and $`m`$ appear always as a union $`rm`$, that is, the tile $`h`$ of Sadoc and Mosseri with three mirror symmetries . In Fig. 2 we show how the decorated tiles $`r`$ and $`m`$ are locally transformed into a decorated tile $`h`$. Hence, the projection class of tilings $`𝒯^{(MS)}`$ has four decorated prototiles $`a`$, $`h`$, $`z`$ and $`s`$. The decoration is such that it breaks all the symmetries of the tiles.
From the inflation rules of the $`𝒯^{(2F)}`$ tiles $`A^{}`$, $`B^{}`$, $`C^{}`$, $`D^{}`$, $`F^b`$, $`F^r`$, $`G^b`$ and $`G^r`$ we derive the inflation rules for the five composite decorated prototiles $`a`$, $`m`$, $`r`$, $`z`$ and $`s`$ and present them in Figs. 3–7. Using the rule of the composition of the decorated tile $`h`$ given in the Fig. 2 one easily concludes how the decorated tile $`h`$ inflates. The derived inflation rules of the four prototiles $`a`$, $`h`$, $`z`$ and $`s`$ in the projection species $`𝒯^{(MS)}`$ are, up to a decoration, the same as those proposed by Sadoc and Mosseri . Hence, we identify the inflation and the projection species and denote them by the same symbol, $`𝒯^{(MS)}`$.
## 3. Dehn Invariants and Steininflation of Mosseri–Sadoc Tiles
The Dehn invariant of a polyhedron $`P`$ takes values in a ring $`𝐑𝐑_\pi `$ where $`𝐑_\pi `$ is the additive group of residues of real numbers modulo $`\pi `$; the tensor product is over $`𝐙`$, the ring of rational integers. Denote by $`l_i`$ the lengths of edges of $`P`$. Denote by $`\alpha _i`$ the corresponding lateral angles and by $`\overline{\alpha _i}`$ – the classes of $`\alpha _i`$ modulo $`\pi `$. The Dehn invariant, $`𝒟(P)`$, of $`P`$ is equal to
$$𝒟(P)=l_i\overline{\alpha _i},$$
(6)
with the sum over all edges of $`P`$. The important property of Dehn invariants is their additivity: if a polyhedron is cut in pieces, the Dehn invariant is the sum of the Dehn invariants of pieces. This property implies that the vector of Dehn invariants of tiles is an eigenvector of the inflation matrix (5) with an eigenvalue $`\tau `$.
The vector of Dehn invariants for the Mosseri–Sadoc tiles is:
$$\stackrel{}{d}_{MS}=𝒟\left(\begin{array}{c}z\\ h\\ s\\ a\end{array}\right)=5\left(\begin{array}{c}\tau \\ 2\\ \tau 1\\ \tau \end{array}\right)\overline{\alpha },$$
(7)
$`\mathrm{cos}\alpha `$ $`=`$ $`{\displaystyle \frac{\tau }{\tau +2}}={\displaystyle \frac{1}{\sqrt{5}}}.`$ (8)
Thus, the space of Dehn invariants of the Mosseri–Sadoc tiles is one–dimensional, there is only one independent lateral angle. For the Dehn invariants applied to the inflation, see Ref.
For the vector of volumes of the Mosseri–Sadoc tiles one obtains
$$\stackrel{}{v}_{MS}=\mathrm{Vol}\left(\begin{array}{c}z\\ h\\ s\\ a\end{array}\right)=\frac{1}{12}\left(\begin{array}{c}4\tau +2\\ 6\tau +4\\ 4\tau +3\\ 2\tau +1\end{array}\right).$$
(9)
We show that the inflation matrix for Mosseri–Sadoc tiles $`z`$, $`h`$, $`s`$ and $`a`$ can be uniquely reconstructed from the Dehn invariants (and the volumes).
Denote the inflation matrix by $`M_{MS}`$.
The vectors $`\stackrel{}{d}_{MS}`$ and $`\stackrel{}{v}_{MS}`$ are eigenvectors of the inflation matrix, with the eigenvalues $`\tau `$ and $`\tau ^3`$ correspondingly (the eigenvalue is equal to the inflation factor to the power which is the dimension of the corresponding invariant).
Explicitely,
$$M_{MS}\left(\begin{array}{c}4\tau +2\\ 6\tau +4\\ 4\tau +3\\ 2\tau +1\end{array}\right)=\left(\begin{array}{c}16\tau +10\\ 26\tau +16\\ 18\tau +11\\ 8\tau +5\end{array}\right)$$
(10)
for the vector of volumes and
$$M_{MS}\left(\begin{array}{c}\tau \\ 2\\ \tau 1\\ \tau \end{array}\right)=\left(\begin{array}{c}\tau +1\\ 2\tau \\ 1\\ \tau 1\end{array}\right)$$
(11)
for the the vector of Dehn invariants.
Assume that the inflation matrix is rational. Then, decomposing eqs. (10) and (11) in powers of $`\tau `$, we obtain four vector equations for $`M_{MS}`$ or a matrix equation
$$\begin{array}{cc}\hfill M_{MS}\left(\begin{array}{cccc}4& 2& 1& 0\\ 6& 4& 0& 2\\ 4& 3& 1& 1\\ 2& 1& 1& 0\end{array}\right)& \\ \hfill =\left(\begin{array}{cccc}16& 10& 1& 1\\ 26& 16& 2& 0\\ 18& 11& 0& 1\\ 8& 5& 1& 1\end{array}\right)& .\end{array}$$
(12)
The solution of this equation is unique and we find the matrix (5).
## Acknowledgements
The work of Z. Papadopolos was supported by the Deutsche Forschungsgemeinschaft. Z. Papadopolos is grateful for the hospitality to the center of the Theoretical Physics in Marseille, where this work has been started. The work of O. Ogievetsky was supported by the Procope grant 99082. We also thank the Geometry–Center at the University of Minnesota for making Geomview freely available. |
no-problem/9911/cond-mat9911004.html | ar5iv | text | # Conformation changes and folding of proteins mediated by Davidov’s soliton
## 1 Introduction
Many properties of biomolecules and their biological functioning can not be fully understood when only their chemical structure is taken into considerations. Nowadays, it is clear that biological activity has important physical and dynamical aspects. One concept which has already made a significant (although somewhat controversial) contribution to the physics of biological systems is that of solitons. The idea that solitons may provide the mechanism for energy and charge transport in proteins was developed by Davidov (1977, 1982). Since then, many papers that followed Davidov’s approach were published (for example: Olsen et al. 1989; Ciblis & Cosic 1997; Cruzeiro-Hansson & Takeno 1997; Förner 1997). In our recent paper we suggested that the folding of proteins to their native state and the conformational changes in folded proteins can both be associated with the propagation of solitons through the backbone of the protein (Caspi & Ben-Jacob 1998). Here we demonstrate that such a process can be mediated by explicit, Davidov’s-like solitons.
A protein is a polymer built from 20 different amino acids ordered in sequence. This sequence is called the primary structure of that protein. The protein’s actual three dimensional structure is, however, much more complex. Under normal conditions it folds to a predetermined and rather static structure. This so called native state can be described in terms of local secondary structure elements such as $`\alpha `$-helices and $`\beta `$-sheets, arranged in a global tertiary structure. Although the secondary and the tertiary structures are dictated by the amino acid sequence, it is still not known today how to predict the 3-D structure from a known sequence (Friesner & Gunn 1996). A possible approach to this problem could be to try and follow the dynamics of the folding process. Full molecular dynamics simulation of a detailed protein can not exceed, using existing computation hardware, the micro-second time scale, while the folding process may take up to minutes, and not less then milliseconds. Moreover, even the most elaborated simulations can not account for every detail of the actual physical molecule (they usually neglect quantum effects, for example). We can never escape the need to determine which properties of the system are essential and which have no significant influence. Therefore, simplified models were used in many cases (e.g. Camacho & Thirumalai 1993; Socci & Onuchic 1994; Shakhnovich 1994; Bruscolini 1997).
The folding dynamics may be viewed as a navigation process in the energy landscape of the protein conformational space. It is believed that the native structure corresponds to a small region in that space located in the vicinity of the ground state. As pointed out by Levinthal, a stochastic search for the ground state over a random landscape might take cosmological time. From this fact one may draw the conclusion that the landscape of natural proteins is not random, but rather directed towards the ground state (Wolynes et al. 1995). Overcoming the energetic barriers is usually ascribed to a stochastic and uncorrelated thermal activation (Karplus & Shakhnovich 1992). We suggested (Caspi & Ben-Jacob 1998) that other processes, which are correlated and deterministic, may take dominant roles in dictating the folding pathway. These processes involve the interaction between solitonic excitations and the conformational degrees of freedom of the protein. In our paper we presented a simple toy model in which we used a Sine-Gordon topological soliton. Here we shall demonstrate that folding can also be mediated using a more realistic non-topological Davidov’s soliton.
Even folded proteins do not always have a single static conformation. In fact, many biological processes in living organisms are associated with conformational changes. The most intensively studied examples are those of heme proteins such as myoglobin, which have different conformational states depending on whether they are bound or not to a CO or O<sub>2</sub> ligand (Frauenfelder 1988). Hemoglobin, for example has four subunits, each containing a heme group which is capable of absorbing an O<sub>2</sub> molecule. Actually, the oxygen binds cooperatively to the hemoglobin (Stryer 1995). The affinity for oxygen grows when there are already oxygen molecules bound. The thing that causes this change in the affinity is a global transition between two conformational states of the hemoglobin, which is induced by the bound oxygen. The exact details of this transition process are still not clear, but it is assumed that the bound oxygen and some external conditions (such as the pH, partial pressure of O<sub>2</sub> and CO<sub>2</sub>, etc.) cause a different conformational state to be energetically preferable. Transition from a metastable to a stable conformational state is usually treated as being thermally induced. We suggest that conformational transition of this sort may be mediated by solitons, via a mechanism similar to the one we describe for protein folding. We shall refer to the process in which conformational changes are induced by the propagation of a solitonic wave along the molecular chain as a Soliton Mediated Conformational Transition (SMCT).
Solitons are localized, non-dispersive excitations, which exist in many non-linear systems. They are very stable, and therefore can propagate without much energy loss or dispersion to much larger distances than wave-packets of linear waves. The role of solitons in the process of folding (or change of conformation) in proteins would be to provide an efficient mechanism for extracting the energy gained in a single event of local conformational transition and transferring it to a distant location. There, it may be used to activate another transition, and the energy released in that process could be extracted again. This picture is very different from the usual assumption that the energy released in each folding step dissipates to the environment. In a SMCT process, large sections of a protein can fold very fast (actually, instantaneously on a time scale of the entire global transition). Moreover, the folding process can be orchestrated deterministically. By this we do not mean that proteins have specific initial states. Indeed, the denatured state is an ensemble of many different conformational states. However, just has we can build an ordered brick wall from a an unordered pile of bricks by methodically placing the bricks at their designed places, so can an unordered conformation be transformed through a well defined path to an ordered one. We shall demonstrate this point by considering a random initial state in our model.
In proteins, the dipole-dipole interaction between neighbouring amide-I (the CO double bond) quantum modes of vibration gives rise to linear collective modes known as exitons. The dipole-dipole interaction is influenced, however, by lattice vibrations, so the exitons interact with acoustical phonones in the protein. This interaction introduces non-linearity into the system. Davidov used a variational approach and obtained equations for the local amplitudes of the quantum modes. In the continuous limit, the equations (after some transformations) can be approximated by the Non-linear Schrödinger equation (NLS):
$$i\dot{a}+a_{xx}+2|a|a=0,$$
(1)
where the variables were rescaled to non-dimensional units. There has been an intensive study of this equation in optics, where it is used to describe the propagation of a wave packet envelope in an optical fiber with a non-linear refractive index. It should be noted, however, that the “time” variable in the NLS equation for optical fibers is actually a large-scale spatial variable. In Davidov’s model, $`t`$ is strictly the time. The NLS equation has solitary wave solutions that look like wave-packets with a sech-shaped envelop. These are, in fact, non-topological solitons, which means that they have no topological constraint. Unlike the topological Sine-Gordon solitons, which have a fixed amplitude and finite rest-mass (the minimal energy needed to create a soliton), the NLS solitons have a velocity-dependent amplitude and can exist at any arbitrarily small energy.
Although the actual equations that describe the energy and charge transport in proteins are probably much more complex than the simple NLS equation, they all have similar properties so we shall use this equation to model the possible solitonic excitations. Davidov used a similar formulation to describe solitonic excitation along the hydrogen bonds in an $`\alpha `$-helix structure, and this is the model that is most frequently cited. We, however, consider here solitons that move along the backbone of the protein. As for Davidov’s model of charge conduction in proteins, we believe that topological solitons of the Sine-Gordon type, analogous to those in single-strand DNA (Hermon et al. 1998), are more appropriate in this case.
## 2 The model
Following our previous paper (Caspi & Ben-Jacob 1998), we shall introduce a toy-model inspired by the generic properties of protein molecules. Proteins have two local angles per residue (usually denoted $`\varphi `$ and $`\psi `$), which are relatively free, and may be considered at this stage as the most important degrees of freedom. For almost all amino acid residues in a polypeptide chain, there are two distinct minima of the local potential energy in the $`(\varphi ,\psi )`$ plane, corresponding to the $`\alpha `$ and the $`\beta `$ local conformation of the chain. The only exceptions are glycin, which has four minima, and proline which has only one. We shall use a scalar variable $`\varphi `$ to represent the local conformation of the protein. In the continuous limit, which we shall employ for simplicity, it corresponds to the local curvature of the molecule. The local potential energy will be simply modeled by an asymmetric $`\varphi ^4`$ double well potential, namely
$$V(\varphi )=\epsilon (\varphi +\delta )^2(\varphi ^2\frac{2}{3}\varphi \delta +\frac{1}{3}\varphi ^22),$$
(2)
where $`\delta `$ is the asymmetry parameter, ranging from $`1`$ to $`1`$. The two minima are positioned at $`\varphi =\pm 1`$. The energy difference between the minima is $`\mathrm{\Delta }E=\frac{16}{3}\epsilon \delta `$. The maxima is positioned at $`\varphi =\delta `$ and its energy is always zero.
We shall now assume that the amide-I vibrations could be influenced by the local conformation of the protein, and therefore the amplitude field $`a(x)`$ interacts with the conformation field $`\varphi `$. In order that solitons propagating through the protein backbone could mediate changes in its conformation, an interaction of an appropriate form should be introduced. Consider the following interaction potential:
$$u(a,\varphi )=\mathrm{\Lambda }|a|^2\varphi ^2,$$
(3)
where $`\mathrm{\Lambda }`$ is a positive parameter. If $`a`$ is non-zero, then, at the minima, the energy of the combined $`V`$ plus $`u`$ potential increases, which effectively lowers the barrier for a folding transition. We therefore expect that a sufficiently energetic soliton may enable transition from a metastable to a stable conformation.
The full Lagrangian density then reads:
$$=ia^{}\dot{a}|a_x|^2+|a|^4+\frac{1}{2}m\dot{\varphi }^2V(\varphi )u(a,\varphi ).$$
(4)
It should be emphasized that this Lagrangian describes only a “toy protein”. We would like to get a qualitative understanding of SMCT processes, rather than give a precise but intractable description of complicated interactions in a realistic model. We ignore global interactions at this initial stage. Those interactions can be incorporated later by letting the parameters of the local potential be dependent on the global conformation. Since we expect that the time scale of global conformational changes would be much longer than of the local SMCT events, the parameters should vary relatively slowly in time. If fact, we imagine that a fast local SMCT event is followed by a relaxation process in which the strains caused by the the local conformational transition are relaxed by a slow global transition. In our model we consider only the fast SMCT events in which the global conformation is almost constant. Our adiabatic approach is therefore justified.
From Lagrangian (4) we obtain the following equations of motion
$`i\dot{a}`$ $`=`$ $`a_{xx}2|a|^2a+\mathrm{\Lambda }\varphi ^2a`$ (5)
$`m\ddot{\varphi }`$ $`=`$ $`4\epsilon (\varphi +\delta )(\varphi ^21)2\mathrm{\Lambda }|a|^2\varphi \mathrm{\Gamma }\dot{\varphi },`$ (6)
that also include a dissipation term. These equations can be solved numerically. We considered initial conditions of a moving soliton, with $`\varphi =1`$ (local minima) for all x. Note that, as in our SG model (Caspi & Ben-Jacob 1998), the form of the interaction potential lowers the barrier between the two local minima of the conformation energy, so it enables transformation between the two states. Moreover, it is by no means obvious that enough energy, which is released during this transition, would be returned to the soliton. In order that the soliton could propagate further and the folding process could go on, the gained energy should balance the energy extracted by the conformation field as it overcomes the energy barrier and the energy lost to radiation of non-solitonic NLS modes induced by the interaction. Unlike the SG model in which the topological constraint ensured the stability of the soliton, here the soliton may be destroyed completely when it loses its energy.
## 3 Numerical results
For the numerical study of our model, it would be useful to define a collective coordinate and momentum for a soliton
$$Q=\frac{1}{M}_{\mathrm{}}^{\mathrm{}}𝑑xx|a|^2P=i_{\mathrm{}}^{\mathrm{}}𝑑x(a^{}a_xa_x^{}a),$$
(7)
where $`M=_{\mathrm{}}^{\mathrm{}}𝑑x|a|^2`$ is the soliton mass (it is a constant of motion). $`Q`$ and $`P`$ are in fact canonical variables which represent the soliton ‘center of mass’ and its conjugate momentum. The relation $`\dot{Q}=P/M`$ holds. In the exact NLS equation, $`P`$ (and therefore $`\dot{Q}`$) is a constant of motion. Here we use $`P`$ as an indicator for the soliton kinetic energy. Simulations reveal that when $`\delta =0`$ (i.e. symmetric barrier) the soliton slows down, its amplitude decreases, and eventually it disappears. Next, $`\delta `$ is decreased so that the $`\varphi =1`$ conformation becomes lower in energy. When the interaction is strong enough, and the soliton has enough energy (actually, when its initial momentum $`P`$ is large enough), the soliton transfers enough energy to the conformation angles to ensure dynamical activation of the folding transition to the local ground state ($`\varphi =1`$) in practically every point it passes (see fig. 1). However, the soliton extracts back some of the conformation energy. After a short period of time it reaches a steady state in which its average velocity is almost constant (it may have some periodic fluctuations), and the energy gained balances the energy lost. A similar result was also obtained with the SG model.
In an actual unfolded state local folding angles are supposed to be distributed somehow between $`\alpha `$ and $`\beta `$ conformation. We have checked the behaviour of a traveling soliton in disordered initial conditions, namely, when the $`\varphi `$ field is randomly distributed between the two minima. Still, a steady state velocity is reached, though lower compared with the ordered case, since less energy is gained as half of the conformation points are already at their ground state. The folding process is, nevertheless, extremely effective even in this case. The collective momentum of the soliton during the folding process is shown on the right side of fig. 2. Four stages of folding are shown on the left side of that figure. The field $`\varphi `$ was interpreted as the curvature (after some rescaling and shifting) of the toy-protein, where the two minima of the potential $`V(\varphi )`$ correspond to two distinct values of curvature. The transformation from a quasi random conformation to an ordered helical structure is mediated by a soliton moving in the upward direction.
## 4 Conclusions
In this paper we demonstrated that non-topological solitons can participate in a SMCT process, at least in our much simplified toy-model of proteins, which, nevertheless, captures some generic properties of the actual molecules. The continuous NLS equation was used as an approximate model for Davidov’s soliton. We introduced an appropriate interaction (of a rather natural form) between the solitons and the conformational degrees of freedom, which correspond to the $`\psi ,\varphi `$ angles of real proteins. The energy that was carried by the soliton is transfered into the conformation field, and this allows it to overcome energy barriers and reach the desired ground state. The soliton then receives back some of the energy that was gained in this process, which, at the “steady state”, balances its energy loss. This process can go on as long as enough local folding energy is available. Unlike a stochastic process, in which all the energy that is gained at a specific local fold transition dissipates, here a much more efficient mechanism allows a fast and well-controlled conformation transition.
We can now suggest a possible scenario for protein folding. This process may be composed from few fast SMCT events which cause creation of local structures (the precise nature of which is determined, nevertheless, also by global interactions and constraints). Every event by itself does not cause an immediate change of the global structure, since large amplitude movements of the entire heavy molecule are much slower. The global reorganization of the molecular conformation should therefore be a slow relaxation of strains which follows the SMCT. This sets up global conditions for a new SMCT process. Note that this scenario provides a rather deterministic folding pathway. Soliton creation is probably induced spontaneously by thermal excitation, or by an external agent such as ‘chaperone’ enzyme. It might be that solitons are generated in locations of preferable amino-acids sequences. Their propagation might be blocked on different sequences. Therefore, the sequence may not only dictate the final conformation, but also the dynamics of the conformational transition.
Conformation changes can probably be mediated by a single SMCT event, followed by slow relaxation. We should note that SMCT processes are always exothermic. Conformation changes which are associated with increase in energy and entropy, such as thermal protein denaturation, are not be explained by this model, but do not need complicated mechanism since denaturated state is not specific. The SMCT transition we described is always from a metastable to a stable state. In order for that transition to occur in the reverse direction, external interactions should raise the energy of the initial state above that of the target state. We shall elaborate on this matter elsewhere.
This research is partly supported by a GIF grant. |
no-problem/9911/hep-ph9911517.html | ar5iv | text | # References
CERN–TH/99-359
hep-ph/9911517
November 1999
A Note on New Sources of Gaugino Masses
Karim Benakli<sup>1</sup><sup>1</sup>1e-mail:Karim.Benakli@cern.ch
CERN Theory Division CH-1211, Geneva 23, Switzerland
ABSTRACT
In IIB orientifold models, the singlet twisted moduli appear in the tree-level gauge kinetic function. They might be responsible for generating gaugino masses if they acquire non–vanishing $`F`$-terms. We discuss some aspects of this new possibility, such as the size of gaugino masses and their non-universalities. A possible brane setting is presented to illustrate the usefulness of these new sources.
Supersymmetry breaking is a major issue in superstring and M-theory. It is for instance necessary to lift the degeneracy of vacua. For phenomenological applications, supersymmetry breaking will provide mass splitting between supersymmetric partners, explaining why these have not been observed in nature yet. The precise dynamics involved in the generation of such masses is still unknown, but one can use a phenomenological parametrization which turns out to be useful for many purposes dealing with low-energy predictions.
For weakly coupled heterotic strings, such a line of ideas was advocated in , where non–vanishing $`F`$-terms were assumed for the moduli fields $`S`$ (dilaton) and $`T_i`$ (associated to the Kahler structure of the compact internal space). The gauge groups originating from reduction of the ten-dimensional gauge symmetry have a universal tree-level coupling. Non-universalities of couplings and gaugino masses arise at one-loop through a $`T_i`$ dependence of threshold corrections.
Another convenient framework to pursue these investigations is provided by type IIB orientifolds. Soft terms for such compactifications have been discussed in . It was noticed that non-universal gaugino masses could be generated if for instance different parts of the standard model gauge group originated from different sets of branes . To allow unification of gauge couplings one would then need to construct models where the moduli, controlling the gauge couplings on different branes, get potentials with minima at the same value. Here we will address another origin for soft terms: twisted moduli related to blowing-up modes.
The IIB orientifolds are obtained as compactifications on three tori $`T^1,T^2,T^3`$ on which different points are identified under a discrete symmetry $`Z_N`$, which leads to a set of fixed points. Requiring $`𝒩=1`$ supersymmetry and Poincaré invariance in four dimensions allows the presence of 9- and 5-branes (equivalently under $`T`$-dualities 3- and 7-branes).
The space group action of the orbifold is defined by some twist eigenvector $`v=(v_1,v_2,v_3)`$. In the sector twisted by $`\theta ^k`$ the orbifold group acts as $`X_i\theta ^kX_i`$, $`\theta =\mathrm{exp}2\pi ivJ`$, where $`J=(J_1,J_2,J_3)`$, with $`X_i`$ and $`J_i`$ the coordinate and generator of rotation in the $`i`$-th torus respectively. For a given twist $`\theta ^k`$ one finds $`_{i=1}^34\mathrm{sin}^2\pi kv_i`$ fixed points that we label by an index $`f`$. In a similar way, the orbifold acts also on the Chan-Paton factors through some twist parametrized by a vector $`V_a`$ with model-dependent fractional entries $`l/N`$. In the case of even $`N`$, some sets of $`D5_i`$-branes are present, sitting at the origin $`X_j=X_k=0`$ in the $`ji`$ and $`ki`$ complex planes. We label by an index $`p_i`$ the $`4\mathrm{sin}^2\pi kv_i`$ fixed points located in the world-volume of the $`D5_i`$-branes.
In addition to the dilaton $`S`$ and to the three moduli $`T_i`$, $`i=1,2,3`$, parametrizing the Kahler structure (volume) of the tori, there are the twisted moduli $`Y_f^k`$ associated to blowing-up the orbifold singularities $`f`$ due to a twist $`\theta ^k`$ <sup>2</sup><sup>2</sup>2 We have changed notation from the usual $`M_f^k`$ to avoid confusion with masses.. The new feature in IIB orientifolds is that these moduli couple at tree-level to gauge kinetic terms. The gauge kinetic functions for gauge fields on the D9- and D5-branes are given by
$`f_b^9`$ $`=`$ $`S+{\displaystyle \frac{1}{N}}{\displaystyle \underset{k=1}{\overset{[N1/2]}{}}}{\displaystyle \frac{\mathrm{cos}2\pi kV_b^9}{_{i=1}^32\mathrm{sin}\pi kv_i}}{\displaystyle \underset{f}{}}Y_f^k`$
$`f_{ia}^5`$ $`=`$ $`T_i+{\displaystyle \frac{1}{N}}{\displaystyle \underset{k=1}{\overset{[N1/2]}{}}}{\displaystyle \frac{cos2\pi kV_{ia}^5}{2\mathrm{sin}\pi kv_i}}{\displaystyle \underset{p_i}{}}Y_{p_i}^k.`$ (1)
In general (1) results in different independent linear combinations of $`Y_{ai}`$ of the $`Y_i^k`$ for each of the gauge kinetic function corresponding to gauge groups $`G_a`$. So $`F`$-terms for the twisted moduli $`Y_i^k`$ will be a new source of tree-level gaugino masses:
$$M_a=\underset{i}{}c_a^iM_{Y_{ai}},$$
(2)
where $`M_{Y_{ai}}`$ are the contributions of different $`Y_{ai}`$ and the coefficients $`c_a^i`$ are model-dependent. We see that the gaugino masses and the associated complex phases could be non-universal in these models.
The cases of odd $`N`$ lead to a drastic simplification. Only one linear combination, which we denote as $`Y`$, of the twisted moduli appears in the gauge kinetic function. The coefficient of the dependence for the group $`G_a`$ is given by the beta-functions $`b_a`$ of the running of the corresponding gauge coupling :
$$f_a^9=S+\frac{b_a}{2}Y.$$
(3)
In the absence of an $`F`$-term for $`S`$ but for $`Y`$, a tree–level gaugino mass proportional to the one-loop beta-function coefficient will be generated (using the convention of ):
$$M_a=\frac{\sqrt{3}}{2}\frac{b_ag_a^2}{16\pi ^2}m_{3/2}e^{\alpha _Y}(K_Y^Y)^{1/2}=\sqrt{\frac{3}{8}}\frac{b_ag_a^2}{16\pi ^2}m_{3/2}e^{\alpha _Y}$$
(4)
where we have used $`\mathrm{Re}f_a^9=8\pi ^2/g_a^2`$ with $`g_a`$ the four-dimensional gauge coupling. In (4), $`\alpha _Y`$ is the complex phase, $`K`$ is the Kahler potential which we assumed in the second equality to be given by $`(Y+\overline{Y}+\mathrm{})^2`$. The fact that the form of the gaugino masses is similar to a one-loop form can be traced back to the fact that the dependence on $`Y`$ is there to cure sigma-model anomalies . One-loop contributions to gaugino masses could be important in this case<sup>3</sup><sup>3</sup>3A nice discussion of such effects might be found for example in ..
The relation (4) means that the gauginos have non-universal masses but a unique phase. The beta-functions coefficients $`b_a`$ take into account all the states that are massless at the string scale. If these are identified with the low-energy ones, one then has the low-energy prediction:
$$\frac{M_3}{b_3}=\frac{M_2}{b_2}=\frac{M_1}{b_1},$$
(5)
where $`M_3`$, $`M_2`$ and $`M_1`$ are the gaugino masses associated with the $`SU(3)`$, $`SU(2)`$ and $`U(1)`$ factors of the standard model.
The presence of both $`F_S`$ and $`F_Y`$ will obviously lead to non-universal gaugino masses with two independent phases, one of which could be chosen to vanish. The $`F_S`$ is expected to dominate because of the coupling constant suppression of the $`F_Y`$.
Does this non-universality also mean that gauge unification is lost? The crucial issue here is that although we have used non-vanishing $`F_{Y_i^k}`$, we have made no assumption on the vacuum expectation values of $`Y_i^k`$ moduli themselves. In fact, to be more precise, the gauge kinetic function is given in the string basis by linear multiplets $`l`$ and $`y_{as}`$:
$`f_a`$ $`=`$ $`{\displaystyle \frac{1}{l}}+{\displaystyle \underset{s}{}}c_{as}y_{as},`$ (6)
where $`c_{as}`$ are model-dependent constants. Under linear-chiral duality, $`l`$ is associated with the dilaton while $`y_{as}`$ are associated with the $`Y_i^k`$ moduli. It was argued in that the latter modulus $`y_{as}`$ should have a vanishing<sup>4</sup><sup>4</sup>4Supersymmetry breaking could lead to vevs $`y_{as}`$, but these should remain very small to keep the orientifold picture valid. vev to be in the orientifold limit, where our results are valid. This ensures automatic unification.
Let us turn to some brane setting to illustrate how this new possibility can be useful.
In general one might have 9-branes and three types of $`5_i`$–branes corresponding to the different choices $`T^i`$ of the torus on which the $`5`$-branes are wrapping. There are three kind of charged states that originate from open strings stretched between 9-branes denoted as (99) states, those stretched between 5-branes denoted as ($`5_i5_j`$) and those stretched between 5-branes and 9-branes denoted as ($`5_i9`$). The (99), $`(5_i5_i)(55)_i`$ strings give rise to gauge vector multiplets of the corresponding gauge group $`G^9`$ and $`G_i^5`$, respectively. They also lead to chiral multiplets ($`99`$) and $`(55)_i`$ charged only under $`G^9`$ and $`G_i^5`$ respectively. In contrast, the ($`5_i5_j`$) lead to chiral fields charged under both $`G_i^5`$ and $`G_j^5`$, while ($`5_i9`$) open strings lead to chiral superfields charged under both $`G_i^5`$ and $`G^9`$.
Suppose that the standard model gauge symmetry originates from 9-branes. We also assume that there are two (or three, but the the last one plays no role) sets of D5-branes: $`5_1`$ located at $`X_2=X_3=0`$ and $`5_2`$ located at $`X_1=X_3=0`$. The gauge coupling on the two sets are given by:
$`f_{1a}^5`$ $`=`$ $`T_1+{\displaystyle \frac{1}{N}}{\displaystyle \underset{k=1}{\overset{[N1/2]}{}}}{\displaystyle \frac{cos2\pi kV_{1a}^5}{2\mathrm{sin}\pi kv_1}}{\displaystyle \underset{p1}{}}Y_{p1}^k`$
$`f_{2a}^5`$ $`=`$ $`T_2+{\displaystyle \frac{1}{N}}{\displaystyle \underset{k=1}{\overset{[N1/2]}{}}}{\displaystyle \frac{cos2\pi kV_{2a}^5}{2\mathrm{sin}\pi kv_2}}{\displaystyle \underset{p2}{}}Y_{p2}^k`$ (7)
Consider the $`5_1`$-brane to be a hidden sector where non-perturbative effects break supersymmetry breaking and generate $`F`$–terms for some of the $`Y_{p1}^k`$ moduli. This could arise from gaugino condensation (or string-scale breaking as the brane–antibrane models of if the string scale is at an intermediate region ), which leads to a potential which goes as $`e^{c/g_{1a}^2}`$, which depends on the $`Y_{p1}^k`$ and could lead to $`F`$-terms for the latter. The $`9`$-brane standard model gauge kinetic function involves all the twisted moduli and will thus have the corresponding gaugino masses generated at tree-level.
We identify the standard model matter fields as coming from ($`5_29`$) open strings. These feel only the $`Y_{p2}^k`$ moduli, which share with the $`Y_{p1}^k`$ set only the modulus $`Y_0^k`$ associated to the blowing-up mode of the origin $`X_1=X_2=X_3=0`$. If $`F_{Y_{p1}^k}0`$ for $`p_10`$ and $`F_{Y_0^k}=0`$, then the scalar soft masses will be generated at one-loop only, mediated by gaugino masses. This might provide a brane realization for the scenario<sup>5</sup><sup>5</sup>5The nice phenomenological peculiarities of soft terms as suggested in were also present in . The low-energy predictions are also similar to . I thank A. Pomarol for stressing these points to me. proposed in . However, here the gaugino masses are generically non-universal. A $`\mu `$-term of the same order as the gaugino masses will be generated through a Kahler potential if there is a coupling $`Y_{p2}^kH_1H_2`$. Such a term is expected for instance for the case of compactifications of the form $`(K3\times T^2)/\mathrm{\Gamma }`$ with a singular $`K3`$ and $`\mathrm{\Gamma }`$ a discrete symmetry as $`Z_N`$. Before compactification on $`T^2`$ and acting with $`\mathrm{\Gamma }`$, it is known from that there are couplings $`Y_q^kF^2`$ with $`F^2`$ the six-dimensional gauge field strength from the (99) sector and $`Y_q^k`$ are the twisted moduli associated with blowing up the $`K3`$ singularities. Now upon the reduction to some of the gauge-field components will lead to chiral fields in four dimensions that could be identified with the Higgs doublets. It is interesting to look for explicit string models with such properties.
In conclusion, we have seen that in addition to contributions from the dilaton $`S`$ and the moduli fields $`T_i`$, there might be new contributions from twisted moduli $`Y_f^k`$ corresponding to blowing-up modes for the singularities of IIB-orientifolds. These are generically present and allow an extension of new possiblities for soft-terms as generic non-universalities of masses and phases, as well as the possibility to naturally restrict the tree-level soft-terms to part of the spectrum while generating other masses at higher orders.
I am grateful to C. Kounnas, Y. Oz, A. Pomarol and A. Uranga for useful discussions. I wish also to thank G. Giudice and M. Quirós for comments on the manuscript. |
no-problem/9911/hep-ph9911423.html | ar5iv | text | # On the Mixing Amplitude of 𝐽/𝜓 and Vector Glueball 𝑂
## 1 Introduction
Among various solutions attempting to resolve the “$`\rho \pi `$ anomaly” in charmonium decays , the resonance enhancement model, as proposed by Hou and Soni and later generalized by Brodsky, Lepage, and Tuan , requires the existence of a vector glueball $`O`$ . Since both the rest energy $`m_Oi\mathrm{\Gamma }_O/2`$ and the mixing amplitudes $`f_{O\psi },f_{O\psi ^{}}`$ with charmonium $`J/\psi `$ and $`\psi ^{}`$ are needed in the analysis of “$`\rho \pi `$ anomaly ”, it is necessary to study these issues within a hadronic model and henceforth provide some quantitative information. A naive approach based on a constituent model of gluons was used earlier to study the glueball spectrum in the pure Yang-Mills gauge theory. In the present study , we extend this framework by including the quark-gluon interaction and calculate the mixing angle $`\theta _{O\psi }`$ within a nonrelativistic approximation.
## 2 Setup of the Problem
In the non-relativistic framework, we can calculate the mixing angle $`\mathrm{tan}\theta _{O\psi }(\stackrel{}{Q})`$ between the physical composites $`J/\psi `$, $`O`$, and un-mixed hadrons $`c\overline{c}`$ and $`ggg`$ via the evolution operator $`U(T,T)=e^{2iHT}`$,
$$\left(\begin{array}{c}|J/\psi (\stackrel{}{Q})\\ |O(\stackrel{}{Q})\end{array}\right)_{NR}=\left(\begin{array}{cc}\hfill \mathrm{cos}\theta _{O\psi }(\stackrel{}{Q})\mathrm{sin}\theta _{O\psi }(\stackrel{}{Q})& \\ \hfill \mathrm{sin}\theta _{O\psi }(\stackrel{}{Q})\mathrm{cos}\theta _{O\psi }(\stackrel{}{Q})& \end{array}\right)\left(\begin{array}{c}|c\overline{c}(\stackrel{}{Q})\\ |ggg(\stackrel{}{Q})\end{array}\right)_{NR},$$
(1)
$$\mathrm{tan}\theta _{O\psi }(\stackrel{}{Q})=\underset{T\mathrm{}e^{iϵ}}{lim}\frac{ggg(\stackrel{}{Q})|U(T,T)|c\overline{c}(\stackrel{}{Q})_{NR}}{c\overline{c}(\stackrel{}{Q})|U(T,T)|c\overline{c}(\stackrel{}{Q})_{NR}}ggg(\stackrel{}{Q})|U(T,T)|c\overline{c}(\stackrel{}{Q})_{NR}.$$
(2)
Neglecting the decay widths of both $`J/\psi `$ and $`O`$, the mixing angle $`\theta _{O\psi }`$ is then related to a relativistically normalized mixing amplitude $`f_{O\psi }`$, by
$`f_{O\psi }(q^2)`$ $`=`$ $`\left[(\gamma \delta \alpha \beta )q^2+(\alpha \beta m_{ggg}^2\gamma \delta m_{c\overline{c}}^2)\right]\mathrm{sin}\theta _{O\psi }\mathrm{cos}\theta _{O\psi }`$ (3)
$``$ $`\left[(\delta \beta )q^2+(\beta m_{ggg}^2\delta m_{c\overline{c}}^2)\right]\theta _{O\psi },`$ (4)
where in the second line we assume a small mixing angle and keep terms only to first order in $`\theta _{O\psi }`$, and the state-dependent normalization factors are
$$\begin{array}{cccccccccccc}\alpha (\stackrel{}{Q})\hfill & & \sqrt{\frac{\omega _{J/\psi }(\stackrel{}{Q})}{\omega _{c\overline{c}}(\stackrel{}{Q})}}\hfill & =& \left(\frac{\stackrel{}{Q}^2+m_{J/\psi }^2}{\stackrel{}{Q}^2+m_{c\overline{c}}^2}\right)^{1/4},\hfill & & \beta (\stackrel{}{Q})\hfill & & \sqrt{\frac{\omega _{J/\psi }(\stackrel{}{Q})}{\omega _{ggg}(\stackrel{}{Q})}}\hfill & =& \left(\frac{\stackrel{}{Q}^2+m_{J/\psi }^2}{\stackrel{}{Q}^2+m_{ggg}^2}\right)^{1/4},\hfill & \\ \gamma (\stackrel{}{Q})\hfill & & \sqrt{\frac{\omega _O(\stackrel{}{Q})}{\omega _{ggg}(\stackrel{}{Q})}}\hfill & =& \left(\frac{\stackrel{}{Q}^2+m_O^2}{\stackrel{}{Q}^2+m_{ggg}^2}\right)^{1/4},\hfill & & \delta (\stackrel{}{Q})\hfill & & \sqrt{\frac{\omega _O(\stackrel{}{Q})}{\omega _{c\overline{c}}(\stackrel{}{Q})}}\hfill & =& \left(\frac{\stackrel{}{Q}^2+m_O^2}{\stackrel{}{Q}^2+m_{c\overline{c}}^2}\right)^{1/4}.\hfill & \end{array}$$
(5)
With these definitions, our task is to calculate the transition amplitude Eq.(2) in the framework of constituent models of charm quarks and gluons. Specifically, we need to construct the state vectors of $`c\overline{c}`$ and $`ggg`$ and then calculate the annihilation amplitude between these constituents. The transition amplitude between $`c\overline{c}`$ and $`ggg`$ composites can then be expressed as a convolution of the bound state wave functions and the annihilation amplitude among quark/gluon constituents.
## 3 Details of the calculations
In this section, we outline the basic ingredients in our calculations of the transition amplitude. Our normalizations and conventions follow that of .
1. The state vector of the $`c\overline{c}`$ particle:
$`|c\overline{c}(\stackrel{}{P},\stackrel{}{\lambda })_{NR}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{V}}}{\displaystyle \frac{d^3\stackrel{}{p}_1}{(2\pi )^3\sqrt{2\omega _c(\stackrel{}{p}_1)}}\frac{d^3\stackrel{}{p}_2}{(2\pi )^3\sqrt{2\omega _c(\stackrel{}{p}_2)}}(2\pi )^3\delta ^{(3)}(\stackrel{}{P}\stackrel{}{p}_1\stackrel{}{p}_2)}`$ (6)
$`\times `$ $`\stackrel{~}{\mathrm{\Psi }}_{c\overline{c}}({\displaystyle \frac{\stackrel{}{p}_1\stackrel{}{p}_2}{2}})S_{rs}(\stackrel{}{\lambda })U^{\alpha \beta }b_r^\alpha (\stackrel{}{p}_1)d_s^\beta (\stackrel{}{p}_2)|0,`$
where $`\stackrel{~}{\mathrm{\Psi }}_{c\overline{c}}(\stackrel{}{p})`$ is the momentum space wave function of $`c\overline{c}`$. The spin and color wave function are given by $`S_{rs}(\stackrel{}{\lambda })\frac{1}{\sqrt{2}}\stackrel{}{\lambda }(\stackrel{}{\sigma }ϵ)_{rs}`$, $`U^{\alpha \beta }\frac{1}{\sqrt{3}}\delta ^{\alpha \beta }`$, respectively.
2. The state vector of the $`ggg`$ particle:
$`|ggg(\stackrel{}{K},\stackrel{}{\zeta })_{NR}`$ $``$ $`{\displaystyle \frac{1}{\sqrt{V}}}\left[\mathrm{\Pi }_{i=1}^3{\displaystyle \frac{d^3\stackrel{}{k}_i}{(2\pi )^3\sqrt{2\omega _g(\stackrel{}{k}_i)}}}\right](2\pi )^3\delta ^{(3)}(\stackrel{}{K}{\displaystyle \underset{i=1}{\overset{3}{}}}\stackrel{}{k}_i)`$ (7)
$`\times `$ $`\stackrel{~}{\mathrm{\Phi }}_{ggg}(\stackrel{}{k}_i;\stackrel{}{K})T_{mpq}(\stackrel{}{\zeta })V^{abc}{\displaystyle \frac{1}{\sqrt{6}}}G_m^a(\stackrel{}{k}_1)G_p^b(\stackrel{}{k}_2)G_q^c(\stackrel{}{k}_3)|0.`$
Without solving the three body problem, we take a trial wave function of vector glueball $`ggg`$,
$$\stackrel{~}{\mathrm{\Phi }}_{ggg}(\stackrel{}{k}_1,\stackrel{}{k}_2,\stackrel{}{k}_3;\stackrel{}{K})\left(\frac{6\pi }{a^2m_g^2}\right)^{3/2}\mathrm{exp}\left[\frac{1}{2a^2m_g^2}\left(\stackrel{}{k}_1\stackrel{}{k}_2\stackrel{}{k}_2\stackrel{}{k}_3\stackrel{}{k}_3\stackrel{}{k}_1+\frac{\stackrel{}{K}^2}{3}\right)\right],$$
(8)
and the variational parameter $`a`$ is fixed at $`0.64`$ to obtain a stable glueball spectrum . The totally symmetric color wave function is given by $`V^{abc}\sqrt{\frac{3}{40}}d^{abc}`$. Please refer to for the definitions of creation operators of the constituent gluons $`G_m^a(\stackrel{}{k})`$ and the spin wave function $`T_{mpq}(\stackrel{}{\zeta })`$.
3. The annihilation amplitude:
Using the Feynman rules for charm quarks and constituent gluons, we can write down the annihilation amplitude corresponding to the process $`c+\overline{c}ggg`$,
$$𝒜_{c\overline{c}ggg}=\overline{v}_n(\stackrel{}{p}_2,s)u_m(\stackrel{}{p}_1,r)G_{\mu i}(\stackrel{}{k}_1)G_{\nu j}(\stackrel{}{k}_2)G_{\rho l}(\stackrel{}{k}_3)S_{rs}(\stackrel{}{\lambda })T_{ijl}(\stackrel{}{\zeta })U^{\alpha \beta }V^{abc}A_{\alpha \beta ,nm}^{abc,\mu \nu \rho },$$
(9)
where
$`A_{\alpha \beta ,nm}^{abc,\mu \nu \rho }`$ $``$ $`(ig_s)^3\left[{\displaystyle \frac{\lambda ^a}{2}}{\displaystyle \frac{\lambda ^b}{2}}{\displaystyle \frac{\lambda ^c}{2}}\right]_{\alpha \beta }\left[\gamma ^\mu {\displaystyle \frac{i}{\mathit{}_1\mathit{}_1m_c}}\gamma ^\nu {\displaystyle \frac{i}{\mathit{}_1\mathit{}_1\mathit{}_2m_c}}\gamma ^\rho \right]_{nm}`$ (10)
$`+`$ 6 permutations ,
and the external legs for constituent gluons are defined as $`G^{\mu i}(\stackrel{}{k}_1)g^{\mu i}+\frac{k_1^\mu k_1^i}{m_g^2}`$.
To simplify the algebra, we perform a momentum expansion,
$`𝒜_{c\overline{c}ggg}(\stackrel{}{p}_i;\stackrel{}{k}_j)`$ $`=`$ $`𝒜_{c\overline{c}ggg}(\stackrel{}{p}_i=0;\stackrel{}{k}_j=0)+\left(\stackrel{}{p}_i{\displaystyle \frac{}{\stackrel{}{p}_i}}+\stackrel{}{k}_j{\displaystyle \frac{}{\stackrel{}{k}_j}}\right)𝒜_{c\overline{c}ggg}|_{\stackrel{}{p}_i=0;\stackrel{}{k}_j=0}`$ (11)
$`+`$ $`\text{higher order terms in }{\displaystyle \frac{p_i}{m_c}},{\displaystyle \frac{k_j}{m_g}}{\displaystyle \frac{1.06g_s^3}{\sqrt{3}m_g(m_g2m_c)}}.`$
4. Mixing angle as a convolution:
A generalized Fermi’s Golden Rules No.2 can be applied to the transition amplitude and the mixing angle formula Eq.(2). With appropriate normalizations, we have
$`\mathrm{tan}\theta _{O\psi }(\stackrel{}{P})={\displaystyle \frac{d^3\stackrel{}{p}_1}{(2\pi )^3\sqrt{2\omega _c(\stackrel{}{p}_1)}}\frac{d^3\stackrel{}{k}_1}{(2\pi )^3\sqrt{2\omega _g(\stackrel{}{k}_1)}}\frac{d^3\stackrel{}{k}_2}{(2\pi )^3\sqrt{2\omega _g(\stackrel{}{k}_2)}}}`$
$`\times (2\pi )\delta ({\displaystyle \underset{i=1}{\overset{2}{}}}\omega _c(\stackrel{}{p}_i){\displaystyle \underset{j=1}{\overset{3}{}}}\omega _g(\stackrel{}{k}_j)){\displaystyle \frac{\stackrel{~}{\mathrm{\Psi }}_{c\overline{c}}(\stackrel{}{p}_1)𝒜(\stackrel{}{p}_1;\stackrel{}{k}_1,\stackrel{}{k}_2)\stackrel{~}{\mathrm{\Phi }}_{ggg}^{}(\stackrel{}{k}_1,\stackrel{}{k}_2)}{\sqrt{6}\sqrt{2\omega _c(\stackrel{}{P}\stackrel{}{p}_1)}\sqrt{2\omega _g(\stackrel{}{P}\stackrel{}{k}_1\stackrel{}{k}_2)}}}.`$ (12)
Since our model is based on a nonrelativistic approximation, we can perform a momentum expansion on a convoluted wave function of the vector glueball $`ggg`$, $`d^3k𝒜\stackrel{~}{\mathrm{\Phi }}_{ggg}^{}`$. The momentum integration of the wave function of the $`c\overline{c}`$ then generates a factor of $`c\overline{c}`$ spatial wave function at origin $`g_s^3\mathrm{\Psi }_{c\overline{c}}(0)`$, which can be extracted from the total hadronic decay rate of $`J/\psi `$ . Using the mass inputs $`m_c=1.5\text{ GeV}`$ and $`m_g=0.7\text{ GeV}`$, and the variational parameter $`a=0.64`$, our calculation give a value of mixing angle with $`\mathrm{tan}\theta _{O\psi }(\stackrel{}{Q}=0)0.015`$.
5. Systematics of our approximations:
In this mixing angle calculations, we have made several approximations:
1. Only nearest states, namely, $`c\overline{c}`$ and $`ggg`$, are considered.
2. Mixing dynamics between quarkonium and glueball is treated perturbatively in the strong coupling constant $`\alpha _s`$.
3. Assuming a nonrelativistic picture, we expand the annihilation amplitude
$`𝒜_{c\overline{c}ggg}`$ in powers of constituent momenta and keep leading term only.
4. A variational solution of the vector glueball wave function is used for the convolution formula Eq.(12).
For a discussion of possible improvement, see our paper for details.
## 4 Summary and conclusions
In this paper, we study the mixing angle $`\mathrm{tan}\theta _{O\psi }`$ and the mixing amplitude $`f_{O\psi }`$ between $`c\overline{c}`$ and vector glueball $`ggg`$ in the potential models of heavy quarks and constituent gluons, including perturbative dynamics of QCD. From this model calculation, we get the mixing angle at $`|\mathrm{tan}\theta _{O\psi }|0.015`$.
If we take the $`J/\psi `$ mass at $`m_{c\overline{c}}m_{J/\psi }3096\text{ MeV}`$ and glueball mass at $`m_{ggg}m_O3168\text{ MeV}`$, the mixing amplitude $`f_{O\psi }(m_{J/\psi }^2)`$, as converted from the mixing angle, is equal to $`0.008\text{ GeV}^2`$. In comparison with the phenomenological analysis , our results are off by a factor of two, which lies in the ballpark within our approximations.
It is unlikely to be a fortuitous coincidence that a naive picture of constituent gluons can give reasonable estimates for both glueball mass spectrum and mixing with quarkonium state, as the later quantity is more sensitive to the actual shape of the glueball wave function. At this stage our result seems to be encouraging, and indicates that a nonrelativistic approximation and more importantly, the constituent gluon picture, does capture some grains of truth behind this phenomenological puzzle. |
no-problem/9911/hep-ex9911022.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Quantum chromodynamics (QCD) predicts the existence of bound states of gluons, called glueballs. Lattice calculations have predicted the lightest scalar glueball to be a scalar resonance with mass $`1600\pm 150\text{ MeV/}c^2`$ with tensor and pseudoscalar glueballs in the $`20002500\text{ MeV/}c^2`$ region .
Experimentally two principal candidates for the scalar glueball have been observed, the $`f_0(1500)`$ and the $`f_J(1710)`$ (a review is given in ). Both are seen in gluon-rich reactions such as $`p\overline{p}`$ annihilation, central production and radiative J/$`\psi `$ decay. The $`f_0(1500)`$ does not fit naturally in the $`q\overline{q}`$ spectrum but could be due to a glueball mixed with $`q\overline{q}`$ states in that mass region .
The spin of the $`f_J(1710)`$ is not yet confirmed, indications for both spin 2 and spin 0 have been reported . If the $`f_J(1710)`$ is indeed spin 0 it becomes a candidate for the $`{}_{}{}^{3}P_{0}^{}(s\overline{s})`$-like state in the $`16002000\text{ MeV/}c^2`$ region, but also for the lightest scalar glueball .
In $`\gamma \gamma `$ interactions, production of a pure gluon state is suppressed. Measuring the two-photon width $`\mathrm{\Gamma }_{\gamma \gamma }`$ of the $`f_0(1500)`$ and the $`f_J(1710)`$, or setting an upper limit on $`\mathrm{\Gamma }_{\gamma \gamma }`$, should indicate whether either is likely to be a pure glueball or has quark content. The CLEO collaboration has recently published a stringent limit on the two-photon width of the $`f_J(2220)`$ resonance (formerly the $`\xi (2220)`$) which is a candidate for the tensor glueball .
At the LEP $`e^+e^{}`$ collider there is a large cross-section for inelastic two-photon scattering, in which each of the incoming electrons acts as a source of virtual photons. In this analysis, the processes $`\gamma \gamma f_0(1500)\pi ^+\pi ^{}`$ and $`\gamma \gamma f_J(1710)\pi ^+\pi ^{}`$ have been studied and upper limits extracted for $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})`$ of both resonances.
## 2 The ALEPH detector
The ALEPH detector and its performance are described in detail elsewhere . Here only a brief description of the detector components relevant for this analysis is given.
The trajectories of charged particles are measured with a silicon vertex detector (VDET), a cylindrical multiwire drift chamber (ITC) and a large time projection chamber (TPC). The three detectors are immersed in a 1.5 T axial magnetic field from the superconducting solenoidal coil and together provide a transverse momentum resolution $`\delta p_{}/p_{}=6\times 10^4p_{}`$ 0.005 ($`p_{}`$ in GeV/$`c`$). The TPC also provides up to 338 measurements of ionisation (dE/dx) used for particle identification.
Between the TPC and the coil, an electromagnetic calorimeter (ECAL) is used to identify electrons and photons and to measure their energy with a relative resolution of 0.18/$`\sqrt{E}+0.009`$ ($`E`$ in GeV). The luminosity calorimeters (LCAL and SiCAL) cover the small polar angle region, 24–190 mrad.
Muons are identified by their characteristic penetration pattern in the hadron calorimeter (HCAL), a 1.2 m thick iron yoke instrumented with 23 layers of limited streamer tubes, together with two surrounding layers of muon chambers. In association with the electromagnetic calorimeter, the hadron calorimeter also provides a measurement of the energy of charged and neutral hadrons with a relative resolution of 0.85/$`\sqrt{E}`$ ($`E`$ in GeV).
The main ALEPH trigger relevant for this analysis, for data taken in 1994 and after, is based on the identification of two track candidates in the ITC, with at least one track pointing to an energy deposit in excess of 200 MeV in the ECAL. The two track candidates are confirmed at the second trigger level using the TPC. For data taken before 1994, the analysis relies on a trigger of two ITC (and TPC) track candidates measured back-to-back within $`11^{}`$.
## 3 Monte Carlo samples
Fully simulated Monte Carlo event samples reconstructed with the same program as the data have been used for the design of the event selection, background estimation and extraction of a limit on two-photon widths $`\mathrm{\Gamma }_{\gamma \gamma }^{f_0(1500)}`$ and $`\mathrm{\Gamma }_{\gamma \gamma }^{f_J(1710)}`$. Samples of the signal processes $`\gamma \gamma f_0(1500)\pi ^+\pi ^{}`$ and $`\gamma \gamma f_J(1710)\pi ^+\pi ^{}`$ were generated using PHOT02 . Each sample was generated as a Breit-Wigner resonance of appropriate mass, width and spin, with the $`f_J(1710)`$ here assumed to be spin zero. The experimental resolution of the masses of both resonances is of the order of $`10\text{ MeV/}c^2`$, compared to full widths in excess of $`100\text{ MeV/}c^2`$.
Expected background processes $`\gamma \gamma e^+e^{}`$, $`\gamma \gamma \mu ^+\mu ^{}`$, and $`\gamma \gamma \tau ^+\tau ^{}`$ were also generated using PHOT02, while background from $`e^+e^{}\mathrm{Z}\tau ^+\tau ^{}`$ was estimated using KORALZ . Both programs are interfaced to the JETSET package for hadronisation, while KORALZ also includes TAUOLA for correct handling of the $`\tau `$ polarisation.
## 4 Event selection
This analysis uses 160.9 $`\mathrm{p}b^1`$ of data taken around $`\sqrt{s}=91`$ GeV from 1990 to 1995. Candidate events for $`\gamma \gamma \pi ^+\pi ^{}`$ are selected according to the following criteria:
* the event contains only two good tracks, of equal and opposite charge. Good tracks are defined to have at least four TPC hits, $`|\mathrm{cos}\theta |<0.93`$, where $`\theta `$ is the polar angle with respect to beam axis, and a minimum distance to the interaction point of less than 8 cm along the beam axis and 2 cm in the radial direction;
* the total energy summed over all objects reconstructed by the energy flow program is equal to the total energy of the two charged tracks;
* the primary vertex is reconstructed within $`\pm 2`$ cm of the nominal interaction point along the beam axis;
* the total energy observed in the event is less than 30 GeV and visible mass less than 10 GeV, to exclude Z events;
* the transverse momentum of the final state with respect to the beam axis does not exceed $`0.1\text{ GeV/}c`$;
* the absolute value of the cosine of the angle $`\theta ^{}`$ between the two tracks in their centre-of-mass system and the boost direction is required to be less than 0.9.
These criteria ensure (quasi-)real photon ($`Q^2=0`$) collisions, and also suppress three- or more body decay.
The dominant background to the two-pion signal is the $`\gamma \gamma \mu ^+\mu ^{}`$ process. At low energies, muons and pions cannot easily be distinguished from each other and the high cross-section for the $`\gamma \gamma \mu ^+\mu ^{}`$ process leads to a substantial contamination of the $`\pi ^+\pi ^{}`$ sample.
To reduce background, events containing identified muons are rejected. With a sample of $`\gamma \gamma \mu ^+\mu ^{}`$ events generated using PHOT02, it was found that the efficiency of rejection is about 100% for events with invariant mass $`W>3\text{ GeV/}c^2`$. Below $`W=3\text{ GeV/}c^2`$, the rejection efficiency is improved by means of simple cuts on the fraction of energy deposited in the ECAL and HCAL, but still the rejection efficiency falls off rapidly, being only $`20`$% for $`W=1\text{ GeV/}c^2`$. For $`W1\text{ GeV/}c^2`$ the average rejection efficiency for $`\gamma \gamma \mu ^+\mu ^{}`$ events is 45%.
After muon rejection, the specific ionisation (dE/dx) of each track, measured in the TPC, is required to be within three standard deviations of the expected ionisation for a pion and more than three standard deviations from that of an electron. Separation of pions and kaons by dE/dx is less efficient, but the rate of kaon pair production is low: from a Monte Carlo sample of $`f_2^{}(1525)K^+K^{}`$, a residual contamination of only $`0.05\%`$ was estimated.
Possible background from beam-gas interactions was investigated. Constraining the reconstructed primary vertex position in $`z`$ removes most of this background. As beam-gas events are uniformly distributed in $`z`$, the number of events falling outside the $`z`$-vertex cut can be used to estimate the remaining background inside the cut: beam-gas contamination is found to be negligible. Residual backgrounds from the processes $`\gamma \gamma \tau ^+\tau ^{}`$ and $`Z\tau ^+\tau ^{}`$ are also negligible.
A sample of 294141 $`\pi ^+\pi ^{}`$ candidate events is selected.
## 5 Fitting the mass spectrum
The $`\pi ^+\pi ^{}`$ invariant mass spectrum obtained is shown in Fig. 1. The steep rise of the spectrum around $`0.5\text{ GeV/}c^2`$ is an artefact of the trigger efficiency. The clear peak in the spectrum above $`1\text{ GeV/}c^2`$ can be identified as the known tensor resonance $`f_2(1270)`$.
No other structure is observed above the $`f_2(1270)`$ peak.
A fit to the invariant mass spectrum is performed. For the resonances $`f_2(1270)`$, $`f_0(1500)`$ and $`f_J(1710)`$, a Breit-Wigner function of the form
$$\frac{mm_0\mathrm{\Gamma }(m)}{(m_0^2m^2)^2+m^2\mathrm{\Gamma }^2(m)}$$
(1)
is used, where the mass-dependent width $`\mathrm{\Gamma }(m)`$, away from the two-pion threshold, has the form
$$\mathrm{\Gamma }(m)=\mathrm{\Gamma }_0\left(\frac{m}{m_0}\right)^{(l+1/2)}\mathrm{exp}\left[\frac{(m^2m_0^2)}{48\beta ^2}\right]$$
(2)
and $`\beta =0.4\text{ GeV/}c^2`$ . For the $`f_2(1270)`$ the mass $`m_0`$, total width $`\mathrm{\Gamma }_0`$ and overall normalisation are free parameters. For the $`f_0(1500)`$ and the $`f_J(1710)`$ the normalisation is a free parameter, while the mass and width are fixed to $`1500\text{ MeV/}c^2`$ and $`112\text{ MeV/}c^2`$ for the $`f_0(1500)`$ and $`1712\text{ MeV/}c^2`$ and $`133\text{ MeV/}c^2`$ for the $`f_J(1710)`$ . The $`f_J(1710)`$ is considered here as a $`J=0`$ state. Although the presence of two objects in the $`1700\text{ MeV/}c^2`$ region has been suggested , no attempt to resolve two objects is made here.
The background spectrum, due to $`\gamma \gamma \mu \mu `$ events and the $`\pi \pi `$ continuum process, is fitted on the data with a $`5^{th}`$ order Chebyshev polynomial.
The mass region 0.8 to $`2.5\text{ GeV/}c^2`$ is used in the fit (below $`0.8\text{ GeV/}c^2`$ the trigger efficiency falls to less than 20%). Four fits are performed, each including a Breit-Wigner for the $`f_2(1270)`$ and the polynomial background and then including
* no additional resonances,
* the $`f_0(1500)`$,
* the $`f_J(1710)`$,
* both the $`f_0(1500)`$ and the $`f_J(1710)`$.
The fit to the data is shown in Fig. 2 for case (i): the Breit-Wigner shape for the $`f_2(1270)`$, the polynomial for background processes and the sum of these are indicated.
The parameters of all fits are summarised in Table 1, including the number of events fitted for the glueball candidate signals. For the fit including the $`f_2(1270)`$ alone, the $`\chi ^2`$ of 75 for 76 degrees of freedom is already very good. The $`\chi ^2`$ per degree of freedom hardly changes with the addition of glueball signals. The fits including the $`f_J(1710)`$ were also performed for $`J=2`$. The results of these fits were essentially the same as for the $`J=0`$ case.
The width of the Breit-Wigner function fitted for the $`f_2(1270)`$ is in all cases in agreement with the world average value of $`185.5_{2.7}^{+3.8}\text{ MeV/}c^2`$ . The fitted mass of $`1214\text{ MeV/}c^2`$ in each case is not consistent with the established value of $`1275.0\pm 1.2\text{ MeV/}c^2`$ , however. This has been previously observed by the MARKII and CELLO collaborations and is believed to be caused by interference of the spin 2 resonant amplitude with other components in the background.
Limited knowledge of the trigger efficiency for this topology prevents an investigation of interference effects using the measured angular distribution. The number of events fitted for the $`f_0(1500)`$ signal is negative, but consistent with zero. Data from the WA76 and WA102 collaborations and recent calculations suggest interference effects in this region can cause the $`f_0(1500)`$ to appear as a dip in the spectrum. However, here limits are set assuming no interference.
The CELLO collaboration has reported possible structure in the $`\pi ^+\pi ^{}`$ invariant mass spectrum around $`1.1\text{ GeV/}c^2`$ . If the fit of case (i) is repeated but excluding a region around $`1.1\text{ GeV/}c^2`$, a clear excess of data over the fitted curve extrapolated through that region can be seen (Fig. 3a). The excess can be described by the introduction of another spin 0 resonance of apparent mass $`1.1\text{ GeV/}c^2`$ and total width $`250\text{ MeV/}c^2`$ (Fig. 3b).
## 6 $`𝚪_{𝜸𝜸}\mathbf{}𝓑𝓡\mathbf{(}𝝅^\mathbf{+}𝝅^{\mathbf{}}\mathbf{)}`$ for $`𝒇_\mathrm{𝟎}\mathbf{(}\mathrm{𝟏𝟓𝟎𝟎}\mathbf{)}`$ and $`𝒇_𝑱\mathbf{(}\mathrm{𝟏𝟕𝟏𝟎}\mathbf{)}`$
The fitted numbers of signal events from the processes $`\gamma \gamma f_0(1500)\pi ^+\pi ^{}`$ and $`\gamma \gamma f_J(1710)\pi ^+\pi ^{}`$ are used to calculate upper limits on $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})`$ for both resonances.
The trigger efficiency is calculated as a function of invariant mass by comparing rates of independent triggers for $`\pi ^+\pi ^{}`$ selected data. In the $`1500\text{ MeV/}c^2`$ mass region the average trigger efficiency is ($`66\pm 7`$)%, while for the $`1700\text{ MeV/}c^2`$ mass region it is ($`77\pm 7`$)%.
The acceptance and selection efficiency for $`\gamma \gamma f_0(1500)\pi ^+\pi ^{}`$ and for $`\gamma \gamma f_J(1710)\pi ^+\pi ^{}`$ are determined from Monte Carlo to be ($`17.5\pm 0.4`$)% and ($`16.3\pm 0.4`$)%, where the quoted errors include the systematic error due to the simulation of the detector (checked by varying the resolution on measured quantities used in the event selection) and the error due to Monte Carlo statistics.
The product branching ratio, $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})`$, of each resonance is calculated directly from the fitted number of signal events in each of the cases (ii)–(iv). The fitted value of $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})`$ (which may be negative) and its error are used to define the mean and width of a Gaussian distribution: the area under the Gaussian is integrated above zero to obtain the 95% C.L. limit on $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})`$.
The upper limits on the product branching ratios at 95% C.L. are, for the individual fits, $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})<0.25`$ keV and $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})<0.59`$ keV for the $`f_0(1500)`$ and $`f_J(1710)`$ respectively, and for the combined fit are $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})<0.31`$ keV and $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})<0.55`$ keV for the $`f_0(1500)`$ and $`f_J(1710)`$ respectively. If an additional spin 0 resonance with a fixed mass of $`1.1\text{ GeV/}c^2`$ (as discussed in section 5) is included in the combined fit, the limit on the product branching ratio for the $`f_0(1500)`$ worsens by $`30\%`$ and for the $`f_J(1710)`$ tightens by $`10\%`$.
The analysis was repeated using the same Breit-Wigner form for the resonances but with an alternative expression for the mass-dependent width that has been used in some previous experimental resonance studies :
$$\mathrm{\Gamma }(m)=\mathrm{\Gamma }_0\frac{2(\frac{m^24m_\pi ^2}{m_0^24m_\pi ^2})^{(l+\frac{1}{2})}}{1+\frac{m^24m_\pi ^2}{m_0^24m_\pi ^2}}.$$
(3)
The fitted parameters and numbers of signal events are given in Table 2. The $`\chi ^2`$ per degree of freedom for each of these fits is considerably worse than for the original fits.
The upper limit on $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})`$ for the case when only the $`f_0(1500)`$ is included is 0.33 keV, and for the $`f_J(1710)`$ only is 0.84 keV. For the case including both resonances, the upper limits are 0.28 keV and 0.69 keV for the $`f_0(1500)`$ and the $`f_J(1710)`$ respectively. These limits are consistent with the original results.
## 7 Stickiness
From limits on $`\mathrm{\Gamma }_{\gamma \gamma }`$ the stickiness of each resonance can be calculated, by comparison of the two photon width with the particle’s width for production in the glue-rich environment of radiative J/$`\psi `$ decay. The ratio is normalised such that the $`q\overline{q}`$ resonance $`f_2(1270)`$ has stickiness equal to 1. Then for a glueball a higher value of stickiness is expected.
Stickiness for a $`0^{++}`$ resonance $`X`$ is defined as
$$S_X(0^{++})=𝒩\frac{m_X}{k_{\mathrm{J}/\psi \gamma X}}\frac{\mathrm{\Gamma }(\text{J/}\psi \text{ }\gamma X)}{\mathrm{\Gamma }(X\gamma \gamma )}$$
where $`m_X`$ is the mass of the resonance; $`k_{\mathrm{J}/\psi \gamma X}=(m_{\mathrm{J}/\psi }^2m_X^2)/2m_{\mathrm{J}/\psi }`$ is the energy of the photon from the radiative J/$`\psi `$ decay, measured in the J/$`\psi `$ rest frame; $`\mathrm{\Gamma }(\text{J/}\psi \text{ }\gamma X)`$ is the width for production of the resonance in radiative J/$`\psi `$ decay; and $`𝒩`$ is the normalisation factor. The branching ratio for $`f_0(1500)\pi ^+\pi ^{}`$ is taken as $`0.30\pm 0.07`$, while the branching ratio for $`f_J(1710)\pi ^+\pi ^{}`$ is $`0.026_{0.016}^{+0.001}`$ , giving, from the fit for both resonances, $`\mathrm{\Gamma }(f_0(1500)\gamma \gamma )`$ = 1.08 keV and $`\mathrm{\Gamma }(f_J(1710)\gamma \gamma )`$ = 21.25 keV. The lower limits on stickiness are then 1.4 and 0.3 for the $`f_0(1500)`$ and the $`f_J(1710)`$ respectively.
## 8 Conclusion
Production of the glueball candidates $`f_0(1500)`$ and $`f_J(1710)`$ in $`\gamma \gamma `$ collisions at LEP1 has been studied via decay to $`\pi ^+\pi ^{}`$. No signal from either resonance is seen, and the upper limits on the product of two-photon width and $`\pi ^+\pi ^{}`$ branching ratio have been calculated as $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})<0.31`$ keV for the $`f_0(1500)`$ and $`\mathrm{\Gamma }_{\gamma \gamma }(\pi ^+\pi ^{})<0.55`$ keV for the $`f_J(1710)`$, both at 95% C.L. , from a simultaneous fit for both resonances.
## Acknowledgements
We wish to thank our colleagues from the accelerator divisions for the successful operation of LEP. We are indebted to the engineers and technicians at CERN and our home institutes for their contribution to the good performance of ALEPH. Those of us from non-member countries thank CERN for its hospitality. |
no-problem/9911/quant-ph9911102.html | ar5iv | text | # A Fundamental Limit of Measurement Imposed by the Elementary Interactions
## Introduction
Quantum information theory is closely related to quantum measurement theory because one must perform measurement to obtain information on a quantum system. In particular, a limit of quantum measurement is also a limit of getting information on a quantum system. Among many possible limits of quantum measurement, the simplest ones were derived directly from the uncertainty principles. However, such simple limits are not the only limits. I here suggest a new limit which comes from the forms and the strengths of the elementary interactions.
It is established that measurement of a quantum system can be analyzed (except for philosophical issues) as interaction processes of the system and the measuring apparatus LP ; QNDmass ; QNDphoton ; PSY ; ozawa ; SF ; pra ; Nano . Many interesting quantities, such as a measurement error and backaction of the measurement, can be calculated from the dynamics of the total quantum system composed of the system and the measuring apparatus LP ; QNDmass ; QNDphoton ; PSY ; ozawa ; SF ; pra ; Nano . The dynamics of a quantum system is governed by the commutation relations and the Hamiltonian. For the interaction Hamiltonian $`H_I`$, it was often assumed in the literature QNDmass ; QNDphoton ; PSY that the form of $`H_I`$ was obtained by “quantizing” some classical Hamiltonian. Or, in some cases ozawa , specific forms were just assumed for $`H_I`$. Discussions based on such $`H_I`$ may be consistent mathematically. Physically, however, such $`H_I`$ can be wrong. For example, quantization of a classical Hamiltonian in general gives a form that is different from the true Hamiltonian, and the difference is typically of the order of $`\mathrm{}`$. However, backaction is also of the order of $`\mathrm{}`$. Hence, physically, any results drawn from such a Hamiltonian cannot be reliable enough. In particular, there are only four types of elementary interactions in nature; their forms are determined by the gauge invariance (and symmetry breaking), and their coupling constants (in the low-energy regime) have definite values ft . In order to draw conclusions which are physically meaningful, one must either use these true interactions, or, if one uses an approximate Hamiltonian, one must confirm that the approximation does not alter the results for the measurement. Unfortunately, however, these points were not examined carefully.
In this paper, I point out that not only the commutation relations but also the limitation of available interactions in nature results in a fundamental limit of quantum measurement. This fundamental limit imposes the fundamental limits of getting information on, preparing, and controlling quantum systems.
## Measurement error and backaction on the measured observable
I consider the relation between the measurement error and the backaction on the measured observable. Although similar arguments should be applicable to other relations, I here present only one example, to explain the basic ideas.
When one measures an observable $`Q`$ of a quantum system, there is a finite measurement error $`\delta Q_{err}`$ in general. Its magnitude is determined by the form and strength of the interaction Hamiltonian $`H_I`$ between the measured system and the measuring apparatus QNDmass ; QNDphoton ; PSY ; ozawa ; SF ; pra ; Nano . There are also backactions of the measurement. A well-known backaction is the one on the conjugate variable $`P`$ of $`Q`$. Its lower limit is determined by the commutation relation between $`Q`$ and $`P`$, i.e., by the uncertainty principle. However, I here discuss backaction on $`Q`$ itself LP ; QNDmass ; QNDphoton ; PSY ; ozawa ; SF ; pra ; Nano , which is denoted by $`\delta Q_{ba}`$. Although the “ideal measurement”, for which $`\delta Q_{ba}=\delta Q_{err}=0`$, is normally assumed in textbook quantum mechanics, real measurements are non-ideal in general, for which $`\delta Q_{ba}`$ and/or $`\delta Q_{err}`$ are finite LP ; QNDmass ; QNDphoton ; PSY ; ozawa ; SF ; pra ; Nano . Since the lower limit of $`\delta Q_{ba}`$ is not (directly) limited by the uncertainty principle, $`\delta Q_{ba}`$ strongly depends on the form and strength of the interaction Hamiltonian $`H_I`$ LP ; QNDmass ; QNDphoton ; PSY ; ozawa ; SF ; pra ; Nano , just as $`\delta Q_{err}`$ does. Therefore, $`\delta Q_{ba}`$ and $`\delta Q_{err}`$ are correlated through $`H_I`$. For this relation, one could not get physically correct results without using a correct Hamiltonian.
For example, if one assumes some $`H_I`$ which satisfies
$$[H_I,Q]=0,$$
(1)
then $`\delta Q_{ba}=0`$, independent of the strength (which may be expressed by the coupling constant $`g`$) of $`H_I`$ QNDmass ; QNDphoton ; SF . Namely, the measurement becomes of the first kind (FK) LP ; SF . On the other hand, $`\delta Q_{err}`$ tends to increase as $`g`$ is decreased: For example, $`\delta Q_{err}\mathrm{}`$ as $`g0`$ QNDphoton ; PSY ; ozawa ; SF ; pra ; Nano . Hence, one would conclude that
$`\delta Q_{ba}`$ and $`\delta Q_{err}`$ are uncorrelated. (2)
However, this is true only for such a hypothetical $`H_I`$. Namely, if
$$[H_I,Q]0$$
(3)
for the true form of $`H_I`$ (i.e., for the elementary interaction), then $`\delta Q_{ba}`$ strongly depends on the form and strength of $`H_I`$. Therefore, the correct conclusion is
$`\delta Q_{ba}`$ and $`\delta Q_{err}`$ are strongly correlated. (4)
in sharp contrast to (2).
Unfortunately, most previous work on measurement QNDmass ; QNDphoton ; PSY ; ozawa assumed some “effective” interactions for $`H_I`$, such as a photon-photon interaction, which is not the true form, the elementary interaction. The validity of such effective forms, when applied to problems of measurement, was not examined carefully.
## measurement of the photon number
From this viewpoint, of particular interest is the measurement of a gauge field such as the photon field. The form of its interaction is completely determined by the requirement of the gauge invariance: the only interaction is the gauge-invariant interaction with the matter field ft . However, most previous work on measurement of a gauge field, particularly work on the “quantum non-demolition (QND) measurement” of photons QNDphoton ; PSY , assumed other forms for $`H_I`$, such as a photon-photon interaction, as “effective” interactions. Here, the QND measurement is the FK measurement for a conserved observable, i.e., for $`Q`$ that is conserved during the free motion (i.e., while the system is decoupled from the measuring apparatus) QNDmass . It therefore seems that most previous discussions on measurement of a gauge field need to be reconsidered.
To demonstrate the dramatic difference between the correct result derived from a proper $`H_I`$ and a wrong result derived from an “effective” $`H_I`$, I will present results for the QND photodetector proposed in Refs. , for both cases where the true $`H_I`$ and an “effective” $`H_I`$ are used.
A schematic diagram of the QND photodetector is shown in Fig.1 pra . Before going to the full analysis in the following sections, I here give an intuitive, semi-classical description of the operation principle SFOY . The device is composed of two quantum wires, N and W, which constitute an electron interferometer. The lowest subband energies (of the $`z`$-direction confinement) $`ϵ_a^N`$ and $`ϵ_a^W`$ of the wires are the same, but the second levels $`ϵ_b^N`$ and $`ϵ_b^W`$ are different. Electrons occupy the lowest levels only. A $`z`$-polarized light beam hits the dotted region. The photon energy $`\mathrm{}\omega `$ is assumed to satisfy $`ϵ_b^Wϵ_a^W<\mathrm{}\omega <ϵ_b^Nϵ_a^N`$, so that real excitation does not occur and no photons are absorbed. However, the electrons are excited “virtually” virtual , and the electron wavefunction undergoes a phase shift between its amplitudes in the two wires. Since the magnitude of the virtual excitation is proportional to the light intensity virtual , so is the phase shift. This phase shift modulates the interference currents, $`J_+`$ and $`J_{}`$. By measuring $`J_\pm `$, we can know the magnitude of the phase shift, from which we can know the light intensity. Since the light intensity is proportional to the photon number $`n`$, we can get information on $`n`$. We thus get to know $`n`$ without photon absorption, i.e., without changing $`n`$; hence the name QND. (More precisely, the distribution of $`n`$ over the whole ensemble is unchanged, whereas $`n`$ of each element of the ensemble is changed due to the “reduction” of the wavefunction. See, e.g., section 7 of Ref. Nano .) In contrast, conventional photodetectors drastically alter the photon number by absorbing photons. Keeping this semi-classical argument in mind, let us proceed to a fully-quantum analysis.
## Results derived from an “effective” Hamiltonian
As explained above, the essential role of the photon field is to modulate the electron phase. It is thus tempting to employ the following form
$$H_I=\underset{\nu =N,W}{}\underset{k}{}g_{k\nu }c_{k\nu }^{}c_{k\nu }n,$$
(5)
as an “effective” interaction Hamiltonian. Here, $`c_{k\nu }`$ denotes the annihilation operator of an electron of wavelength $`k`$ in quantum wire $`\nu `$, and $`g_{k\nu }`$ is an effective coupling constant. This Hamiltonian commutes with the photon number $`n`$;
$$[H_I,n]=0.$$
(6)
This equation, and the fact that $`n`$ is conserved during the free evolution, meet the condition for QND detectors that were claimed in Refs. QNDmass ; QNDphoton .
It is then easy to show that the backaction $`\delta n_{ba}=0`$, whereas the measurement error $`\delta n_{err}`$ is finite and dependent on $`g_{k\nu }`$. Namely, we obtain conclusion (2) for the interaction Hamiltonian (5).
## Results derived from the correct Hamiltonian
The correct interaction (i.e., the elementary interaction of, in this case, quantum electrodynamics) is not the form of Eq. (5), but the following one ft ;
$$H_I=\frac{e}{\mathrm{}c}d^4x\overline{\psi }\gamma ^\mu A_\mu \psi .$$
(7)
Here, $`\psi `$ is the (relativistic) electron field, $`\gamma ^\mu `$ ($`\mu =0,1,2,3`$) are the gamma matrices, $`A_\mu `$ is the electromagnetic potential (the set of the vector potential and scalar potential), and $`\overline{\psi }=\psi ^{}\gamma ^0`$. This form is uniquely determined by the local gauge invariance ft . Furthermore, the value of the coupling constant $`e/\mathrm{}c`$ (more precisely, the renormalized value at the low energy region) is uniquely determined by experiment ft . Since $`n`$ is quadratic in $`A_\mu `$, we find
$$[H_I,n]0,$$
(8)
in contrast to Eq. (6). Hence, the QND condition claimed in Refs. QNDmass ; QNDphoton is not satisfied by the correct Hamiltonian. According to the detailed analysis on QND measurement SF , the condition claimed in Refs. QNDmass ; QNDphoton is the strongest one among various possibilities, which are called the strong, moderate, and weak conditions SF .
The weak condition given in Ref. SF can be slightly generalized to the following form unpub (which reduces to that of Ref. SF as $`\epsilon _{ba}0`$): For a small positive number $`\epsilon _{ba}`$,
$$\left|\underset{j}{}\left|\underset{k,\mathrm{}}{}a_kb_{\mathrm{}}u_{ij}^k\mathrm{}\right|^2|a_i|^2\right|\frac{\epsilon _{ba}}{2}\left|\underset{j}{}\left|\underset{k,\mathrm{}}{}a_kb_{\mathrm{}}u_{ij}^k\mathrm{}\right|^2+|a_i|^2\right|\text{for all }i,$$
(9)
for some set of $`\{a_k\}`$’s and for some $`\{b_{\mathrm{}}\}`$. Here, I have used the same notations as in Ref. SF ; $`\{a_k\}`$ and $`\{b_{\mathrm{}}\}`$ represent the initial states of the measured system and the “probe” system, respectively, i.e., the expansion coefficients in terms of the eigenfunctions of the measured and “readout” variables SF . The $`u_{ij}^k\mathrm{}`$ is the matrix representation of the unitary evolution induced by the interaction with the probe system, hence is a function of the Hamiltonian SF . Physically, this condition means that the backaction can be made small enough for a particular set of measured states if the measuring apparatus is appropriately designed. The magnitude of the backaction is represented by $`\epsilon _{ba}`$, which gives the upper limit of the change of the probability distribution over the ensemble SF . Note that the limitation of measured states corresponds to a limitation of the response range of the measuring apparatus. Such a limitation is quite realistic, because any existing apparatus do have finite response ranges. Furthermore, as is clear from Eq. (8), only by accepting such a limitation can we realize the FK or QND measurement of wide classes of observables including photon number.
The condition satisfied by the QND photodetector of Fig. 1 is this (generalized) weak condition. In fact, the backaction $`\delta n_{ba}`$ which is induced by the QND detector of Fig. 1 is evaluated as unpub ; nonrel
$$\delta n_{ba}\frac{\gamma ^2nN}{|\mathrm{\Delta }|^4\tau _p},$$
(10)
where $`\gamma `$ ($`e/\mathrm{}c`$) denotes a constant that is determined by the structures of the quantum wires and the optical waveguide, $`N`$ is the number of electrons detected as the interference current, $`\mathrm{\Delta }=ϵ_bϵ_a\mathrm{}\omega `$ is the detuning energy, and $`\tau _p`$ denotes the optical-pulse duration (which is assumed, for simplicity, to be longer than $`\tau _t`$ of Ref. pra ). It is seen that for a small positive number $`\epsilon _{ba}`$, one can make
$$\frac{\delta n_{ba}}{n}\epsilon _{ba},$$
(11)
by taking $`1/|\mathrm{\Delta }|`$, $`1/\tau _p`$, $`\gamma ^2`$, and/or $`N`$ small enough. This corresponds to Eq. (9). Namely, the (generalized) weak condition (9) for the QND measurement is indeed satisfied. In other words, the QND measurement is possible for a certain set of photon states (i.e., for photon states whose pulse duration $`\tau _p`$ is longer than some value) if the structure of the device is appropriately designed (which determines $`\gamma ^2`$ and $`N`$).
On the other hand, the measurement error $`\delta n_{err}`$ is evaluated as pra ; Nano ; nonrel
$$\delta n_{err}\frac{|\mathrm{\Delta }|}{\gamma ^2\sqrt{N}}.$$
(12)
This dependence on $`N`$ is a general result for quantum interference devices SS . It is seen that for a small positive number $`\epsilon _{err}`$, one can make
$$\frac{\delta n_{err}}{n}\epsilon _{err},$$
(13)
by taking $`1/|\mathrm{\Delta }|`$, $`\gamma ^2`$, $`N`$, and/or $`n`$ large enough. However, as seen from Eq. (10), their increase results in the increase of the backaction. Therefore, the backaction (not on the conjugate observable, but on the photon number itself) and the measurement error are strongly correlated: one cannot make $`\delta n_{err}`$ ($`\delta n_{ba}`$) smaller without accepting the increase of $`\delta n_{ba}`$ ($`\delta n_{err}`$). This correct conclusion, derived from the correct Hamiltonian (7), should be contrasted with the wrong result obtained from the “effective” Hamiltonian (5).
## Optimization of the measuring apparatus and the upper limit of the information entropy
Because of this unavoidable correlation, one must make some optimization. For example, if one wishes to achieve
$$\epsilon _{ba},\epsilon _{err}10^2\text{for }\tau _p10\text{ ps},$$
(14)
then the QND detector can be designed in such a way that unpub
$`{\displaystyle \frac{\delta n_{ba}}{n}}`$ $``$ $`10^2,`$ (15)
$`{\displaystyle \frac{\delta n_{err}}{n}}`$ $``$ $`{\displaystyle \frac{10^2}{n}}\text{(i.e., }\delta n_{err}10^2\text{)}.`$ (16)
In this case, Eq. (14) is satisfied if
$$n10^4.$$
(17)
Hence, the lower boundary $`n_{min}`$ of the response range of the QND detector is found to be $`n_{min}=10^4`$. On the other hand, the upper limit $`n_{max}`$ of the response range is evaluated in this case as $`n_{max}10^6`$, which is derived from the requirement that the phase shift of an electron should be less than $`\pi `$ pra . Then, Eq. (14) is satisfied for all values of $`n`$ between $`n_{min}`$ and $`n_{max}`$. Moreover, $`\delta n_{err}`$ is smaller than the “standard quantum limit” $`\sqrt{n}`$ QNDphoton ; PSY , throughout this range.
It is interesting to evaluate the information entropy $`I`$ that can be obtained by the QND detector. Since $`\delta n_{err}`$ is independent of $`n`$, the number of different values of $`n`$ that can be distinguished by this QND detector is simply given by
$$\frac{n_{max}n_{min}}{\delta n_{err}}10^4.$$
(18)
Hence, $`I`$ is estimated as
$$I\mathrm{log}_210^413.3.$$
(19)
This should be contrasted with the wrong conclusion $`I1`$ that would be obtained from the “effective Hamiltonian” (5). We have obtained the correct value of $`I`$ because we have taken account of the strong correlation between $`\delta n_{ba}`$ and $`\delta n_{err}`$. Since this correlation comes from the form of the gauge interaction, similar correlation should also be present for any measuring apparatus of any gauge field. Therefore, similar optimization of the apparatus is necessary, and the response range and the information entropy would be finite, for any measuring apparatus of any gauge field.
## Concluding remarks
The above example clearly demonstrates that the limitation of available interactions in nature gives rise to a fundamental limit of quantum measurement. I discuss possible consequences of this fundamental limit.
Clearly, it imposes a fundamental limit of getting information on quantum systems because the “ideal measurement” (i.e., the measurement with $`\delta Q_{err}=\delta Q_{ba}=0`$, although the backaction on the conjugate variable is of course finite) is forbidden for a certain set of observables such as the number of gauge quanta. Namely, one cannot get information on $`Q`$ without disturbing $`Q`$ itself. If one wishes to make this backaction on $`Q`$ small, then the response range of the apparatus and the information entropy obtained by the measurement are limited, as demonstrated in the previous section.
It should be pointed out that this limitation of measurement leads to many other limitations. For example, it imposes a limit on the controllability of quantum systems: One must perform a measurement to control a system, but an accurate measurement gives rise to large backactions on some of the measured quantities, and the control becomes imperfect. Another example is the limitation of the preparation of quantum states. To prepare a system in a desired state, one must perform ideal measurements on some set of observables, e.g., on a “complete set of commuting observables” dirac . However, as I have shown in this paper, an ideal measurement is impossible for some observables. Therefore, the problems of the preparation of quantum states should be reconsidered.
As seen from these examples, the limitation of the measurement means that accurate specification of quantum systems is impossible in general. This seems to be related to the foundation of the statistical mechanics. For example, imperfect knowledge about the initial state of the system leads to rapid loss of knowledge as the time evolves. This results in the ergodic property and the mixing property dset , although quantum systems cannot have these properties if the initial state is accurately known BB .
### Acknowledgment
This work has been supported by the Core Research for Evolutional Science and Technology (CREST) of the Japan Science and Technology Corporation (JST). |
no-problem/9911/astro-ph9911457.html | ar5iv | text | # On the interpretation of the multicolour disc model for black hole candidates
## 1 Introduction
X-ray spectra of galactic black hole candidates (GBHC) are becoming an increasingly important tool in determining both the physical properties of the source surroundings and, as the ultimate goal, the parameters (mass and specific angular momentum) of the black hole. Two main spectral states, as defined by their spectral components and flux level (typically in the $`110`$ keV band), have been observed in GBHC. In the Hard/Low state, the sources emit most of their power in a hard power law tail with photon index $`\mathrm{\Gamma }1.31.7`$ and exponential cutoff at about $`100`$ keV. In the generally more luminous Soft/High state, most of the energy comes from a thermal component with characteristic temperature $`kT1`$ keV, while a power law component (with $`\mathrm{\Gamma }2.02.5`$) dominates above a few keV. In addition, sources have been reported to be in a Very high state, spectrally intermediate between the Soft/High and the Hard/Low states, when the power law component has a flux comparable to the thermal one and no sign of a cutoff is seen in the high energy tail.
We will focus our discussion on the Soft/High state. According to the common paradigm in interpreting the observed spectra, the ultra-soft thermal component is the product of the emission from an optically thick accretion disk. On the other hand, the hard power law component is probably due to multiple Compton scattering of soft photons by a population of hot electrons which reside in an active (or flaring) coronal region surrounding the accretion disc.
In the zeroth order approximation for describing emission from the accretion disc, not only is the disc assumed to be geometrically thin and optically thick , but also the vertical temperature structure is neglected, as is the effect of Comptonization on the emergent spectrum. Each point of the disc is then assumed to radiate like a blackbody at an effective temperature which scales with the radius as $`r^{3/4}`$. This is the so-called multicolour disc model (MCD; Mitsuda et al. 1984). Such a simple model has the advantage of being easy to use in trying to fit spectral data. It has only two adjustable parameters: a (colour) temperature, independent on the source distance, and a normalization factor, which in turn depends on the inner radius of the disc, on the distance of the source and on the inclination of the disc from the line of sight. When these two latter quantities are independently known, or can be reasonably inferred, fitting the ultra-soft component of a GBHC spectrum with the MCD model immediately gives estimates of the temperature in the inner region of the disc and its inner radius.
The aim of this letter is to demonstrate that the values one can obtain with this fitting procedure, which is now a standard one, are not in general directly related (or at least not in a reliable way) to the actual disc parameters. We will use a self-consistent model for the radiative transfer in Shakura-Sunyaev accretion discs about a compact source, including first-order corrections for gravitational redshift and Doppler blurring, as developed by Ross & Fabian . The observable spectrum is computed for a number of different physical situations, each determined by the values of the viscosity parameter $`\alpha `$, the accretion rate and the fraction of the accreted power dissipated in the corona . In all the cases we consider, the value of the inner radius of the disc will be kept fixed. Nevertheless, when trying to fit the computed spectra with the MCD model, we will obtain results at variance with this assumption. That is because changes in the normalization factor of the MCD model are produced, in a very complicated way, by the variations in the disc structure (and hence in the Comptonized spectrum) induced by variability of the accretion rate and coronal activity.
## 2 Structure of the disc
We assume the basic structure for spatially thin accretion discs around Schwarzschild black holes given by the standard theory of Shakura & Sunyaev . The disc is assumed to have a fixed inner boundary at the innermost stable orbit at radius $`R_{\mathrm{in}}=3R_S=6GM/c^2`$. We consider both an inner, radiation pressure dominated region and an outer, gas pressure dominated region of the disc. In determining the structure of the disc, the opacity is assumed to be dominated by Thomson scattering,
$$\kappa \kappa _T=0.2(1+X)\mathrm{cm}^2\mathrm{g}^1,$$
(1)
where $`X`$ is the mass fraction of hydrogen. Following Svensson & Zdziarski , we slightly modify the standard set of disc equations, allowing a fraction $`f`$ of the disc accretion power to be dissipated in the corona instead of in the cold disc itself. Furthermore, at every given radius $`R`$, the density is taken to be uniform in the vertical direction (which is the correct solution for the disc structure when radiation pressure dominates).
Choosing dimensionless parameters
$$m=\frac{M}{M_{}},\dot{m}=\frac{\dot{M}}{\dot{M}_{\mathrm{Edd}}},r=\frac{R}{R_S},$$
where $`\dot{M}_{\mathrm{Edd}}=3.1\times 10^8mM_{}\mathrm{yr}^1`$ is the Eddington accretion rate for a disk efficiency $`\eta =0.083`$, the flux emerging from the surface of the disc is given by
$$F_0=1.2\times 10^{27}m^1r^3[\dot{m}J(r)](1f),$$
(2)
with $`J(r)=(1\sqrt{3/r})`$. The radial structure of the disc is summarized by the values of the half-thickness of the disc (in units of the Schwarzschild radius) $`h`$ and the uniform gas density $`\rho _0`$:
1. Radiation pressure dominated region
$`h={\displaystyle \frac{H}{R_S}}=5.3(1+X)[\dot{m}J(r)](1f)`$ (3)
$`\rho _0=3.4\times 10^7\alpha ^1m^1r^{3/2}`$
$`\times [\dot{m}J(r)]^2(1f)^3\mathrm{g}\mathrm{cm}^3,`$ (4)
2. Gas pressure dominated region
$`h=3.4\times 10^2(1+X)^{1/10}\alpha ^{1/10}m^{1/10}r^{21/20}`$
$`\times [\dot{m}J(r)]^{1/5}(1f)^{1/10}`$ (5)
$`\rho _0=4(1+X)^{3/10}\alpha ^{7/10}m^{7/10}r^{33/20}`$
$`\times [\dot{m}J(r)]^{2/5}(1f)^{3/10}\mathrm{g}\mathrm{cm}^3.`$ (6)
The radiative transfer and the vertical temperature structure are treated self-consistently at each radius, as described in Ross & Fabian , using the Fokker-Planck/diffusion equation of Ross, Weaver & McCray in plane-parallel geometry. The local temperature profile, $`T(z)`$, is found by balancing the heating rate due to dynamic heating, Compton scattering and free-free absorption with the cooling rate due to inverse Compton scattering and free-free emission.
Incoherent Compton scattering within the accretion disc must be treated properly whenever it has an important effect on the emergent spectrum. The competition between Compton scattering of higher-energy photons and inverse Compton scattering of lower energy photons often results in the emergence of a Wien-law hard tail. The effective optical depth for absorption is given by
$$\tau _{}(\nu )=\sqrt{3\tau _{\mathrm{ff}}(\nu )(\tau _\mathrm{T}+\tau _{\mathrm{ff}}(\nu ))},$$
(7)
where $`\tau _\mathrm{T}`$ is the Thomson depth below the surface, and $`\tau _{\mathrm{ff}}(\nu )`$ is the optical depth due to free-free absorption. For high energy photons the thermalization depth (where $`\tau _{}(\nu )=1`$) can be reached at very high Thomson depths, where the temperature is considerably higher than near the surface. Compton scattering in the outer layers of the disc downscatters these photons to lower energies and produces a Wien-law tail.
Decreasing the flux emerging from the disc itself, either by lowering the accretion rate ($`\dot{m}`$) or increasing the fraction ($`f`$) of the accretion power dissipated in the corona, results in ever larger regions of very high density, both in the outermost portions of the disc and, for very low fluxes, in the very innermost portions as well. In those cases Comptonization is not complete or ‘saturated’ because
$$y=\frac{4kT}{m_ec^2}\tau _T^2<1$$
(8)
at the thermalization depth. In these regions we treat Compton scattering as coherent, dropping the Fokker-Planck term in the radiative transfer equation (see discussion in Ross & Fabian ).
In our calculations we fix the value of the black hole mass $`m=10`$ and assume a composition $`X=0.71`$. Following Ross & Fabian we divide the region $`3<r<200`$ of the disc into 20 annuli that make comparable contributions to the total luminosity. For each set of the parameters $`\alpha `$, $`\dot{m}`$ and $`f`$, we find the radius at which radiation pressure equals gas pressure, which is the root of the equation
$$\frac{r_{AB}}{J(r_{AB})^{16/21}}370(m\alpha )^{2/21}\dot{m}^{16/21}(1f)^{6/7}.$$
(9)
We use equations (3) and (4) for all the annuli for which $`3<r<r_{AB}`$, and equations (5) and (6) for $`r>r_{AB}`$. The typical emergent spectrum from each annulus is calculated and then multiplied by the area of the annulus to find the contribution to the spectral luminosity. The resulting spectra are added together, taking into account blurring due to gravitational redshift and transverse Doppler effect using the method described by Chen, Halpern & Filippenko . Finally, in order to allow direct comparison with observations in which a dramatic change in the inner radius has been reported (GRS 1915+105, see e.g. Belloni et al. \[1997b\]), we choose an inclination angle $`i=70^{}`$ and a distance $`D=12.5\mathrm{kpc}`$.
## 3 The multicolour disc model
For each set of parameters the emergent disc spectrum is used as a table model to create a fake set of data (with RXTE response matrix p012\_LR1\_970804.rsp, integration time 1000 s). These data are then fitted with the standard multicolour blackbody model (DISKBB in XSPEC) in the $`220`$ keV band. The model assumes that the local emission from the disc is Planckian, with a temperature profile $`T(r)r^{3/4}`$. Therefore the observed flux from the disc is approximated by
$$F_d(E)=\frac{8\pi R_{\mathrm{col}}^2\mathrm{cos}i}{3D^2}_{T_{\mathrm{out}}}^{T_{\mathrm{col}}}\left(\frac{T}{T_{\mathrm{col}}}\right)^{11/3}B(E,T)\frac{dT}{T_{\mathrm{col}}},$$
(10)
where $`B(E,T)`$ is the Planck function and $`T_{\mathrm{out}}`$ is the temperature at the outer radius of the disc (and it is assumed that $`T_{\mathrm{out}}T_{\mathrm{col}}`$). The two fit parameters of the model are the colour temperature of the inner accretion disc ($`T_{\mathrm{col}}`$) in keV and the normalization factor
$$n=\frac{R_{\mathrm{col}}^2\mathrm{cos}i}{D^2},$$
where $`R_{\mathrm{col}}`$ is expressed in km and the distance $`D`$ in units of $`10\mathrm{k}\mathrm{p}\mathrm{c}`$.
The parameter $`R_{\mathrm{col}}`$ is not the effective inner disc radius (i.e. the radius at which the temperature of the disc is the highest). The assumption behind using the MCD model is that the Comptonized emergent spectrum can be approximated by a diluted blackbody spectrum with a colour temperature ($`T_{\mathrm{col}}`$) higher than the effective temperature ($`T_{\mathrm{eff}}`$). The ratio $`f_{\mathrm{col}}=T_{\mathrm{col}}/T_{\mathrm{eff}}`$, the spectral hardening factor , is then assumed to be constant throughout the disc and for varying physical parameters ($`\dot{m}`$, $`\alpha `$, $`f`$, etc.). Thus the actual inner edge of the disc would be given by
$$R_{\mathrm{in}}=\eta R_{\mathrm{eff}}=\eta g(i)R_{\mathrm{col}}(T_{\mathrm{col}}/T_{\mathrm{eff}})^2,$$
(11)
where $`\eta `$ is the ratio of the inner radius of the disc to the radius at which the emissivity actually peaks, typically $`\eta 0.60.7`$ . The factor $`(T_{\mathrm{col}}/T_{\mathrm{eff}})^2`$ comes from the assumed dilution of the blackbody spectrum, while $`g(i)`$ takes into account general relativistic corrections and is of the order $`g0.70.8`$ .
In the next section we will show that these assumptions are not justified in general due to the differences in Compton processes induced by changes in $`\dot{m}`$, $`f`$, and $`\alpha `$.
## 4 Results
We have simulated disc spectra in eleven different physical situations represented by the ten sets of parameters S1-S11 (see Table 1). The sets are listed in order of decreasing value of the radius $`r_{\mathrm{AB}}`$, the boundary between the radiation pressure and gas pressure dominated regions. Table 1 also lists the value of the radial coordinate $`r_{\mathrm{coh}}`$ for the boundary of the region(s) where Compton scattering has been treated as coherent (see section 2). As the power dissipated in the disc decreases and the density increases, from the combined effects of changes in $`\dot{m}`$ and $`f`$, this boundary moves inwards. That is, for an increasing portion of the disc Comptonization is incomplete and the local spectrum will be somewhat harder. For the S10 model, the density in the innermost region of the disc is so high (due to the $`J(r)^2`$ term) that we have to consider incomplete Compton scattering also in the first inner annuli, where a significant fraction of the X-ray power is emitted. Consequently, for this set of parameters, two values of $`r_{\mathrm{coh}}`$ are listed. Finally, for the model S11, when the disc is entirely gas pressure dominated, we considered incomplete Compton scattering throughout all the disc (but see next section for a caveat on the applicability of our model to this extreme case).
Figure 1 shows the observed spectra for S1, S4, S7 and S10, taking into account Doppler blurring and gravitational redshift. For comparison, Fig. 1 also shows the blurred spectra that would be observed if each point on the surface of the accretion disc emitted a blackbody spectrum at the local effective temperature. The observed spectra are harder than such effective blackbody spectra, and the discrepancy increases as we increase the density of the disc and decrease the power released in the disc itself (going from S1 to S10).
Table 2 lists the values of the fit parameters $`T_{\mathrm{col}}`$ and $`R_{\mathrm{col}}`$ for the ten cases, as well as the corresponding reduced $`\chi ^2`$ of the fits and the spectral hardening factors obtained from eq. (11) by assuming $`\eta g(70^{})=0.5`$ and $`R_{\mathrm{in}}=88.6\mathrm{km}`$, as appropriate for the $`10M_{}`$ Schwarzschild black hole we are considering. The results are also shown in Figures 2a and 2b, where we plot the spectral hardening factor and the colour radius as functions of $`\dot{m}(1f)`$, which is a measure of the flux emergent from the disc itself.
To check the sensitivity of our results to the energy range fitted, we also fitted the spectra in the $`310`$ keV band. The values obtained for the colour radii were, at most, just one or two kilometers smaller than those obtained in the 2-20 keV band (i.e. a change of about $`34\%`$).
It is clear that the multicolour model gives systematically lower values for the “inner radius” than the actual ones, a point which is well known. More interestingly, our results show that the hardening factor is by no means constant when we change the physical parameters of the disc. For example, assuming the conventional value $`f_{\mathrm{col}}=1.7\pm 0.2`$ and taking Doppler blurring and gravitational redshift into account, we would have underestimated the value for the inner disc radius by up to a factor of $`2`$ (S11) using the fits with a multicolour model. On the other hand the model gives quite stable and acceptable results for high accretion rates and/or lower values of the fraction of the power eventually dissipated in the corona. This in turn can help to understand the cases in which a nearly constant value of $`R_{\mathrm{col}}`$ is observed even when $`T_{\mathrm{col}}`$ varies .
## 5 Discussion
We have shown that MCD models systematically underestimate the value of the inner radius of the accretion disc for a black hole candidate. We also have shown that the spectral hardening factor $`f_{\mathrm{col}}`$, which is needed to correct the results of the fits with a multicolour model, is not, as is usually assumed, constant when the accretion rate and/or fractional coronal activity change.
Recent observations of galactic black hole candidates seem to point towards an extreme dependence of the observed colour radius on $`f`$.
In the work of Sobczak et al. , where the authors report on the RXTE spectral observations of the 1996-97 outbursts of the microquasar GRO J1655-40, a sudden jump inward of the color radius (see their Fig. 7) occur for the five very high state observations of the source during which the power-law (hard) flux was extremely high and the blackbody–to–total flux ratio was less than $`0.5`$ (see their Fig. 5).
A similar trend has been reported by Muno et al. in the case of the microquasar GRS 1915+105. In their observations, the smallest values of the inner disc radius obtained by fitting the spectrum with the MCD model are associated with the highest values of the power-law–to–blackbody flux ratio, which in turn should be directly related to the value of $`f`$.
In both these cases the inferred inner disc radii can change by more than a factor of four.
Belloni et al. (1997a,1997b), analyzing RXTE data of GRS 1915+105, interpreted the sudden change measured in the inner colour radius of roughly a factor of four in terms of the rapid disappearence of the whole inner part of the geometrically thin accretion disc. Once again, in that case the inner edge of the disc appears to shrink to its smallest values during an outburst in which the power law flux dominates the blackbody, and the blackbody–to–total flux ratio is less than $`0.5`$.
This behaviour is exactly what our results predict we should expect from the MCD model when we increase $`f`$ and/or reduce the accretion rate. In this case, every time we the observations imply that the coronal activity is dominant, the multicolour fits should be corrected with a varying hardening factor $`1.7<f_{\mathrm{col}}<3`$ in order to recover the actual value of the inner disc radius.
It should be clear, however, that the issue of determining the exact shape of accretion disc spectra is very complicated and is very sensitive to the actual physical models used to describe it. In particular, key elements are the vertical structure of the disc, its density profile and, in particular, the surface density.
It is quite possible, and perhaps easier to conceive, that dramatic changes in the disc spectrum are produced by changes in the surface properties of the accretion disc rather than by the disappearance of the entire inner regions. This, unfortunately, is hard to model theoretically. It would be interesting, for example, to assess the problem in the extreme case when almost all the accretion power is dissipated in the corona (i.e. in the limit $`f1`$). As we believe that even in our most extreme simulation (S11) with $`f=0.8`$ the simple treatment of the vertical disc structure we used here can be inadequate, and as this must be true a fortiori for larger values of $`f`$, we do not push our study further.
To do that we would need to take into account the X-ray reflection spectra of discs illuminated by a hot corona and to model carefully the heating of the disc atmosphere (e.g., see Sincell & Krolik , Ross & Fabian , who show how a hard tail in the disc spectrum is produced in this case). It is natural to believe, given the trend we observe in the simulations presented above and the observational results we referred to, that in such a situation the oversimplified multicolour model will give even smaller values of the colour radius.
A detailed treatment of this kind requires further numerical work which is beyond the aim of this Letter.
## Acknowledgments
This work was done in the research network “Accretion onto black holes, compact stars and protostars” funded by the European Commission under contract number ERBFMRX-CT98-0195’. AM, RRR and ACF thank the PPARC, the College of the Holy Cross and the Royal Society for support, respectively. |
no-problem/9911/cond-mat9911105.html | ar5iv | text | # Instability due to long range Coulomb interaction in a liquid of polarizable particles (polarons, etc.)
## Abstract
The interaction Hamiltonian for a system of polarons a la Feynman in the presence of long range Coulomb interaction is derived and the dielectric function is computed in mean field. For large enough concentration a liquid of such particles becomes unstable. The onset of the instability is signaled by the softening of a collective optical mode in which all electrons oscillate in phase in their respective self-trapping potential. We associate the instability with a metallization of the system. Optical experiments in slightly doped cuprates and doped nickelates are analyzed within this theory. We discuss why doped cuprates matallize whereas nickelates do not.
An electron moving in a highly polarizable lattice can distort the environment and create a potential which self-traps the electron. The resulting object can move by dragging the distortion resulting in a heavy particle called polaron. Apart from its motion as a whole the polaron has an extra degree of freedom which can be thought as the electron oscillating in the self-trapping potential. Because of this internal mode a polaron can be polarized in an external electric field.
In this Letter a model Hamiltonian that takes into account the interactions due to mutual polarization is derived. It is shown that a polaron liquid or more generally a liquid of polarizable particles becomes unstable for high enough concentrations due to the dipole-dipole interaction. The onset of the instability is indicated by a softening of the internal mode of the polaron which can be measured in optical experiments. Recent experimental results in cuprates and nickelates are analyzed within this scenario.
For simplicity the present treatment is semiclassical (in the spirit of the Drude model) however we expect that the same effect will show up in a fully quantum mechanical computation.
We concentrate on the case of dielectric polarons but we believe our results are valid also for other kinds of heavily dressed particles in solids as long as they have a polarizable internal degree of freedom.
Consider an electron in a dielectric with static dielectric constant $`ϵ_0`$ dominated by lattice polarization. If the electron is not allowed to move it will generate a radial polarization of the lattice towards itself. This distortion implies that some ionic positive charge ($`q`$) will accumulate in the vicinity of the electron. The charge induced in the lattice can be estimated as
$$q=e\left(1\frac{1}{ϵ_0}\right).$$
(1)
Now let us free the electron. If the electron binds to the induced lattice charge we have a dielectric polaron. In this case for the electron to move it has to drag the distortion of the lattice.
To describe this situation we use Feynman’s effective model for the polaron. The effect of the distortion is mimic by a heavy fictitious particle of mass $`M`$. The electron, of mass $`m`$, is coupled to the heavy particle with a harmonic potential. The result is a composite particle. The non interacting Hamiltonian for a liquid of such particles reads $`H_0=_iH_i`$ with
$$H_i=\frac{p_i^2}{2m}+\frac{P_i^2}{2M}+\frac{k}{2}|𝒓_i𝑹_i|^2.$$
(2)
The first term in $`H_i`$ is the kinetic energy of the electron. The second term is the kinetic energy of the heavy mass representing the surrounding deformation. The third term is the coupling between the electron and the distortion. $`𝒑_i`$, $`𝒓_i`$ ($`𝑷_i`$, $`𝑹_i`$) are the momentum and position of the electron (fictitious mass) $`i`$.
In the original formulation by Feynman the action corresponding to Eq. (2) is regarded as an effective action for the dynamics of a polaron. The parameters are obtained variationally from the Fröhlich model of a dielectric polaron. Hare we take as granted that $`H_0`$ is a first approximation to describe the dynamics of a polaron liquid and explore the consequences of adding a Coulomb term. On physical grounds we will assume the charge $`q`$ is distributed with spherical symmetry centered on $`R_i`$.
To derive the interaction Hamiltonian consider two fixed electrons in the dielectric at $`𝒓_1,𝒓_2`$. As before each electron will induce a close-by charge in the lattice of magnitude $`+q`$. The total electrostatic energy can be computed as usual. One gets self-energy terms plus the interaction term:
$$E_{\mathrm{int}}=\frac{e^2}{ϵ_0|𝒓_1𝒓_2|}.$$
(3)
This energy includes the vacuum electrostatic energy between charges plus the elastic energy stored in the dielectric:
$$E_{\mathrm{int}}=E_{\mathrm{Coul}}+E_{\mathrm{ela}}.$$
(4)
The former is:
$$E_{\mathrm{Coul}}=\frac{e^2}{ϵ_0^2|𝒓_1𝒓_2|}$$
(5)
i.e. the electrostatic energy between the screened charges. Eq. (4) can be solved for the elastic energy:
$$E_{\mathrm{ela}}=\frac{e^2}{|𝒓_1𝒓_2|}\frac{1}{ϵ_0}\left(1\frac{1}{ϵ_0}\right).$$
(6)
The factors with the dielectric constant in Eq. (6) ensure that the elastic energy is zero for an infinitely rigid lattice ($`ϵ_01`$) or an infinitely soft lattice ($`ϵ_0\mathrm{}`$).
Suppose now that we move the electrons from the equilibrium position keeping ions fixed so that the distortion does not move. i.e. $`𝒓_i𝑹_i`$. Since the elastic energy depends only on the configuration of the lattice we should replace in Eq. (6) $`𝒓_i𝑹_i`$. So that in general the elastic energy is:
$$E_{\mathrm{ela}}=\frac{(eq)q}{|𝑹_1𝑹_2|}$$
(7)
where we eliminated $`ϵ_0`$ using the definition of the induced lattice charge $`q`$ \[Eq. (1)\].
The Coulomb energy can be decomposed into elementary Coulomb interactions. For general position of charges and distortions it reads:
$$E_{\mathrm{Coul}}=\frac{e^2}{|𝒓_1𝒓_2|}\frac{eq}{|𝒓_1𝑹_2|}\frac{eq}{|𝑹_1𝒓_2|}+\frac{q^2}{|𝑹_1𝑹_2|}$$
(8)
which reduces to Eq. (5) for $`𝒓_i=𝑹_i`$.
Adding again the elastic and Coulomb energy we can write the interaction Hamiltonian for a liquid of such particles:
$$H_{\mathrm{int}}=\underset{ij,ij}{}\frac{1}{2}\frac{e^2}{|𝒓_i𝒓_j|}\frac{eq}{|𝒓_i𝑹_j|}+\frac{1}{2}\frac{eq}{|𝑹_i𝑹_j|}$$
(9)
where indexes $`i,j`$ run over particles.
A similar argument can be used to derive the interaction Hamiltonian with an external electric field. The Coulomb contribution of the induced charge cancels with the elastic part and one obtains,
$$H_E=\underset{i}{}e𝒓_i.𝑬.$$
(10)
The many-particle Hamiltonian is $`H=H_0+H_{\mathrm{int}}+H_E`$.
It is convenient to change to center of mass variables, $`𝝆_i=(𝒓_im+𝑹_iM)/(M+m)`$, $`𝒖_i=𝒓_i𝑹_i`$. Making a Taylor expansion for small $`𝒖_i`$ we obtain the following interacting Hamiltonian in the dipole approximation,
$`H_{\mathrm{int}}={\displaystyle \underset{ij,ij}{}}`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{e(eq)}{|𝝆_i𝝆_j|}}{\displaystyle \frac{e(eq)}{1+m/M}}{\displaystyle \frac{𝒖_i.(𝝆_i𝝆_j)}{|𝝆_i𝝆_j|^3}}`$ (11)
$``$ $`{\displaystyle \frac{e(eq)}{2(1+m/M)^2}}𝒖_i.\mathit{\varphi }(𝝆_i𝝆_j).𝒖_i`$ (12)
$`+`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{e(eq)}{(1+m/M)^2}}+eq\right)𝒖_i.\mathit{\varphi }(𝝆_i𝝆_j).𝒖_j`$ (13)
where we defined the matrix $`\varphi _{\mu \nu }(𝝆)=\delta _{\mu \nu }/\rho ^33\rho _\mu \rho _\nu /\rho ^5`$ and $`\mu ,\nu `$ are Cartesian indexes.
Hamiltonian Eq. (3) have the essential ingredients to treat the interacting polaron problem at concentrations such that the interparticle distance exceeds the polaron radius.
Now we compute the dielectric constant of such a system. To obtain equations of motion for one particle in the mean field of the others we compute the forces $`𝑭_{𝒖_i}=H/𝒖_i`$ and $`𝑭_{𝝆_i}=H/𝝆_i`$. The former is the force that polarizes particles and the latter is the force acting on the center of mass. The equations of motion in mean field read,
$`\mu \ddot{𝒖}`$ $`=`$ $`k𝒖+4\pi Lneq𝒖{\displaystyle \frac{e}{1+m/M}}𝑬`$ (14)
$`(m+M)\ddot{𝝆}`$ $`=`$ $`e𝑬`$ (15)
where $`𝒖<𝒖_i>`$, etc. $`n`$ is the density of particles, $`1/\mu =1/m+1/M`$, and $`L`$ is a geometric factor discussed below.
A crucial point in the derivation is the evaluation of the force due to dipole-dipole interactions $`𝑭_𝑳=eq𝒖.<_{j,ji}\mathit{\varphi }(𝝆_i𝝆_j)>`$. This is the well known Lorentz-Lorenz local field i.e. the dipolar field at $`𝝆_i`$ due to dipoles at positions $`𝝆_j`$ averaged over the position of the dipoles. Using well known results from the theory of dielectrics we obtain the second term in Eq. (14). Either for a random isotropic distribution of dipoles or for a cubic array $`L=1/3`$. For other distributions $`L`$ can be computed using the results of Ref. (see below).
By putting $`𝑬=𝑬_0e^{i\omega t}`$ we solve for the dipole moment $`𝚷`$ and compute the polarization vector $`𝑷n𝚷`$ where in the original variables the dipole moment is $`𝚷_ie𝒓_i+q𝑹_i`$ .
Finally from the relation between $`𝑷`$ and $`𝑬`$ one obtains the following dielectric function for the interacting polaron liquid
$$ϵ(\omega )=1\frac{\mathrm{\Omega }_p^2}{\omega ^2+i\gamma \omega }\frac{\omega _p^2\mathrm{\Omega }_p^2}{\omega ^2\omega _{\mathrm{coll}}^2+i\gamma ^{}\omega }$$
(16)
where we have introduced phenomenological inverse relaxation times $`\gamma ,\gamma ^{}`$ and defined the bare plasma frequency $`\omega _p^24\pi ne^2/m`$ and the polaron plasma frequency $`\mathrm{\Omega }_p^24\pi ne(eq)/(M+m)=4\pi ne^2/ϵ_0(M+m)`$. In the latter $`M+m`$ can be identified with the polaron effective mass. We have also defined the renormalized frequency of the internal mode
$$\omega _{\mathrm{coll}}^2\omega _0^2\frac{q}{e}L(1+\frac{m}{M})\omega _p^2$$
(17)
where $`\omega _0\sqrt{k/\mu }`$ is the frequency of the internal mode of an isolated polaron.
The dielectric function Eq. (16) takes a very simple form: the sum of a Drude term plus a collective excitation at frequency $`\omega _{\mathrm{coll}}`$. The latter consists of all electrons oscillating in phase inside their respective self-trapping potentials. In this oscillation polarons develop time-dependent in-phase dipole moments. The dipole-dipole interaction tends to soften this mode and the effect increases as interparticle distance decreases i.e. density increases ($`\omega _p^2`$). Eq (16) can be seen as a generalization of both Drude model and Clausius-Mossotti equation. Notice that the f-sum rule is exactly satisfied.
At some critical density given by $`4\pi n_c/e^2m=\omega _0^2e/qL(1+m/M)`$ the energy cost to displace the electrons away from the respective lattice distortion vanishes and an instability occurs. Within this simplified model we cannot describe the new phase that arises for $`n>n_c`$. In fact we assumed above that electron and distortion are bound and, even more, that the binding potential is harmonic. Both approximations will break down close to $`n_c`$. Interestingly, if a positive quartic term were present in the distortion-electron potential, the system would become a liquid crystal of ferroelectric polarons above the critical density.
A less exotic possibility is that polarons start to dissociate. This can occur either abruptly or in a continuous way. In the former case all polarons collapse at $`n_c`$ whereas the latter case can be realized by a two component system of coexisting polarons and free electrons.
Recently the softening of a polaron band has been observed as a function of doping in the Nd<sub>2</sub>CuO<sub>4-y</sub> system. We have used Eqs. (16),(17) to analyze those data.
Slightly doped cuprates show a polaronic band, phonon bands, a mid-IR band and a charge transfer band (we neglect much weaker magnetic bands). Following Calvani and collaborators we fitted reflectivity data on Nd<sub>2</sub>CuO<sub>4-y</sub> with a Drude-Lorentz model for the dielectric constant (a sum of Lorentzians plus a Drude term, see Ref. ). The model dielectric function is of the same form as in Eq. (16) but for the addition of phonons and of high energy electronic contributions, and with an appropriate $`ϵ_{\mathrm{}}`$ replacing 1. Because of the electronic screening $`\omega _p`$ and $`\mathrm{\Omega }_p`$ in Eq. (16) should be considered screened plasma frequencies. Two Lorentzians were used above $`10^4`$ cm<sup>-1</sup> to fit the charge transfer band and higher energy contributions. Four Lorentzians were used below 800 cm<sup>-1</sup> to model the TO phonons, one Lorentzian above 5000 cm<sup>-1</sup> (depending on sample) were used to fit the mid-IR band (MIR) and finally one Lorentzian below 2000 cm<sup>-1</sup> were used to fit the internal mode of polarons. In Fig. (1) we show the resulting optical conductivity $`\sigma (\omega )\mathrm{Im}ϵ(\omega )\omega /4\pi `$, excluding the electronic contributions and the phonons. Neither $`n`$ nor $`y`$ are well controlled variables so we use $`\omega _p^2(n)`$ as our control parameter which can be determined directly by adding the polaron and the Drude oscillator strength \[Eq. (16)\]. In the inset of Fig. 1 we show $`\omega _{\mathrm{coll}}^2`$ vs. $`\omega _p^2`$. Both quantities are obtained from the fits.
As found by Lupi et al. in this doping range we see a rapid decrease of the internal mode energy. To estimate the rate of decrease we neglect the higher doping point (it is in the region where our approximations break down) and we do a linear regression. We obtain a slope of $`0.5`$.
To get a theoretical estimate of the slope in Eq. (17) we need the Lorentz-Lorenz local field factor $`L`$. Using the results of Ref. we estimate $`L=.6`$ for Nd<sub>2</sub>CuO<sub>4-y</sub> and for the electric field of the radiation parallel to the Cu-O plane.
Also the induced charge in the lattice will not be given any more by Eq. (1). In fact Eq. (1) gives the total induced charge but part of this charge which we call $`q_e`$ is induced in the electronic degrees of freedom. This should be subtracted from $`+q`$ and added to $`e`$ since it will follow antiadiabatically the electron and hence is positioned at $`𝒓_i`$, not at $`𝑹_i`$. This effect is taken into account by replacing $`q/e`$ by $`(qq_e)/(eq_e)`$ in Eq. (17). $`q_e`$ can be estimated, using long wave length Lyddane-Sachs-Teller like arguments, to be of the order of $`q_e=e(11/ϵ_{\mathrm{}})(1+2/ϵ_0)/(1+2/ϵ_{\mathrm{}})`$. Using $`ϵ_0=30`$ and $`ϵ_{\mathrm{}}=5`$ one gets $`(qq_e)/(eq_e)=0.91`$. Finally, as a first approximation, we can take $`(1+M/m)1`$ so we obtain for the slope -0.55 in excellent agreement with the experiment. The linear regression extrapolates to $`\omega _{\mathrm{coll}}=0`$ at a critical value $`\omega _{p,c}^2=7.6\times 10^6`$ cm<sup>-2</sup>.
For $`n>n_c`$ a sudden increase of Drude weight is observed (not shown), that we associate with metallization. This occur at much lower doping than the insulator superconducting transition observed at lower temperatures. We believe that more complicated phenomena like charge ordering may change the picture at low temperatures. Quantum corrections will also be important close to $`n_c`$ and at low temperatures.
Lupi et al. find that the collective mode subsists at a small finite frequency for $`n>n_c`$ and then monotonously decreases as doping is increased at a much slower rate. This suggests a scenario where the polarons stabilize at a concentration slightly below $`n_c`$ and added carriers go to free states. Different versions of such two-fluid model has been considered in the literature. Interestingly, this coexistence implies that for a finite doping range the system is at the verge of a dielectric anomaly with a soft electronic mode. At larger doping, polarons are expected to become unstable due to short-range interactions and screening and a first order transition will occur to a one component Fermi liquid. We speculate that this scenario can explain the phase diagram of cuprates. In fact the soft collective mode can explain many of the anomalous Fermi liquid properties that these systems show at optimum doping levels including pairing formation and superconductivity due to exchange of the soft mode boson. At overdoping the experimental optical conductivity shows a single component Drude behavior which suggest a first order transition. A related scenario including the softening of an optical electronic mode associated with matallization has been proposed previously.
The present theory also explains why a material like doped La<sub>2</sub>NiO<sub>4</sub> never becomes metallic. In this system the observed energy of the internal mode for small doping is roughly by a factor of 3 larger than in Nd<sub>2</sub>CuO<sub>4-y</sub> . Then if we assume that the slopes of $`\omega _{\mathrm{coll}}^2`$ vs. $`\omega _p^2`$ are similar in the two systems we expect, according to Eq. (17), that the critical plasma frequency $`\omega _{p,c}^2`$ in the nickelate is by a factor of 9 larger than in the cuprate. On the other hand, the observed spectral weight in this system reaches a maximum and decreases again as a function of doping. We estimate the maximum $`\omega _p^2`$ by integrating the fitted line shapes in Ref. to be roughly $`9\times 10^6`$cm<sup>-2</sup> which is almost by an order of magnitude smaller than the estimated $`\omega _{p,c}^2`$. This means that on undimensionalized scales ($`\omega _{\mathrm{coll}}^2/\omega _0^2`$ vs. $`\omega _p^2/\omega _{p,c}^2`$) only the first $`10\%`$ of the curve in the inset of Fig. 1 is physically accessible and the system never reaches the point in which it should metallize. Meaningfully a modest softening of the internal mode is actually observed in this relatively small $`\omega _p^2/\omega _{p,c}^2`$ variation range consistent with Eq. (17).
To conclude we have derived a many body Hamiltonian that takes into account dipole-dipole interactions in a polaron system in a general way. We obtained the dielectric constant in mean field for a liquid of such particles and showed that the system becomes unstable for large concentration. We argue that the theory explains a rapid softening of the internal mode of the polaron recently observed in cuprates and discussed the doping-induced matallization of cuprates and nickelates. We speculate that the same framework can be used to explain the occurrence of matallization or not as a function of doping in a wide range of materials where the electron phonon interaction dominates.
We thanks P. Calvani, S. Lupi, M. Capizzi, S. Fratini and P. Quémerais for enlightening discussions and Lupi et al. for providing us their reflectivity data before publication. |
no-problem/9911/hep-ex9911026.html | ar5iv | text | # CP VIOLATION IN 𝐾⁰ DECAYS
## 1 MOTIVATION
There are a few reasons why pursuit of experimental studies of CP violation are important. The first is that we only have one established effect. There are many manifestations, all in the decays of the neutral kaon: $`2\pi `$, $`\pi l\nu `$, $`\pi \pi \gamma `$, and, most recently<sup>a</sup><sup>a</sup>aObservation of CP Violation in $`K_L\pi ^+\pi ^{}e^+e^{}`$ Decays, KTeV Collaboration, hep-ex 9908020 in $`\pi \pi ee`$. But these are all traceable to the $`ϵ`$ impurity in the $`K_L`$ state: the sizes of the asymmetries are precisely predicted and this effect is dubbed “indirect” in that it arises from a (presumably) second order CP- and T- violating interaction in the K-K transition.
The effect is also of cosmological significance. We live in a world of matter: we find no anti-planets, no significant component of anti-matter in extra-galactic cosmic rays, and no evidence for anti-galaxies. In our galaxy, there are approximately $`10^{69}`$ protons and $`10^{79}`$ photons whereas, if we extrapolate this volume back to a time of $`10^6`$ sec from the big-bang, its temperature was about $`10^{13}`$ degrees and it would have contained $`10^{79}`$ protons, anti-protons, and gammas. Nevertheless, it is thought that the baryon asymmetry was present even at this early time so the question remains as to its origin, and whether the CP violation we see in the neutral kaon decays is at all connected.
The standard model “explanation” of the effect (in the kaon system) is compelling and it predicts a direct effect in K decays as well as larger effects in the mixing and decays of B mesons. Now, we have one effect and one parameter ($`\delta `$, the phase in the CKM matrix, or $`\eta `$, the Wolfenstein parameter): we badly need more effects, both within the K and B systems, to see if the model is consistent. Indeed the future of K decay experiments is aimed at such studies as we will see later.
## 2 MEASURING $`ϵ^{}/ϵ`$
The best avenue to see direct CP violation is the study of $`ϵ^{}/ϵ`$ in the neutral kaon decays to two pions. All four modes need to be studied and the double ratio,
$$R=\frac{\mathrm{\Gamma }(K_L2\pi ^0)/\mathrm{\Gamma }(K_S2\pi ^0)}{\mathrm{\Gamma }(K_L\pi ^+\pi ^{})/\mathrm{\Gamma }(K_S\pi ^+\pi ^{})},$$
gives $`ϵ^{}/ϵ`$ through the relation $`ϵ^{}/ϵ`$ = (1-R)/6.
The experimental situation in the early 1990’s led to three new efforts to further search for direct CP violation, these being KTeV at Fermilab (an evolution of the E731 experiment), NA48 at CERN (an evolution of the NA31 experiment), and KLOE at DA$`\mathrm{\Phi }`$NE, an entirely new effort. At the present time, there are first results from both KTeV and NA48, and KLOE is just beginning to take data.
For the present generation of experiments, there are certain common features:
* collection of all four modes simultaneously by means of two beams, one with $`K_L`$ decays and the other with $`K_S`$
* use of high-precision electromagnetic calorimetry for excellent reconstruction of the 2$`\pi ^0`$ decays
* precision magnetic spectrometry.
This leads to relatively low and understandable backgrounds as well.
The areas where there remain differences are the following:
* how do the experiments make $`K_S`$?
* how do the experiments account for the acceptance difference (arising from the differing lifetimes) between $`K_S`$ and $`K_L`$?
* is the analysis done blind?
## 3 KTEV
The first result from KTeV has been published<sup>b</sup><sup>b</sup>bA. Alavi-Harati et al., Phys. Rev. Lett. 83, 22 (1999) so here I will just briefly discuss some of its key features. The KTeV apparatus is shown schematically in Figure 1.
KTeV uses a regenerator to derive its $`K_S`$ beam. Shown in Figure 2 is the reconstructed decay distribution for $`\pi ^+\pi ^{}`$ decays downstream of the regenerator, for events with momentum between 30 and 35 GeV/c. In this bin, the $`K_S`$ lifetime is only about 1.7 m so that, to normalize the $`ϵ^{}/ϵ`$ measurement, only a few meters would be needed. As the $`K_S`$ component decays, it becomes of comparable magnitude to that of the much smaller $`K_L`$ component and their interference is most noticable, at about 15 m downstream of the target. Thus using all the decays downstream of the regenerator gives access to the $`K_S`$ lifetime, the mass difference between $`K_S`$ and $`K_L`$, and the phase difference between the two decays, this for both $`\pi ^+\pi ^{}`$ and 2$`\pi ^0`$ decays. And the ability to understand these distributions gives confidence that the detector and beam are well understood.
There are two drawbacks to the use of a regenerator. The first is the ambient rate that it causes in the detector. Essentially all the kaons and neutrons incident upon it interact<sup>c</sup><sup>c</sup>cThe KTeV regenerator is about 1.8 kaon interaction lengths and about 3.5 neutron interaction lengths. and therefore potentially spray unwanted “accidental” particles into the detection elements. The beam which strikes the regenerator is filtered and attenuated to enhance the K/n ratio and to reduce the overall rate. Nevertheless, the interaction rate in the regenerator is about 1.5 MHz and this source contributes the majority of the ambient rate in the detectors.
This turns out to not be a major problem for several reasons. The first is that the regenerator is fully active, consisting of 85 separate scintillator pieces each read out by two photomultipliers. The FNAL beam structure is 53 MHz, meaning that approximately 3% of the buckets are occupied, with either a neutron or a kaon, or, rarely, more than one such particle<sup>d</sup><sup>d</sup>dIt is important that the intensity from bucket to bucket be uniform so attention must be paid to monitoring beam microstructure.. Whenever there is an inelastic interaction producing particles that could hit the detector, it is effectively vetoed by the signals in the regenerator. So the activity will be “out of time.”
The KTeV detector with the longest latency, and therefore most sensitive to such activity, is the drift chamber system where drifts up to approximately 200 ns occur. Out-of-time activity can interfere with the real track information if it occurs on the relevant wires. This effect has been extensively studied and to a very high order is properly simulated by simply superimposing accidental events upon monte-carlo generated “pure” events. By such studies, we have determined that the overall effect of the ambient rate from the regenerator shifts the ratio of $`K_S`$ to $`K_L`$ (“single ratios”) by less than 0.001 in each ($`\pi ^+\pi ^{}`$, 2$`\pi ^0`$) case. There is a loss of events, at the level of 2%, due to such accidental activity in the chambers, but the loss is quite symmetric between the two beams. The main reason for this is that the “spray” from the regenerator is broad enough that it effectively is uniform over the drift chambers; were the ambient particles confined to a small region about the (regenerator) beam region, the difference would be greater.
The second drawback from the use of the regenerator is incoherent scattering of $`K_S`$. When such scatters produce extra particles (inelastic events), they are effectively self-vetoed. But elastic scattering off Carbon nuclei produce negligible recoil energy so that these events smear out the coherently regenerated beam and even can “cross-over” to the vacuum side. Figure 3 shows this effect for neutral events at the calorimeter.
Fortunately this effect can be directly measured using the $`\pi ^+\pi ^{}`$ events themselves. For the angular distribution of this process is independent of how the kaon decays. Hence the distribution is determined with the charged sample (using the magnetic spectrometer) and this can be used to accurately give the effects in the neutral decays. Small uncertainties in the charged acceptance for this scattered component, however, give a systematic uncertainty of order 0.0005 in the double ratio.
We show in the table below the performance of each of the KTeV systems in comparison to that for the previous generation of experiments at FNAL. The calorimeter, made of pure CsI, has achieved a resolution of 0.75% averaged over the momentum spectrum.
The Monte-Carlo simulation in KTeV is most important because the vertex distributions are so different between $`K_S`$ and $`K_L`$. Accepted $`K_S`$ events reconstruct on average about 5m upstream of $`K_L`$ so, to control the ratio of recorded events to better than 0.001 means that we should understand any instrumental induced “slope” in the acceptance at the level of about 0.02% per meter.
The best means to study the acceptance in the data is by using the high statistics decay modes that are taken simultaneously with the 2$`\pi `$ decays. Figure 4 shows the decay distribution for about 40 million $`Ke_3`$ decays together with the monte-carlo simulation. Fitting the ratio of the two to a linear departure shows that the distrubutions match at the level of about 0.005% per meter. It is relevant to point out that the lifetime of the $`K_L`$ itself would contribute a slope in this plot of about 0.05% per meter. However, the same distribution for our $`\pi ^+\pi ^{}`$ sample shows a slope three times greater, about a 2.5 standard deviation effect; accordingly, we use this larger slope as a measure of the possible systematic uncertainty associated with acceptance.
Making the acceptance corrections, we then look at the momentum dependence of the “regeneration” amplitude for each of the modes. This is expected to exhibit a power-law behavour and a very important check is that the power be the same for the two modes. We find indeed that the regeneration amplitude can be well represented by the form $`ff/k\alpha p^\alpha `$ and for the parameter $`\alpha `$, we find 0.5890(15) and 0.5884(19) for the charged and neutral modes respectively.
We fit for the $`K_S`$ lifetime and find consistent results for each mode. The KTeV (preliminary) value is 0.8967(7) $`\times 10^{10}`$ s and this is shown in comparison with other recent measurements in Figure 5. Again, in fitting for the mass difference, we again find internally consistent values with the KTeV value being 0.5280(13) which is shown in the Figure. Finally, we determine the potentially CPT violating phase difference between the two CP violating amplitudes, finding $`\mathrm{\Delta }\mathrm{\Phi }=0.09(46)^0`$, as is shown in Figure 6. And, at the suggestion of Alan Kostelecky<sup>e</sup><sup>e</sup>eSensitivity of CPT Tests with Neutral Mesons, V.A. Kostelecky, Phys. Rev. Lett. 80, 1818 (1998)., we have examined our data for a diurnal variation in $`\mathrm{\Phi }_+`$ as might be predicted in certain CPT and Lorentz violating interactions, finding no day-night effect at the level of about 1/3 degree<sup>f</sup><sup>f</sup>fTo report a value for $`\mathrm{\Phi }_+`$, one needs to worry about certain systematic uncertainties in the phase of the regeneration amplitude. But such are irrelevant if one is just interested in the time dependence of $`\mathrm{\Phi }_+`$.
### 3.1 Looking at the Answer
The $`ϵ^{}/ϵ`$ analysis is done blind in KTeV.
We use as much as possible the high statistics modes for study of the detector and to improve our modeling of it, rather than the 2$`\pi `$ modes from which the answer is calculated. The acceptance calculations are validated using $`Ke_3`$ and 3$`\pi ^0`$ decays. Then the 2$`\pi `$ samples are examined and the $`\tau _s`$ and $`\mathrm{\Delta }m`$ parameters are determined and with high precision are found to be consistent between the two modes. A precision CPT test is done, to the level of 1/2 degree. And the regeneration powers are found to be equal with high precision. Finally, all systematic studies are completed and the value for each error is tabulated.
It is only after all of these checks, studies, and evaluations that the answer is uncovered – the decision to commit to a result is made beforehand.
The result was Re($`ϵ^{}/ϵ`$) = (28.0 $`\pm `$ 4.1) $`\times 10^4`$ where the statistical error was 0.0003 and the systematic 0.00026. At nearly seven standard deviations, this establishes direct CP violation.
## 4 NA48
The NA48 experiment has given its first preliminary result very recently. A sketch of their very elegant beam arrangement is shown in Figure 7. They too use two beams but the $`K_S`$ beam is derived from a small fraction of the primary beam that is diverted and then targetted upon a close-by target. The two beams have different angular divergences but cross at the detector. (In KTeV, the beams are precisely the same divergence but are separated at the detector.)
The performance of the NA48 calorimenter is excellent, allowing 0.7% resolution at high energies.
In NA31, the $`K_S`$ target was serially stepped through the decay region to approximate $`K_L`$ decays; in this way, acceptance corrections were minimal. NA48 does not use this technique but rather has elected to weight their $`K_L`$ events according to the distribution of $`K_S`$ events. This results in a significant statistical loss but has the advantage that the corrections that need to be applied, now to the weighted event ratios, are less than 1% whereas those for KTeV are of order 5%.
One potential problem that the group has studied extensively is an energy dependence in the double ratio. When fit for a linear slope, a 3 standard deviation departure is found with the value changing by about 5% over the span of energies from 70 to 170 GeV/c. But extensive studies of the data, including examining the double ratio in energy bins beyond their nominal fiducial region, has convinced the group that this is a statistical fluctuation.
The preliminary result they report, based upon about 10% of their anticipated sample, is: Re($`ϵ^{}/ϵ`$) = (18.5 $`\pm `$ 7.3) $`\times 10^4`$ where the systematic error is larger than the statistical one. However, most of the systematic error is dominated by statistics so that, for awhile anyway, the total error will roughly scale with added data.
## 5 THE GRAND AVERAGE
The most recent results are shown graphically in Figure 8. The grand average of these last measurements is (21.2 $`\pm `$ 2.8) $`\times 10^4`$, with a confidence level of about 7%.
The KTeV result is in better agreement with that from NA31 rather than from E731 but the experimenters have not found any reason other than a fluctuation to account for this difference. The E731 beam, detector, and analysis were extensively documented in a long article<sup>g</sup><sup>g</sup>gCP and CPT Symmetry Tests from the Two-Pion Decays of the Neutral Kaon with the Fermilab E731 Detector with L.K. Gibbons et al., Physics Review D55, 6625 (1997) for Physical Review D. At present the NA48 result is in good agreement with those from E731, KTeV, and NA31.
The theoretical situation is still developing. Figure 9 shows this grand average along with some recent predictions. It is still too early to say if the rather large value points to new physics or to inadequacies in the standard model calculations. The goals of the next round of lattice calculations, aiming at 10% determinations of the matrix elements, would mean that $`ϵ^{}/ϵ`$ would become a “powerful precision test and new physics probe<sup>h</sup><sup>h</sup>hBill Marciano, summary talk of the Chicago Kaon Conference, June 1999, editors Jon Rosner and Bruce Winstein, to be published by the University of Chicago Press..”
## 6 RARE DECAYS
In this section, I will briefly mention other studies of rare K decays that have an implication for CP violation.
### 6.1 $`K_L\pi ^0ll`$
A theoretically clean channel<sup>i</sup><sup>i</sup>iL. Littenberg, Phys. Rev. D39, 3322 (1989). is the decay of the long-lived Kaon to a neutral pion, neutrino, anti-neutrino pair. This is mediated by the Z-penguin, with a branching ratio given by<sup>j</sup><sup>j</sup>jB. Winstein, and L. Wolfenstein, Rev. Mod. Phys. 65, 1113 (1993).:
$$BR(K_L\pi ^0\nu \overline{\nu })=8\times 10^{11}(M_t/M_W)^{2.2}A^4\eta ^2$$
where $`\eta `$ is the parameter in the CKM matrix.
This mode has been the by-product of a number of searches by Fermilab experiments. The latest result comes from the KTeV experiment which reports<sup>k</sup><sup>k</sup>kJ. Adams et al., hep-ex/9806007.:
$$BR(K_L\pi ^0\nu \overline{\nu })<5.9\times 10^7(KTeV,90\%\text{confidence}).$$
Less clean but still interesting is the ee mode. The KTeV 90% confidence value is 6.64 $`\times 10^{10}`$ and the similar limit<sup>l</sup><sup>l</sup>lJ. Whitmore, Chicago Kaon Conference, ibid. for the $`\mu \mu `$ mode is 3.4 $`\times 10^{10}`$. These are both an order of magnitude improved over the corresponding E731 results and while not quite at the level expected in the Standard Model, are beginning to rule out certain parameter regions in an extended SUSY scheme.<sup>m</sup><sup>m</sup>mSupersymmetric contributions to rare kaon decays: beyond the single mass-insertion approximation, G. Colangelo and G. Isidori, hep-ph/9808487.
## 7 FUTURE RARE DECAY STUDIES
The prime goal, now that direct CP violation has been seen, in the K system is the study of the $`\pi ^0\nu \nu `$ decay and there are several groups seriously considering this mode. Many institutions in KTeV have written an expression of interest to pursue this mode at Fermilab, using the much higher intensity of the Main Injector. There is a similar proposal at BNL. The goal is the collection of about 100 events. The idea is that with this mode and the corresponding $`K^+`$ decay, one can accurately determine the angle $`\beta `$ of the unitarity triangle in the K sector and this can be compared to a similar determination in the B sector.
The E787 experiment at Brookhaven has the best information on the $`K^+`$ decay: they have seen one unambiguous event which corresponded to a branching ratio<sup>n</sup><sup>n</sup>nGeorge Redlinger, Chicago Kaon Conference, ibid. in the range of about 2 $`\times 10^{10}`$. They have a proposal to increase their sensitivity to allow about 10 events to be collected. And the CKM proposal at Fermilab would collect about 100 such events.
These experiments, if successful, would allow a test at the level of a few degrees and thus will be quite sensitive to any new physics.
## 8 CONCLUSION
KTeV has reported on just 20% of the data collected in 1996/7. And more data is being collected now in the 1999 run, with a variety of systematic checks, so that the accuracy on $`ϵ^{}/ϵ`$ should improve significantly.
NA48 has also collected more data and should run as well in 2000. And KLOE is collecting its first samples now. So, during the next few years, we should see several high precision determinations of Re($`ϵ^{}/ϵ`$). The theorists are actively looking to improve their predictive power so that we should know on that time scale if we are seeing new physics. One thing is certain: a new CP violating effect has been established. It will be quite interesting to see if such an effect can be isolated in B decays and in any case new CP violating signatures in that sector, in comparison with studies to be made in the kaon decays, will tell us if the CKM picture is complete, or if new mechanisms are needed.
## 9 ACKNOWLEDGEMENT
The author would like to acknowledge the conference organizers for setting up a most stimulating environment, and Kathy Visak for help in the preparation of this manuscript. |
no-problem/9911/astro-ph9911348.html | ar5iv | text | # Photometry of pulsating stars in the Magellanic Clouds as observed in the MOA project
## 1. Introduction
The MOA project Microlensing Observations in Astrophysics is a collaboration of about 30 astronomers/physicists in New Zealand and Japan working in about 10 different institutions (headquarters Nagoya and Auckland universities). The project was established in 1995. The MOA project goals are (a) the search for dark matter by microlensing in the galactic halo, (b) the search for planets by microlensing, in halo and galactic bulge, (c) the study of variable stars in the Magellanic Clouds and bulge and (d) the optical identification of $`\gamma `$-ray bursters (GRBs)
The observational program is undertaken on the 61-cm Boller & Chivens telescope at the University of Canterbury’s Mt John University Observatory in the centre of the South Island of New Zealand. Here there is a dark sky background and typically 1–3 arcsec seeing; between a third and a half of the weather is suitable for CCD photometry. At the latitude of $`44^{}`$ S, the Magellanic Clouds are circumpolar and can be observed all year.
The telescope is a Ritchey-Chrétien cassegrain instrument modified to accommodate wide-field f/6.25 optics and a modern computer control system.
## 2. Observing program
Three series of MOA observations have been undertaken as follows
1. Jan 1996 – Dec 1996 f/13.5 Cass. $`30^{}\times 30^{}`$, MOAcamI
2. Jan 1997 – Aug 1998 f/6.25 Cass. $`1^{}\times 1^{}`$, MOAcamI
3. Aug 1998 – present f/6.25 Cass. $`0.92^{}\times 1.39^{}`$, MOAcamII
In all cases, very broad-band $`B`$ and $`R`$ filters have been used, covering respectively about 395–620 nm ($`\lambda _{\mathrm{eff}}500`$ nm) and 620–1050 nm ($`\lambda _{\mathrm{eff}}700`$ nm) (see Fig. 1). Series 1 used the f/13.5 optics and encompassed $`3\times 10^5`$ stars in three LMC fields. Series 2 covered three fields (nlmc1,2,3) in the LMC bar ($`1\times 10^6`$ stars) plus two SMC fields (smc1,2) ($`4\times 10^5`$ stars).
For series 3, which is currently in progress, the new MOAcamII is used to observe 16 LMC fields (all year) ($`3\times 10^6`$ stars) plus 8 SMC fields (all year) ($`1\times 10^6`$ stars) plus 16 galactic bulge fields (winter only)($`5\times 10^6`$ stars). Each of the MOAcamII fields covers about 1.25 square degrees. Exposure times are 5 min. in $`B`$ and $`R`$ in the Clouds, 3 minutes in the Galactic bulge.
The CCD camera MOAcamI comprises a mosaic of nine TI TC215 $`1000\times 1018`$ pixel CCDs whose pixel size is 12 $`\mu `$m ($`0.3^{\prime \prime }`$ at f/13.5, $`0.65^{\prime \prime }`$ at f/6.25). The chips are not butted, so four consecutive interleaved exposures are required to cover a $`1^{}\times 1^{}`$ field.
The recently acquired MOAcamII has now replaced the first camera. It consists of three butted SITe $`2048\times 4096`$-pixel thinned, back-illuminated CCDs with 15-$`\mu `$m pixels ($`0.81^{\prime \prime }`$ at f/6.25). The peak QE is 85%. The area of the CCD surface is $`6.1\times 9.2`$ cm, and there are 24 Mpixels. Both cameras operated at a temperature of $`100^{}`$ C with liquid N<sub>2</sub>.
Generally we expose each Magellanic Cloud field once or sometimes twice per night in both $`B`$ and $`R`$, and each Galactic bulge field about four times per night in $`R`$ and once in $`B`$, unless a known high magnification microlensing event is in progress, in which case frequent successive exposures are made.
## 3. Data reduction
MOA observations of CCD images are recorded to 8-mm Exabyte tape for sending to Japan, while an archival copy on DLT tape is retained at Mt John. For series 1 and 2 observations, all the reductions have been completed using the DoPhot software package. One of us (I.B.) completed an archived database of all series 1 and 2 frames giving magnitudes in $`B`$ and $`R`$. In the case of series 2 images, there are some 200 data points in each colour and about $`1.4\times 10^6`$ stars.
Series 3 images using MOAcamII have hitherto not been reduced. I. Bond is currently perfecting implementation of the Alard & Lupton (1997) difference imaging technique, which already has shown that a substantial improvement in photometric precision is possible for crowded fields in only moderate seeing. It is hoped to reduce series 3 images early in 2000 and then to be able to reduce all images on site at the observatory in nearly real time. A Sun Enterprise 450 computer with 90 Gbyte hard disk drive is used for on-site data analysis at Mt John.
Mt John $`B,R`$ MOA photometry gives colours which transform linearly to Johnson colours, as calibrated by Reid, Dodd & Sullivan (1997), as follows:
$$(BV)_\mathrm{J}=0.356+1.036(BR)_{\mathrm{MOA}}$$
$$(RI)_\mathrm{J}=0.150+0.612(BR)_{\mathrm{MOA}}$$
The calibration of MOA magnitudes gives the following transformation from the MOA instrumental system to that of the HST Guide Star Catalogue.
$$m_R(\mathrm{GSC})=m_R(\mathrm{MOA})+25.10$$
$$m_B(\mathrm{GSC})=m_B(\mathrm{MOA})+24.82$$
The typical precision of the series 2 MOA magnitudes is $`\pm 0.13`$ mag. at $`B`$ or $`R19`$; $`\pm 0.04`$ mag. at $`B`$ or $`R14`$ – 15.
## 4. Variable stars in the MOA database
Three surveys on variable stars in the MOA database have been undertaken or are in progress. One of these has been by S. Noda (Nagoya Univ., MSc) and M. Takeuti (Tohoku Univ., Sendai), and is a study of variables in series 2 observations. The stars are selected by several criteria (see Noda et al. in these proceedings). Periods have been obtained by the phase dispersion method (PDM). Important aspects of this study include:
* Study of red variables – mainly SR; 188 variables in the LMC, 35 in the SMC
* Study of eclipsing binaries; 69 EB in the LMC, 136 in the SMC
* Study of Cepheids; 78 in the LMC, 38 in the SMC
* Study of blue variables (many detected); may be Be stars, at least one is a so-called bumper, resembling a microlensing light curve
By using another selection, 282 AGB variables were found in the LMC and 69 in the SMC (see Takeuti et al. in these proceedings). The periods and the $`K`$-band magnitudes of most of these stars are comparable with those in the literaure. The relation between the period and the mean colour for the AGB variables is reported.
Another study has been by N. Rattenbury (Auckland Univ., MSc), using selected MOA fields for series 1 and 2 data. Variable stars were detected by deriving the Welch-Stetson variability index (Stetson, 1996; Welch & Stetson, 1993). Only about 2% of the stars with greatest variability were selected for further study in this preliminary survey. Typically these stars have a range in the $`R`$ magnitude exceeding $`12\sigma `$.
Periods were analysed in several ways, the most successful algorithm being the date compensated discrete Fourier transform (DCDFT) algorithm of Ferraz-Mello (1981). The following summarizes the preliminary results.
* Study of Cepheids (periods 1–10 d): 70 were identified in the LMC and 420 in the SMC
* Study of Cepheids or semi-regular variables ($`P`$ 10–50 d): 433 in the LMC, several more in the SMC
* Eclipsing binaries and RR Lyrae stars ($`P`$ 1–20 h): 7 in LMC, 71 in SMC
* Blue variables ($`(BV)_\mathrm{J}<0.36`$): 37 were identified in the SMC including 7 of long or irregular period, 2 possible “bumpers”, and at least 3 eclipsing binaries.
* Red variables with $`(BV)_\mathrm{J}>1.65`$ and $`P>100`$ d (LPV, SR): 89 were found in the SMC and 5 in the LMC
Fig. 2 shows $`B`$ and $`R`$ phased light curves for four of the Cepheids in the field smc2 (series 2). Fig. 3 shows unphased quasi-periodic light curves for four long period variables from field nlmc2 (series 2).
Finally G. Bayne, W. Tobin and J. Pritchard (Univ. Canterbury, N.Z.) are searching the MOA database for eclipsing binaries and analysing the light curves (series 1, 2). This work is still in progress, using the period-searching algorithm of Grison (1994).
## References
Alard, C. and Lupton, R. 1998, ApJ 503, 325
Grison, P. 1994, A&A 289, 404
Reid, M., Dodd, R.J. and Sullivan, D.J. 1997, Australian J. Astron. 7, 79
Ferraz-Mello, S., 1981, AJ 86, 619
Stetson, P. B., 1996, PASP 108, 851
Welch, D.L. and Stetson, P.B., 1993, AJ 105, 1813 |
no-problem/9911/cond-mat9911036.html | ar5iv | text | # References
A Model for
Double-Stranded Excitations of DNA
V.L. Golo<sup>+</sup>, E.I. Kats<sup>++</sup>, and Yu.M. Yevdokimov<sup>+++</sup>
<sup>+</sup>Moscow University, Department of Mechanics and Mathematics, Moscow, Russia
<sup>++</sup> Landau Institute, Moscow, Russia
<sup>+++</sup>Institute of Molecular Biology,Moscow, Russia
ABSTRACT
We calculate the spectrum of torsional vibrations of a double-stranded structure that models the double helix of the DNA. We come to the conclusion that within the framework of the model elementary exictations may display an asymmetry as regards their winding and direction of the propagation, depending on initial polarization. The asymmetry could have a bearing on processes that take place in molecules of the DNA.
Conformational excitations of the DNA molecule are important for understanding transformations of DNA, for example during transcription, and therefore there is an urgent need for their investigation. But, on the other hand, the latter is a difficult problem, even on the assumption that the excitations are weak enough to preserve the hydrogen bonds between base-pairs, so that the double-stranded structure is deformed but not disrupted. As to our knowledge, it was Cappelmann and Beim, , who were the first to attempt such a study. They gave a model of DNA that accommodates a number of modes of elastic vibrations with sophisticated spectrum, and the analysis of their bearing on the actual properties of DNA needs further studying. As usual in such cases, the most important thing is the wise choice of an initial model, which determines the final outcome; taking into account the qualitative nature of the approach, one should employ the minimal number of assumptions so as to obtain simple characteristics of the dynamics of a molecule. In our work we aim at studying elastic modes of vibrations of the double-stranded structure, which are similar to the phonon modes of crystalline lattices. We show that the asymmetry of the double helix, due to its having an orientation of the right or the left screw, may result in the propagation of pairs of waves travelling in opposite directions and producing twists of different signs in the double-stranded structure.
According to the results of , , the number of open base pairs at room temperature is very small, so that the system of H-bonds may be considered intact, within a good approximation. On the other hand the H-bonds are more fragile than the sugar-phosphate frame, and therefore it appears to be a reasonable proposition to consider excitations for either of them separately. Thus, at least for sufficiently small excitational energies, we may assume the conformation of the sugar-phophate frame being static, while the H-bonds subject to small deformations. Hence, the minimal list of basic properties of the DNA molecule under the circumstances may be drawn as follows:
* a molecule of DNA is an elastic system at the spatial scale of the order $`500\AA `$ , or less;
* the discrete nature of the system is to be taken into account at the intra base-pair distances, that is $`3.4\AA `$;
* the conformation of the double-stranded structure of DNA may be described, qualitatively, by the vector field $`\stackrel{}{Y}`$ defined at points of the central line of a molecule, with $`\stackrel{}{Y}=0`$ corresponding to the unperturbed state, (cf. );
* the double helix of the two strands means that the system has a twisted ground state characterized by the twist vector $`\stackrel{}{\mathrm{\Omega }}`$.
Consequently, the dynamics of the double helix is to be determined by the twist, and therefore the vector $`\stackrel{}{\mathrm{\Omega }}`$.
In accord with the requirements formulated above we may cast the energy of a molecule in the form
$$E=\underset{i=1}{\overset{N}{}}\frac{1}{2M}(_t\stackrel{}{P}_i)^2+\underset{i=1}{\overset{N}{}}\frac{K}{2}(\stackrel{}{Y}_i)^2+\underset{i=1}{\overset{N}{}}\frac{L}{2}\stackrel{}{Y}_i^2$$
where $`\stackrel{}{Y}_i`$ is the value of $`\stackrel{}{Y}`$ at the site corresponding to the i-th base-pair, and the covariant derivative $`\stackrel{}{Y}_i`$ is given by the equation
$$\stackrel{}{Y}_i=\frac{1}{a}(\stackrel{}{Y}_{i+1}\stackrel{}{Y}_i)+\frac{1}{a}\widehat{\mathrm{\Omega }}_i\stackrel{}{Y}_i$$
in which $`a`$ is the distance between adjacent base-pairs and the matrix $`\widehat{\mathrm{\Omega }}_i`$ is determined by the formula
$$\widehat{\mathrm{\Omega }}_i=\widehat{R}_i^1(\widehat{R}_{i+1}\widehat{R}_i)$$
Here the matrix $`\widehat{R}_i`$ is the matrix of a rotation that transforms the local coordinates at the site $`i`$ towards the laboratory one. It is important that the matrix $`\widehat{\mathrm{\Omega }}_i`$ corresponds to the twist of the double strands of the helix, and one can describe it by the vector $`\stackrel{}{\mathrm{\Omega }}`$, mentioned above. One can derive the dynamical equations for the field $`\stackrel{}{Y}`$ from the energy given above as usual. The set of values for the matrix $`\widehat{\mathrm{\Omega }}_i`$ is supposed to be fixed by the equilibrium ground state of the double helix. It is worth noting that, to a certain extent, the distribution of base-pairs of different shapes can be mimicked by the choice of $`\widehat{\mathrm{\Omega }}_i`$.
Assuming that the matrices $`\widehat{\mathrm{\Omega }}_i`$ do not depend on the site $`i`$, which means that some spatial homogeneity could be accepted, we are looking for eigenmodes that can be cast in the form
$$\stackrel{}{Y}_n=e^{inqa}Y_q$$
and obtain the modes having the spectrum
$`\omega _{1,2}^2`$ $`=`$ $`v^2\left[4A^2\mathrm{sin}^2\left({\displaystyle \frac{aq}{2}}\right)\pm 2\mathrm{sin}\varphi \mathrm{sin}(aq)+4\mathrm{sin}^2\varphi +\mathrm{\Delta }^2\right]`$
$`\omega _3^2`$ $`=`$ $`4v^2\mathrm{sin}^2\left({\displaystyle \frac{aq}{2}}\right)`$
$`v`$ $`=`$ $`\sqrt{{\displaystyle \frac{K}{Ma^2}}}`$
in which $`A^2`$ is given by the equation
$$A^2=1+2\mathrm{sin}^2\left(\frac{\varphi }{2}\right)$$
the angle $`\varphi `$ is that of the rotation matrix $`\widehat{\mathrm{\Omega }}`$. The spectrum is illustrated in Fig.1. It is worth noting that a similar form of the spectrum is obtained for the linearized expression for matrices $`\widehat{\mathrm{\Omega }}_i`$, that is the rotations
$$\widehat{R}_i^1\widehat{R}_{i+1}$$
being considered infinitesimal (see ). The agreement between the two approaches is in accord with the qualitative nature of the model we consider.
One can infer an important conclusion from the shape of the spectrum .
Let us consider the first mode $`\omega _1`$ which corresponds to the minus sign in the equation given above. The frequency $`\omega _1`$ has a pronounced minima at $`q_{}`$, so that there are two sets of waves travelling in opposite directions of the molecule. It should be noted that the double helix has an orientation, the left or the right one, due to its screw structure. Thus, we may talk about the sign of the propagation of a wave. The value of $`\omega _1`$ at $`q_{}`$ is not zero, and it results in a winding motion of the vector $`\stackrel{}{Y}`$ with an orientation determined by the sign of the propagation, so that for a pair of waves belonging to the eigenmode $`\omega _1`$ the windings must be of different orientaion. We see that the effect depend on the polarization of a wave, in fact we have considered an eigenmode. In case we have a mixture of them, the asymmetry of the winding can be blurred or totally destroyed.
One may cast these arguments in a more quantitative form. Let us consider the first mode; its spectrum can be roughly approximated by the expression
$$\omega =c|qq_0|+\omega _0$$
or
$$\omega =\{\begin{array}{cc}cq\omega _1\hfill & \text{if }q>q_0\hfill \\ cq+\omega _2\hfill & \text{otherwise}\hfill \end{array}$$
(see Fig.2), in which the frequencies $`\omega _1,\omega _2`$ read
$$\omega _1=\omega _0cq_0,\omega _2=\omega _0+cq_0$$
It is important that for the numerical values involved, that is corresponding to the DNA data, the frequencies $`\omega _1,\omega _2`$ are positive, as is seen in Fig.2.
If we are looking for a solution with the help of the Fourier transform, we may write its expression in the form of the decomposition
$$u(x,t)=_{\mathrm{}}^{\mathrm{}}𝑑qu_q(t)$$
in which $`u_q(t)`$ reads
$$u_q(t)=u_q^0e^{i(\omega tqx)}$$
and $`\omega `$ depends on $`q`$. It is easy to convince oneself that the solution can be cast in the form
$$u(x,t)=_{\mathrm{}}^{q_0}𝑑q[u_q^0e^{i(t+x)q}]e^{i\omega _2t}+_{q_0}^{\mathrm{}}𝑑q[u_q^0e^{i(tx)q}]e^{i\omega _1t}$$
or, otherwise
$$u(x,t)=u_+(t+x)e^{\omega _2t}+u_{}(tx)e^{\omega _1t}$$
The latter desribes the two waves travelling in opposite directions and having different windings.
We made a numerical simulation of the dynamics of our model. The asymmetry is seen in that two waves travelling from an initial disturbance in opposite direction may oscilate having equal phases of different sign. As is expected, the effect of the asymmetry depends on an initial polarization, which corresponds to a superposition of the eigenmodes indicated above, so that for a certain direction it can be totally suppressed, whereas for those belonging to the eigenmodes it is pronounced.
On considering the estimates for the values of the density per unit of length, approximately $`10^{14}gr/cm`$, and according to the papers , ,the elastic constant, $`10^6erg/cm`$, me may suggest, therefore, that the velocity of the waves considered above is of the order $`10^5cm/sec`$. Consequently, the frequencies of the modes, even for the region of wavenumbers corresponding to the dips of the spectrum seen in Fig.1, are those of hypersound, that is $`10^{11}`$ or $`10^{12}`$ or even more, depending on the values of the elastic constant. The values are by many orders of magnitude above those involved in biological process. Their relation to the latter presents a very interesting problem. In fact, we shall investigate the influence of super-high frequency dynamics on the very low frequncy one. It should be noted that so far no electronic properties have been taken into account in conjunction with the elastic ones, and in this respect the approach of elasticity theory, generally employed now, might be incomplete. Studying the elastic excitattions in the DNA could be a means for finding the answer.
Another important question is the influence of ambient media that should result in a very strong dissipation of excitations. Whether they will be overdamped, and to all practical purposes unobservable , as appears to be the case in a similar situation in proteins, remains to be seen. A strong argument in favour of their playing an important part in the dynamics of the DNA, is the regidity of the latter displayed by its large persistence length.
At any rate, the investigation of elementary excitations should be a necessary prerequisite for investigating conformational changes due to the action of external agents, e.g. enzymes. The salient feature is that in these circumstances a preferential direction of the promoter’s action is often observed , as well as a torsional dynamics of the double helix. Thus, the anysotropy of the propagation of a disturbance due to the screw orientation of the double helix for which we make out a case in this paper, appears to be very interesting and worth further studying.
This research was supported by RFFR. One of the authors (VLG) is thankful to CNRS for the financial support, and ENS-Lyon for the hospitality. The authors are thankful to Michel Peyrard for many fruitful discussions and comments. |
no-problem/9911/astro-ph9911014.html | ar5iv | text | # 1 The figure on top shows the measured EGRET flux for PKS 1622-297 (Mattox et al. 1997) with the large error bars [green in a color rendition]. The solid [blue] line is a plausible hypothesis for the actual flux. Assuming this flux, simulated GLAST measurements are shown with much smaller error bars [red]. The figure on the bottom shows the simulated GLAST measurements [red] along with simulated ATN optical observations [green]. It is assumed that the 𝛾-ray flux varies linearly in relationship to the optical with a lag of 2000 seconds. The optical observation interval is 30 minutes.
To appear in the Proceedings of the OJ-94 Annual Meeting 1999 “Blazar monitoring towards the third millennium”, held in Torino, Italy, May 19–21, 1999
The Automatic Telescope Network
J.R. Mattox
Institute for Astrophysical Research, Boston University; Email: Mattox@bu.edu
Abstract
Because of the scheduled GLAST mission by NASA, there is strong scientific justification for preparation for very extensive blazar monitoring in the optical bands to exploit the opportunity to learn about blazars through the correlation of variability of the gamma-ray flux with flux at lower frequencies. Current optical facilities do not provide the required capability.Developments in technology have enabled astronomers to readily deploy automatic telescopes. The effort to create an Automatic Telescope Network (ATN) for blazar monitoring in the GLAST era is described.
1. Introduction
The EGRET $`\gamma `$-ray telescope aboard the Compton Observatory has detected $``$1 GeV emission from $``$70 blazars (Hartman et al. 1999, Mattox 1999a). The apparent $`\gamma `$-ray luminosity seen for the EGRET blazars is as much as one hundred times larger than that at all other wavelengths for some flaring EGRET blazars. Variability of the $`\gamma `$-ray flux from some blazars on a time-scale as short as 4 hours has been observed (Mattox et al. 1997). This implies that the region of $`\gamma `$-ray emission must be very compact. Because the opacity for $`\gamma `$-ray to $`\gamma \gamma `$ pair production with x-rays must not prevent $`\gamma `$-rays from escaping, relativistic beaming with a Lorentz factor of $``$10 for the bulk of material in the jet is required (Mattox et al. 1997). This conclusion is reinforced by the observation of a high x-ray state during the 1996 $`\gamma `$-ray flare of 3C 279 (Wehrle et al. 1998) which implies that the x-rays originate in the same volume as the $`\gamma `$-rays). With substantial accretion onto a $``$$`10^8`$ M blackhole, there is sufficient power to create the relativistic jets. However, the physics involved in the conversion of gravitational potential to kinetic luminosity is not understood.
It is widely believed that this GeV emission is due to inverse-Compton scattering by shock-accelerated leptons within the relativistic jet. However, there is disagreement over the origin of the $``$1 eV photons which are scattered. Some modelers believe that they originate in the synchrotron emission of the leptons, so the $`\gamma `$-rays are a result of the synchrotron self-Compton (SSC) process (Bloom & Marscher 1993).
Another possibility is that the low energy photons come from outside of the jet. This is designated as the external Compton scattering (ECS) process. Dermer, Schlickeiser, & Mastichiadis (1992) suggested that they come directly from an accretion disk around a black hole at the base of the jet. It was subsequently proposed that the dominant source of the low energy photons for scattering could be due to re-processing of disk emission by broad emission line clouds (Sikora, Begelman, & Rees 1994).
The correlation of optical, x-ray, and $`\gamma `$-ray emission is expected to provide a definitive test of these models. If the variation in the synchrotron flux is due to a change in the electron density, the SSC emission, which depends on the second power of electron density, will be observed to vary quadratically in comparison to the synchrotron with no lag (for a single homogeneous emission zone). If the high-energy emission is ECS due to reprocessed in the broad line region before inverse Compton scattering (Sikora Begelman and Rees 1994), the ECS flux will lag the optical disk flux by at least 1 day if the ECS flare is due to increased optical emission from the disk. If the ECS flare is due to an enhancement in the relativistic particle content of the jet, the optical and $`\gamma `$-ray flux will vary in a linear fashion with no lag if the optical synchrotron and ECS emission occur in the same region of the jet. Ghisellini & Madau (1996) suggest that the dominant source of low energy photons for ECS scattering is broad-line-region re-processing of jet synchrotron emission — the “mirror model”. In this case, there could be linear correlation between synchrotron emission and ECS emission with a $`\gamma `$-ray lag of $``$2$`\times 10^3`$ seconds, if jet plasma entering a region of enhanced magnetic field resulting in more synchrotron emission, and the re-processing region was $``$$`10^3`$ light-seconds away. In this scenario, a change in the synchrotron emission spectrum would occur, and could be discerned with accurate photometry in multiple optical bands (e.g., B and R). Thus, good sampling of multiple energy ranges offers the opportunity to distinguish models for the continuum emission of blazar jets.
As we gain understanding of this emission, it potentially can then be used to study the origin of the jet. In addition to providing for discrimination between models, simultaneous multiwavelength observations also have the potential to determine of properties of the jet; e.g., Takahashi et al. (1996) infer the strength of the magnetic field in the jet of blazar Mrk 421 by examining the rate of Synchrotron energy losses with the ASCA satellite.
NASA’s GLAST mission, the next generation GeV gamma-ray telescope, will provide excellent GeV sensitivity. It is currently scheduled to be launched in 2005 and to operate for a minimum of 5 years. NASA’s URL for GLAST is http://glast.gsfc.nasa.gov. A simulated blazar GLAST light curve is shown in Figure 1 for PKS 1622-297 (which produced the largest point source flux observed by EGRET, Mattox et al. 1997). Compared to EGRET, a dramatic enhancement in the $`\gamma `$-ray light curve resolution due to the increased aperture of GLAST is apparent.
2. The Optical Monitoring of Blazars in the GLAST Era
If the $`\gamma `$-ray luminosity function of blazars is roughly similar to the radio luminosity function of the parent population of extragalactic flat-spectrum radio sources as often assumed (e.g., Stecker and Salamon 1996), GLAST will detect $``$5000 blazars. To observe just 20% of this population in the optical band just once per month will require 33 observations per day. For the expected magnitude distribution ($``$14 to $``$24 — deduced from the distribution of the V magnitudes given by NED for robust 3EG EGRET identifications, Mattox 1999a), a telescope of at least $``$1m aperture will be required for many of these sources. Assuming an average of 10 minutes to set up, expose, and read-out for each observation, 80% of all available time with a 1m class telescope will be required for just monthly monitoring.
Since $`\gamma `$-ray variability was observed with EGRET on sub-day time scales, (see Figure 1), more frequent monitoring is appropriate for bright sources which will produce sufficient GLAST counts to be more rapidly resolved. Table 1 specifies the required density of optical monitoring based on the number of sources which are expected to be time resolved by GLAST on a timescale shorter than $``$10 days. The average number of optical observations is 222 per day, and requires four 1m class telescopes at a variety of longitudes. The prospect of making an average of 222 optical observations per day for the duration of the GLAST mission (4$`\times 10^5`$ observations over 5 years) provides a strong incentive to consider the use of automated, ground-based telescopes.
This monitoring could be accomplished from space by a single facility with capabilities similar to the Hubble Space Telescope. However, the construction of four 1m class telescopes on the ground can be accomplished for a cost $``$2 orders of magnitude less than a space telescope with HST capabilities.
In addition to the $`\gamma `$-ray flux, Figure 1 shows simulated high-temporal-density optical observations. It was assumed that the $`\gamma `$-ray flux is varies linearly in relationship to the optical with a lag of 2000 seconds, consistent with the “mirror model” for $`\gamma `$-ray emission. The linear dependence and the lag are both clearly resolved. Thus, we can expect that dense optical sampling in conjunction with GLAST observations will produce a very good opportunity to test blazar models in detail. Because PKS 1622-297 is a 20th magnitude source, a network of $``$1m telescopes would be required to obtain this data. Although a large multiwavelength campaign was organized for 3C 279 in 1996 (Wehrle et al. 1998) which explicitly included over 20 optical observers, only one optical observation was obtained in the 3 days of maximum $`\gamma `$-ray flux. Without extensive preparation prior to the GLAST mission, more extensive optical coverage may not be available during the GLAST mission.
3. Automatic Telescopes
The automatic operation of optical telescopes began with a 50-inch telescope on Kitt Peak. This telescope was constructed with NASA funding as the Remotely Controlled Telescope (RCT). The initial intention was to develop techniques for controlling telescopes in space. It was soon apparent that this was not a useful approach to learning to control a space telescope — the dynamics were very different, as were the scales of the budgets (personal communication, Steve Maran, 1999). The RCT telescope focus then shifted to an attempt to demonstrate the operation of an automated telescope — something which the Whitford Committee suggested in 1964 as a means to enhance the productivity of small telescopes (Maran 1967). A decade of effort resulted in one astronomical paper (Hudson et al. 1971), and the realization that a human telescope operator was much more cost effective than telescope automation with the technology available in 1969 (personal communication, S. Maran, 1999).
During the 3 decades which have transpired since the RCT telescope experiment, remarkable advances in technology have occured. Modern technology now makes telescope automation straight forward. It is likely that vision of the Whitford Committee of substantial gains in productivity through the automation of telescopes may soon be realized. The most important technological advances have been: (1) the development of powerful, reliable, inexpensive, and compact computers; (2) the development of intelligent controllers for mechanical motions; (3) the development of charge coupled devices (CCDs), and (4) the accumulation of experience in the most effective ways to control and use fully automated and unattended observatories.
A telescope equipped with servo motors and rotation encoders on both axes, and driven by a computer with an accurate model of telescope flexure and pointing aberrations, can point anywhere on the sky with an open loop accuracy of $`<`$10”. Thus, source acquisition is straightforward, and easily automated. The CCD camera has liberated astronomers from the drudgery of the darkroom, and the anguish of the interpretation of non-linear photographic media. The CCD based camera produces digital data with linear response, and with a quantum efficiency as high as $``$90%, two orders of magnitude better than photographic film. Also, CCDs can provide simultaneous measures of sky brightness and comparison star brightness, which permits accurate differential photometry even with a partly cloudy sky.
A number of groups are operating automatic telescopes and some are developing plans for networks of automatic telescopes (hypertext links to those with web pages are maintained at http://gamma.bu.edu/atn/auto\_tel.html). At least three manufacturers have designed telescopes of aperture 60 cm or larger which are capable of automated operation, Torus Technologies and DFM in this country, and TTL in England.
A coordinated network of automated telescopes at diverse sites will facilitate optical monitoring of selected GLAST sources on sub-day timescales, a task which is not otherwise routinely feasible — although we have done this experimentally with a miniscule duty-cycle with the Whole Earth Blazar Telescope (WEBT) which is described in a separate paper in this volume (Mattox, 1999b). The ATN project has been undertaken to promote the development of a network of automated telescopes to support blazar monitoring. A web site has been established at http://gamma.bu.edu/atn/ to coordinate effort, and disseminate information.
4. Automatic Telescope Standards
Much remains to be done in the realm of software before automatic telescopes can execute a program such as the GLAST blazar monitoring. There are currently no standards in place for automated telescopes, to permit them to be used coherently.
Therefore, an international working group is in formation to work with the IAU Commission (number 9) on Instruments to develop standards for automatic telescopes. The web site is http://gamma.bu.edu/atn/standards/. These standards will expedite the creation and utilization of networks of telescopes for science and education.
The existence of a standard command set will form an interface between a telescope specific TCS (Telescope Control System) and a higher level Observatory Control System (OCS). This will promote the development of telescope-independent OCS software, which will provide for instant robotocization of additional new and refurbished telescopes which comply with the TCS standard.
The standards will also include appropriate protocol for Internet control of automatic telescopes. The existence of such a standard protocol will promote cooperative development and utilization of networks of robotic telescopes. A standard protocol for Internet control will facilitate cooperative utilization of these telescopes. This will provide for the utilization of more diverse facilities by all participants, increasing the range of projects possible, and the efficiency of telescope utilization.
5. Other Scientific Applications of the Networks of Automatic Telescopes
It is also expected that networks of automatic telescopes will be useful for studying other transient phenomena, e.g., binary stellar systems, gamma-ray bursts, quasar/galaxy lensing systems, microlensing events, and asteroseismology. A network which is sized to provide observing time for other areas of investigation will include more telescopes. Therefore, it will be more efficiently scheduled, and will provide better multi-longitude coverage for blazars, and gamma-ray burst follow-up. Therefore, it is of interest to collaborate with other astronomers who can use automatic optical telescopes.
Automatic telescopes can also serve as a very valuable facility for science education. A network of automatic telescopes is being proposed by the Hands-On Universe Project (http://hou.lbl.gov/) to provide abundant, high-quality, CCD data for education. The possibility of integrating blazar monitoring into this effort is being explored.
References
Bloom, S.D. & Marscher, A.P., 1993, CGRO Sym, AIP Conf. Proc. #280, p. 578
Dermer, C. D., Schlickeiser, R., & Mastichiadis, A., 1992, A&A 256, L27
Ghisellini, G., & Madau, P., 1996, MNRAS, 280, 67
Hartman 1999, et al., “3rd EGRET Catalog”, ApJS, in press
Hudson, K.I., Chiu, H.Y., Maran, S.P., Stuart, F.E., Vokac, P.R., 1971, ApJ, 165, 573
Maran, S.P., 1967, Science, 158, 867
Mattox, J.R., 1999a, “The Identification of 3EG EGRET Sources with Flat-Spectrum Radio Sources”, ApJS, in preparation
Mattox, J.R., 1999b, “The Whole Earth Blazar (WEB) Telescope”, this volume
Mattox, J.R., et al., 1997, ApJ, 476, 692
Sikora, M., Begelman, M. C., & Rees, M. J. 1994, ApJ 421, 153
Stecker, F.W. and Salamon, M.H., 1996, ApJ, 464, 600
Takahashi, T., et al., 1996, ApJ, 470, L89
Wehrle, A.E., et al., 1998, ApJ, 497, 178 |
no-problem/9911/hep-ph9911218.html | ar5iv | text | # Quantum Magnetic Collapse
## Acknowledgments
Three of us (S. S. M., A. P. M. and H. P. R.) would like to thank Professor Virasoro, the ICTP High Energy Group, IAEA and UNESCO for hospitality at the International Centre for Theoretical Physics. H.P.R. thanks A.E. Shabad for a discussion and the University of Helsinki for hospitality. The partial support by the Academy of Finland under Project No. 163394 is greatly acknowledged. |
no-problem/9911/astro-ph9911337.html | ar5iv | text | # Simultaneous intensive photometry and high resolution spectroscopy of 𝛿 Scuti stars IV. An improved picture of the pulsational behaviour of X Caeli. Based on observations collected at European Southern Observatory (Proposal 58.E-0278)
## 1 Introduction
X Cae has been recently the target of two campaigns: the first photometric in 1989 (Mantegazza & Poretti, 1992, Paper I) and the second both photometric and spectroscopic in 1992 (Mantegazza & Poretti 1996, Paper II).
From the photometric data obtained in these campaigns it has been possible to single out 14 terms due to at least 6 independent pulsation modes and to some harmonics and non–linear couplings among the strongest ones. With the additional information supplied by the high resolution spectroscopy, an attempt has been performed to identify at least some of these modes. However, due to the rather short baseline of spectroscopic data (4 consecutive nights), only very preliminary results have been obtained. The only really confident result was that the dominant mode (7.39 cd<sup>-1</sup> ) has $`m=1`$ and $`\mathrm{}=`$ 1 or 2. A successive improved analysis of these data (Mantegazza & Poretti, 1998) favoured the choice of $`\mathrm{}=1`$ and $`m=1`$, and suggested that the rotational axis inclination is larger than $`45^o`$. About the other modes the conclusion was that they are probably non-radial with $`\mathrm{}=2\pm 1`$, while high–degree sectorial modes, which seem to be rather common in fast rotating $`\delta `$ Scuti stars (Mantegazza, 1997; Kennelly et al., 1998), were not observed. The comparison between the photometric data of the two campaigns shows that while the dominant mode has a very stable amplitude, those of other modes show a certain degree of variability. The line profiles showed the presence of a narrow absorption core, which was attributed to the presence of a circumstellar shell, even if the possibility of a companion could not be completely ruled out.
In this paper we shall present and discuss the results of a new simultaneous photometric–spectroscopic campaign performed in 1996 in which the spectroscopic baseline has been doubled with respect to the previous one (8 consecutive nights). The longer spectroscopic baseline and the simultaneous collection of photometry and spectroscopy allowed us to perform the analysis through a synergic approach, first formalized in Bossi et al (1998): such an approach seems the most rewarding in the analysis of non-radial multi-mode pulsators.
## 2 Physical parameters
The stellar physical parameters were derived in Paper II using $`uvby\beta `$ photometric indices. However the stellar luminosity, as estimated from the Hipparcos parallax ($`M_{bol}=1.30\pm 0.15`$), does not agree with that derived from this approach ($`M_{bol}=0.73\pm 0.3`$), therefore we decided to reconsider the estimation of all these parameters.
Moon & Dworetsky (1985) calibration applied to the $`uvby\beta `$ data extracted from the catalogue by Hauck & Mermilliod (1990) supplies $`T_{eff}=6900^oK`$ and $`\mathrm{log}g=3.50`$ (this includes also the correction for metallicity effects, Dworetsky & Moon , 1986); at the same time the very recent calibration of Geneva photometric indices by Kunzli et al. (1997) applied to the data by Rufener (1988) gives $`T_{eff}=6860\pm 60^oK`$ and $`\mathrm{log}g=3.76\pm 0.14`$. We see that, while we can assume $`T_{eff}=6900\pm 100^oK`$, there is a certain disagreement on $`\mathrm{log}g`$. From the Hipparcos luminosity and the above derived photometric temperature we can estimate a radius $`R=3.43\pm 0.32R_{}`$, and we can also see that, according to the theoretical models by Shaller et al. (1992), these two quantities fit a stellar mass of $`M=2.0\pm 0.1`$. Hence we derive $`\mathrm{log}g=3.66\pm 0.09`$, which is just in between the two photometric estimates, and therefore in the following we shall adopt this value. With these physical parameters we can estimate that the frequency of the fundamental radial mode should be $`\nu _F=6.7\pm 0.7`$cd<sup>-1</sup> .
## 3 Observations and data processing
### 3.1 Photometry
Photometry of the star X Caeli was performed with the ESO 50 cm telescope at la Silla, Chile, for 14 consecutive nights (November 15–28, 1996), covering 100 hours of useful observation time. HR 1766 and HR 1651 where chosen as comparison stars as their stability was already confirmed by the previous campaigns (Mantegazza and Poretti, 1996).
Although air transparency in la Silla was rather stable during the observing nights, the data were processed using the instantaneous extinction coefficients technique developed by Poretti and Zerbi (1993) in order to deal with the residual intra-night transparency variations. The reduced photometric data consist of 916 and 889 differential measurements in Johnson’s $`V`$ and $`B`$ bands respectively.
The analysis of the magnitude differences between the two comparison stars confirmed their stability and showed that the photometric quality of the nights was acceptable. The standard deviation of the $`B`$ and $`V`$ differential time series is 3.67 and 3.57 mmag respectively and the frequency analysis of these series by means of the least square technique (Vaniĉek, 1971) showed a flat spectrum: the highest peaks show amplitudes not exceeding 1.4 % and 2.2% (0.32 and 0.37 mmag semi–amplitude in the corresponding sinusoidal signal) respectively.
### 3.2 Spectroscopy
The spectroscopic observations have been performed in the Remote Control Mode at the La Silla Observatory (ESO) using the Coudè Echelle Spectrograph (CES) attached to the Coudè Auxiliary Telescope (CAT) during 8 consecutive nights (November 21-28, 1996). The spectrograms, which cover the 4483–4533 Å region, were acquired by means of the ESO CCD #38. The resulting resolution was of 0.018 Å/pixel and the effective resolving power, as measured from spectrograms, was 55000.
225 spectrograms of X Cae with exposure times of 15 minutes have been gathered in total. They cover about 61 hours of observing time. The data reduction was performed using the MIDAS package developed at ESO. The spectrograms have been normalized by means of internal quartz lamp flat fields, and calibrated into wavelengths by means of a Thorium lamp. The normalization at the continuum has been performed by fitting a third degree polynomial to the continuum windows in the region containing the three adjacent spectral lines which already have been used in the previous study (i.e. the normalization was limited to the 4497–4517 Å region ). Finally the spectra have been shifted and rebinned in order to remove the observer’s motion.
The mean standard deviation of the pixel values in the continuum regions allows the estimation of the signal–to–noise ratio of the spectrograms. The resulting average value at the continuum level is about 230.
Fig. 1 shows in the upper panel the average spectrum. We can notice that the spectral lines have at the center the narrow absorption core already present in the spectrograms of the previous campaign. The lower panel shows the standard deviation of the individual spectrograms about the average one. The figure clearly shows the variability of the line profiles; we can also notice the close similarity with Fig. 4 of paper II.
The analysis of the spectroscopic variations has been made on the two lines TiII 4501 and FeII 4508, which are the best defined in our spectra and free of strong blends. The two equilibrium equivalent widths are 15.8 and 11.9 km s<sup>-1</sup> respectively. In the following, when averages of quantities derived from the two lines are computed, a weight proportional to their equivalent widths has been assigned.
A non–linear least squares fit of a rotationally broadened gaussian profile to the average profiles of the two lines, allowed us to estimate the projected rotational velocity and the intrinsic widths: $`v\mathrm{sin}i=69.0\pm 1.0`$ km s<sup>-1</sup> and $`W_{4501}=11.9\pm 1.0`$ km s<sup>-1</sup> and $`W_{4508}=15.4\pm 1.0`$ km s<sup>-1</sup>. This $`v\mathrm{sin}i`$ estimate is in excellent agreement with the value of $`70\pm 2`$ km s<sup>-1</sup> derived in Paper II.
## 4 Analysis of photometric data
The frequency analysis of the photometric data was performed by means of the least–squares technique (Vaniĉek 1971, see also Mantegazza 1997). The data in the $`V`$ and the $`B`$ band were first analyzed separately and yielded a solution with 16 and 15 terms respectively.
In order to improve the signal–to–noise ratio, $`V`$ and $`B`$ data were put together by shifting and rescaling the $`V`$ time series following a procedure similar to that described by Breger et al. (1998). The resulting file with 1805 data points has been frequency analyzed in the way previously described. 17 terms were singled out and are listed in the last column of table 1.
The modes are listed in order of detection and the $`S/N`$ ratios were computed using the Period98 package (Sperl, 1998). According to Breger et al. (1995) peaks with $`S/N>4`$ have a high probability of being intrinsic to the star. We see that this criterion is met by 15 out of 17 detected periods. For some of them there is the uncertainty of 1 cd<sup>-1</sup> in particular for $`f_{11}`$, $`f_{14}`$, $`f_{17}`$. The solution reported here is the combination of values which give the least–squares best fit. In the same table we report for comparison the corresponding frequency values detected by our previous campaigns. The comparison between the frequencies in common in different seasons tells us that their values agrees within 0.005 cd<sup>-1</sup> (in the worst cases their difference is of about 0.01 cd<sup>-1</sup>). For $`f_{11}`$ and $`f_{17}`$ this comparison confirms the values here assumed, while for $`f_{14}`$ the closest connection is with an 1 cd<sup>-1</sup> alias of 7.17 cd<sup>-1</sup> term. However the distance between the two terms (0.06 cd<sup>-1</sup>) is an order of magnitude larger than the above reported average separation between all the other terms, so this connection is very doubtful. All the other terms are in agreement, with the exception of $`f_5`$, whose old value was 7.667 cd<sup>-1</sup>. Since the amplitude in the more recent data is larger and the analysis of the spectroscopic data (see below) favours the choice of the 6.66 cd<sup>-1</sup>, in the following we shall adopt this as the true frequency.
About the two weakest terms, which are only marginally significant, we have to note that they have been independently detected from the analysis of line profile variations (see below), and furthermore the 11.26 cd<sup>-1</sup> term was independently detected in the 1992 photometric data: this considerably strengthens their reliability. Regarding the 9.003 cd<sup>-1</sup> term, its spectroscopic counterpart was detected at 10.01 cd<sup>-1</sup> . Since the diagram describing the changing of the phase of this perturbation across the line profiles looks cleaner by considering 10.01 instead of 9.003 cd<sup>-1</sup> , we adopted the former value as the correct one.
Fig. 2 shows in the upper part of each panel the differential V measurements with respect to the star’s mean brightness, and the best fitting 17–sinusoid model: no significant systematic differences between fitting curve and data points are apparent.
The availability of three V datasets allows us to compare the evolution in time of the mode amplitudes. In Table 2 we report the amplitudes in the three seasons of the 9 strongest modes. It is apparent that $`f_3`$, $`f_5`$, $`f_6`$ and $`f_7`$ have shown conspicuous variations, while $`f_1`$, $`f_2`$ and $`f_4`$ seem substantially stable.
## 5 Analysis of line profile variations
### 5.1 Radial velocities
Radial velocities were derived by the computation of the line barycenters (this is equivalent to evaluate the first moment of the line profile). They have been derived for both lines and then a weighted mean has been performed. The resulting time series has been Fourier analyzed by means of the same least–squares technique adopted for the light curve analysis. 13 frequency terms have been detected, all corresponding to analogous terms found from the analysis of photometric data or to their 1 cd<sup>-1</sup> aliases and one more term with a frequency close to 1 cd<sup>-1</sup> , which was also detected in the subsequent analysis of the line profile variations (see below). Table 3 reports in successive columns, arranged in order of increasing frequency: the adopted frequency, the frequency detected in the radial velocity power spectrum, the amplitude of the term, the order of detection.
The observed radial velocities are shown in the lower part of each panel of Fig. 2 together their best fitting curve. In this case too no systematic differences between the data and the computed curves are apparent.
### 5.2 Mode detection
The search for periodicities in the line profile variations has been performed by applying the least squares power spectrum technique generalized to study line profile variations. A detailed description of the technique can be found in Mantegazza and Poretti (1999). Fig. 3 reports in the top panels the power spectra of the line profile variations in the TiII 4501 line (left panel) and in the FeII 4508 line (right panel) and in the bottom ones the residual power spectra, obtained giving as “known constituents” (defined in Mantegazza & Poretti, 1999) all the detected terms. Possibly some more modes are still present at frequencies below 15 cd<sup>-1</sup> , but at present they are not detectable without ambiguity.
Table 4 reports the detected terms in order of detection for the TiII 4501 line (which has the stronger $`S/N`$). In successive columns we report: the derived frequency in the TiII 4501 line, that in the FeII 4508 line, the frequency of the corresponding photometric term, the finally adopted frequency for the mode, and the rms amplitude along the line profile of the spectroscopic term. This amplitude has been derived from a weighted average between the amplitudes of the two lines normalized to the intensities of the TiII line (the rms amplitudes in the two lines are about in the ratio 1.8:1) and it is measured in per mille of the continuum intensity.
We see that the three strongest terms are the same found from photometry, namely 7.39, 13.98, and 6.04 cd<sup>-1</sup>. There is also a low frequency term (2.0 c/d) which is an alias of the reciprocal of the sidereal day and probably it is due to an instrumental effect (a similar term has been found and discussed by Mantegazza and Poretti, 1999, analysing the spectrograms of BB Phe). The analysis of the line profile variations of the FeII 4508 line shows also a low frequency term at 0.56 cd<sup>-1</sup> , which however has not been detected in the other higher $`S/N`$ line. It is therefore not very likely that this signal can be attributed to the star.
We can see that all the spectroscopically detected terms have been also photometrically detected. This tells us that in X Cae we don’t have to expect the presence of high–degree modes as found in other $`\delta `$ Scuti stars (see e.g. the cases of BV Cir, Mantegazza et al. in preparation, HD 101158 (Mantegazza 1997) and $`\tau `$ Peg, Kennelly et al. 1998).
An indepenent check of these results has been performed using the CLEAN algorithm generalized to the analysis of the line profile variations. Examples of the use of this technique to analyse line profile variations in $`\delta `$ Scuti stars can be found in De Mey et al.(1998) and Bossi et al.(1998).
The original spectra have been computed up to 30 c/d, but for clarity the figures are truncated to 20 c/d, since there is no significant signal above this value, and have been obtained with a gain of 0.3 and 500 iterations. Figure 4 shows the spectra of the two lines with the frequency identification of the highest peaks which correspond to those detected by the least–squares analysis or to their 1 cd<sup>-1</sup> aliases. Finally Fig. 5 shows a comparison between the relative amplitudes of the modes detected from photometry (top panel), radial velocity variations (middle panel) and line profile variations (bottom panel).
### 5.3 Amplitude and phase diagrams
After the detection of all the relevant modes affecting the line profiles, it is possible to obtain their amplitudes and phases across the line profile ($`A_j(\lambda )`$, $`\varphi _j(\lambda )`$ $`j=1,\mathrm{\#}modes`$ with their formal errors $`\delta A_j(\lambda )`$ and $`\delta \varphi _j(\lambda )`$) by means of a simultaneous least–squares fit ( Mantegazza & Poretti, 1999). We also get the estimate of the unperturbed average line profile (the zeropoints of each pixel time series fit: $`A_0(\lambda )`$).
Figure 6 shows the behaviour of amplitude and phase of each mode across the profile of the TiII 4501 line. To obtain this figure we used all the terms detected in the spectra with the addition of the 7.47 c/d one. This is among those photometrically strongest, and it has been also detected in the radial velocity data. Probably its spectroscopic detection is hampered by its closeness to the dominant mode, so that it is drowned in its peak because the ratio between the amplitudes of the two modes is more unfavourable than in the light and radial velocity curves.
Very similar results have been obtained from the fit of the FeII 4508 line. These diagrams, in particular the phase ones, give already some clues about some possible mode identifications. We note that a mixture of different pulsation modes is present. The phase curve of the 5.88 cd<sup>-1</sup> term is typical of a low degree retrograde mode, while those of 7.47, 13.52 and 13.99 cd<sup>-1</sup> indicate axisymmetric modes and that of 7.39 cd<sup>-1</sup> indicates a low–degree prograde mode.
### 5.4 Model fit to line profile variations
It is possible to try to estimate the $`\mathrm{},m`$ parameters of each detected mode by fitting the variations it induces on the line profile shape (Bossi et al., 1998; Mantegazza & Poretti, 1999).
In order to do this it is necessary to separate the contributions of the different modes. The perturbations $`\mathrm{\Delta }p_j(\lambda ,t)`$ induced by mode $`j`$ on the line profile can be approximated as:
$$\mathrm{\Delta }p_j(\lambda ,t)=A_j(\lambda )\mathrm{cos}(2\pi \nu _jt+\varphi _j(\lambda ))$$
$`(1)`$
with the corresponding error $`\delta \mathrm{\Delta }p_j(\lambda ,t)`$ derived from error propagation by $`\delta A_j(\lambda )`$ and $`\delta \varphi _j(\lambda )`$. These last quantities as well as $`A_j(\lambda )`$ and $`\varphi _j(\lambda )`$ have been derived in the previous subsection. We can try to fit these functions ($`\mathrm{\Delta }p_j(\lambda ,t)`$) with perturbations computed with a model of a non-radial pulsating star viewed at a certain inclination $`i`$. So for each plausible choice of $`\mathrm{},m,i`$ we can build a discriminant
$$\sigma _p(\mathrm{},m,i)=\underset{\lambda }{}\underset{t}{}\frac{(\mathrm{\Delta }p_j(\lambda ,t)\mathrm{\Delta }p_c(\lambda ,t,\mathrm{},m,i))^2}{\delta \mathrm{\Delta }p_j(\lambda ,t)^2}$$
$`(2)`$
where $`\mathrm{\Delta }p_c(\lambda ,t,\mathrm{},m,i)`$ are the computed profile variations which best fit the observed ones.
If moreover we have simultaneous photometric observations, we can obtain, by simultaneously fitting all the terms detected in the light curve, amplitude and phase with respective errors for the $`j`$ mode. Therefore we can calculate its light variations and relative errors ($`l_j(t)`$ and $`\delta ł_j(t)`$) and compare them with those predicted by the best fitting models and obtain the discriminant:
$$\sigma _l(\mathrm{},m,i)=\underset{t}{}(l_j(t)l_c(t,\mathrm{},m,i))^2/\delta ł_j(t)^2$$
$`(3)`$
A global discriminant is then defined as:
$$\sigma _T(\mathrm{},m,i)=\sigma _p(\mathrm{},m,i)+\sigma _l(\mathrm{},m,i)$$
$`(4)`$
This is the function minimized with a non–linear least–squares fit for each detected $`j`$ mode and for any choice of $`\mathrm{},m,i`$.
In summary, since by means of eq. 1 we estimated the variations induced on the line profile by each individual mode, we can compare them with the variations computed from a model of a pulsating star for each possible non–radial mode. Such a comparison is done computing the weighted sum of the squares of the differences between the two quantities: we called such a sum the discriminant. This approach is similar to the one developed to identify the pulsation modes with the “moment method” (Balona, 1986, 1987; Aerts, 1996).
In addition, since the model allows us to compute the light variations related with each mode, we can then adjust its free parameters to best fit at the same time observed line profile and light variations. This can be done with a non–linear least–squares algorithm that minimizes the global discriminant described in eq. 4.
The process of deriving the line profile and light variations for each mode (eq.1) carries on an error estimation together with the results. The squared reciprocal of such estimated errors are the weights to be used in the discriminants computation, with the additional advantage to give to the dicriminants themselves the adimensional character needed to add them in eq. 4.
The model used to compute the synthetic line profile variations ($`\mathrm{\Delta }p_c(\lambda ,t,\mathrm{},m,i)`$) and light variations ($`l_c(t,\mathrm{},m,i)`$) is the one described by Balona (1987). For each assigned $`\mathrm{},m,i`$ the computed profile variations can be modeled according to the amplitude and phase of vertical ($`v_r`$) and horizontal ($`v_r`$) velocities and flux variations. For $`\delta `$ Scuti stars usually $`v_hv_r`$ and therefore in order to not introduce into the model too many free parameters, we keep the usual theoretical relation (e.g. Heynderickx et al. 1994) $`v_h=74.4Q^2v_r`$ ($`Q`$ pulsation constant) and $`\psi _h=\psi _r`$.
The observed light variations constrain strongly the computed flux variations, so it is very useful to have simultaneous spectroscopic and photometric data, otherwise in order to get meaningful physical results it is better to fit a simplified model which considers velocity variations only, neglecting flux variations.
We applied the method to the data by computing eq. 1 for each identified mode. In order to do this ten equi-spaced phases which cover the corresponding pulsation period were used.
We have explored for each detected term all the possible modes with $`0\mathrm{}4`$, and considered inclinations between 20 and 90 degrees. The independent analysis of the two spectral lines gives concordant indications about the better fitting modes, and moreover the estimated physical parameters of these modes are similar. Fig. 7 shows the discriminants of the best fitting modes for all the detected frequency terms as a function of the inclination of the rotational axis. In order to allow a comparison between the different panels the discriminant scale is the same for all of them, but it has been shifted in order that the best fitting mode discriminant in each panel was close to the lower border. The discriminants are the weighted means of those derived from the fits of line profile variations in the two spectral lines. The figure contains also the panels for the terms at 12.09, 13.43, 13.52 and 14.79 cd<sup>-1</sup> which, as already remarked, could not be independent pulsation modes.
In order to evaluate the inclination, a total discriminant has been computed by adding for each inclination the minimum global discriminants of all the detected modes. As a matter of fact, the 7.39 cd<sup>-1</sup> mode dominates this discriminant and all the other modes supply only perturbations. Their presence affects appreciably this discriminant only for $`i`$ close to $`90^o`$. The result is reported in Fig. 8, where we see that there is a rather broad minimum centred at about $`70^o`$.
Assuming this inclination, the equatorial rotational velocity results of about 73 km s<sup>-1</sup>, and therefore the rotational period is about 2.4 d and the ratio between rotational and pulsation frequencies is lower than about 0.06 for all the modes.
Having established a reasonable value for the inclination angle, we can look at the discriminants of the individual modes in order to decide which are the more plausible identifications for this inclination. The proposed $`\mathrm{},m`$ identifications for each detected mode are reported in the $`6^{th}`$ column of Table 7. For some modes these identifications are rather uncertain or there are different solutions which fit almost equivalently well the line profile variations. The more uncertain results are given between brackets.
In Fig. 9 we show the variations of the Ti 4501 line profile due to the 7.39 cd<sup>-1</sup> term (solid line) phased on a complete pulsation cycle and the corresponding best fitting variations generated by a model of a $`\mathrm{}=1,m=1`$ mode. This model explains the 93% of the line profile variance, fits the $`B`$ light variations with a standard deviation of 0.4 mmag and the radial velocity variations with an accuracy of $`\pm `$0.2 km s<sup>-1</sup>.
In Fig. 10 we show the observed amplitude (upper panel) and phase variations (bottom panel) of the 7.39 cd<sup>-1</sup> term across the TiII line profile with the error bars derived from the least–squares fit, together with the corresponding curves derived from the best fitting model (solid lines). If we take into account all the approximations present in our model of pulsating star, we have to be more than satisfied by the quality of these results. Furthermore the figure makes it clear that the bump in the amplitude curve at the left of the line profile, that was already present in the 1992 data (see Fig. 4 of paper II), is a signature of the pulsation mode.
### 5.5 Color phase shifts
In order to add weight to the mode identification a further independent tessera can be included in the mosaic by considering the phase differences between $`V`$ and $`B`$ light variations. Figure 11 shows this quantity versus the respective observed frequency for the 6 strongest photometric modes (namely, in order of frequency, 6.04, 6.66, 7.49, 7.47, 13.52 and 13.58 cd<sup>-1</sup>). From the line profile variation analysis we have derived that apart from the 7.47 cd<sup>-1</sup> mode, all the others have as preferred choice $`\mathrm{}=1`$ (see Table 5). We see from the diagram that all of them fall in a strip of $`\pm 3^o`$ about the zero shift (dotted lines), just as expected. On the contary the 7.47 cd<sup>-1</sup> mode, for which the best option is $`\mathrm{}=0`$, has the highest fase difference and falls just outside this strip, again as expected.
## 6 Discussion
### 6.1 Physical quantities related to pulsation modes
In the Table 5 we present the physical quantities obtained by fitting line profile and light variations for the better identified modes. In the successive columns we give: observed frequency ($`\nu `$), $`\mathrm{},m`$, frequency in the stellar reference frame ($`\nu _0`$), pulsation constant ($`Q`$), amplitude of local radial velocity ($`v_r`$), percentual local flux variation ($`\mathrm{\Delta }F/F`$) at the line wavelength, phase shift between flux and radial variation ($`\psi `$), percentual local radius variation ($`\mathrm{\Delta }R/R`$), and the ratio between relative flux and radius variation ($`f`$). The reported amplitudes are sinusoidal amplitudes, so that the full amplitudes are twice these values.
The mode with the best defined parameters is obviously 7.39 cd<sup>-1</sup>. We believe that these parameters are rather reliable: the formal uncertainties that we got from the simultaneous fit of its line profile and light variations are $`\pm 1`$ km s<sup>-1</sup> on $`v_r`$, $`\pm 1.7\%`$ on $`\mathrm{\Delta }F/F`$, $`\pm 7^o`$ on $`\mathrm{\Psi }`$, $`\pm 0.09\%`$ on $`\mathrm{\Delta }R/R`$ and $`\pm 1.0`$ on $`f`$. It is interesting to compare the $`f=9`$ and $`\mathrm{\Psi }=114^o`$ values with those predicted by theoretical models for a star with physical parameters similar to X Caeli. For low–degree $`\mathrm{}`$ modes of low radial order we have, for a mixing length parameter $`\alpha =0.5`$, $`f11`$ and $`\mathrm{\Psi }115^o`$ and, for $`\alpha =1.0`$, $`f8`$ and $`\mathrm{\Psi }150^o`$ (theses data were kindly supplied us by W. Dziembowsky, and the $`f`$ values have beeen rescaled at the line wavelength). The agreement with the theory is rather good, and it seems that our data favour a mixing length parameter of about 0.5.
The parameters of the other modes are more uncertain, given the small amplitudes of the perturbations produced by these modes on the line profiles.
It is worth noticing that the two lowest frequency modes (5.89, 6.04 cd<sup>-1</sup> ) are retrograde, as probably is also the case for the 6.66 cd<sup>-1</sup> one . Up to now there was not a large evidence of the presence of retrograde modes in $`\delta `$ Scuti stars: Kennelly et al. (1998) suspected that two high–degree modes in $`\tau `$ Peg could be retrograde, and Mantegazza and Poretti (1999) showed the presence of a rather certain low–degree retrograde mode in BB Phe.
We observe too, that the 6.66 and 7.39 cd<sup>-1</sup> terms have, within the approximations, due to the uncertaintinies on $`i`$ and on the Ledoux factor $`C`$, the same frequency in the stellar reference frame. Hence, having $`\mathrm{}=1`$, and $`m=1`$ and $`m=1`$ respectively, they could be the rotational splitting of the same mode.
### 6.2 Fundamental stellar parameters as derived from pulsation
The identification of the 7.47 cd<sup>-1</sup> mode allows us to estimate some physical parameters of the star. As we have seen, this mode has been clearly detected in the light and in the radial velocity curves, but its detection is more uncertain in the line profile variations. Moreover its phase curve shows a clear rotation of $`180^o`$ in the line center, and the amplitude of the line profile variations are larger in the wings. These facts suggest that we are in presence of a mode that shifts the line rather than affecting its shape. Since in addition the mode is preferably seen in the integrated quantities and the inclination of the rotational axis is rather high, this mode is probably a radial mode. The same suggestion is supplied by its colour phase shift, even if on this basis alone we cannot rule out the $`\mathrm{}=1`$ possibility (Fig. 11).
If this is the correct identification then its $`Q`$ value as derived from the assumed physical parameters ($`0.029\pm 0.004`$ d) would put it midway between fundamental ($`Q=0.033\pm 0.001`$ d) and first overtone mode ($`Q=0.025\pm 0.001`$ d). In the first hypothesis the 7.39 cd<sup>-1</sup> term should be a $`\mathrm{}=1`$ $`f`$ mode, while according to the second it should be a $`\mathrm{}=1`$ $`p_1`$ mode (see the theoretical models by Fitch, 1981). In Table 6 we report the stellar fundamental parameters and their uncertainties as derived from three different hypotheses: a) adopting the Hipparcos parallax, b) 7.47 cd<sup>-1</sup> is the radial fundamental mode, c) 7.47 cd<sup>-1</sup> is the $`1^{st}`$ overtone radial mode. All the hypotheses assume the effective temperature derived from multicolor photometry, and the validity of the theoretical models by Shaller et al. (1992). In this table $`\nu _F`$ is the fundamental radial mode frequency. The uncertainty on the 7.47 cd<sup>-1</sup> mode ($`\pm `$0.02 cd<sup>-1</sup>) is larger than on the other detected modes (see section 4) because of its proximity to the 7.39 cd<sup>-1</sup> mode.
The bolometric magnitude obtained with the fundamental mode assumption is $`1\sigma `$ higher than derived from Hipparcos parallax, while that obtained with the first overtone assumption is $`1.4\sigma `$ lower. On these bases it is difficult to get a firm conclusion even if it is slightly favoured the fundamental mode option for the 7.47 cd<sup>-1</sup> term. This without considering the possible presence of a circumstellar envelope that, introducing some extinction, could shift the result toward the first overtone mode hypothesis. In any case it can be appreciated how the use of the astroseismological information leads to a shrinking of the error bars.
## 7 Conclusions
The analysis of the simultaneous photometric and spectroscopic observations of X Caeli has supplied the following results:
* 17 terms have been detected in the photometric data, most of which were already detected in the 1989 and 1992 campaigns. The comparison of the amplitudes of the strongest terms has shown that while the dominant mode (7.39 cd<sup>-1</sup> ) looks rather stable other terms have large amplitude variations.
* The analysis of the radial velocity curve detected 13 terms, and again 13 (but not exactly the same) were detected from the analyis of the line profile variations. All these terms are among those detected from photometry. Therefore no high–degree modes have been detected, even if the line profiles are sufficiently broadened by rotation to allow it.
* By means of simultaneous least–squares fits of the line profile and light variations induced by each mode we were able to supply $`\mathrm{},m`$ identifications for many modes, and to estimate the inclination of the rotational axis (about $`70^o`$).
* The two shortest period modes are retrograde ones, while the 6.66 and 7.39 cd<sup>-1</sup> modes are probably the result of the rotational splitting of a $`\mathrm{}=1`$ mode.
* The physical parameters of the modes with the most reliable identifications have been given (Table 5). Rather accurate values have been obtained for the dominant mode, which has $`\mathrm{}=1`$, $`m=1`$. When compared with those predicted by theoretical models they allow the estimation of the mixing length parameter which results about 0.5.
* The 7.47 cd<sup>-1</sup> mode is probably radial. It remains uncertain if it is the fundamental or the first overtone. The stellar physical parameters derived in both hypotheses are given and compared and discussed with those obtained using the Hipparcos parallax.
Table 7 summarizes all the detected modes in the present campaign with the different techniques and in order of increasing frequency. In the successive columns the $`V`$, $`BV`$, RV amplitudes and the rms one across the line profile are given. The last three columns give furthermore the most probable $`\mathrm{},m`$ identification, the eventual alternative frequency due to aliasing, and the possible relationships with the observed frequencies of other modes.
The research presented in this paper has supplied a lot of useful and interesting results concerning the pulsational behaviour and the physical characteristics of the $`\delta `$ Scuti star X Caeli. However a number of questions remains open and needs further investigations. First of all some of the detected terms still suffer of the 1 cd<sup>-1</sup> uncertainty which could probably be removed through a multisite photometric campaign.
Furthermore a better resolution both in spectroscopy and photometry would be useful to have a better plot of some close terms as the 7.39 and 7.47 cd<sup>-1</sup> . This would allow us to get a more accurate evaluation of the contribution of each mode to the observed variations and consequently would allow us to perform a more reliable mode identification. For instance the confirmation of the radial nature of the 7.47 cd<sup>-1</sup> mode would be of particular relevance in order to nail the fundamental stellar physical parameters.
Longer baselines would moreover allow us to detect more pulsation modes whose presence can be guessed from the present data, and to improve the accuracy of the phase shifts in different photometric bands and hence to allow an independent check of the $`\mathrm{}`$ identifications supplied by spectroscopy.
Given the brigthness of the object the use of a larger telescope in photometry would not significantly improve the results. However, since some modes produce variations of only a few thousands of the continuum intensity (see Fig. 6), a larger telescope in spectroscopy would allow us to obtain more accurate estimates of the line profile variations through higher $`S/N`$ data.
Nonetheless we emphasize that in order to improve the mode detection an effort should be made on the theoretical side too. Indeed, as Fig. 10 clearly shows, the computed amplitudes across the line profile of the best fitting mode ($`1,1`$) for the dominant term (7.39 cd<sup>-1</sup>), for which the spectroscopic $`S/N`$ is amply adequate, have still considerable systematic deviations from the observed amplitude variations.
###### Acknowledgements.
We are grateful to Dr. W. Dziembowsky for supplying us the theoretical data, to Dr. M. Bossi for some enlightening discussions, to Dr. E. Poretti for a critical reading of the manuscript and to Dr. V. Fraschini for the help in the data reduction. Finally we wish to thank the referee, Dr. D. Sullivan, for his useful comments and suggestions. |
no-problem/9911/cond-mat9911170.html | ar5iv | text | # REFERENCES
Comment on “Density Matrix Renormalization Group Study of the Haldane Phase in Random One-Dimensional Antiferromagnets” In a recent Letter, Hida presented numerical results indicating that the Haldane phase of the Heisenberg antiferromagnetic (HAF) spin-1 chain is stable against bond randomness, for box distributions of the bond strength, even when the box distribution stretches to zero bond strength. The author thus concluded that the Haldane phase is stable against bond randomness for any distribution of the bond strength, no matter how broad. This conclusion contradicts our earlier prediction that a sufficiently broad bond distribution drives the HAF spin-1 chain through a critical point into the Random Singlet phase. In this Comment, we (i) point out that the randomness distributions studied in Ref. do not represent the broadest possible distributions, and therefore these numerical results do not lead to the conclusion that the Haldane phase is stable against any randomness; and (ii) provide a semiquantitative estimate of the critical randomness beyond which the Haldane phase yields to the Random Singlet phase, in a specific class of random distribution functions for the bond strength.
(i) The box distribution function for bond strength $`J`$ used in Ref. takes the form $`P_{box}(J)=1/W`$ for $`1W/2J1+W/2`$, and $`P_{box}(J)=0`$ otherwise. There is no randomness for $`W=0`$, and the randomness increases with $`W`$. The maximum randomness for this class of distribution functions is reached at $`W=2`$, as further increasing $`W`$ would introduce ferromagnetic bonds to the system and change the physics completely. On the other hand, the $`W=2`$ distribution does not represent the broadest possible distribution among all possible bond distributions. In fact, it is a member of the class of power-law distributions: $`P_{pow}(J)=\alpha J^{1+\alpha }`$ for $`0J1`$, and $`P_{pow}(J)=0`$ otherwise. The case $`W=2`$ of $`P_{box}`$ corresponds to $`\alpha =1`$ in $`P_{pow}`$. The power-law distributions are of particular interest and importance in studies of random spin chains, because under real space renormalization group (RSRG) transformations, all generic distribution functions will flow to power-law distributions in a HAF spin-1/2 chain. The broadness of a power-law distribution is parameterized by the single positive exponent $`\alpha >0`$. The smaller $`\alpha `$ is, the broader the distribution is on a logarithmic scale (the width is $`1/\alpha `$). Clearly, for any $`0<\alpha <1`$, the distribution function $`P_{pow}(J)`$ is broader than those studied in Ref. . Thus the numerical results of Ref. do not necessarily lead to the conclusion that the Haldane phase in HAF spin-1 chain is stable against any bond randomness, no matter how broad the distribution.
Hida also argues that the persistence of the string order is a consequence of the absence of translation symmetry in the effective spin-1/2 description of the random HAF spin-1 chain, similar to what happens in the random dimerized spin-1/2 chain. We point out that the Random Singlet phase in the effective spin-1/2 model does not have translational symmetry, and that broken translational symmetry and string order are not synonymous.
(ii) While a naive extension of the original form of the RSRG cannot be applied directly to the random HAF spin-1 chain unless the initial distribution function is sufficiently broad, we have developed a RG scheme that is suitable for the random HAF spin-1 and other random spin chains. The key ingredient of this scheme is to project out the highest energy state(s) in the strongest bond and keep all lower energy states at each RG step. Details of this scheme will be presented elsewhere. Using this scheme to simulate the HAF spin-1 chain with power law distributions, we have found two phases separated by a critical point $`\alpha _c0.67`$. For $`\alpha >\alpha _c`$ (which includes the flat distributions of Ref. ) the system is in the Haldane phase characterized by the string order, while for $`\alpha <\alpha _c`$ it is in the Random Singlet phase. While there are approximations involved and our estimate of $`\alpha _c`$ may not be very accurate numerically, it clearly suggests that one needs bond distributions that are broader than those studied to destabilize the Haldane phase.
K.Y. was supported by NSF DMR-9971541 and the Sloan Foundation.
Kun Yang
NHMFL and Physics Department
Florida State University
Tallahassee, Florida 32310
R. A. Hyman
Department of Physics
DePaul University
Chicago, IL 60614-3504 |
no-problem/9911/cs9911006.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Question answering task has been done in TREC8 using English documents . Here, we examine question answering task in Japanese sentences<sup>1</sup><sup>1</sup>1With respect to Japanese sentences, domain-dependent work such as that on dialogue systems and help systems has been done , but little work has been done on detecting the answer from natural-language databases as in question answering task. However, much has been done on English sentences, such as work on detecting sentences in written answers to work on detecting answers themselves . <sup>,</sup><sup>2</sup><sup>2</sup>2This paper outlines one part of a question answering system that we have been developing for a long time .. Our approach is to use syntactic information.
## 2 Question Answering System
### 2.1 Outline
1. The system detects keywords in question sentences, and then detects sentences in which the sum of the keywords’ IDF values is high<sup>3</sup><sup>3</sup>3In this paper, the system detects one sentence by one sentence. However, it would be better to detect a series of sentences and detect the answer from a series of sentences. If a series of sentences is used to detect the answer, context information can be used..
2. The question sentences and the detected sentences are parsed by the Japanese syntactic analyzer . (This allows us to obtain the dependency structures.)
3. The answer is selected by matching a question sentence and the detected sentences using syntactic information. How this is done is described in the next section.
### 2.2 Matching a Question and Detected Sentences Using Syntactic Information
We use the syntactic information when matching a question sentence and the detected sentences.
The score of a detected sentence $`s`$ is as follows.
$$Score(s)=B1(s)+\alpha B2(s)\beta DNUM(s)$$
(1)
$$B1(s)=\underset{\text{all bunsetsus b in the question sentence}}{}BNST1(b)$$
(2)
$`B2(s)`$ $`=`$ $`{\displaystyle \underset{\text{all pairs of two bunsetsus (b1, b2) in the question sentence, where b1 depends on b2 (i.e. b2 is the head of b1.)}}{}}BNST2(b1,b2)`$ (3)
Each of the bunsetsus in the question sentence can be paired with one of the bunsetsus in the sentence $`s`$ in order to maximize the value of $`Score(s)`$. (A bunsetsu in Japanese corresponds to a phrasal unit such as a noun phrase or a prepositional phrase in English.) $`BNST1(b)`$ is the similarity between the bunsetsu $`b`$ in the question sentence and the bunsetsu in the detected sentence $`s`$ paired with the bunsetsu $`b`$. $`BNST2(b)`$ is the similarity between the set of the two bunsetsus, ($`b1`$, $`b2`$), and the set of the two bunsetsus in the detected sentence $`s`$ paired with $`b1`$ and $`b2`$. $`DNUM(s)`$ is the number of the bunsetsus of the sentence $`s`$. $`\alpha `$ and $`\beta `$ are constants, and are set by experiment. (Although we use only monomial and binomial syntactic information in Eq. 1, we can also use trinomial or polynomial syntactic information.)
We calculate the similarity between two words by using the EDR dictionaries <sup>4</sup><sup>4</sup>4 The similarity between words can be handled by using thesauri. But the similarity between long expressions such as clauses is difficult to handle. To solve this problem, we have already considered the method of using rewriting rules . This method will be described in later papers.. In the case of the bunsetsu containing an interrogative pronoun, the similarity is calculated according to the situations. For example, when a bunsetsu in the question sentence is “where” and the paired bunsetsu in the sentence $`s`$ has the meaning of location<sup>5</sup><sup>5</sup>5Specifying bunsetsus whose meanings are locations is done by using thesauri such as the EDR dictionaries., the similarity between them is set to high.
Our system performs the above matching process and selects the answer from the sentence having the highest score. The answer is selected by considering a bunsetsu paired with a bunsetsu containing an interrogative pronoun as the desired answer.
In general, the answer of the question sentence can be obtained by matching the question sentence and the database sentences. In the case of YES-NO questions, the system has only to match the question sentence and the database sentence, and outputs YES if matched (or NO otherwise). In the case of fill-in-the-blanks-type questions<sup>6</sup><sup>6</sup>6 The process of solving fill-in-the-blanks-type questions can be considered as a case of ellipsis resolution if the blanks are considered as ellipses. We have already discussed how corpora can be used in ellipsis resolution . So we should be able to use corpora to fill blanks in the fill-in-the-blanks-type questions., the system has only to consider the element of the database sentence, paired with an interrogative pronoun such as “What” as the desired answer. Our approach here is an implementation of this idea using syntactic information.
## 3 Example
This section shows three examples of when our system obtained correct answers.
We used as question sentences the English-to-Japanese translations of sample sentences in TREC8. We used as database sentences the Daijirin Japanese word dictionary and the Mainichi Japanese newspaper (1991-1998). When we use the Daijirin dictionary, we added the strings “entry word $`+`$ wa (topic-marking functional word)” to the beginning of each sentence.
First, we inputted the following Japanese sentence into our system.
| uganda | no | shuto | wa | doko | desu | ka. |
| --- | --- | --- | --- | --- | --- | --- |
| (Uganda) | (of) | (capital) | topic | (where) | (be) | (?) |
| (What is the capital of Uganda?) | | | | | | |
As a result of calculating the score of Eq. 1, the following sentence in the Daijirin dictionary had the highest score and “Kampala” was correctly selected.
| kanpara | wa | uganda | kyouwakoku | no | shuto | desu. |
| --- | --- | --- | --- | --- | --- | --- |
| (Kampala) | topic | (Uganda) | (republic) | (of) | (capital) | (be) |
| (Kampala is the capital of the Uganda republic.) | | | | | | |
The score is calculated in the following.
$`Score`$ $`=`$ $`9.7\text{(Matching between “Uganda” and “Uganda republic”)}`$
$`+`$ $`5.9\text{(Matching between “capital” and “capital”)}`$
$`+`$ $`1.6\text{(Matching between “capital of Uganda” and “capital of Uganda republic”)}`$
$`\mathrm{}`$
$`=`$ $`17.2`$
Next, we inputted the following Japanese sentence into our system.
| magunakaruta | ga | tyouin | sareta | no-wa | nan | nen | desu | ka. |
| --- | --- | --- | --- | --- | --- | --- | --- | --- |
| (Magna Carta) | subject | (sign) | passive | topic | (what) | (year) | (be) | (?) |
| (What year was the Magna Carta signed?) | | | | | | | | |
As a result of calculating the score of Eq. 1, the following sentence in the Daijirin dictionary had the highest score and “1215” was correctly selected.
| magunakaruta | wa | 1215 | nen | igirisu | no | houken | shokou | ga |
| --- | --- | --- | --- | --- | --- | --- | --- | --- |
| (Magna Carta) | topic | (1215) | (year) | (England) | (of) | (feudal) | (lords) | subject |
| kokuou | jon | ni | semari, | ouken | no | seigen | to |
| --- | --- | --- | --- | --- | --- | --- | --- |
| (king) | (John) | object | (press) | (royal authority) | (of) | (limitation) | (and) |
| shokou | no | kenri | wo | kakunin | saseta | bunsho. |
| --- | --- | --- | --- | --- | --- | --- |
| (lords) | (of) | (right) | object | (confirm) | causative | (document). |
| (Magna Carta is the document in which feudal lords of England made King John | | | | | | |
| confirm the limitation of the royal authority and their rights in 1215.) | | | | | | |
The score is calculated in the following.
$`Score`$ $`=`$ $`32.0\text{(Matching between }\text{nan nen}\text{ “what year” and }\text{1215 nen}\text{ “1215 year”)}`$
$`+`$ $`14.6\text{(Matching between “Magna Carta” and “Magna Carta”)}`$
$`\mathrm{}`$
$`=`$ $`48.1`$
Finally, we inputted the following Japanese sentence into our system.
| paakinson | byou | wa | nou | no | dono | bubun | ni-aru | saibou | no |
| --- | --- | --- | --- | --- | --- | --- | --- | --- | --- |
| (Parkinson) | (disease) | topic | (brain) | (of) | (what) | (area) | (in) | (cell) | (of) |
| shi | ni | kankei-shite-imasu | ka. |
| --- | --- | --- | --- |
| (demise) | (to) | (be linked) | (?) |
| (The symptoms of Parkinson’s disease are linked to the demise of cells in what area | | | |
| of the brain?) | | | |
As a result of calculating the score of Eq. 1, the following sentence in the Mainichi newspaper had the highest score and “substantia nigra” was correctly selected.
| paakinson | byou | wa | tyuunou | no | kokushitsu | ni-aru | meranin |
| --- | --- | --- | --- | --- | --- | --- | --- |
| (Parkinson) | (disease) | topic | (midbrain) | (of) | (substantia nigra) | (in) | (melanin) |
| saibou | ga | hensei-shi, | kokusitsu | saibou | nai-de | tsukurareru |
| --- | --- | --- | --- | --- | --- | --- |
| (cell) | subject | (degenerate) | (substantia nigra) | (cell) | (in) | (be made) |
| shinkei-dentatsu-busshitsu | no | doupamin | ga | nakunari | hatsubyou-suru, |
| --- | --- | --- | --- | --- | --- |
| (neurotransmitter) | (of) | (dopamine) | subject | (run out) | (be taken ill) |
| to-sarete-iru. |
| --- |
| (be recognized) |
| (Parkinson’s disease is recognized when melanin cells in the substantia nigras of the midbrain |
| --- |
| degenerate. The neurotransmitter dopamine, which is made in substantia nigra cells, runs |
| out, and Parkinson’s disease arises.) |
The score was calculated in the following.
$`Score`$ $`=`$ $`10.6\text{(Matching between “Parkinson’s disease” and “Parkinson’s disease”)}`$
$`+`$ $`6.3\text{(Matching between “cell” and “melanin cell”)}`$
$`+`$ $`1.5\text{(Matching between “brain” and “midbrain”)}`$
$`\mathrm{}`$
$`+`$ $`0.4\text{(Matching between “area of brain” and “substantia nigra of midbrain”)}`$
$`+`$ $`0.3\text{(Matching between “in area” and “in substantia nigra”)}`$
$`\mathrm{}`$
$`=`$ $`32.2`$
“cells in interrogative pronoun of brain” and “cells in substantia nigra of midbrain” were matched, and “substantia nigra” was correctly selected<sup>7</sup><sup>7</sup>7 It would be good to modify Eq. 1 in order to increase the similarity of bunsetsus around an interrogative pronoun..
## 4 Conclusion
We have outlined our question answering system using syntactic information. We intend to run more experiments, to make our system more robust.
We think that the human sentence-reading process involves a matching process between the sentence being read now and data recalled in the brain . Our question answering system matches question sentences and sentences in its database, and may therefore provide some clues to shed light on the human reading process. Future work will involve extending this current work to work on the human reading process. |
no-problem/9911/cond-mat9911275.html | ar5iv | text | # Diffusion of gold nanoclusters on graphite
## I Introduction
Nanometer-size clusters — or simply nanoclusters — are intrinsically different from bulk materials. Yet, understanding of several of their most fundamental physical properties is just beginning to emerge (see for instance Refs. ), thanks largely to rapid progresses in the technology of fabrication and analysis, but also considerable advances in computational tools and methodology.
It has recently been demonstrated that depositing clusters (rather than single atoms) on surfaces allow the fabrication of interesting nanostructured materials whose properties can be tailored to specific technological applications, e.g., micro-electronic, optoelectronic, and magnetic devices. If single-atom deposition is used, the nanostructures have to be grown directly on the substrate through diffusion and agregation, which depends in a detailed (and in general very complicated) manner on the interactions between surface atoms and adatoms. By contrast, for cluster deposition, the clusters are prepared before they hit the surface, giving considerably more flexibility in assembling or organizing clusters for particular applications. It has been shown, for instance, that by changing the mean size of the incident carbon clusters, it was possible to modify the structure of the resulting carbon film from graphitic to diamond-like. This however requires that sufficient control over the cluster deposition and subsequent growth process be achieved.
Diffusion evidently plays a central role in the fabrication of thin films and self-organized structures by cluster deposition. It has been demonstrated experimentally that gold or antimony clusters diffuse on graphite surfaces at a surprisingly high rate of about $`10^8`$ cm<sup>2</sup>/s at room temperature, quite comparable to the rates that can be achieved by single atoms in similar conditions. This was confirmed theoretically by Deltour et al. using molecular-dynamics simulations: clusters consisting of particles which are incommensurate with the substrate exhibit very rapid diffusion. The cluster diffuses “as a whole”, and its path is akin to a Brownian motion induced by the internal vibrations of the clusters and/or the vibrations of the substrate. This is in striking contrast with other cluster diffusion mechanisms, whereby the motion results from a combination of single-atom processes, such as evaporation-condensation, edge diffusion, etc. The latter mechanisms are more appropriate to clusters which are in epitaxy with the surface, and are likely not significant in cases where the mismatch is large and/or the substrate-cluster interactions are weak, such as in Refs. and (see Ref. for a review).
In the present paper, we re-examine the problem of cluster diffusion in the cluster-substrate-mismatched case, now using a much more accurate model: indeed, in the work of Deltour et al., cluster-cluster, cluster-substrate and substrate-substrate interactions were all assumed to be of the Lennard-Jones form, which cannot be expected to correctly describe “real materials”. Here, we consider a simple, but realistic model for the diffusion of gold clusters on a graphite surface (HOPG). We are concerned with gold because it has been the object of several experimental studies, but also because realistic semi-empirical, many-body potentials are available for this material. The energetics of gold atoms is described in terms of the embedded-atom-method (EAM), while carbon atoms are assumed to interact via Tersoff potentials; the (weak) interactions between gold and carbon atoms are modeled with a simple Lennard-Jones potential. A comparable model was used recently by Luedtke and Landman to study the anomalous diffusion of a gold nanocluster on graphite; diffusion was found to proceed via a stick-slip mechanism, resulting in an apparent Lévy-flight type of motion. In the present work, we examine closely the variations with temperature of the rate of diffusion, as well as the microscopics of cluster dimers (diclusters).
We find the diffusivity of monoclusters to be entirely comparable to that for single adatoms. Likewise, and most important, diclusters are also found to diffuse at a rate which is comparable to that for adatoms and monoclusters. It is therefore expected that large islands, formed by the aggregation of many clusters, should also be mobile. Based on this observation, we carried out kinetic Monte Carlo simulations of island diffusion and coalescence assuming a proper scaling law for the dependence on size of the diffusivity of large clusters. We find that islands consisting of as many as 100 monoclusters exhibit significant mobility; this is consistent with the observation on graphite of large (200 monoclusters) gold islands. The morphology of cluster-assembled materials is profoundly affected by the mobility of multi-cluster islands.
## II Computational details
Diffusion coefficients for clusters can only be obtained at the expense of very long MD runs: there exists numerous possible diffusion paths, and there is therefore not a single energy barrier (and prefactor) characterizing the dynamics. These systems, further, do not lend themselves readily to accelerated MD algorithms. Brute-force simulations – long enough for statistically-significant data to be collected – therefore appear to be the only avenue. This rules out ab initio methods, which can only deal with very small systems (a few tens of atoms) over limited timescales (tens of picoseconds at best): empirical or semi-empirical potentials must be employed.
As mentioned above, we describe here the interactions between Au particles using the embedded-atom method (EAM), an $`n`$-body potential with proven ability to describe reliably various static and dynamic properties of transition and noble metals in either bulk or surface configurations. The model is “semi-empirical” in the sense that it approaches the total-energy problem from a local electron-density viewpoint, but using a functional form with parameters fitted to experiment (equilibrium lattice constant, sublimation energy, bulk modulus, elastic constants, etc.).
The interactions between C atoms are modelled using the Tersoff potential, an empirical $`n`$-body potential which accounts well for various conformations of carbon. The Tersoff potential for carbon is truncated at 2.10 Å, which turns out to be smaller than the inter-plane distance in graphite, 3.35 Å. Thus, within this model, there are no interactions between neighbouring graphite planes. This is of course an approximation, but not a bad one since basal planes in graphite are known to interact weekly. (This is why it is a good lubricant!). A pleasant consequence of this is that the substrate can be assumed to consist of a single and only layer, thus reducing formidably the (nevertheless very heavy) computational load of the calculations.
Last — and most problematic — is the Au-C interaction, for which no simple (empirical or semi-empirical) model is to our knowledge available. One way of determining this would be to fit an ab initio database to a proper, manageable functional potential. However, since Au-C pairs conform in so many different ways in the present problem, this appears to be a hopeless task, not worth the effort in view of the other approximations we have to live with. We therefore improvised this interaction a little bit and took it to be of the Lennard-Jones form, with $`\sigma =2.74`$ Å and $`ϵ=`$ 0.022 eV, truncated at 4.50 Å. The parameters were determined rather loosely from various two-body models for Ag-C and Pt-C interactions. Overall, we expect our model to provide a qualitatively correct description of the system, realistic in that the most important physical characteristics are well taken into account. It is however not expected to provide a quantitatively precise account of the particular system under consideration, but should be relevant to several types of metallic clusters which bind weekly to graphite.
We consider here gold nanoclusters comprising 249 atoms, a size which is close to that of clusters deposited in the experiments. The graphite layer has dimensions 66.15$`\times `$63.65 Å<sup>2</sup> and contains 1500 atoms. Calculations were carried out for several temperatures in the range 400–900 K. It should be noted that a free-standing 249-atom Au cluster melts at about 650 K in this model. This temperature is not affected in a significant manner by the graphite substrate as the interaction between Au and graphite-C atoms are weak. However, the dynamics of the cluster is expected to be different in the high-temperature molten state and the low-temperature solid state. All simulations were microcanonical, except for the initial thermalisation period at each temperature; no drift in the temperature was observed.
Simulations were carried out in most cases using a fully dynamical substrate. In two cases, one for the single cluster and the other for the dicluster, extremely long runs using a static (frozen) substrate were performed: it has been found by Deltour et al. that diffusion is quantitatively similar on both substrates (however, see Section III B below). The equations of motion were integrated using the velocity form of the Verlet algorithm with a timestep of 1.0 and 2.5 fs for dynamic and static substrates, respectively. (Carbon being a light atom, a smaller timestep is needed in order to properly describe the motion). The dynamic-substrate simulations ran between 10 and 14 million timesteps (depending on temperature), i.e., 10–14 ns. The static-substrate simulation for the monocluster, in comparison, ran for a total of 50 million timesteps, i.e., a very respectable 125 ns = 0.125 $`\mu `$s; the corresponding dicluster simulation ran for 75 ns. All calculations were performed using the program groF, a general-purpose MD code for bulk and surfaces developed by one of the authors (LJL).
## III Results
### A Dynamic-substrate simulations
We first discuss diffusion on a dynamic substrate, i.e., with all parts of the system explicitly dealt with in the MD simulations. Fig. 1 gives the (time-averaged) mean-square displacements (MSD’s) of the cluster’s center-of-mass at the various temperatures investigated, which will be used to calculate the diffusion constant, $`D=lim_t\mathrm{}r^2(t)/4t`$. As indicated above, the simulations extend over 10–14 ns, but the MSD’s are only shown for a maximum correlation time of one ns in order to “ensure” statistical reliability. It is evident (e.g., upon comparing the results at 700 and 800 K) that the diffusion coefficients that can be extracted from these plots will carry a sizeable error bar. Nevertheless, it is certainly the case that (i) diffusion is very significant and (ii) it increases rapidly with temperature.
There is no evidence from these plots that the MSD’s obey a non-linear power law behaviour (i.e., that the cluster undergoes superdiffusion) which could be associated with “Lévy flights”: the statistical accuracy of the data is simply not sufficient to draw any conclusions. The cluster does however undergo long jumps during the course of its motion. We will return to this point below when we discuss diffusion on a frozen substrate.
In lack of a better description of the long-time behaviour of the diffusion process, we simply assume that $`r^2(t)4Dt`$ as $`t`$ gets large. The resulting diffusion coefficients are plotted in the manner of Arrhenius, i.e., $`\mathrm{log}D`$ vs $`1/k_BT`$, in the inset of Fig. 1. If the process were truly Arrhenius, all points would fall on a single straight line. This is evidently not the case here. Though we could probably go ahead and fit the data to a straight line, attributing the discrepancies to statistical error, there is probably a natural explanation for the “break” that a sharp eye can observe between 600 and 700 K: As noted above, the free Au<sub>249</sub> cluster melts at about 650 K in the EAM model. The presence of the substrate raises the melting point, but very little since the interactions between the cluster and the graphite surface are small. Thus, the cluster is solid at the lowest temperatures (400, 500 and 600 K), but liquid above (700, 800 and 900 K). The statistics are evidently insufficient to allow firm conclusions to be drawn; there nevertheless appears to be a discontinuity near the cluster melting point temperature, with activation energies on either side of about 0.05 eV. We discuss in Section III D the implications of these findings on the kinetics of growth.
### B Static-substrate simulations
The static-substrate simulations, carried out at a single temperature (for the cluster), viz. 500 K, serve many purposes: (i) re-assess the equivalence with dynamic-substrate MD runs reported by Deltour et al.; (ii) provide accurate statistics for a proper comparison of the diffusive behaviour of mono- and diclusters; (iii) examine the possible superdiffusive character of the trajectories.
We focus, first, on a comparison between static- and dynamic-substrate simulations. As can be appreciated from the MSD’s given in the inset of Fig. 2, there is a rather substantial difference between the two calculations: for the dynamic substrate at 500 K, the diffusion constant is $`3.71\times 10^5`$ cm<sup>2</sup>/s, while for the frozen substrate we have $`1.09\times 10^5`$ cm<sup>2</sup>/s. (This value is actually significantly smaller than that for the dynamic substrate at 400 K — 100 K lower temperature — viz. $`1.70\times 10^5`$ cm<sup>2</sup>/s). Again, statistical uncertainties cannot be totally excluded to account for this discrepancy, but it is difficult to imagine that it could explain all of the observed difference (cf. inset to Fig. 1 for a better appreciation of this difference). The explanation might however be quite simple.
As noted above, the cluster-substrate interactions are weak, and this likely plays an important role in determining the characteristics of the motion. Visual inspection of the $`xy`$ paths in the two different situations makes it apparent that the motion has a much stronger “stick-and-jump” character on the frozen substrate than on the dynamic one. On the frozen substrate, further, the trajectory is more compact on a given timescale. This can in fact be characterized in a quantitative manner by considering, following Luedtke and Landman, the function $`P_\tau (d)`$, which gives the distribution of displacements of length $`d`$ over a timescale of $`\tau `$. The motion is best characterized using a value of $`\tau `$ corresponding to the period of vibration of the cluster in a sticking mode (see below).
The function $`P_\tau (d)`$ (normalized to unity) is displayed in Fig. 3 for the dynamic substrate at three different temperatures (400, 500, and 900 K) and for the static substrate at 500 K. The value of $`\tau `$ was determined from the frozen-substrate simulations by simply counting the number of oscillations over a given period of time; we found $`\tau =20`$ ps to within about 10%. We note that, for the dynamic substrate, the period of oscillations at 400 K is about 38 ps, while no oscillations can be found at 500 and higher temperatures, i.e., the sticking mode is absent above 500 K or so.
The difference between static and dynamic substrates is striking: On the dynamic surface, $`P_\tau (d)`$ is a broad featureless distribution, which gets broader as temperature increases. The maximum of the distribution at low temperature lies at about 1.6–1.8 Å — roughly the distance between equilibrium sites on the graphite surface — clearly establishing that the motion proceeds in an quasi continuous manner via “sliding hops” to nearest-neighbours; the hops get longer as temperature increases. On the static substrate, in contrast, a “sticky” vibrational mode, of amplitude roughly 0.25 Å, is clearly visible. This is followed by a broad tail which corresponds, again, to the sliding jumps that are characteristic of the motion on the dynamic substrate.
Sticking, therefore, is much more likely to take place on the static than on the dynamic substrate, thereby contributing to decrease the average distance traveled by the cluster over a given period of time. This conclusion is however not general: The system under consideration here is perhaps a bit peculiar in that the cluster-substrate interactions are especially weak. (In comparison, Luedtke and Landman’s $`ϵ`$ for the Au-C interaction is 0.01273 eV, even smaller than our own value.) One may conjecture that the vibrations of the surface are enough, in such cases, to overcome completely the barrier opposing diffusion, which might not be true of systems where the interactions are stronger (as in the case, e.g., of Deltour et al.’s simulations, Ref. ).
It also appears that our diffusion data do not cover a timescale long enough to warrant firm conclusions to be drawn on the possibility that superdiffusion might be taking place. This is certainly true, as we have seen above, of the dynamic-substrate simulations, which cover “only” 10–14 nanoseconds, but also of the static-substrate simulations (assuming, in view of the above discussion, that they are relevant to the problem under study), which extend to 125 ns. Certainly, the position of the cluster’s center-of-mass does exhibit something of a self-similar character, as reported by Luedtke and Landman, and as can be seen in Fig. 4. Evidently, one cannot trust statistics here over more than a decade or two in time. One might hope that superdiffusion would be more apparent in the long-time behaviour of the MSD’s. To this effect, we plot in Fig. 2 $`\mathrm{log}r^2(t)`$ vs $`\mathrm{log}t`$, for a maximum correlation time of (here) 40 ns; the slope of such a plot is the diffusivity exponent $`\gamma `$. The statistical quality of the data decreases with correlation time, and becomes clearly insufficient over 5 ns or so; the large dip at about 15 ns can testify. Our best estimate of the slope $`\gamma `$ at “large” (more than $``$1 ns) correlation times is anywhere between 0.9 and 1.2, i.e., mild underdiffusion or mild superdiffusion… or no superdiffusion at all! This is consistent with the value reported by Luedtke and Landman, who find $`\gamma =1.1`$ based on an analysis of sticking and sliding times. One point worth mentioning is that the velocity-autocorrelation function for adatom diffusion in the intermediate and high-friction regimes has been shown to follow a power-law behaviour at intermediate times; the exponential dependence resumes at very long times.
### C Diclusters
The morphology of films grown by cluster deposition depends critically on the coefficient of diffusion of monoclusters, as we have just seen, but also, because clusters aggregate, on the coefficient of diffusion of multiclusters. From simple geometric arguments, it might be argued that the rate of diffusion should scale as $`N^{2/3}`$, where $`N`$ is the number of atoms in the cluster, as was in fact observed by Deltour et al. for Lennard-Jones clusters. However, it can be expected that the morphology of the films depends, as well, on the shape of the multi-clusters following the aggregation of monoclusters, i.e., on the kinetics of coalescence.
In a previous publication, we examined the coalescence of gold nanoclusters in vacuum and found it to be much slower than predicted by macroscopic theories. This state of affairs can be attributed to the presence of facets and edges which constitute barriers to the transport of particles required for coalescence to take place . The “neck” between two particles was however found to form very rapidly. We conjectured that these conclusions would apply equally well to the particular case of gold nanoclusters on graphite since the gold-graphite interactions are weak.
We have verified this in the context of the present work: indeed, coalescence is little affected by the presence of the substrate, as demonstrated in Fig. 5. We considered both a free-standing and a supported pair of 249-atom gold clusters. Starting at very low temperature (50 K), temperature was slowly and progressively (stepwise) raised to 600 K. (As noted above, the 249-atom gold cluster melts at about 650 K in this EAM model and we therefore did not go beyond this point). We plot, in Fig. 5, the evolution with time-temperature of the three moments of inertia of the dicluster. Since the cluster can rotate, the moments of inertia provide a more useful measure of the shape of the object than, e.g., the radii of gyration. A side view of the dicluster at 200 K, i.e., after the neck between the two monoclusters has formed completely, is shown in Fig. 6. It is evident that the dicluster does not wet the surface, and therefore the substrate plays a relatively minor role in the coalescence process.
As can be seen in Fig. 5, the behaviour of the free-standing and supported diclusters are almost identical, except for the initial phase of coalescence: the supported cluster forms a neck much more rapidly than the free-standing cluster, presumably because the substrate offers, through some thermostatic effect, an additional route via which coalescence (by plastic deformation) can be mediated; it is conceivable also that the substrate “forces” the atomic planes from the two clusters to align. We have not explored these questions further; it remains that the end points of the two coalescence runs are identical within statistical uncertainty. Thus, again, coalescence is hampered by the presence of facets and edges; the timescale for complete coalescence is much longer than predicted by continuous theories. The shape of islands on the graphite surface will be strongly affected, and it is also expected that the rate of diffusion will be affected (since it is determined by the contact area between substrate and cluster).
The MSD of the dicluster (after proper equilibration at 500 K) is displayed in the inset of Fig. 2. As mentioned earlier, this was calculated from a static-substrate run covering 75 ns. The same limitations as noted above for the monocluster should therefore hold in the present case. It is a very remarkable (and perhaps even surprising) result that the rate of diffusion of the dicluster is quite comparable to that of the monocluster, inasmuch as the frozen-substrate simulations are concerned. (We expect the diffusion constants on the dynamic substrate to be different — and larger — but in a proportion that would be quite comparable to that found here). The value of $`D=1.38\times 10^5`$ cm<sup>2</sup>/s we obtain for the dicluster is in fact a bit larger than that for the monocluster ($`1.09\times 10^5`$ cm<sup>2</sup>/s). The difference is probably not meaningful; what is meaningful, however, is that the the mono- and the dicluster have comparable coefficients of diffusion; this has profound implications on growth, as we discuss in Section III D, below.
The function $`P(d)`$ for the dicluster at 500 K is displayed in Fig. 3; here we estimate that $`\tau =40`$ ps (vs about 20 ps for the monocluster). The distribution is quite similar to that found for the single cluster on the frozen substrate, though broader and shifted to slighlty larger displacements. This last result is likely due to the fact that, being larger, the dicluster is not as easily able to accomodate itself with the substrate as the monocluster; in this sense, it is more loosely bound to the substrate.
### D Comparison with experimental results
Deposition of gold clusters on graphite experiments were carried out in Lyon recently. Several models have been proposed to extract the microscopic cluster diffusion coefficients from the measured island densities. Of course, in order to provide a meaningful interpretation of the data, the models must take into account the precise conditions in which the experiments are performed. In Lyon, for instance, the flux of clusters is chopped, rather than continuous, and this affects the kinetics of diffusion and growth considerably. Previous estimates of the rates of diffusion of Au on graphite, which overlooked this important detail, are therefore in error. In Ref. , a diffusion coefficient of 10<sup>-3</sup> cm<sup>2</sup>/s at 400 K is given; for a discussion, see Ref. . The “correct” number, including flux chopping, would be 1.0 cm<sup>2</sup>/s if monoclusters only were assumed to be mobile. However, as we have seen above, cluster dimers diffuse at a rate which is quite comparable to that for monoclusters, suggesting that larger clusters would diffuse as well. The Lennard-Jones simulations of Deltour et al. indicate that the rate of diffusion of compact $`N`$-atom clusters scales roughly as the inverse of the contact area between the cluster and the substrate: $`D_N=D_1N^{2/3}`$. (Compact clusters are expected to form through aggregation and coalescence; see Ref. ).
Experimentally, however, it is almost impossible to determine whether or not multiclusters do diffuse, and at which rate. In view of this, and the expected importance of multicluster mobility on growth, we have carried out a series of kinetic Monte Carlo (KMC) simulations in order to estimate the largest island which must be allowed to diffuse in order to account for the experimentally-observed gold island density on graphite at 400 K, viz. $`4\times 10^8`$ islands/cm<sup>2</sup>, or $`1.1\times 10^5`$ per site. To do so, we assume that the diffusion constant for monoclusters found in the present simulations is correct, and that the rate of diffusion of $`N`$-clusters scales according to the law given above. All other parameters (incident cluster flux, temperature, chopping rate, etc.) are fixed by experiment.
Figure 7 shows the results of the KMC simulations: we plot here the island density that would be observed if the largest mobile island were of size $`N_{\mathrm{max}}`$. The computational load increases very rapidly with $`N_{\mathrm{max}}`$ and we therefore only considered islands of sizes less than or equal to 35. The data points follow very closely a power-law relation and we can thus extrapolate to larger values of $`N_{\mathrm{max}}`$, i.e., smaller island densities. We find in this way that islands up to a maximum size of about 100 mono-clusters must be mobile in order to account for the observed island density of $`1.1\times 10^5`$ per site. In what follows, we discuss in more detail the connection of this observation with experiment.
We first note that, in the gold-on-graphite experiments, large islands form which are “partially ramified”, in the sense that the branch width is much larger than the size of the deposited clusters, each branch being formed by the coalescence of up to 200 monoclusters. In contrast, for antimony cluster deposition on graphite at room temperature, the islands are fully ramified, i.e., have a branch width identical to the diameter of the monoclusters; this establishes unambiguously that cluster coalescence is not taking place in this case. It has been shown, further, that the mobility of the islands is negligible in antimony. Our results suggest, therefore, when taken together with the work of Deltour et al., that compact islands, which form through diffusion and coalescence, are mobile according to a $`N^{2/3}`$ law. In contrast, ramified islands, which form when coalescence does not take place, have much reduced mobility – certainly much less than would be expected from a $`N^{2/3}`$ law. $`N_{\mathrm{max}}`$, therefore, signals the crossover point between the two mobility regimes or, equivalently, the multicluster size at which the morphology of the islands crosses over from compact to ramified (or vice-versa). The physical reasons underlying the relation between mobility and morphology are not clear, but there appears to be no other ways to interpret the experimental results. This problem clearly deserves further studies.
To summarize this section, the mobility of large islands is evidently a necessary ingredient to account for the experimentally observed island density. Our simulations suggest that these islands can be as large as 100 monoclusters; while this is consistent with experiment, the exact value, as well as the precise dependence of the diffusion rate on size, cannot at present be estimated.
## IV Concluding remarks
Cluster-deposition techniques are of great potential interest for assembling materials with specific, tailor-made applications. Yet, the fabrication process depends critically on the possibility for the clusters to diffuse on the surface in order to settle in appropriate positions, thus forming self-organized structures, or to aggregate/coalesce with other clusters in order to form larger-scale structures and eventually continuous layers. In this article, we have demonstrated, using molecular-dynamics simulations with realistic interatomic potentials, that the diffusion of large metallic clusters on graphite can take place at a pace which is quite comparable to that for single adatoms. We have also established that the rate of diffusion of cluster dimers can be very sizeable, comparable in fact to that for monoclusters. An extremely important consequence of this is that islands formed by the aggregation of clusters are also expected to be mobile. Using kinetic Monte Carlo simulations and assuming a proper scaling law for the dependence on size of the diffusivity of large clusters, we estimate that islands containing as much as 25 000 atoms (100 monoclusters) are expected to undergo diffusion at a significant rate on graphite surfaces. These findings have profound consequences for the morphology of cluster-assembled thin films.
###### Acknowledgements.
We are grateful to Laurent Bardotti, Art Voter and Tapio Ala-Nissila for useful discussions. This work was supported by the Natural Sciences and Engineering Research Council of Canada and the “Fonds pour la formation de chercheurs et l’aide à la recherche” of the Province of Québec. LJL is grateful to the Département de physique des matériaux de l’Université Claude-Bernard-Lyon-I, where part of this work was carried out, for hospitality, support, and pleasant weather. |
no-problem/9911/cond-mat9911022.html | ar5iv | text | # Acoustic Nature of the Boson Peak in Vitreous Silica
\[
## Abstract
New temperature dependent inelastic x-ray (IXS) and Raman (RS) scattering data are compared to each other and with existing inelastic neutron scattering data in vitreous silica ($`vSiO_2`$), in the 300 $`÷`$ 1775 K region. The IXS data show collective propagating excitations up to $`Q=3.5`$ $`nm^1`$. The temperature behaviour of the excitations at $`Q=1.6`$ $`nm^1`$ matches that of the boson peak found in INS and RS. This supports the acoustic origin of the excess of vibrational states giving rise to the boson peak in this glass.
\]
A peculiar characteristic of disordered solids, not present in their crystalline counterpart (Phillips, 1981), is an excess of low frequency states with respect to the predictions of the Debye theory. Inelastic neutron scattering (INS) studies on the density of states, $`g(\omega )`$, reveal, in fact, a broad peak (the boson peak ) (BP) in the quantity $`g(\omega )/\omega ^2`$ (Buchenau et al. 1986, Buchenau et al. 1988, Sokolov et al. 1995) at energies 2$`÷`$10 meV (15$`÷`$80 cm<sup>-1</sup>), where the Debye theory would predict a constant. This excess is responsible for the presence, in any glass studied so far, of a bump in the temperature dependence of $`C_p/T^3`$ (Sokolov et al. 1993, Brodin et al. 1994).
The broad boson peak feature is also observed in Raman scattering (RS) spectra. This can be understood considering that the first order Raman scattering intensity, $`I(\omega )`$, in the harmonic approximation, is connected to $`g(\omega )`$ (Galeneer and Sen 1978): $`I(\omega )=C(\omega )g(\omega )[n(\omega ,T)+1]/\omega .`$ Here $`n(\omega ,T)`$ is the Bose population factor and $`C(\omega )`$ is the photon-excitation coupling function, assumed to have a smooth behaviour in the considered energy region (Fontana, Rocca and M. P. Fontana 1987, Fontana et al. 1997).
The nature of this excess of states is still subject to speculation and, at present, there are two different prevailing hypotheses (See Philosophical Magazine, special issue: V international workshop on disordered systems, Andalo, 1994). In the first one, the excess of states is explained by the localization of the high energy vibrational modes induced by the static disorder in the glass. In the second case, collective propagating modes are thought to persist at high frequency, and the BP reflects their density of states. In the first view the BP energy corresponds to the energy of the high $`Q`$ excitations, while in the second one the excitations are expected to propagate at energies above the BP energy.
Information on the character of the excitations contributing to the boson peak can be obtained from measurements of the dynamic structure factor $`S(Q,\omega )`$. Using inelastic x-ray scattering (IXS) (Sette et al. 1995), it has been possible to measure $`S(Q,\omega )`$ of different glasses in the momentum-energy region of the BP, and to support the hypothesis of the acoustic-like character of the collective modes and therefore of the peak itself (Masciovecchio et al. 1996a, Masciovecchio et al. subm.). In the specific case of $`v`$-silica, however, the IXS data have been alternatively interpreted either with the phonon-localization model (Foret et al. 1997), or as propagating excitations (Benassi et al 1997, Masciovecchio et al. 1997).
A better insight on the relation between the excitations seen in $`S(Q,\omega )`$ and the BP could be gained by comparing the temperature dependence of their energies. The basic idea is to utilise the observation that in vitreous silica, in contrast to the majority of glasses, the low frequency sound velocity, measured by Brillouin Light Scattering (BLS), (Vacher and Pelous 1976) and the boson peak energy (Wischnewski et al. 1998) shift towards higher values with increasing temperatures, and, more importantly, the rates of change reported in the two cases are different. It is not the aim of this work to account for these peculiarities, likely to be due to the specific shape of the interatomic potential in silica. Rather, one can utilize the different temperature dependence of excitations with different $`Q`$, as seen by IXS, INS, RS and BLS, to establish a link between specific vibrational states and the features of the density of states. In particular, one can assess whether the excitations observed by IXS at energies comparable to the BP energy have either a temperature behaviour similar to that of the BP, or to that of the low $`Q`$ excitations measured by BLS, or even another dependence.
In this Letter we report temperature dependent IXS and RS data of vitreous silica. In the IXS spectra, propagating collective excitations are found at $`T`$ 1400 K, thus confirming that the high frequency longitudinal dynamics has an acoustic-like nature in the whole temperature region below melting, and up to a momentum transfer $`Q`$=3.5 nm<sup>-1</sup>, corresponding to energies more than twice the BP energy ($`E_{_{BP}}4`$ meV at room temperature). More importantly, we find the same temperature shift for the BP energy and for the energy, $`\mathrm{\Omega }(Q)`$, of the modes at $`Q^{}=1.6`$ nm<sup>-1</sup> ($`\mathrm{\Omega }(Q^{})E_{_{BP}}`$). This supports the argument that the acoustic modes, probed by the IXS, give a significant contribution to the BP.
The IXS experiment was carried out at the very high energy resolution inelastic x-ray scattering beamline (BL21/ID16) at the European Synchrotron Radiation Facility. The total resolution function, measured using a Plexiglas scatterer at the maximum of its static structure factor ($`Q`$=10 nm<sup>-1</sup>), has a full width at half maximum (fwhm) of 1.5 $`\pm `$ 0.1 meV. The $`Q`$ values were selected between 0.75 and 4 nm<sup>-1</sup>, and the $`Q`$ resolution was set to 0.2 nm<sup>-1</sup> fwhm. Energy scans were performed by varying the relative temperature between the monochromator and the analyser crystals by $`\pm `$0.45 K with a step of 0.0075 K. Each scan took about 200 min, and each $`Q`$-value was obtained by averaging five scans (the total integration time was 500 s per point). The data were normalized to the intensity of the incident beam. Further details on the IXS beamline are reported elsewhere (Masciovecchio et al. 1996a, b, Verbeni et al. 1996). The Raman scattering measurements were performed using a standard Raman laser system. Depolarized spectra were collected in the -300$`÷`$2000 cm<sup>-1</sup> frequency range with 1 cm<sup>-1</sup> frequency resolution. The $`SiO_2`$ suprasil sample, purchased from Goodfellow, was the same 2 $`mm`$ diameter rod used in (Benassi et al 1997, Masciovecchio et al. 1997).
From the linear dispersion curve determined by a first set of IXS spectra taken at $`T`$=1375 K, it was possible to establish that the excitations have the same propagating nature as that previously observed in the 300$`÷`$1000 K region. In Fig. 1, we report as an example the IXS spectrum taken at $`Q=1`$ nm<sup>-1</sup>, together with the fitting function. As in previous works, the IXS spectra were fitted by the experimentally determined resolution function convoluted with a $`\delta `$-function for the elastic peak and a Damped Harmonic Oscillator (DHO) model (Fak and Dorner, 1992) for the inelastic signal. The excitation energies are reported in the inset. The data show an acoustic like behaviour, linearly extrapolating at small $`Q`$ with a slope $`v=6800\pm 200`$ m/s. This value, larger than the sound velocity measured by Brillouin light scattering (Vacher et al. 1976) implies a deformation of the acoustic branch not recognized before, which will provide an interesting subject to further investigations. The dispersion relation covers the whole BP energy region ($`5÷7`$ meV), and reaches energies as high as 15$`÷`$20 meV. These observations, as in previous lower temperature measurements, show that, even at temperatures approaching $`T_g`$, it exists an acoustic-like propagating mode up to energies that lie well above $`E_{_{BP}}`$.
A second set of IXS measurements was performed as a function of temperature at the fixed value $`Q^{}=1.6`$ nm<sup>-1</sup>, which was chosen since $`\mathrm{\Omega }(Q^{})`$ is in the range of $`E_{_{BP}}`$ at room temperature. The IXS data and the corresponding fits are shown in Fig. 2. It is possible to observe directly from the row data that the intensity and energy of the inelastic excitation increase with $`T`$. This is emphasized in the inset of Fig. 2, where the $`T`$-dependence of the inelastic-to-elastic intensity ratio, $`R(T)`$, and the excitation energy, $`\mathrm{\Omega }(Q^{},T)`$, are reported.
To compare the $`T`$dependence of $`\mathrm{\Omega }(Q^{},T)`$ with that of other spectroscopic features, we define the scaling factor $`a_{_{IXS}}(T)`$ as $`\mathrm{\Omega }(Q^{},T)/\mathrm{\Omega }(Q^{},T=0K)`$. Here $`\mathrm{\Omega }(Q^{},T=0K)`$ is the $`T=0`$ extrapolation of the measured $`\mathrm{\Omega }(Q^{},T)`$. The values of $`a_{_{IXS}}(T)`$ are reported in Fig. 3, together with the scaling factor for the low $`Q`$ ($`Q0.036`$ nm<sup>-1</sup>) excitation energy determined by Brillouin light scattering (Scopigno 1997), $`a(T)_{_{BLS}}`$. The difference in the $`T`$-dependence of the low- and high-$`Q`$ excitations indicates again that the acoustic modes dispersion relation is deformed from a simple linear law. This behaviour was hidden in the error bars in previous IXS and BLS studies (Benassi et al 1997, Masciovecchio et al. 1997). and is evidenced here thanks to increased sensitivity and temperature range. It is not the aim of the present work to discuss the origin of the observed deformation of the dispersion relation <sup>*</sup><sup>*</sup>*This effect could, for example, be ascribed to the presence of a positive fourth order anharmonicity in the next-nearest neighbours interaction potentials (Scopigno 1997) or to a relaxation process active in the glass. .
This non trivial $`T`$-dependence can help us to assess the origin of the BP by the study of the temperature dependence of its energy. In Fig. 3 we report, in fact, the scaling factor $`a_{_{INS}}(T)`$ of the BP energy maximum, as determined from existing INS measurements (Wischnewski et al. 1998). The striking similarity between $`a_{_{INS}}(T)`$ and $`a_{_{IXS}}(T)`$ provides a first strong indication that the states probed by IXS contribute to the BP. This similarity is further confirmed by the temperature dependence of the BP energy as measured by RS in the 300$`÷`$1100 K region. In Fig. 4 we report $`I(\omega )\omega /C(\omega )/[n(\omega ,T)+1]`$, which corresponds directly to $`g(\omega )/\omega ^2`$. We used the coupling coefficient $`C(\omega )`$ derived from the ratio of the INS (Wischnewski et al. 1998) and RS (this work) data measured in $`vSiO_2`$ at room temperature: this $`C(\omega )`$ is reported in the inset a) of Fig. 4, and it was used to scale all the RS spectra at different temperatures It has been recently shown that the coupling function is not dependent on temperature in the energy region spanned by the BP, (A. Fontana et al. subm.).. The energy of the maximum of the spectra of Fig. 4, derived by a local quadratic fit, is reported in inset b) and the scaling factors $`a_{_{RS}}(T)`$ are reported in Fig. 3. The two data sets (INS and RS) are equivalent within the error bar.
In presence of the deformation of the dispersion relation, it is striking to observe the identical temperature behaviour for two spectroscopic features that in principle may share only a similar energy: the BP maximum and the acoustic excitations at $`Q^{}=`$1.6 nm<sup>-1</sup>. This demonstrates their common origin. Consequently, the high frequency acoustic excitations, in spite of their propagating nature, must have a density of states exceeding the Debye prediction.
A model, that is consistent with a picture of linearly dispersing propagating excitations with an excess of states, must take into account the effect of the topological disorder on the eigenvectors of the vibrational modes. It has been shown by molecular dynamics simulation that, in a glass up to the BP energy region, the eigenvector of a given mode $`j`$ can be thought as the sum of two very different components. i) A plane wave like part, with an average momentum $`Q_j`$, related to the eigenvalue by the linear dispersion $`\omega _jvQ_j`$. The $`Q`$ distribution around $`Q_j`$ gets broader increasing the mode energy. ii) A random uncorrelated component whose spatial Fourier transform extends from $`Q_j`$ up to $`Q`$ values much higher than the first sharp diffraction peak position with a flat spectral distribution (Mazzacurati, Ruocco and Sampoli 1996). The first component accounts for the peaks in the dynamic structure factor and for the existence of a dispersion relation. The finite width of the inelastic peaks in the $`S(Q,\omega )`$, as that in Fig. 1, is due to the finite projection of eigenmodes of different energy at the considered $`Q`$. The presence of the second component allows to satisfy the eigevectors orthogonality conditions without the constrain on mode counting holding for plane waves, i. e. without imposing the $`Q^2`$ dependence of the modes density at a given frequency. Specifically, with respect to the Debye behaviour, this can give rise to an accumulation of states at low energy.
In conclusion, a study on the temperature dependence of the dynamics in vitreous silica has allowed to identify equivalent temperature behaviours for the maximum of the BP and for the high frequency continuation of the sound branch at the same energy. This provides arguments in favour of the acoustic-like origin for the excess in the density of states in vitreous silica. Reminding that the BP is observed basically in any disordered material, and that IXS has shown the presence of a propagating collective dynamics in all the glasses and liquids studied so far, it is natural to speculate on the generality of the conclusion reached here on the nature of the BP.
We acknowledge A. P. Sokolov and G. Viliani for useful discussions.
REFERENCES
For a review see Amorphous Solids : Low-Temperature properties, edited by W.A. Phillips, (Springer, Berlin, 1981).
BENASSI P., KRISH M., MASCIOVECCHIO C., MONACO G., MAZZACURATI V., RUOCCO G., SETTE F. and VERBENI R., 1997, Phys. Rev. Lett. 77, 3835.
BRODIN A., FONTANA A., BORJESSON L., CARINI G. and TORELL L. M., 1994, Phys. Rev. Lett. 73, 2067.
BUCHENAU U., PRAGER M., NUKER N., DIANOUX A. J., AHMAD N. and PHILLIPS W. A., 1986, Phys. Rev. 34, 5665.
BUCHENAU U., ZHOU H. M., NUKER N., GILROY K. S., and PHILLIPS W. A., 1988, Phys. Rev. Lett. 60, 1318.
FAK B. and DORNER B., 1992, Institute Laue Langevin (Grenoble, France), technical report No. 92FA008S.
FONTANA A., ROCCA F. and FONTANA M. P., 1987, Phys. Rev. Letters 58, 503.
FONTANA A., ROSSI F., CARINI G., D ’ANGELO G., TRIPODO G. and BARTOLOTTA A., 1997, Phys. Rev. Letters 58, 1078.
FONTANA A., DELL ’ANNA R., MONTAGNA M., ROSSI F., VILIANI G., RUOCCO G., SAMPOLI M., BUCHENAU U. and WISCHNEWSKI A., submitted to Europhys. letters
FORET M., COURTENS E., VACHER R. and SUCK J. B., 1997, Phys. Rev. Lett. 77, 3831.
GALENEER F. L. and SEN P. N., 1978, Phys. Rev. B 17, 1928.
MASCIOVECCHIO C., RUOCCO G., SETTE F., KRISH M., VERBENI R., BERGMAN U. and SOLTWISCH M., 1996, Phys. Rev. Lett. 76, 3356.
MASCIOVECCHIO C., BERGMAN U., KRISH M., RUOCCO G., SETTE F. and VERBENI R., 1996, Nucl. Inst. and Meth. B-111, 181 and B-117, 339
MASCIOVECCHIO C., RUOCCO G., SETTE F., BENASSI P., CUNSOLO A, KRISH M., MAZZACURATI V., MERMET A., MONACO G. and VERBENI R., 1997, Phys. Rev. B55, 8049.
MASCIOVECCHIO C., MONACO G., RUOCCO G., SETTE F., CUNSOLO A., KRISH M., MERMET A., SOLTWISCH M. and VERBENI R., submitted to Phys. Rev. Lett.
MAZZACURATI V., RUOCCO G. and SAMPOLI M., 1996, Europhys. Lett. 34, 681.
SCOPIGNO T., 1997, Thesis, University of L’Aquila. Unpublished.
SETTE F., RUOCCO G., KRISH M., BERGMAN U., MASCIOVECCHIO C., MAZZACURATI V., SIGNORELLI G. and VERBENI R., 1995, Phys. Rev. Lett. 75, 850.
SOKOLOV A. P., BUCHENAU U., STEFFEN W., FRICK B. and WISCHNEWSKI A., 1995, Phys. Rev. B 52, R9815.
SOKOLOV A. P., KISLIUK A., QUITMANN. D. and DUVAL E., 1993, Phys. Rev. B 48, 7692.
See for example Philosophical Magazine, 1995, B 71, 471-811. Special issue: V international workshop on disordered systems, Andalo, 1994, A. Fontana and G. Viliani Editors.
VACHER R. and PELOUS J., 1976, Phys. Rev. B14, 823.
VERBENI R., SETTE F., KRISH M., BERGMAN U., GORGES B., HALCOUSSIS C., MARTEL K., MASCIOVECCHIO C., RIBOIS J.F., RUOCCO G. and SINN H., 1996, J. of Synchrotron Radiation 3, 62.
WISCHNEWSKI A., BUCHENAU U., DIANOUX A. J., KAMITAKARHA W. A. and ZARESTKY J. L., 1998, Phil. mag. B77, 579.
FIGURE CAPTIONS
FIG. 1 - Inelastic x-ray scattering spectrum of $`vSiO_2`$ at T=1375 K and $`Q=1`$ $`nm^1`$ ($``$). The full line is the best fit to the data as discussed in the text. The dashed and dotted lines represent the elastic and inelastic contributions to the fit. The inset reports the energy of the excitations as derived from the fits. The dashed line is the linear extrapolation to small $`Q`$, and its slope corresponds to 6800 m/s.
FIG. 2 - X-ray spectra of $`vSiO_2`$ at $`Q`$=1.6 nm<sup>-1</sup> taken at different temperatures. The data are shown together with their best fits (full line) and the individual contributions to the fitting function: elastic peak (dotted line), and inelastic components (dashed line). In the inset are reported: 1) ($``$) the inelastic-to-elastic intensity ratio, $`R(T)`$; it follows the expected linear behaviour; and 2) ($``$) the excitation energy, $`\mathrm{\Omega }(T)`$, derived from the DHO model. The reported full lines are their linear fits.
FIG. 3 - Temperature dependence of the scaling factor $`a(T)`$, normalized to $`a(0)=1`$, for the different spectra. Open symbols refer to incoherent measurements of the BP: INS (open circles) and Raman scattering (open squares). Full symbols are the scaling factors for the excitations energies determined from the dynamics structure factor $`S(Q,\omega )`$ measured at $`Q`$=1.6 nm<sup>-1</sup> by IXS (full circles) and at $`Q`$=0.036 nm<sup>-1</sup> by BLS (full squares). The dashed lines are linear guides to the eye to emphasise differences and similarities among the temperature behaviours of the four data sets.
FIG. 4 - Examples of RS spectra taken, from to top to bottom, at 333, 523, 823, and 1073 K. The reported data correspond to the reduced Raman intensities divided by the $`C(\omega )`$ (Fontana et al. submitted) and shown in inset a). These spectra correspond to the $`g(\omega )/\omega ^2`$. In inset b), the energy position of the maximum intensity at each temperature ($``$) is reported togheter with its linear fit in the low temperature region (dashed line). |
no-problem/9911/math9911119.html | ar5iv | text | # On contractible curves on normal surfaces
## 1. Introduction
The goal of this paper is to characterize contractible curves on proper normal algebraic surfaces.
The corresponding question for compact normal complex-analytic surfaces was completely solved by Grauert : A curve $`R`$ on such a complex-analytic surface $`X`$ is contractible to another complex-analytic surface $`Y`$ if and only if the curve $`R`$ is negative definite. Later, Artin showed that a similar result holds in the category of proper normal 2-dimensional algebraic spaces.
Usually, if one starts with a proper normal algebraic surface $`X`$ and a negative definite curve $`RX`$, the resulting algebraic space $`Y`$ is not a scheme anymore. We want to characterize those curves $`RX`$ for which $`Y`$ remains a scheme. For simplicity, we call such curves contractible. Our question was already considered by Artin in a series of two papers , where he gives several necessary or sufficient conditions for smooth surfaces. For example, the classical Castelnuovo criterion tells us that an exceptional curve of the first kind is contractible.
Our main result is a characterization of contractible curves $`RX`$ in terms of complementary Weil divisors, see theorem (3.4). Roughly speaking, one has to find a Weil divisor which defines the contraction numerically and behaves reasonably on the formal completion $`𝔛`$ of $`RX`$. From this characterization we will easily derive some generalizations of the classical results of Castelnuovo and Artin. Furthermore, our approach will explain certain differences between characteristic $`p>0`$ and characteristic zero.
The results will be useful in the classification à la Enriques of proper normal algebraic surfaces, since they allow us to introduce the technique of extremal rays from Mori theory without conditions on the singularities. Since our ground field $`K`$ is arbitrary, the results might also be applied to generic hyperplane sections or generic fibres in higher dimensional geometry.
As an application we will show that for every Weil divisor $`DZ^1(X)`$, the homogeneous spectrum $`P(X,D)`$ of the graded algebra $`_{n0}H^0(X,nD)`$ is always a scheme of finite type. Here the interesting point is that the algebra usually fails to be finitely generated.
This article is divided into six sections, including the introduction. In the second section we will set up the notation and recall some facts about the cone of curves for proper normal surfaces. The third section contains a characterization of almost affine open subsets and our main results on contractible curves. The fourth section is a bit technical. We introduce the notion of numerical $``$-factoriality and look for situations in which our criteria are applicable. The fifth section contains some sufficient conditions for contractibility. In the last section we prove that the homogeneous spectra $`P(X,D)`$ are always schemes of finite type.
I would like to thank the referee for pointing out an error in a preliminary version and several useful suggestions.
## 2. The cone of curves
This section is preparatory in nature. We will set up the notation and discuss the cone of curves for proper normal surfaces. Throughout the paper, we will work over an arbitrary base field $`K`$. The word surface always refers to a 2-dimensional, irreducible, separated $`K`$-scheme of finite type. In this section, $`X`$ will always be a proper normal surface.
###### (2.1)
Let $`Z^1(X)`$ be the group of Weil divisors. If $`X`$ is a proper regular surface, the group of Weil divisors comes along with the $``$-valued intersection pairing $`A,BAB`$. If $`X`$ is just a proper normal surface, Mumford pointed out that there is still a $``$-valued intersection form on $`Z^1(X)`$ with the usual properties. Mumford’s approach uses resolutions of singularities for surfaces, which has been established in full generality by Lipman .
###### (2.2)
We will call a curve $`RX`$ negative definite if the intersection matrix $`\mathrm{\Phi }=(R_iR_j)`$ for the irreducible components $`R_iR`$ is negative definite. For example, if $`f:XY`$ is a birational morphism of proper normal surfaces, the contracted curve $`RX`$ is negative definite, see for example , X, 1.9. Since $`R_iR_j0`$ holds for $`ij`$, it is a well known fact from linear algebra that the inverse matrix $`\mathrm{\Phi }^1`$ has negative entries. Moreover, if $`R`$ is connected, all entries are strictly negative. We will use this fact several times.
###### (2.3)
A Weil divisor $`AZ^1(X)`$ which lies in the radical of the rational intersection pairing is called numerically trivial. Let us write $`N(X)`$ for the quotient of $`Z^1(X)`$ modulo the radical of Mumford’s rational intersection pairing. We obtain a non-degenerate intersection pairing
$$N(X)\times N(X).$$
It is well-known that $`N(X)`$ is a free $``$-module of finite type. The Hodge index theorem tells us that this intersection form has exactly one positive eigenvalue. In Mori theory, one traditionally works with the pairing $`N^1(X)\times N_1(X)`$ induced by $`\mathrm{Div}(X)\times Z^1(X)`$. But $`\mathrm{Pic}(X)=0`$ might easily happen if $`X`$ is non-projective , in which case this pairing has lost its significance.
###### (2.4)
We call a subset of a real vector space a cone if it is closed under multiplication with positive scalars and addition. The real vector space $`N(X,)=N(X)`$ contains two important cones, which we should recall.
The closed cone $`\overline{\mathrm{NE}}(X)`$ generated by all curves $`CX`$ is called the pseudo-effective cone.
A Weil divisor $`A`$ meeting the condition of the Nakai criterion for ampleness, which means $`A^2>0`$ and $`AC>0`$ for all curves $`CX`$, will be called ample. The pseudo-ample cone $`\overline{\mathrm{NA}}(X)`$ is the closed cone generated by the ample cycles. Note that an ample Weil divisor does not define a closed embedding into some $`^n`$ unless it is a $``$-Cartier divisor.
It follows from , XIII, 7.1 that $`\overline{\mathrm{NE}}(X)`$ is polar to $`\overline{\mathrm{NA}}(X)`$.
Since the pseudo-effective cone is closed and contains no lines, it is generated by its extremal rays. One calls a closed subcone $`P\overline{\mathrm{NE}}(X)`$ extremal if $`e+e^{}P`$ for $`e,e^{}\overline{\mathrm{NE}}(X)`$ implies $`e,e^{}P`$. The extremal closed subcones have the following geometric significance:
###### (2.5) Proposition.
The extremal closed subcones $`P\overline{\mathrm{NE}}(X)`$ are precisely the following subsets:
1. Cones of the form $`P=_+R_i`$, generated by the irreducible components $`R_i`$ of a unique reduced negative definite curve $`RX`$.
2. Cones of the form $`P=\overline{\mathrm{NE}}(X)F^{}`$ for certain real pseudo-effective class $`F\overline{\mathrm{NA}}(X)`$ with $`F^2=0`$.
Moreover, the cones $`P=_+R_i`$ of the first sort satisfy the following finiteness condition: If $`QN(X,)`$ is the closed cone generated by all irreducible curves $`CX`$ not supported by $`R`$, then $`PQ=0`$ holds.
###### Proof.
An extremal closed subcone $`P\overline{\mathrm{NE}}(X)`$ is of the form $`\overline{\mathrm{NE}}(X)a^{}`$ for some support function $`a\overline{\mathrm{NA}}(X)`$. If $`a^2=0`$ holds, we are in case $`(ii)`$. Assume that $`a^2>0`$ holds. Then the intersection form on the vector subspace $`PP`$ is negative definite. Write $`P=P_i`$ as the convex hull of extremal rays. The argument in , lemma 4.12 for smooth surfaces also applies for normal surfaces, consequently each $`P_i`$ is generated by a unique integral, negative definite curve $`R_iX`$. If there would be a relation $`e_j=_{ij}\lambda _ie_i`$, the coefficients must be positive, and we obtain the contradiction
$$0>R_j^2=\underset{ij}{}\lambda _iR_iR_j0.$$
So the $`R_i`$ are linearly independent, hence finite in number, and $`R=R_i`$ is the negative definite curve whose irreducible components generate $`P`$. Conversely, one easily sees that the irreducible components of a negative definite curve generate an extremal cone.
To verify the finiteness condition, let $`ePQ`$. We have $`e=\lambda _iR_i`$ with positive coefficients and $`e=limC_n`$ for certain real effective 1-cycles $`C_n`$ whose components are not contained in $`R`$. Hence we have
$$0>e^2=lim(eC_n)=lim(\lambda _iR_iC_n)0,$$
contradiction. QED.
###### (2.6) Remark.
It is possible to extend Mumford’s intersection pairing fruitfully to unibranched surfaces. Let $`X`$ be a unibranched surface, which means that the normalization $`X^{}X`$ is a bijective map , EGA 0, 23.2.1. Obviously we have $`Z^1(X)=Z^1(X^{})`$. If $`\eta X`$ is the generic point, let $`d`$ be the length of the Artin ring $`𝒪_{X,\eta }`$. For two Weil divisors $`A,B`$ on $`X`$ the result , EGA IV, 21.10.4 forces us to put
$$AB=d(AB)_X^{},$$
where the right hand side is computed on the normalization. The results of this paper remain true for unibranched surfaces, but we are content with the normal case. Nevertheless, it should be noted that from a conceptual point of view, unibranched surfaces form a better category, since this category is stable under extensions of the ground field $`K`$, which is not true for normal surfaces.
## 3. Characterization of contractible curves
This section contains our main results on contractible curves. Throughout, $`X`$ will be a proper normal surface. We first make the trivial observation that it suffices to treat the case of connected curves:
###### (3.1) Lemma.
A curve $`RX`$ is contractible if and only if all its connected components $`R_iX`$ are contractible.
###### Proof.
Use patching. QED.
For the purpose of this paper, the following terminology will be useful. We call an open subset $`UX`$ almost affine if the affine hull $`U^{\text{aff}}=\mathrm{Spec}\mathrm{\Gamma }(U,𝒪_X)`$ is of finite type over the ground field $`K`$ and the canonical morphism $`UU^{\text{aff}}`$ is proper and birational. We remark that these are precisely the semi-affine open subsets of Goodman and Landman with a 2-dimensional ring of global sections $`\mathrm{\Gamma }(U,𝒪_X)`$. There is a kind of dual notion for curves: We say that a curve $`AX`$ is ample on itself if $`AA_i>0`$ holds for all irreducible components $`A_iA`$. These concepts are related in the following way:
###### (3.2) Proposition.
Let $`UX`$ be an open subset, $`C=XU`$ its complement, and $`C_iC`$ the integral components. Then the following conditions are equivalent:
1. The open subset $`UX`$ is almost affine.
2. The closed subset $`CX`$ is a connected curve and some Weil divisor $`A`$ supported by $`C`$ has $`A^2>0`$.
3. There is a curve $`AX`$ with $`\mathrm{Supp}(A)=C`$ which is ample on itself.
###### Proof.
We verify the implications $`(a)(b)(c)(a)`$. Assume that $`U`$ is almost affine. Passing to the contraction $`f:XY`$ defined by $`UU^{\text{aff}}`$, we can assume that $`U`$ is affine. Then $`𝒪_U`$ is ample, and we easily find a section $`s\mathrm{\Gamma }(U,𝒪_X)`$ defining a Cartier divisor $`D_UU`$ whose closure $`\overline{D_U}X`$ intersects each component $`C_i`$. According to , II, 2.2.6, each $`C_i`$ is a curve. Multiplying $`s`$ by a suitable power of $`t\mathrm{\Gamma }(U,𝒪_X)`$ defined by the canonical homomorphism $`𝒪_X𝒪_X(C)`$ we can assume that the domain of definition $`\mathrm{dom}(s)X`$ equals $`U`$. Let $`D=\mathrm{div}(s)`$ be the corresponding principal divisor on $`X`$ with $`DU=D_U`$ and write $`\mathrm{cyc}(D)=D_1D_2`$ with $`D_1=\overline{D_U}`$; then $`D_2C_i=D_1C_i>0`$ holds for each component $`C_i`$. Since $`\mathrm{dom}(s)=U`$, the Weil divisor $`D_2`$ is effective with $`\mathrm{Supp}(D_2)=C`$. Suppose $`C^{}C`$ is a connected component. Let $`D_2^{}D_2`$ be the corresponding connected component. Then $`(D_2^{})^2=D_2^{}D_2=D_2^{}D_1>0`$. Moreover, the curve $`C`$ is be connected by the Hodge index theorem. Hence $`A=D_2`$ satisfies condition $`(b)`$.
Now assume that $`(b)`$ holds. Decompose $`A=A_+A_{}`$ into positive and negative part. Then $`0<A^2=A_+^22A_+A_{}+A_{}^2A_+^2+A_{}^2`$. So there is a curve $`A_0X`$ supported by $`C`$ with $`A_0^2>0`$. Now we rename the irreducible components of $`C`$ and find a complete list $`C_1,\mathrm{},C_n`$, possibly with repetitions, with $`A_0C_1>0`$ and $`C_iC_{i+1}`$ nonempty. Inductively we define curves $`A_jX`$ with support $`A_0C_1\mathrm{}C_j`$ and $`A_jC_i>0`$ for $`1ij`$ as follows: If $`A_j`$ is already defined, then $`A_{j+1}=\lambda _jA_j+C_{j+1}`$ meets our conditions for $`\lambda _j>0`$ sufficiently large. Now $`A=A_n`$ is the desired curve which is ample on itself.
Finally, assume that $`(c)`$ holds. Let $`g:X^{}X`$ be a resolution of all the singularities $`x\mathrm{Sing}(X)C`$. Then $`g^{}(A)`$ is an effective $``$-Cartier divisor with support $`A^{}=g^1(A)`$. Let $`E^{}\mathrm{Div}(X^{})`$ be a relatively ample divisor whose support is contracted by $`g`$. Then $`A^{}=g^{}(nA)+E^{}`$ is effective with support $`A^{}`$, provided $`n>0`$ is sufficiently large, and ample on itself. Hence it suffices to treat the case that $`A`$ is a Cartier divisor. The associated invertible sheaf $`=𝒪_X(A)`$ is ample on $`A`$. According to , theorem 1.10, a multiple $`^n`$ is globally generated, and the homogeneous spectrum $`Y`$ of $`\mathrm{\Gamma }(X,\mathrm{Sym})`$ yields a birational contraction $`f:XY`$. We infer that $`U=f^1(V)`$ is the preimage of an affine open subset $`VY`$, hence is almost affine. QED.
From this we easily derive a geometric characterization of contractible curves in terms of complementary curves ample on themselves:
###### (3.3) Theorem.
A connected negative definite curve $`RX`$ is contractible if and only if there is a curve $`AX`$ disjoint to $`R`$ with $`AC>0`$ for every curve $`CX`$ not supported by $`R`$.
###### Proof.
Assume there is a contraction $`f:XY`$ mapping $`R`$ to a closed point $`yY`$. Then the complement $`YV`$ of an affine open neighborhood $`VY`$ can be regarded as a curve on $`X`$. According to (3.2) it is the support of a curve $`AX`$ ample on itself with the desired property. Conversely, if there is such a curve $`AX`$, the open subset $`U=XA`$ is almost affine and defines the contraction $`f:XY`$ of $`R`$. QED.
We have seen that one can always find a possibly non-effective Weil divisor $`A`$ satisfying the above condition, and the real problem is to choose an effective one. Our main result has the advantage that it does not involve any conditions of effectivity:
###### (3.4) Theorem.
A connected negative definite curve $`RX`$ is contractible if and only if there is a Weil divisor $`AZ^1(X)`$ satisfying the following three conditions:
1. The Weil divisor $`A`$ is Cartier near $`RX`$.
2. We have $`AC0`$ for all curves $`CX`$, with equality if and only if $`CR`$ holds.
3. For every integer $`m>0`$ there is an integer $`n>0`$ and a numerically trivial Weil divisor $`NZ^1(X)`$ which is Cartier near $`RX`$ such that the linear class of $`nA+N`$ is trivial on $`mR`$.
###### Proof.
According to (3.3) the conditions are necessary, and the real issue is to show sufficiency.
We first get rid of the singularities. Let $`g:X^{}X`$ be a resolution of all singularities $`x\mathrm{Sing}(X)R`$. According to , EGA II, 8.11.1 we have to show that $`R^{}=g^1(R)`$ is contractible. A straightforward argument shows that $`A^{}=g^{}(A)`$ satisfies the three conditions on $`X^{}`$, hence we can assume that $`R\mathrm{Reg}(X)`$ holds. Now let $`g:X^{}X`$ be a resolution of all remaining singularities. According to (3.1) we have to show that $`R`$ viewed as a curve $`R^{}X^{}`$ is contractible. Let $`P^{}N(X^{},)`$ be the cone generated by the irreducible components of $`R^{}`$ and $`Q^{}N(X^{},)`$ the closed cone generated by all other integral curves. Choose a relatively ample exceptional divisor $`D^{}Z^1(X^{})`$ contracted by $`g`$. If $`U^{}N(X^{},)`$ is a compact neighborhood of zero, the linear form associated to $`D^{}`$ is bounded on $`U^{}Q^{}`$. Since $`P^{}Q^{}=0`$ holds by (2.5), the $``$-divisor $`A^{}=D^{}+g^{}(nA)`$ is strictly positive on the punctured neighborhood $`U^{}Q^{}\left\{0\right\}`$ provided $`n`$ is sufficiently large. Consequently, the numerical class of $`A^{}`$ is a support function of the pseudo-effective cone with respect to the extremal subcone $`P^{}`$, and condition $`(i)`$ holds true on $`X^{}`$. The other conditions trivially remain unaffected, hence we can assume that $`X`$ is a regular surface.
Obviously $`A^2>0`$ holds. Since $`R`$ is negative definite, there exists a divisor $`DZ^1(X)`$ supported by $`R`$ which is anti-ample on $`R`$; according to (2.2), the divisor is effective and its support equals $`R`$. We have
$$(tAD)^2=t^2A^22tAD+D^2>0$$
for all $`t>0`$ sufficiently large. Arguing as above we also see that $`(tAD)C>0`$ holds for all curves $`CY`$, provided $`t>0`$ is sufficiently large. We now replace $`A`$ by a suitable multiple and assume, using the Nakai criterion, that $`AD`$ is ample.
Set $`=𝒪_X(A)`$ and $`=𝒪_X(D)`$; thus $``$ is pseudo-ample and $``$ is ample. According to Fujita’s vanishing result , theorem 5.1, there is a natural number $`t_0>0`$ such that
$$H^1(X,^s()^t𝒩)=0$$
holds for all integers $`s0`$, $`tt_0`$ and all numerically trivial invertible sheaves $`𝒩`$. If we replace $``$ and $``$ by $`^{t_0}`$ and $`^{t_0}`$ we can assume that
$$H^1(X,^s𝒩)=0$$
holds for all integers $`s>0`$ and all numerically trivial invertible sheaves $`𝒩`$. Thus the right hand term in the exact sequence
$$H^0(X,^s𝒩)H^0(D,^s𝒩D)H^1(X,^s𝒩)$$
is zero. Now choose an integer $`m`$ with $`DmR`$. According to condition $`(c)`$, we can pick a numerically trivial invertible sheaf $`𝒩`$ and an integer $`n>0`$ such that the restriction $`^n𝒩D`$ is trivial. Let us replace $``$ by $`^n𝒩`$. Consider the open subset $`UX`$ on which $``$ is globally generated; by construction, this open set contains the generic points of $`R`$, hence $``$ is ample on the complement $`XU`$. According to , theorem 1.10, some multiple $`^t`$ is globally generated, and the homogeneous spectrum
$$Y=\mathrm{Proj}(\mathrm{\Gamma }(X,\mathrm{Sym}))$$
is a projective normal surface yielding the desired contraction $`f:XY`$ of $`R`$. QED.
## 4. Improvement of cycles
How can one ensure that the conditions of theorem (3.4) are fulfilled? It is easy to find a Weil divisor $`AZ^1(X)`$ satisfying the condition (ii) of (3.4). In this section we discuss the possibilities to improve the Weil divisor $`A`$ in such a way that it also satisfies the other two more delicate conditions (i) and (iii). For this purpose the following notion is useful:
###### (4.1) Definition.
Let $`S`$ be an arbitrary subset of a proper normal surface $`X`$. We say that $`X`$ is numerically $``$-factorial with respect to $`S`$ if each Weil divisor $`D`$ is numerically equivalent to a $``$-Weil divisor which is $``$-Cartier near $`S`$.
Of course, only the points $`xS`$ whose local rings $`𝒪_{X,x}`$ are not $``$-factorial are relevant for this notion. If the conditions holds for $`S=X`$ we call the surface $`X`$ numerically $``$-factorial. According to the Nakai criterion, such a surface is projective; more precisely, the canonical map $`\mathrm{Pic}(X)N(X)`$ has finite cokernel.
We have the following behaviour under birational morphisms:
###### (4.2) Proposition.
Let $`f:XY`$ be a birational morphism of proper normal surfaces, $`RX`$ the contracted curve, $`𝔛X`$ the corresponding formal completion, $`SX`$ a subset containing $`R`$, and $`TY`$ its image. If the cokernel of $`\mathrm{Pic}^0(X)\mathrm{Pic}^0(𝔛)`$ is a torsion group and if $`X`$ is numerically $``$-factorial with respect to $`S`$, then $`Y`$ is numerically $``$-factorial with respect to $`T`$.
###### Proof.
Given a cycle $`DZ^1(Y)`$; then there is an integer $`n>0`$ and a numerically trivial Weil divisor $`N`$ on $`X`$ such that $`f^{}(nD)+N`$ is Cartier near $`S`$. Passing to a multiple and adding a numerically trivial Cartier divisor, we can assume that the corresponding reflexive $`𝒪_X`$-module $``$ is trivial on $`𝔛`$. Hence $`=f_{}()`$ is invertible at $`T`$ and is represented by $`nD+f_{}(N)`$. QED.
Let $`nRX`$ be the infinitesimal neighborhoods of the contracted curve $`R`$. Then the inverse system of groups $`nH^1(R,𝒪_{nR})`$ is eventually constant, and we obtain a well-defined group scheme of finite type $`\mathrm{Pic}_{𝔛/K}^0=\mathrm{Pic}_{nR/K}^0`$, provided $`n`$ is sufficiently large. We can use the group scheme $`G`$ defined by the exact sequence
$$\mathrm{Pic}_{X/K}^0\mathrm{Pic}_{𝔛/K}^0G0$$
to check the hypothesis of the previous result, using the following observation:
###### (4.3) Lemma.
Let $`RS`$ be a closed subscheme of a proper $`K`$-scheme $`S`$. If $`H^1(S,𝒪_S)H^1(R,𝒪_R)`$ is surjective, or if the base field $`K`$ is of characteristic $`p>0`$ and the cokernel $`G`$ of the homomorphism $`\mathrm{Pic}_{S/K}^0\mathrm{Pic}_{R/K}^0`$ is a unipotent group scheme, then the cokernel of $`\mathrm{Pic}^0(S)\mathrm{Pic}^0(R)`$ is a torsion group.
###### Proof.
Let $``$ be an numerically trivial invertible $`𝒪_R`$-module and $`l\mathrm{Pic}_{R/K}^0`$ the corresponding rational point. The map $`H^1(S,𝒪_S)H^1(R,𝒪_R)`$ is the tangential map for the homomorphism $`\mathrm{Pic}_{S/K}^0\mathrm{Pic}_{R/K}^0`$; if the first condition holds, this map is surjective, and we find a closed point $`m\mathrm{Pic}_{S/K}^0`$ mapping to $`l`$. Assume that the second condition holds. From the definition of unipotent group schemes , SGA 3, p. 534, it follows immediately that $`G(K)`$ is a torsion group. Replacing $``$ by a multiple we also find a closed point $`m\mathrm{Pic}_{S/K}^0`$ mapping to $`l`$.
If the base field is algebraically closed, the point $`m`$ is represented by a numerically trivial invertible $`𝒪_S`$-module $``$ restricting to $``$. In general, we find a finite field extension $`KK^{}`$ and an invertible sheaf $`^{}`$ on $`X^{}=XK^{}`$ restricting to $`^{}=K^{}`$ on $`R^{}=RK^{}`$. Since $`p_{}(𝒪_S^{})`$ is a locally free $`𝒪_S`$-module, say of rank $`n>0`$, there is a commutative diagram of norm homomorphism , EGA II, 6.6.8
$$\begin{array}{ccc}\mathrm{Pic}(S^{})& & \mathrm{Pic}(R^{})\\ N& & N& & \\ \mathrm{Pic}(S)& & \mathrm{Pic}(R).\end{array}$$
Now $`^n=N(^{})`$ extends to $`=N(^{})`$. It remains to check that $``$ is numerically trivial. Using base change, we can assume that $`S`$ is an integral curve. Then $`^{}`$ is represented by a Cartier divisor $`D^{}`$ whose cycle $`n_x^{}x^{}`$ satisfies $`n_x^{}dim\kappa (x^{})=0`$. Obviously, the cycle $`\mathrm{cyc}(f_{}(D))`$ is of degree zero. QED.
###### (4.4) Corollary.
Let $`Y`$ be a proper normal surface over a finite ground field $`K`$ or with $`H^2(Y,𝒪_Y)=0`$. Then $`Y`$ is numerically $``$-factorial, and in particular projective.
###### Proof.
Let $`XY`$ be a resolution of singularities, and $`𝔛X`$ the formal completion of the contracted curve $`RX`$. If $`H^2(Y,𝒪_Y)=0`$ holds, then the restriction $`H^1(X,𝒪_X)H^1(𝔛,𝒪_𝔛)`$ is surjective, and the claim follows from the above lemma and (4.2). If the ground field $`K`$ is finite, then $`\mathrm{Pic}_{𝔛/K}^0`$ contains only finitely many rational points, and $`\mathrm{Pic}^0(𝔛)`$ must be a finite group. Consequently, $`Y`$ is even $``$-factorial. QED.
Now we turn to the following situation: Let $`RX`$ be a curve on a proper normal surface $`X`$, and $`AZ^1(X)`$ a Weil divisor which is already Cartier near $`R`$ and numerically trivial on $`R`$. We would like to know whether we can make the linear class of $`A`$ trivial on $`R`$. Here the canonical class $`K_X`$ becomes useful:
###### (4.5) Proposition.
Let $`AZ^1(X)`$ be Cartier near $`R`$ and numerically trivial on $`R`$. If the curve $`mRX`$ is contained in the base scheme of $`K_X+mR`$, then there is an integer $`n>0`$ and a numerically trivial Cartier divisor $`N`$ such that the linear class of $`nA+N`$ is trivial on the infinitesimal neighborhood $`mR`$.
###### Proof.
We have an exact sequence
$$H^1(X,𝒪_X)H^1(R,𝒪_{mR})H^2(X,𝒪_X(mR))H^2(X,𝒪_X),$$
and the map on the right is dual to the canonical map $`H^0(X,𝒪_X(K_X))H^0(X,𝒪_X(K_X+mR))`$. Recall that the base scheme of $`K_X+mR`$ is the intersection of all effective curves $`CX`$ linearly equivalent to this class, hence the map is surjective. We conclude that $`H^1(X,𝒪_X)H^1(R,𝒪_{mR})`$ is surjective, and the claim follows from (4.3). QED.
In positive characteristics the infinitesimal neighborhoods are irrelevant for our problem:
###### (4.6) Proposition.
Let $`AZ^1(X)`$ be Cartier near $`R`$ and trivial on $`R`$. Assume that the base field $`K`$ is of characteristic $`p>0`$. Then the linear class of $`p^mA`$ is trivial on $`mR`$ for all integers $`m>0`$.
###### Proof.
We make induction on the integer $`m`$. Set $`_m=𝒪_X(mR)`$. The exact sequence
$$0_m/_{m+1}𝒪_{(m+1)R}^\times 𝒪_{mR}^\times 1$$
yields an exact sequence
$$H^1(R,_m/_{m+1})\mathrm{Pic}((m+1)R)\mathrm{Pic}(mR)0.$$
By induction, $`p^mA`$ is trivial on $`mR`$. Hence the class of $`p^mA`$ on $`(m+1)R`$ lies in the image of $`H^1(R,_m/_{m+1})`$. Since this group is annihilated by $`p`$, we conclude that $`p^{m+1}A`$ is trivial on $`(m+1)R`$. QED.
## 5. Applications
In this section we will collect some sufficient conditions for contractibility and discuss an example. We assume that $`X`$ is a proper normal surface, $`RX`$ is a connected negative definite curve, and denote by $`𝔛X`$ the corresponding formal completion. Our most general criterion is the following:
###### (5.1) Theorem.
Assume that $`X`$ is numerically $``$-factorial with respect to $`R`$. If the curve $`mRX`$ is contained in the base scheme of the class $`K_X+mR`$ for all integers $`m>0`$, or if the base field $`K`$ is of characteristic $`p>0`$ and the cokernel of $`\mathrm{Pic}_{X/K}^0\mathrm{Pic}_{R/K}^0`$ is a unipotent group scheme, then $`R`$ is contractible.
###### Proof.
Let $`A`$ be a Weil divisor which is a support function of the pseudo-effective cone $`\overline{\mathrm{NE}}(X)`$ with respect to the extremal subcone $`P=_+R_i`$ generated by the irreducible components $`R_iR`$. Since $`X`$ is numerically $``$-factorial with respect to $`R`$, we can assume that $`A`$ is Cartier near $`R`$. Given an integer $`m>0`$. If the first condition holds, we invoke (4.5) and find an integer $`n>0`$ and a numerically trivial divisor $`N\mathrm{Div}^0(X)`$ such that the linear class of $`nA+N`$ is trivial on $`mR`$. If the second conditions holds, we first find such $`n`$ and $`N`$ such that $`nA+N`$ is trivial on $`R`$. Hence $`p^m(nA+N)`$ is trivial on $`mR`$ by (4.6). Now the claim follows from theorem (3.4). QED.
The following will be useful for the classification of proper normal surfaces of Kodaira dimension $`\kappa (X)=\mathrm{}`$:
###### (5.2) Corollary.
If the linear class $`K_X+mR`$ is not effective for all integers $`m>0`$, or if the base field $`K`$ is of characteristic $`p>0`$ and $`K_X+R`$ is not effective, then $`R`$ is contractible.
###### Proof.
Since $`K_X+R`$ is not effective, the same holds for $`K_X`$, and $`H^2(X,𝒪_X)`$ must vanish. According to (4.4), the surface $`X`$ is numerically $``$-factorial. If the first condition holds, the base scheme of $`K_X+mR`$ is the whole surface $`X`$, thus contains $`mR`$. If the second condition holds, the map $`H^1(X,𝒪_X)H^1(R,𝒪_R)`$ is surjective, hence $`\mathrm{Pic}_{X/K}^0\mathrm{Pic}_{R/K}^0`$ is an epimorphism. The claim now follows from (5.1). QED.
We can generalize the criterion of Castelnuovo-Enriques from regular to normal proper surfaces:
###### (5.3) Corollary.
Assume that $`X`$ is numerically $``$-factorial with respect to $`R`$. If $`RX`$ is irreducible with $`RK_X0`$, then $`R`$ is contractible.
###### Proof.
Assume that $`mR`$ is not in the fixed scheme of $`K_X+mR`$ for some integer $`m>0`$. Choosing $`m`$ minimal, we can represent $`K_X+mR`$ by a curve $`CX`$ not containing $`R`$. But
$$0K_XR=CRmR^2>0$$
gives a contradiction. According to (5.1), $`R`$ is contractible. QED.
###### (5.4) Remark.
For log-terminal surfaces $`X`$ and $`K_XR<0`$, this becomes a special case of the contraction theorem from Mori theory (, theorem 3.2.1).
The following is already contained in , theorem 2.9:
###### (5.5) Proposition.
Assume that the $`K`$-surface $`X`$ is already defined over a subfield $`K^{}K`$, such that $`K^{}`$ is a finite field and that $`K^{}K`$ is the composition of purely transcendental and radical extensions. Then every negative definite curve $`RX`$ is contractible.
###### Proof.
Let $`X^{}`$ be the normal surface over $`K^{}`$ with $`X=X^{}_K^{}K`$. According to , X, 7.17.4, the mapping $`Z^1(X^{},)Z^1(X,)`$ is surjective up to linear equivalence. Hence there is a negative definite curve $`R^{}X^{}`$ with $`R=R^{}_K^{}K`$, and we can assume that $`K`$ is finite. Then $`X`$ is $``$-factorial, and $`\mathrm{Pic}^0(mR)\mathrm{Pic}_{mR/K}^0(K)`$ are finite groups. According to (3.4), the curve $`R`$ is contractible. QED.
###### (5.6) Example.
The case of geometrically ruled surfaces is already instructive. Let $`C`$ be a normal proper connected curve and $`p:XC`$ a geometrically ruled surface. Each section $`RX`$ determines an extension
$$0𝒪_C0$$
with $`=p_{}(𝒪_X(R))`$ and $`=p_{}(𝒪_R(R))`$. Hence we have $`X=()`$ and $`R=()`$, and $`𝒪_R(R)p^{}()R`$ holds. Assume there is a section with $`R^2<0`$. Let $`A`$ be a divisor representing $`p^{}()𝒪_X(R)`$. Then $`A`$ is trivial on $`R`$, and in characteristic $`p>0`$ we deduce from (3.4) and (4.6) that $`R`$ is contractible. However, the situation is more complicated in characteristic zero. We have
$$\mathrm{Pic}(2R)=\mathrm{Pic}(R)H^1(C,^{});$$
decomposing the class of $`A`$ into $`(0,\alpha )`$, one can show that $`\alpha `$ corresponds to the Yoneda class in $`\mathrm{Ext}^1(,𝒪_X)`$ of the extension $``$. In characteristic zero, we conclude that $`R`$ is contractible if and only if the extension $``$ is split. In this case a splitting $`𝒪_C`$ yields another section $`AX`$ disjoint to $`R`$ with $`A^2>0`$, defining the contraction.
## 6. A finiteness result for models
In this section we show that the model of a normal surface $`X`$ defined by an arbitrary Weil divisor $`D`$ is a scheme of finite type. This generalizes results of Zariski , Proposition 11.5, Fujita , p. 235, and Russo .
###### (6.1)
Let $`X`$ be a proper normal surface and $`DZ^1(X)`$ a Weil divisor. This yields a graded $`K`$-algebra
$$R(X,D)=_{n0}H^0(X,nD),$$
which in turn defines a homogeneous spectrum
$$P(X,D)=\mathrm{Proj}(R(X,D)).$$
According to , EGA II, 3.7.4 the open subset $`UX`$ of all points $`xX`$ such that there is a homogeneous $`sR_+(X,D)`$ with $`s(x)0`$ is the largest open subset on which the canonical map $`𝒪_XR(X,D)_{n0}𝒪_X(nD)`$ defines a morphism $`r:UP(X,D)`$. We call the scheme $`P(X,D)`$ the $`D`$-model of the surface $`X`$.
Using , EGA I, 6.8.2, we see that for each homogeneous element $`sR_+(X,D)`$ the open subset $`D_+(s)P(X,D)`$ equals the affine hull $`X_s^{\text{aff}}=\mathrm{Spec}\mathrm{\Gamma }(X_s,𝒪_X)`$, hence if $`X_{s_1}\mathrm{}X_{s_n}`$ is a covering of $`U`$, then $`D_+(s_1)\mathrm{}D_+(s_n)`$ is a covering of $`r(U)P(X,D)`$. In this section we will prove the following
###### (6.2) Theorem.
Let $`X`$ be a proper normal surface, and $`DZ^1(X)`$ an arbitrary Weil divisor. Then the model $`P(X,D)`$ is a separated normal $`K`$-scheme of finite type of dimension $`2`$.
It is well known that the algebra $`R(X,D)`$ might fail to be finitely generated, see for example , p. 562. We will explain the geometric reason for this below. First, we record the following
###### (6.3) Corollary.
Let $`U`$ be a normal surface, not necessarily proper. Then $`\mathrm{\Gamma }(U,𝒪_U)`$ is an integrally closed $`K`$-algebra of finite type and of dimension $`2`$.
###### Proof.
By the Nagata compactification theorem , we can find a proper scheme $`X`$ containing $`U`$ as an open subset. Making a blow-up and a normalization we can assume that $`X`$ is a proper normal surface and that $`D=XU`$ is a Cartier divisor. Now $`U^{\text{aff}}`$ is an affine open subset of $`P(X,D)`$, and the claim follows from (6.2). QED.
###### (6.4)
Proof of theorem (6.2). We start with some preliminary reductions. The $`D`$-model does not change if we replace $`D`$ by a positive multiple. If $`H^0(X,nD)=0`$ holds for all integers $`n>0`$, then $`P(X,D)`$ is empty, hence it suffices to treat the case that $`D`$ is effective. For each $`n>0`$ let $`F_nX`$ be the fixed curve of $`nD`$. The effective Weil divisor $`M_n=nDF_n`$ has no fixed curve. Following Kawamata , definition 1.1, we call such Weil divisors *movable*, and refer to $`M_n`$ as the movable part of $`nD`$. The canonical map $`F_n:H^0(X,M_n)H^0(X,nD)`$ is bijective, and $`M_nC0`$ holds for all curves $`CX`$. Let $`B_nX`$ be the base scheme of the class $`nD`$, that is the scheme-theoretical intersection of all curves $`CX`$ representing the class $`nD`$. Clearly, $`B_n`$ contains $`F_n`$ and all closed points $`xX`$ where $`nD`$ is not Cartier. Passing to a multiple of $`D`$, we can assume that the supports of $`B_n`$ and $`F_n`$ are independent of $`n`$.
If $`M_n=0`$ hold for all integer $`n>0`$, we have $`P(X,D)=X^{\text{aff}}`$, hence we can assume that $`D`$ has a non-zero movable part.
Now let $`F_n^{}X`$ be the union of all connected components $`CF_n`$ with $`CM_n>0`$, and let $`F_n^{\prime \prime }=F_nF_n^{}`$ be the union of the remaining connected components. Since $`M_n`$ is movable, the condition $`M_nF_n=0`$ is equivalent to $`F_n^{}=0`$. For simplicity, we set $`M=M_1`$ and $`F=F_1`$. The proof of (6.2) will be completed by the next two propositions:
###### (6.5) Proposition.
Under the above assumptions, the scheme $`P(X,D)`$ is a normal projective curve if and only if $`M^2=0`$ and $`MF=0`$ holds.
###### Proof.
Assume that $`M^2=0`$ holds. Then for each curve $`\lambda _iC_iX`$ representing $`M`$ we must have $`MC_i=0`$. Since $`M`$ is movable, it can also be represented by a curve $`CX`$ disjoint to $`C_i`$, and we deduce that $`M`$ must be a globally generated Cartier divisor. Thus $`Y=P(X,M)`$ is a proper curve, defining a fibration $`f:XY`$, and $`M`$ is the preimage of an ample class on $`Y`$.
Now assume that $`MF=0`$ also holds; since $`M`$ is movable, $`F^{}`$ must vanish. Thus $`F=F^{\prime \prime }`$ is supported by certain fibres $`X_y`$ of $`f:XY`$. Assume that $`\mathrm{Supp}(F_y)=\mathrm{Supp}(X_y)`$ holds for some point $`yY`$. Decompose $`X_y=\lambda _iE_i`$ and $`F_y=\mu _iE_i`$ into prime cycles. Rearranging the terms, we can assume that $`\lambda _1/\mu _1\lambda _i/\mu _i`$ holds for all indices $`i`$. We have
$$\lambda _1F_y=\mu _1X_y+\underset{i>1}{}(\lambda _1\mu _i\mu _1\lambda _i)E_i,$$
where the first summand $`\mu _1X_y`$ is movable, whereas the second summand $`_{i>1}(\lambda _1\mu _i\mu _1\lambda _i)E_i`$ is effective. Hence $`E_1`$ is not contained in the base locus of $`\lambda _1D`$, but by our assumption (6.4), the supports of $`F_n`$ are constant, contradiction. So $`\mathrm{Supp}(F)`$ contains no fibre of $`f:XY`$. Consider the open subset $`V=XF`$. I claim that the canonical map $`𝒪_Yf_{}(𝒪_V)`$ is bijective. This is local in $`Y`$; passing to the henselization $`𝒪_{Y,y}𝒪_{Y,y}^{}`$, we can contract the components $`F_yX_y`$, according to , Proposition 4, p. 169. By Serre’s condition $`(S_2)`$ we have $`f_{}(𝒪_V)𝒪_{Y,y}^{}=𝒪_{Y,y}^{}`$. Consequently, the open subsets $`D_+(s)=X_s^{\text{aff}}`$ form a covering of $`Y`$ when $`s`$ ranges over the homogeneous elements of $`R_+(X,D)`$, and we infer $`P(X,D)=Y`$.
Conversely, assume that $`Y=P(X,D)`$ is a projective curve, and consider the morphism $`r:UY`$, where $`UX`$ is the maximal open subset as in (6.1). If $`M^2>0`$ holds, then $`MF^{}`$ supports a curve ample on itself, and its complement is almost affine (3.2); thus $`P(X,D)`$ would contain a 2-dimensional open subset, contradiction. So $`M^2=0`$ holds, and we obtain a fibration $`f:XY`$ extending $`r`$. If $`F^{}0`$ holds, the generic fibre $`X_\eta `$ would be affine, contradicting , EGA I, 9.3.4. Hence we have $`MF=0`$. QED.
###### (6.6) Proposition.
With the assumptions in (6.4), the model $`P(X,D)`$ is a normal surface if and only if either $`M^2>0`$, or $`M^2=0`$ and $`MF0`$ holds.
###### Proof.
The condition is sufficient: Let $`CX`$ be a curve representing the class $`M+F^{}`$ and $`V=XC`$. In case $`M^2>0`$ the curve $`C`$ supports a curve ample on itself. If $`M^2=0`$ holds, $`Y=P(X,M)`$ is a proper curve, and no connected component of $`F^{}`$ is vertical with respect to the corresponding fibration $`f:XY`$. Again $`C`$ supports a curve ample on itself. In both cases $`V`$ is almost affine and $`F^{\prime \prime }V`$ is contracted in $`V^{\text{aff}}`$, thus $`V^{\text{aff}}`$ is an open subset of $`P(X,D)`$ which is 2-dimensional and of finite type over $`K`$. According to (6.5), we have $`M_n^2>0`$ or $`M_n^2=0`$ and $`M_nF_n0`$ for all $`n>0`$, consequently $`P(X,D)`$ is a 2-dimensional scheme of finite type.
Conversely, the condition is necessary by (6.5). QED.
If the scheme $`P(X,D)`$ is not proper, the algebra $`R(X,D)`$ is not finitely generated. Concerning this, we have the following characterization:
###### (6.7) Proposition.
Assume that the model $`P(X,D)`$ is a surface. With the assumptions in (6.4), the surface $`P(X,D)`$ is proper if and only if $`M`$ is $``$-Cartier and $`MF=0`$ holds.
###### Proof.
Assume that $`M`$ is $``$-Cartier and $`MF=0`$ holds. Thus $`F^{}=0`$ and $`M^2>0`$ holds, and $`F^{\prime \prime }X`$ is a negative definite curve. Let $`f:XY`$ be the contraction of the negative definite curve $`RX`$ orthogonal to $`M`$. Since $`M`$ is $``$-Cartier, $`Y=P(X,M)`$ holds. Choose homogeneous $`s_1,\mathrm{},s_nR_+(X,M)`$ with $`X=X_{s_1}\mathrm{}X_{s_n}`$ and $`RX_{s_i}`$. Thus we have $`U_{s_i}^{\text{aff}}=X_{s_i}^{\text{aff}}`$ and deduce $`P(X,D)=P(X,M)`$.
Conversely, assume that $`Y=P(X,D)`$ is a proper surface. Then each curve $`CX`$ representing the class $`M_n+F_n^{}`$ is the support of a curve ample on itself, hence $`V=XC`$ is almost affine, and $`Y`$ is covered by the open subsets $`V^{\text{aff}}`$. Hence we can extend $`r:UY`$ to a proper morphism $`r:VY`$ on the larger open subset $`V=UF^{\prime \prime }`$, thus $`X=V`$ and $`F^{}=0`$ hold, and $`MF=0`$ follows. Assume that $`M`$ is not $``$-Cartier at some point $`xX`$; then $`D`$ is also not $``$-Cartier at $`x`$, and we have $`xUF^{\prime \prime }`$, contradiction. QED.
Mathematisches Institut
Ruhr-Universität
44780 Bochum
Germany
E-mail: s.schroeer@ruhr-uni-bochum.de |
no-problem/9911/cond-mat9911209.html | ar5iv | text | # References
Gap anisotropy in the angle-resolved photoemission spectroscopy of Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+δ</sub>
Tulika Maitra<sup>1</sup><sup>1</sup>1email: tulika@phy.iitkgp.ernet.in and A. Taraphder<sup>∗†</sup><sup>2</sup><sup>2</sup>2email: arghya@phy.iitkgp.ernet.in, arghya@cts.iitkgp.ernet.in
Department of Physics & Meteorology and Centre for Theoretical Studies,
Indian Institute of Technology, Kharagpur 721302 India
## Abstract
The gap anisotropy in Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+δ</sub> is revisited in the framework of a d-wave scenario in view of the recent angle-resolved photoemission experiment. Based on a tight-binding fit to the normal state dispersion, a detail analysis on the effects of the inclusion of the next harmonic in the d-wave has been presented. Significant effect has been observed in the superconducting T<sub>c</sub>. The density of states is linear at the nodes with enhanced weight, caused by a marked increase in the low energy excitaions which affect the thermodynamics considerably. The slope of the $`\rho _sT`$ curve in the low temperature regime increases and the specific heat reflects the enhanced entropy at low temperatures. The leading edge of the ARPES energy distribution curves have been calculated and found to shift towards higher energy. The effect of scattering by non-magnetic impurities in this context are also outlined.
PACS Nos. 74.72-h, 74.20.Fg
Introduction
The nature of the superconducting gap anisotropy in the high temperature superconductors has attracted a lot of attention in the last few years. A considerable progress has been made in the recent past with the resolution and clarity of the angle-resolved photoemission spectroscopy (ARPES) having reached a very high level backed by careful and thorough analysis of the data. A general consensus seems to have emerged that the symmetry of the order parameter (OP) in most of the high temperature superconductors is predominantly d-wave. In a very recent ARPES experiment on the underdoped Bi2212 system Mesot et. al. have reported an observable departure from a simple interpretation in terms of the usual d-wave symmetry. Inclusion of the next higher harmonic in the superconducting pairing function seems to give a better fit to their data. We develop the idea in order to look for other observable effects of such a term, and calculate its effects on various physical quantities in the framework of the usual phenomenological theory in the weak coupling limit and suggest further experiments. We have calculated the changes in superconducting transition temperature, nodal structure on the fermi surface (FS), the density of states (DOS) in the superconducting state and its effect on the specific heat. The energy distribution curves (EDC) (that are seen in the ARPES experiments) are also obtained. Fluctuations are included in one loop level and the superfluid density (and hence the penetration depth) is obtained as a function of temperature. We observe the differences in slopes of $`\rho _s(T)`$ in the low temperature regime with increasing level of mixing of the higher harmonic. We also calculate the effect of both doping and the mixing of higher order terms on the specific heat.
Model and Calculations
In the experiments of Mesot et al, ARPES results have been fitted to a d-wave gap function for several Bi2212 systems. The maximum gap values have been adjusted at each of the doping levels for the best fit, and the EDCs at different angles and the slope of the gap function close to the node have been carefully fitted. Clear deviation from the usual $`d_{x^2y^2}`$ behaviour is observed and attempts were made to fit the data with an admixture of about 4-10% of the next harmonic cos(6$`\varphi `$). A much improved fit was indeed obtained with the inclusion of this higher order term. As one underdopes the system, the normal state resistivity is known to increase and effective screening becomes weaker. Under such a situation it is reasonable to assume that the higher harmonics in the effective interactions have to be brought in, and highly sensitive experiments would reveal the effects due to the growing range of interaction. In the spin fluctuation models too the pairing interaction grows sharply in the momentum space with underdoping and long range effects in real space become increasingly relevant.
We undertake to examine the ramifications arising out of the addition of the higher harmonic with a model band structure that reproduces the observed FS and the location of the van-Hove singularity in Bi2212 accurately. The six parameters describing the hopping on the BiO plane used to obtain the fit were \[0.131, -0.149, 0.041, -0.013, -0.014, 0.013\] (in eV), corresponding to a doping of $`\delta =0.17`$. The van-Hove singularity, due to the saddle points $`𝐤=(\pi ,\mathrm{𝟎}),(\mathrm{𝟎},\pi )`$ is 30meV below the FS, as observed in ARPES.
Superconductivity occurs through attractive long range interactions of which we keep only the near-neighbour part. Combined with the fact that there is a reasonably strong on-site repulsion between the electrons in the cuprates, forcing the pairing function $`\mathrm{\Psi }(r=0)=0`$, such an interaction is known to support d-wave pairing. The vanishing of the on-site part of the pairing function implies that the $`_q\mathrm{\Delta }(𝐪)=\mathrm{𝟎}`$, where the $`𝐪`$-sum runs over the entire Brillouin zone (BZ), thereby contributing equal regions of positive and negative sign to the OP. The usual $`d_{x^2y^2}`$ (or $`cos(2\varphi )`$) OP admits of such constraints. All the higher harmonics, $`cos[(2+4n)\varphi ]`$ are also admissible under the symmetry restrictions of the d-wave. In the underdoped systems, it is likely that higher neighbour interactions contribute increasingly, but in the present analysis we kept only the first of such terms. We also include only the singlet component of the order parameter in our analysis, as there is no reason so far to include the triplet part in the parameter regions that one works with in these system.
The superconducting gap equation that is numerically solved is the usual mean-field factorized gap equation
$$\mathrm{\Delta }_𝐤=\frac{1}{N}\underset{𝐤^{}}{}V(𝐤𝐤^{})\frac{\mathrm{\Delta }_𝐤^{}tanh(\beta E_𝐤^{}/2)}{2E_𝐤^{}},$$
where $`V(𝐤𝐤^{})`$ is written in the separable form
$$V(𝐤𝐤^{})=g\eta (𝐤)\eta (𝐤^{}).$$
Here we have chosen the basis function $`\eta (𝐤)`$ to be the $`B_1`$ representation of the one dimensional irreducible representations of $`C_{4v}`$, $`\eta (𝐤)=\frac{1}{2}(cosk_xcosk_y)`$ (the usual $`d_{x^2y^2}`$ symmetry); $`g`$ measures the strength of the attractive interaction. The triplet channels of pairing have not been considered in the foregoing. The value of $`g`$ was chosen (320 meV) such that the transition temperature $`T_c`$ at $`\delta =0.17`$ remains close to 80K (without any admixture of $`cos(6\varphi )`$ term). In the spirit of Mesot et al., we introduced the next order term $`cos(6\varphi )`$ in $`\eta (𝐤)`$ as $`\eta (𝐤)=\alpha cos(2\varphi )+(1\alpha )cos(6\varphi )`$, where $`\alpha `$ measures the relative contributions of the two terms.
The superconducting gap has been calculated for each filling ($`\delta `$) at different levels of mixing of the higher harmonic. As representative curves, we show in Fig.1 the ones with 0, 4, 10 and 20% of mixing for $`\delta =0.17`$. The solutions of the gap function for different $`\alpha `$ show typical square root behaviour as $`TT_c`$. The transition temperature (and $`\mathrm{\Delta }(0)`$) reduces considerably as $`\alpha `$ deviates from one (the inset to Fig. 1). The gap function with the same set of values of $`\alpha `$ are drawn in the Y-quadrant (Fig. 2) for a demonstration of how they change with the mixing of $`cos(6\varphi )`$. Similar shifts have been seen by Mesot et al. as well.
Calculation of the density of states (DOS) in the superconducting state with and without the higher harmonic term is straightforward. For $`\alpha =1`$, one gets the usual d-wave DOS with $`\rho (E)|E|`$ for $`E0`$. Addition of the higher harmonic term makes the rise steeper as there are now more excitations available at lower energy though the low energy behaviour of the DOS remains the same (linear). In Fig. 3 the DOS is shown for the normal and the broken symmetry states (with only $`\alpha =1`$, as the $`\alpha =0.8`$ curve is indistinguishable in the scale of that figure). The typical d-wave V-shaped DOS is obtained with the shoulders at $`\pm \mathrm{\Delta }`$ around $`E=0`$. The inset shows for a comparison, in an enhanced scale of energy, the effect of the addition of the higher harmonic term. The pile-up of states at lower energy is quite evident.
In order for an analysis of the ARPES spectra we take the point of view that the spectral function $`A(𝐤,\omega )`$, convoluted with the fermi function $`f(\omega )`$ and the resolution of the detector, gives the EDC. The ARPES intensity (without the detector resolution) is given by $`I_0f(\omega )A(𝐤,\omega )`$ where, $`I_0`$ is a prefactor weakly dependent on $`T`$ and $`\omega `$. It depends on the incident photon energy, momentum and the (electron-phonon) matrix element between the initial and final states. The momentum resolution of the detectors used in experiments considered is about one degree in the BZ and is assumed constant over a circular window of 1<sup>o</sup> radius. With the incident photon energy around 20eV, the energy resolution of the detector is taken to be a Gaussian of standard deviation 7meV, a value consistent with the present day experimental resolutions.
With all these taken into account, the ARPES intensity is given by
$$I(𝐤,\omega )=I_0_𝐤^{}_\nu \stackrel{~}{G}(\omega \nu )f(\nu )A(𝐤^{},\nu ).$$
Here $`\stackrel{~}{G}(\omega )`$ is the Gaussian energy resolution function discussed above and the $`𝐤^{}`$-integration is within a circular radius of 1<sup>o</sup>. The spectral function is the usual mean-field one
$$A(𝐤,\omega )=u_𝐤^2\delta (\omega E_𝐤)+v_𝐤^2\delta (\omega +E_𝐤)$$
where, $`E_𝐤=\sqrt{ϵ_𝐤^2+\mathrm{\Delta }_𝐤^2}`$, and the coherence factors $`u_𝐤^2=\frac{1}{2}(1+ϵ_𝐤/E_𝐤),v_𝐤^2=\frac{1}{2}(1ϵ_𝐤/E_𝐤).`$
Taking the normal state dispersion $`ϵ_𝐤`$ discussed above, $`A(𝐤,\omega )`$ at five representative angles (on the FS) $`\varphi =0,10,23,35,45`$ have been calculated and their frequency and momentum-resolved behaviour (Fig. 4) obtained. The angles are measured with respect to the line $`(\pi ,\pi )(\pi ,0)`$ as is the practice. We have restricted ourselves to the Y-quadrant of the Brillouin zone in order to avoid the complications with the shadow bands that appear in the X-quadrant due to the presence of an incommensurate superlattice (along $`\mathrm{\Gamma }Y`$ direction) in the BiO planes. We choose a point on the FS (at the above angles), and perform the resolution averaging around that point in momentum and frequency. Extreme care is required in evaluating the averages on the FS where the band is most dispersive, i.e., around 45<sup>o</sup>. Panel 1 in Fig. 4 shows frequency averaged $`A(𝐤,\omega )`$ and panel 2, the momentum-averaged one. The procedure has been repeated for $`\alpha =1.0`$ and 0.90 to show the effects of the higher harmonic on the gap function as one moves along the FS.
With the momentum and frequency averaged $`A(𝐤,\omega )`$ obtained, it is easy to calculate the ARPES intensity around different points on the FS. The curves at different angles are shown in panel 3 of Fig. 4. The maximum gap is taken to be 30 meV, the same as used by Mesot et al, for demonstration. The temperature used is 10K, a typical value used in experiments at low temperatures.
It easily follows from the above discussions that the thermodynamic properties will be affected by the addition of the higher harmonics. We calculate the specific heat as a function of temperature (Fig. 5) in the superconducting state by taking a derivative of the entropy with temperature. Figs. 5 (a), (b) and (c) show the specific heat for different values of $`\alpha `$ at several levels of doping. It is clear from the graphs (a) - (c) that in order to account for the extra entropy, the curves move up as higher harmonics are brought in and then the area is conserved with reduced transition temperatures. As the doping increases, the curvatures of the Fermi surface affect the specific heat as seen in the normal state specific heat curve in (c). At T<sub>c</sub> the specific heat jumps to its normal state value as is typical of a second order transition.
The calculation of superfluid density $`\rho _s`$ has been performed using the standard techniques of many body theory. The diamagnetic and paramagnetic contributions to the current (and hence the phase stiffness) are calculated in the linear response by first making a Peierls substitution $`t_{ij}=t_{ij}exp(ie/c\mathrm{})_{𝐫_j}^{𝐫_i}𝐀.d𝐥`$ in the hopping matrix element. The paramagnetic contribution at long wavelengths (via excitations above the condensate) in the linear response theory is obtained in terms of the correlation function
$$𝐣_{para}^x(𝐪)=\frac{i}{c}lim_{q0}lim_{\omega 0}𝑑\tau \theta (\tau )e^{i\omega \tau }[j_x^{para}(𝐪,\tau ),j_x^{para}(𝐪,0)]𝐀_𝐱(𝐪).$$
The correlation function on the RHS is calculated at the one-loop level. Calculation of the diamagnetic contribution (from the Meissner condensate) is straightforward: $`𝐣_{dia}^x(𝐪)=\frac{e^2}{N\mathrm{}^2c}_{𝐤,\sigma }c_{𝐤,\sigma }^{}c_{𝐤,\sigma }𝐀_𝐱(𝐪).`$ In the resulting expression, the OP values were taken as their mean-field unrenormalized (by the fluctuations) ones, a procedure that is known to work except very close to T<sub>c</sub>. The resulting $`\rho _sT`$ curves are shown in Figs. 6 for $`\alpha =1,0.9`$ and 0.8 at different doping levels. Note that the calculations were done in the gauge $`A_y=0`$ and the gauge invariance is restored if vertex corrections are included. Such an approximation entails neglecting the vortex-like fluctuations in the 2D model. The above expression for $`\rho _s`$ is derived for an isotropic order parameter but is expected to work quite well even in the anisotropic case at hand as the asymptotic form of the vortex-vortex interaction (at high vortex density) is logarithmic and corrections to the expression above due to these fluctuations are indeed small.
Results and Discussion
The nature of the OP gleaned from Fig. 1 shows a sensitive dependence on $`\alpha `$, going down as mixing increases. The cos($`6\varphi `$) term changes sign four times now in each quadrant and such rapid changes average out to a smaller value. This is more pronounced if the next higher harmonic ($`cos(10\varphi )`$) is introduced. As $`\alpha `$ decreases, the $`\mathrm{\Delta }`$ versus $`\varphi `$ curve becomes flatter around the node (Fig.2) enhancing the quasiparticle excitations above the condensate. Such excitations will reflect in a reduced T<sub>c</sub> (and $`\mathrm{\Delta }(0)`$) and pile up of states in the DOS at low energies (Fig. 3). This will, of course, affect the thermodynamics considerably. For example, though the specific heat follows the typical (mean-field) T<sup>2</sup> behaviour of a d-wave superconductor at low temperature, the increased entropy at lower energy (for $`\alpha `$ deviating from one) manifests itself in the specific heat curves (Fig. 5).
The ARPES line shapes are shown in Fig. 4. The frequency and momentum- averaged spectral function is plotted in panel 1 and panel 2 at different angles relative to ($`\pi ,\pi `$) on the FS. $`A(𝐤,\omega )`$ (not shown) has the usual delta function behaviour, two sharp peaks separated by 2$`\mathrm{\Delta }`$ at $`\varphi =0`$ and closing in as $`\varphi `$ increases and finally merging at $`\varphi 45^o`$. The frequency broadening (panel 1) is a mere consequence of the Gaussian averaging procedure on the $`A(𝐤_F,\omega )`$. Strong angle dependence is observed in panel 2 where momentum averaged $`A(𝐤_F,\omega )`$ is plotted for different $`\varphi `$. The OP does not change within the $`𝐤`$-window for small angles, while for large angles, the effects are very strong. At large angles the band is highly dispersive and large changes in energy occur within the $`𝐤`$window. The $`𝐤`$-dependence of the OP further enhances this angle dependence since close to the node $`\mathrm{\Delta }_𝐤`$ varies linearly with $`𝐤`$ while at the gap-maximum the variation is a weaker quadratic one. The $`𝐤`$-averaged spectral weight for $`\alpha =0.9`$ shows perceptibly smaller gap as $`\varphi `$ increases. Such a reduction of the gap has been observed by Mesot et al. The effect of $`\alpha `$ is clearly visible in the EDCs in panel 3. The leading edges for the curves corresponding to $`\alpha =1`$ and 0.90 move continuously away from the fermi energy as $`\varphi `$ decreases, the curve for $`\alpha =1`$ moving more rapidly. This is the scenario depicted in Mesot et al. for seven different angles, where the best fit is obtained with $`0.89\alpha 0.96`$ for different samples. The peaks observed in our EDCs are due to the mean-field pile-up of states, and not due to electronic correlations.
The superfluid density $`\rho _s`$ is proportional to $`\lambda ^2`$ ($`\lambda `$ is the penetration depth) and has a power law dependence on temperature at low temperatures. Figs. 6 (a)-(c) show $`\rho _s`$ for three representative $`\alpha `$ (1.0, 0.90 and 0.80) and is seen to fall off faster with the inclusion of the higher harmonics. As observed earlier, the curves are linear to a high degree close to zero temperature. We observe that the slope of $`\rho _s(T0)`$ decreases monotonically with increasing $`\alpha `$ (i.e., increasing T<sub>c</sub>). As temperature rises, excitations from the condensate tend to decrease $`\rho _s`$ at the expense of normal quasiparticles above the condensate. The gradual flattening of the $`\mathrm{\Delta }\varphi `$ curve with decreasing $`\alpha `$ around the node makes quasiparticle excitations more accessible at lower temperatures, causing a faster descent of $`\rho _s`$ with temperature. We have also calculated $`\rho _s`$ at different doping levels and find that the slope increases as one underdopes, consistent with the observations of Mesot et al.
It is known that non-magnetic impurities act as pair breaking scatterers for a d-wave superconductor. Repeated scattering even with small momentum transfer on the fermi surface between lobes of opposite sign in the Brillouin zone effectively reduces the average gap value thereby reducing the T$`_c.`$ Owing to the presence of the $`cos(6\varphi )`$ term, such pair-breaking processes are clearly going to be more efficient and will reduce the transition temperature as the higher harmonic components increase. The pair breaking effect of momentum dependent impurity potential and anisotropic gap including the higher harmonics have been studied recently by solving the Abrikosov-Gorkov equations in the T-matrix representation where it has been found that states begin to appear in the gap as the impurity potential as well as the higher harmonic component is increased thereby reducing both the superconducting gap and the transition temperature.
In conclusion, we have worked out the effects due to the increasing presence of higher harmonics observed recently in a d-wave superconductor on underdoping. Based on the standard phenomenological theory for d-wave superconductors, we observe that the transition temperature, density of states, specific heat, superfluid density, ARPES intensity and the slope of the OP at the node bear clear and detectable signature of this higher order term. Although the interaction between quasiparticles have not been included in the above, the conclusions drawn remain valid qualitatively on strong physical grounds as has been shown in a number of occasions earlier. The predictions made here are easily verifiable experimentally and will shed light on the nature and strength of the higher order term claimed to be present in these superconductors. More experiments are also required to fully understand the origin and the physics behind such additional terms.
Figure captions
* Normalized (by $`\mathrm{\Delta }(0)`$ at $`\alpha =1`$) gap versus temperature for four different values of $`\alpha `$. The inset shows variation of the transition temperature with $`\alpha `$.
* The order parameter as a function of angle (measured with respect to the line $`(\pi ,\pi )(\pi ,0)`$ in the first BZ as in refs.) with increasing mixture of the higher harmonic.
* The Density of States in the superconducting state (dotted line) with $`\alpha =1`$ and in the normal state (solid line) at $`\delta =0.17`$. The inset shows the difference in the DOS for $`\alpha =1`$ and $`\alpha =0.8`$ in the superconducting state close to the node at very low temperature (the gaps were 15 meV and 11.8 meV as in Fig. 1).
* The frequency- and momentum-averaged $`A(𝐤,\omega )`$ (panel 1 & 2) and the corresponding EDCs (panel 3, see text) at various angles (measured from $`Y\overline{M}`$ direction) on the FS (shown outside the panel in the Y-quadrant). Figures (a)-(e) are for angles 45, 35, 23, 10 and 0 degree. The solid and dotted lines correspond to $`\alpha =1.0`$ and 0.9.
* The specific heat curves at three different levels of mixing of the higher harmonic ((a), (b) and (c) correspond to $`\delta =0.10,0.17`$ and 0.28). The figures clearly reveal the enhanced quasiparticle excitations as $`\alpha `$ deviates from one.
* The superfluid density is shown against temperature for three different values of $`\delta =0.10,\mathrm{\hspace{0.17em}0.17}`$ and 0.28 (Figs. (a)-(c)). The change in slope at low temperatures is clearly visible. |
no-problem/9911/quant-ph9911022.html | ar5iv | text | # Kochen-Specker theorem and experimental test on hidden variablesTo appear in Int. J. Mod. Phys. A.
## I The Kochen-Specker theorem
The Kochen-Specker (KS) theorem contains one of the most fundamental findings in quantum mechanics (QM): Yes-no questions about an individual physical system cannot be assigned a unique answer in such a way that the result of measuring any mutually commuting subset of these yes-no questions can be interpreted as revealing these preexisting answers. To be precise, the KS theorem asserts that, in a Hilbert space with a finite dimension, $`d3`$, it is possible to construct a set of $`n`$ projection operators, which represent yes-no questions about an individual physical system, so that none of the $`2^n`$ possible sets of “yes” or “no” answers is compatible with the sum rule of QM for orthogonal resolutions of the identity (i.e., if the sum of a subset of mutually orthogonal projection operators is the identity, one and only one of the corresponding answers ought to be “yes”). This conclusion holds irrespective of the quantum state of the system. Implicit in the KS theorem is the assumption of noncontextuality: Each yes-no question is assigned a single unique answer, independent of which subset of mutually commuting projection operators one might consider it with. Therefore, the KS theorem discards hidden-variable theories with this property, known as noncontextual hidden-variable (NCHV) theories. Local hidden-variable theories, such as those discarded by Bell’s theorem , are a particular type of NCHV theories, so in this sense, the KS theorem is more general than Bell’s theorem. However, while Bell’s theorem has been successfully tested in the laboratory , the translation of any proof of the KS theorem into an experiment on an individual system seems to be an impossible task. This is so because the $`n`$ projection operators appearing in a proof of the KS theorem can be combined in $`t`$ different orthogonal resolutions of the identity, and each of them represents a maximal test which is incompatible with the other $`t1`$ maximal tests. However, we have recently presented an experiment which seems to challenge the idea that the KS theorem cannot be tested in the laboratory: A test on an individual system of a two-part two-state system (prepared in whatever quantum state) exists for which the predictions of any NCHV theory always differ with those of QM .
This test would be related somehow with some of the proofs of the KS theorem. The smallest proof of the KS theorem currently known contains eighteen projection operators (yes-no questions) which can be combined in nine resolutions of the identity (maximal tests) of a four-dimensional Hilbert space, $`_4`$ . Both results, and , concern to physical systems described in $`_4`$, and it would be interesting to establish which is the connection between them . That is precisely the aim of this paper.
## II The proof with 18 vectors
The proof of the KS theorem with eighteen projection operators in $`_4`$ is given in Table I .
| $`1000`$ | $`1111`$ | $`1111`$ | $`1000`$ | $`1001`$ | $`1001`$ | $`111\overline{1}`$ | $`111\overline{1}`$ | $`100\overline{1}`$ |
| --- | --- | --- | --- | --- | --- | --- | --- | --- |
| $`0100`$ | $`11\overline{1}\overline{1}`$ | $`1\overline{1}1\overline{1}`$ | $`0010`$ | $`0100`$ | $`1\overline{1}1\overline{1}`$ | $`1\overline{1}00`$ | $`0101`$ | $`0110`$ |
| $`0011`$ | $`1\overline{1}00`$ | $`10\overline{1}0`$ | $`0101`$ | $`0010`$ | $`11\overline{1}\overline{1}`$ | $`0011`$ | $`10\overline{1}0`$ | $`11\overline{1}1`$ |
| $`001\overline{1}`$ | $`001\overline{1}`$ | $`010\overline{1}`$ | $`010\overline{1}`$ | $`100\overline{1}`$ | $`0110`$ | $`11\overline{1}1`$ | $`1\overline{1}11`$ | $`1\overline{1}11`$ |
TABLE I: Proof of the Kochen-Specker theorem in $`_4`$.
Table I contains eighteen vectors combined in nine columns. Each vector represents the projection operator onto the corresponding normalized vector. For instance, $`001\overline{1}`$ represents the projector onto the vector $`{\scriptscriptstyle \frac{1}{\sqrt{2}}}(0,0,1,1)`$. Each column contains four mutually orthogonal vectors, so that the corresponding projectors sum the identity in $`_4`$. Therefore, in a NCHV theory, each column must have assigned the answer “yes” to one and only one vector. But it is easily seen that such an assignment is impossible since each vector in Table I appears twice, so that the total number of “yes” answers must be an even number.
If we examine this proof without noticing which particular physical system it refers to, all we see is that each projector in Table I is orthogonal to other seven projectors and belongs to two distinct resolutions of the identity. In this sense, all involved projectors and resolutions of the identity play the same role in the proof. This absence of privileged yes-no questions or tests is also characteristic of every proof of the KS theorem in $`_3`$ .
However, in $`_4`$ this situation of apparent symmetry changes if $`_4`$ can be viewed as a product of two tensor factors, $`_2_2`$, corresponding to some subsystems of the physical system. Examples of systems in which such a decomposition is physically meaningful (in a sense that will be specified in Sec. III) are two-part two-level systems such as two spin-$`\frac{1}{2}`$ particles without translational motion, the polarization state of two photons, or two internal levels of a pair of trapped ions. Examples of systems in which such a decomposition is mathematically possible but will not have the same physical meaning mentioned above are the spin state of a single spin-$`\frac{3}{2}`$ particle or two different two-level degrees of freedom in a single ion.
In the following I will suppose that $`_4`$ represents the spin state of two spin-$`\frac{1}{2}`$ particles. Then the translation from the proof in Table I into a proof in $`_2_2`$ can be easily achieved by realizing that the eighteen vectors in Table I are eigenvectors of some products of the usual representation of the Pauli matrices $`\sigma _z`$ and $`\sigma _x`$ for the spin state of spin-$`\frac{1}{2}`$ particles. Then Table I can be rewritten as Table II.
| $`zz`$ | $`xx`$ | $`xx`$ | $`zz`$ | $`zzxx`$ | $`zzxx`$ | $`zxxz`$ | $`zxxz`$ | $`zz\overline{xx}`$ |
| --- | --- | --- | --- | --- | --- | --- | --- | --- |
| $`z\overline{z}`$ | $`\overline{x}x`$ | $`x\overline{x}`$ | $`\overline{z}z`$ | $`z\overline{z}`$ | $`x\overline{x}`$ | $`z\overline{x}`$ | $`x\overline{z}`$ | $`\overline{zz}xx`$ |
| $`\overline{z}x`$ | $`z\overline{x}`$ | $`\overline{x}z`$ | $`x\overline{z}`$ | $`\overline{z}z`$ | $`\overline{x}x`$ | $`\overline{z}x`$ | $`\overline{x}z`$ | $`zx\overline{xz}`$ |
| $`\overline{z}\overline{x}`$ | $`\overline{z}\overline{x}`$ | $`\overline{x}\overline{z}`$ | $`\overline{x}\overline{z}`$ | $`zz\overline{xx}`$ | $`\overline{zz}xx`$ | $`zx\overline{xz}`$ | $`\overline{zx}xz`$ | $`\overline{zx}xz`$ |
TABLE II: Proof of the Kochen-Specker theorem in $`_2_2`$.
The notation of Table II is the following: $`z\overline{x}`$ represents the yes-no question “are the spin component of first particle positive in the $`z`$ direction and the spin component of second particle negative in the $`x`$ direction?”, and $`\overline{zx}xz`$ denotes the yes-no question “are the products $`zx:=\sigma _{1z}\sigma _{2x}`$ and $`xz:=\sigma _{1x}\sigma _{2z}`$ negative and positive respectively?”, etc. The first is an example of a factorizable yes-no question, since it can be answered after separate tests on the first and second particles. The latter is an example of an entangled yes-no question, since it cannot be answered after separate tests on both particles. Therefore, in Table II there are two types of yes-no questions and, consequently, three types of maximal tests: those involving factorizable yes-no questions only, such as those in columns 1 to 4; those involving both factorizable and entangled yes-no questions, such as those in columns 4 to 8; and those involving entangled yes-no questions only, such as the one in the ninth column. Taking into account this new hierarchy of experiments, the relevant elements of the proof of the KS theorem in $`_2_2`$ can be illustrated as in Fig. 1.
FIG. 1: Hierarchy of tests in the proof of the Kochen-Specker theorem in $`_2_2`$. Each dot represents a yes-no question. The upper and the lower dots both represent the same yes-no question, and the far left and the far right dots both represent another yes-no question. Dots in the same straight line or in the same circumference represent mutually compatible yes-no questions, and therefore each straight line and circumference represents a maximal test. Each straight line belonging to the square represents a test containing only factorizable yes-no questions. The other straight lines represent tests containing both factorizable and entangled yes-no questions. The circumference represents a test containing only entangled yes-no questions.
In the next section I will show that the algebraic contradiction contained in this proof of the KS theorem in $`_2_2`$ is of a different kind to that of those appearing in the proofs of the KS theorem in $`_3`$. While in $`_3`$ there are no conflicting predictions between QM and NCHV theories for any of the single tests appearing in the proofs (basically because NCHV theories do not make specific predictions apart from those of QM), in $`_2_2`$ NCHV theories can make specific predictions for single tests that may disagree with those of QM.
## III Noncontextual hidden-variable theories
A NCHV theory must satisfy the following assumptions:
(i) Any one-particle observable must have a definite value. This is not in contradiction with QM since the Kochen-Specker theorem is not valid for $`_2`$. In fact, in $`_2`$ specific NCHV models exist reproducing all statistical predictions of QM . This is one of the reasons why the decomposition of $`_4`$ into $`_2_2`$ makes more physical sense in some systems than in others. For our purposes we will assume that the observables $`z_1:=\sigma _{1z}`$, $`x_1:=\sigma _{1x}`$, $`z_2:=\sigma _{2z}`$, and $`x_2:=\sigma _{2x}`$ must have predefined values, either $`+1`$ or $`1`$ (that will be denoted simply by “$`+`$” or “$``$”). Therefore, there are 16 possible distinct states (in a NCHV theory), corresponding to the different combinations of possible values for these four one-particle observables.
(ii) The value of a two-particle observable which is a product of two one-particle observables corresponding to different particles, such as $`z_1z_2:=\sigma _{1z}\sigma _{2z}`$, $`x_1x_2:=\sigma _{1x}\sigma _{2x}`$, $`z_1x_2:=\sigma _{1z}\sigma _{2x}`$, and $`x_1z_2:=\sigma _{1x}\sigma _{2z}`$, is the product of values of the corresponding one-particle observables. This assumption is a consequence of the general assumption of noncontextuality. Consider, for instance, the obsevable $`zz`$. One particular way of measuring $`zz`$ is by measuring $`z_1`$ and $`z_2`$ separately and multiplying their results. The independece of the results of the separated measurements is garanteed (if such measurements are spacelike separated) by the highest form of noncontextuality: locality (so this the second reason why the decomposition of $`_4`$ into $`_2_2`$ makes more physical sense in some systems than in others). Then the definition of noncontextuality entails that the value for $`zz`$ must be the same whatever the experimental context one might choose .
(iii) The answer to a yes-no question is logically related with the values of the involved observables. For instance, the answer to the question $`zz\overline{xx}`$ is “yes” if $`z_1=z_2`$ and $`x_1=x_2`$, and “no” in any other case.
Let us now examine the predictions of a NCHV theory for the nine tests. Table III contains all possible values for one-particle and two-particle observables involved in the proof.
| one-particle | | | | two-particle | | | | two-particle | | | |
| --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- |
| observables | | | | observables | | | | yes-no questions | | | |
| $`z_1`$ | $`x_1`$ | $`z_2`$ | $`x_2`$ | $`z_1z_2`$ | $`x_1x_2`$ | $`z_1x_2`$ | $`x_1z_2`$ | $`zz\overline{xx}`$ | $`\overline{zz}xx`$ | $`zx\overline{xz}`$ | $`\overline{zx}xz`$ |
| $`\pm `$ | $`\pm `$ | $`\pm `$ | $`\pm `$ | $`+`$ | $`+`$ | $`+`$ | $`+`$ | no | no | no | no |
| $`\pm `$ | $`\pm `$ | $`\pm `$ | $``$ | $`+`$ | $``$ | $``$ | $`+`$ | yes | no | no | yes |
| $`\pm `$ | $`\pm `$ | $``$ | $`\pm `$ | $``$ | $`+`$ | $`+`$ | $``$ | no | yes | yes | no |
| $`\pm `$ | $`\pm `$ | $``$ | $``$ | $``$ | $``$ | $``$ | $``$ | no | no | no | no |
| $`\pm `$ | $``$ | $`\pm `$ | $`\pm `$ | $`+`$ | $``$ | $`+`$ | $``$ | yes | no | yes | no |
| $`\pm `$ | $``$ | $`\pm `$ | $``$ | $`+`$ | $`+`$ | $``$ | $``$ | no | no | no | no |
| $`\pm `$ | $``$ | $``$ | $`\pm `$ | $``$ | $``$ | $`+`$ | $`+`$ | no | no | no | no |
| $`\pm `$ | $``$ | $``$ | $``$ | $``$ | $`+`$ | $``$ | $`+`$ | no | yes | no | yes |
TABLE III: Possible values in a NCHV theory for the observables and yes-no questions involved in the ninth test of the KS theorem in $`_2_2`$.
As can easily be seen studying Table III:
(a) For each of the four tests involving only factorizable yes-no questions (columns 1 to 4 in Table II) NCHV theories predict that one and only one of the answers must be “yes”. The same prediction as in QM. Note that if, instead of $`z`$ and $`x`$, we choose any other spin component, then the predictions of both QM and NCHV theories will still agree.
(b) For each of the four tests involving factorizable and entangled yes-no questions (columns 5 to 8 in Table II) NCHV theories also predict that one and only one of the answers must be “yes”. As QM does. Note that the choice of spin components $`z`$ and $`x`$ allows the corresponding two-particle observables to have a common eigenvalue so the corresponding yes-no question can be represented in QM by a projection operator.
(c) Finally, consider the last test, involving only entangled yes-no questions (column 9 in Table II). If one checks Table III, one reaches the conclusion that in a NCHV theory the four yes-no questions appearing in column 9 in Table II are not mutually exclusive (since the answers to two of them can be “yes”) nor not exhaustive (since the answers to all of them can be “no”). The quantum prediction is radically different. In QM, these four yes-no questions (or, to be precise, their corresponding projectors) form an orthogonal resolution of the identity, so they represent mutually exclusive and exhaustive questions: the answers must be one “yes” and three “no”.
## IV Conclusion and further developments
In brief, the predictions of QM and NCHV agree for the first eight tests of the eighteen vector’s proof of the KS theorem in $`_2_2`$, but disagree for the ninth test. This clarifies the relationship between this proof, originally proposed in , and the test on NCHV theories proposed in .
A joint measurement on a single system of the four projection operators appearing in the ninth test is closely related to the problem of performing a measurement of a nondegenerate Bell operator . Such a measurement is also required for reliable (i.e., with 100% theoretical probability of success) double density quantum coding and for reliable teleportation . It has recently been proved that a measurement of a nondegenerate Bell operator cannot be achieved without using quantum systems interacting one with the other, a condition which is not fulfilled in recent experiments , but which could be achieved with atoms and electro-magnetic cavities in near-future experiments.
## Aknowledgments
This paper is the response to a question raised by Lucien Hardy after a talk at Oxford University in May 1998. I would like to thank Mark S. Child and Gonzalo García de Polavieja for inviting me to give that talk. I also acknowledge useful conversations on this subject with Ignacio Cirac, David Mermin, Guillermo García Alcaine, Asher Peres, and Emilio Santos, and stimulating comments by Helle Bechmann Pasquinucci, Sibasish Ghosh, Gebhard Grübl, and R. L. Schafir. This work was financially supported by the Universidad de Sevilla (OGICYT-191-97) and the Junta de Andalucía (FQM-239). |
no-problem/9911/physics9911068.html | ar5iv | text | # Two-Dimensional Sideband Raman Cooling and Zeeman State Preparation in an Optical Lattice *footnote **footnote *Contribution of NIST, not subject to copyright
## I Introduction
Laser-cooled atoms play a critical role in modern frequency standards such as atomic fountains . As is well-known, Sisyphus-type cooling in optical molasses with polarization gradients results in atoms with temperatures corresponding to tens of the single-photon recoil energies $`\epsilon _r=(\mathrm{}k)^2/2M`$ (for example, $`T30\epsilon _r/k_B3\mu K`$ in the case of $`Cs`$ ). Even lower temperatures can be achieved by velocity-selective methods . These methods, however, require more complicated technical implementations .
Recently, Poul Jessen and co-workers demonstrated an elegant and efficient method of cooling atoms to the vibrational ground state of a far-off-resonance two-dimensional optical lattice. Their method is a variant of Raman sideband cooling based on transitions between the vibrational manifolds of adjacent Zeeman substates. A static magnetic field is used to tune the Zeeman levels so that Raman resonance occurs on the red sideband and results in cooling. Two circularly polarized fields are then used to recycle the atoms for repetitive Raman cooling. The cooling operates in the Lamb-Dicke regime with cw laser beams and does not require phase-locked lasers; a transverse temperature of about $`950nK`$ was achieved.
Stimulated by the concepts and results from Jessen , we propose a new variant of transverse sideband cooling. The basic difference from the method of Ref. is that a linearly polarized pumping field, detuned from resonance, now plays a two-fold role. It provides both optical pumping back to the $`m=0`$ magnetic sublevel, and it causes a uniform ac Stark shift that replaces the external magnetic field-induced Zeeman shift that was used in Ref. . The main improvement consists in the flexibility of optical methods of the atomic state control. For example, in our scheme, changing the polarization from linear to circular and the direction of the pumping beam, we can accumulate all atoms in the outermost Zeeman substate $`m=F`$.
One promising application of this technique is for two-dimensional cooling of atoms in an atomic fountain, or for a space-based atomic clock. More precisely, the method can be used to cool an atomic sample down to sub-$`\mu K`$ temperatures in the transverse directions before launching the cold atoms into the Ramsey interaction region. For the long Ramsey periods that could be used in an atomic clock in space, reducing the transverse spread would increase substantially the number of atoms in the detection region. In this context, Jessen’s original scheme has some disadvantages. Namely, atoms are accumulated in the stretched $`m=F`$, substate of the $`F=4`$ ground-state hyperfine level of $`Cs`$, a perturbing magnetic field is used, and that system also requires that the pumping and repumping beams propagating orthogonal to the cooling plane (i.e. down to axis of the clock), which would perturb atoms drifting through the Ramsey region. Newly proposed atomic-fountain clocks on Earth and clocks proposed for space will use multiple balls of atoms in the Ramsey region, so there would be atoms in the Ramsey cavity undergoing state interrogation while other atoms are state prepared and detected. In this case the out-of-plane light beams and magnetic field would produce unacceptable frequency shifts. Some of these problems are of a technical character and could be solved by different methods. For instance, atoms might be transfered from $`|F=4,m=4`$ to $`|F=4,m=0`$ without additional heating by the adiabatic passage technique , and the Ramsey region might be adequately isolated from the magnetic field. However, the problem of the light beams down the clock axis is more fundamental problem. Our scheme avoids these problems without additional technical complications, while maintaining most of the attractive features. In particular, in the present version, only cw lasers lying in the cooling plane are used, there is no magnetic field, and atoms are prepared in the $`m=0`$ substate. Since the linear polarization of the pumping field coincides (must coincide) with the orientation of the clock’s magnetic field (C-field) in the Ramsey region, switching between the Stark shifted and Zeeman shifted substates will not produce any mixing.
In the following section we discuss the proposed lattice and cooling parameters in the framework of a simple theoretical model. The optimal magnitudes of the Raman transition amplitude, the pumping field intensity, and the detuning are found. These analytical results are then confirmed by numerical calculations for a more realistic model of the $`FF^{}=F`$ transition. In addition, we find that coherence between degenerate (or nearly degenerate) lower vibrational levels can lead, under certain conditions, to significant changes in the cooling efficiency and cooling time.
The proposed cooling method may also be useful for atom optics by providing a high-brightness well-collimated source of atoms, or for general purposes of quantum-state control in a non-dissipative optical lattice.
## II Optical Lattice
The field configuration used for the optical lattice consists of three linearly polarized beams having equal amplitudes and propagating in the $`xy`$-plane with angles of $`2\pi /3`$ between them (Fig. 1). The polarization vectors of these beams are tilted through a small angle $`\varphi `$ with respect to the $`z`$-axis. The resulting field can be written as
$`𝐄(𝐫,t)`$ $`=`$ $`E_0(𝐫)\mathrm{exp}(i\omega _Lt)+\mathrm{c}.\mathrm{c}`$ (1)
$`(𝐫)`$ $`=`$ $`𝐞_z{\displaystyle \underset{i=1}{\overset{3}{}}}\mathrm{exp}(i𝐤_i𝐫)+\mathrm{tan}(\varphi ){\displaystyle \underset{i=1}{\overset{3}{}}}𝐞_i\mathrm{exp}(i𝐤_i𝐫),`$ (2)
where $`𝐤_i`$ and $`\mathrm{tan}(\varphi )𝐞_i`$ are respectively the wave vectors and the in-plane components of the polarization of the $`i`$-th beam. All the beams have the same frequency, $`\omega _L`$, far-detuned to the red of the $`D_2`$ resonance line.
As was shown in Ref. , if the detuning is much greater than the hyperfine splitting of the excited state, then the optical potential for the ground state takes the form
$$\widehat{U}_F=\frac{2}{3}u_s|(𝐫)|^2+\frac{i}{3}u_sg(F)[(𝐫)^{}\times (𝐫)]\widehat{𝐅}.$$
(3)
Here $`g(F)=[F(F+1)+J(J+1)I(I+1)]/[F(F+1)]`$ where $`F`$, $`J`$ and $`I`$ are respectively the total, the electron and the nuclear angular momenta of the ground state, and $`\widehat{𝐅}`$ is the angular-momentum operator. The single-beam light shift $`u_s=A/\mathrm{\Delta }`$, defined as in Ref. , is proportional to the single-beam light intensity $``$ and inversely proportional to the detuning $`\mathrm{\Delta }=\omega _L\omega _{F,F_{max}^{}}`$. For the $`D_2`$ line of $`{}_{}{}^{133}Cs`$ the constant $`A1.5\epsilon _rGHz/(mWcm^2)`$.
To zeroth order in $`\mathrm{tan}(\varphi )1`$, the field (1) is linearly polarized along $`𝐞_z`$ everywhere and vector term in Eq. (3) vanishes, resulting in the isotropic optical potential
$$\widehat{U}^{(0)}=\frac{4}{3}u_s\left[\frac{3}{2}+\mathrm{cos}(\sqrt{3}kx)+\mathrm{cos}(\frac{\sqrt{3}kx3ky}{2})+\mathrm{cos}(\frac{\sqrt{3}kx+3ky}{2})\right].$$
(4)
In other words, contrary to the field configuration of Ref. , all the Zeeman sublevels have the same optical shift. For red detunings $`\mathrm{\Delta }<0`$, the minima of the potential (4) form a lattice consisting of equilateral triangles with a side $`2\lambda /3`$ (one of them has the coordinates $`x=y=0`$).
In the general case, the atomic motion in a periodic potential leads to a energy band-structure. However, for potentials with a periodicity of the order of the light wavelength $`\lambda `$ and with the depth much larger than the recoil energy $`\epsilon _r`$ ($`6u_s`$ in the case under consideration), both the tunneling probability and the width are exponentially small for bands close to a potential minimum. Hence, instead of a lattice with energy bands we can consider vibrational levels arising from independent potential wells. The spectrum of the lower levels can be defined, with good accuracy, from the harmonic expansion in the vicinity of the well’s bottom:
$$\widehat{U}^{(0)}u_s[6+3k^2(X^2+Y^2)],$$
where $`X`$ and $`Y`$ are the displacements from the minimum. This expansion corresponds to a 2D isotropic harmonic oscillator with the frequency $`\mathrm{}\omega _v=\sqrt{12u_s\epsilon _r}`$. Due to the isotropy, the $`n`$-th energy level is $`n+1`$ times degenerate. If the energy separation between adjacent vibrational levels is much greater than the recoil energy, the characteristic size of lower vibrational states is $`l=\sqrt{\mathrm{}/M\omega _v}\lambda `$. In this case we have strong localization, and the Lamb-Dicke regime holds.
## III Sideband Raman Cooling
Raman transitions between vibrational levels of adjacent magnetic substates are induced by the small in-plane component of the field (1). To first order in $`\mathrm{tan}(\varphi )`$, the vector part of Eqn. (4) gives the correction
$$\widehat{U}^{(1)}=\frac{1}{3}u_sg(F)\mathrm{tan}(\varphi )𝐌(𝐫)\widehat{𝐅},$$
(5)
where $`𝐌`$ has the components $`M_x=2\sqrt{3}[\mathrm{cos}(3ky/2)\mathrm{sin}(\sqrt{3}kx/2)+\mathrm{sin}(\sqrt{3}kx)]`$ and $`M_y=6\mathrm{sin}(3ky/2)\mathrm{cos}(\sqrt{3}kx/2)`$. Since this term conserves the symmetry of the main potential (4), each well in the lattice obeys the same conditions for the Raman transitions. For the lower vibrational levels we use a first-order approximation with respect to the displacements $`X,Y`$ from the minimum
$`\widehat{U}^{(1)}`$ $``$ $`3u_sg(F)\mathrm{tan}(\varphi )k(X\widehat{F}_x+Y\widehat{F}_y)`$ (6)
$`=`$ $`{\displaystyle \frac{3}{2}}u_sg(F)\mathrm{tan}(\varphi )k\left((XiY)\widehat{F}_++(X+iY)\widehat{F}_{}\right),`$ (7)
where $`\widehat{F}_\pm `$ are the standard raising and lowering angular momentum operators. The operator $`\widehat{U}^{(1)}`$ has off-diagonal elements both for the vibrational and for the magnetic quantum numbers, inducing transitions with the selection rules $`\mathrm{\Delta }n=\pm 1`$ and $`\mathrm{\Delta }m=\pm 1`$ (for a quantization axis along $`𝐞_z`$). In order of magnitude, the Raman transition rate between the lower vibrational levels is $`U_R=u_s\mathrm{tan}(\varphi )kl`$. As was shown in Refs. , sideband cooling and coherent quantum-state control require this rate to be much greater than the spontaneous scattering rate of lattice photons $`\gamma _s=6\mathrm{\Gamma }u_s/\mathrm{\Delta }`$, where $`\mathrm{\Gamma }`$ is the natural width. In our lattice $`U_R/\gamma _s0.2\mathrm{tan}(\varphi )\mathrm{\Delta }/\mathrm{\Gamma }(\epsilon _r/u_s)^{1/4}`$.
Two other important requirements for efficient Raman sideband cooling are a spatially independent energy shift of the magnetic sublevels and optical pumping. To achieve these, we propose to use another optical field, known as the pump beam, linearly polarized along the $`z`$-axis, propagating in the cooling plane, and detuned by several $`\mathrm{\Gamma }`$ to the blue of the $`FF^{\prime \prime }=F`$ transition of the $`D_1`$ line (Fig. 1). In this case the $`m=0`$ sublevel is dark and unshifted, while the others undergo the light shifts
$$\delta _m=m^2\frac{\mathrm{\Delta }_p\mathrm{\Omega }_p^2}{\mathrm{\Gamma }^2/4+\mathrm{\Delta }_p^2},$$
where $`\mathrm{\Omega }_p`$ is the Rabi coupling for the $`|F,m=\pm 1|F^{\prime \prime }=F,m^{\prime \prime }=\pm 1`$ transitions and $`\mathrm{\Delta }_p`$ is the detuning of the pump field. With a proper choice of $`\mathrm{\Omega }_p`$ and $`\mathrm{\Delta }_p`$, the states $`|m=0,n+1`$ and $`|m=\pm 1,n`$ will have the same energy, which leads to efficient transition between them due to the Raman coupling. The cooling picture is completed by the optical pumping, which provides the relaxation from $`|m=\pm 1,n`$ to $`|m=0,n`$ (see Fig. 2.a). The vibrational quantum number $`n`$ is conserved in this process due to the fact that atoms are in the Lamb-Dicke regime. It is important to note that, contrary to Ref. , in our case, several levels are cooled simultaneously due to the isotropy of the potential $`\widehat{U}^{(0)}`$ (4). If $`\omega _vU_R`$ the state $`|m=0,n=0`$ is approximately dark and the majority of the atoms are eventually pumped into this target state. Thus, the described cooling method can be viewed as a version of dark-state cooling. Aspects of this cooling scheme that distinguish it from that of Ref. are: the cooling will be more efficient because more levels involved, the final pumping is to dark state $`m=0`$, and only laser fields in a plane perpendicular to the quantization axis are used.
To insure that there are no additional constraints and to estimate the cooling parameters, we consider a simple theoretical model based on the double $`\mathrm{\Lambda }`$-system (see Fig. 2.b). This simplified case allows an analytical treatment of the problem. We find the steady-state solution to the generalized optical Bloch equations involving the light-induced and spontaneous transitions and the Raman coupling. We are primarily interested in the limits
$$\mathrm{\Delta }_p\mathrm{\Gamma };\omega _vU_r,$$
(8)
because in this case the light-shift exceeds the field broadening and we can shift the states $`|2`$ and $`|5`$ into degeneracy with negligible perturbation of state $`|1`$. Under these conditions, the solution gives the following results: (i) the population of the target state $`|1`$ is maximal at exact resonance $`\mathrm{\Omega }_p^2/\mathrm{\Delta }_p=\omega _v`$ (see Fig. 3.a), (ii) on resonance, the total population of the states coupled with light is small, and equal to $`(U_R/\omega _v)^2`$, the probability for off-resonance Raman transitions $`|1|6`$ multiplied by a factor 4, and (iii) the population of state $`|2`$ contains two terms: $`\pi _2=1/2(U_R/\omega _v)^2+1/16(\gamma _p/\omega _v)^2`$. The second term is determined by the ratio of the width imposed by light, $`\gamma _p=\mathrm{\Gamma }\mathrm{\Omega }_p^2/\mathrm{\Delta }_p^2`$, to the vibrational frequency $`\omega _v`$. As a result, the target state population is close to unity:
$$\pi _11a(U_R/\omega _v)^2b(\mathrm{\Gamma }/\mathrm{\Delta }_p)^2.$$
(9)
The coefficients are $`a=3/2`$ and $`b=1/16`$ in the case of the double-$`\mathrm{\Lambda }`$ system model.
We now turn to an estimate of the cooling dynamics. Instead of looking for a temporal solution of the Bloch equations $`(d/dt)\rho =\widehat{}\rho `$ for atomic density matrix $`\rho `$ and $`\widehat{}`$ the corresponding Liouvillian superoperator, we find the statistically averaged transition time $`\tau =_0^{\mathrm{}}(\rho (t)\rho (\mathrm{}))𝑑t`$ . This matrix obeys the equations $`\widehat{}\tau =\rho (\mathrm{})\rho (0)`$, where $`\rho (\mathrm{})`$ is the steady-state solution and $`\rho (0)`$ is the initial distribution (we set $`\pi _1=\pi _2=\pi _5=\pi _6=1/4`$ and the other elements equal to zero at $`t=0`$). The cooling rate can be associated with the inverse transition time for the $`|1`$ state $`\gamma _{cool}=\tau _1^1`$. As a function of the optical frequency shift, the cooling rate is a Lorentzian curve with a width $`\sqrt{1/4\gamma _p^2+7U_R^2}`$ (see Fig. 3.a). Exactly on resonance, and in the limits (8) $`\mathrm{\Delta }_p=\mathrm{\Omega }_p^2/\omega _v`$, and $`\gamma _{cool}`$ takes the form
$$\gamma _{cool}=\alpha \frac{\gamma _pU_R^2}{\gamma _p^2+\beta U_R^2}.$$
(10)
Calculations within the framework of the double $`\mathrm{\Lambda }`$-system give $`\alpha =8`$ and $`\beta =28`$. This dependence of the cooling rate on $`\gamma _p`$ and $`U_R`$ can be understood qualitatively if we consider the cooling as optical pumping into the dark state. Obviously, because other parameters do not appear in the conditions (8), the cooling rate is determined entirely by the optical pumping rate $`\gamma _p`$ and the Raman transitions rate $`U_R`$. If $`U_R\gamma _p`$, an atom passes from $`|2`$ to $`|5`$ very quickly, and the cooling rate is proportional to the rate of the slower process of repumping from $`|5`$ to $`|1`$. In the inverse limit $`U_R\gamma _p`$, the slowest process is the transition $`|2|5`$. The corresponding rate, however, is not equal to $`U_R`$, but is suppressed by the factor $`U_R/\gamma _p`$. That can be explained as the inhibition of quantum transitions due to continuous measurements on the final state $`|5`$ (quantum Zeno effect ). The cooling rate $`\gamma _{cool}`$ as a function of $`\mathrm{\Omega }_p`$ (on resonance) is shown in Fig.3.b; $`\gamma _{cool}`$ achieves a maximum $`\gamma _{cool}^{max}=U_R\alpha /(2\sqrt{\beta })`$ at the optimal Rabi coupling $`\mathrm{\Omega }_p^{opt}=(\beta )^{1/4}\omega _v\sqrt{\mathrm{\Gamma }/U_R}`$.
The above described laws for the target-state population and for the cooling dynamics are confirmed by numerical calculations for a more realistic model of the $`FF^{}=F`$ cycling transition with a limited number of vibrational levels of the 2D oscillator taken into account. More precisely, we consider the systems of equations for the steady-state density matrix $`\rho (\mathrm{})`$ and for the transition time matrix $`\tau `$ taking into account for all possible coherences and populations. Including all terms gives a large system of algebraic equations, which can be solved for a fixed set of parameters by numerical methods. We use the LU decomposition method. The numerical results are fitted by the formulae (9,10) very well. The fitting coefficients $`a`$, $`b`$, $`\alpha `$ and $`\beta `$ depend on the angular momentum $`F`$ and on the initial distribution among the vibrational levels. The results, corresponding to the three lowest vibrational levels (with initially equal populations), are presented in Table 1.
In principle, two factors limit the number of vibrational levels which participate efficiently in the cooling: both the anharmonicity and the violation of the Lamb-Dicke regime become appreciable for higher vibrational levels. The second factor is the more stringent limitation and gives $`n^{}0.1\mathrm{}\omega _v/\epsilon _r`$ as an estimate for the maximal vibrational number.
It should be noted that a stray magnetic field can limit the cooling efficiency by breaking the level degeneracy. To be negligible, the Zeeman shift due to any stray magnetic field should be less than the Raman transition rate $`U_R`$, or the width imposed by the pumping field $`\gamma _p`$, which are of the same order of magnitude for optimal cooling.
To provide the cyclic interaction of the atoms with the pump field, repumping from the other hyperfine level is necessary. For this we propose using another light beam tuned in resonance with a $`FF^{}=F+1`$ transition of the $`D_2`$ line. This beam is linearly polarized along $`e_z`$ and runs in the $`xy`$-plane. For example, if the pumping field operates on the $`F=4F^{\prime \prime }=4`$ of the $`D_1`$ line of $`Cs`$, the repumping field is applied to the $`F=3F^{}=4`$ transition of the $`D_2`$ line. To minimize effects of optical pumping on the other hyperfine level, the intensity of the repumping field should be chosen close to the saturation intensity. It is noteworthy that in our lattice the potentials for both hyperfine levels have the same spatial dependence, and consequently the requirement on the repump intensity is not as stringent as in Ref. .
For purposes of illustration we can give numerical estimations for $`{}_{}{}^{133}Cs`$ ($`\mathrm{\Gamma }2\pi \mathrm{\hspace{0.17em}5}MHz`$ and $`\epsilon _r/\mathrm{}2\pi \mathrm{\hspace{0.17em}2}kHz`$). If we take the lattice beams detuning, $`\mathrm{\Delta }=2\pi \mathrm{\hspace{0.17em}10}GHz`$ (from the $`F=4F^{}=5`$ transition of the $`D_2`$ line) and intensity $`=500mW/cm^2`$, then the single-beam Stark shift $`u_s75\epsilon _r2\pi \mathrm{}\mathrm{\hspace{0.17em}150}kHz`$. The lattice has the depth $`6u_s=450\epsilon _r2\pi \mathrm{}\mathrm{\hspace{0.17em}900}kHz`$, and thus supports approximately $`15`$ bound levels with the energy separation $`\mathrm{}\omega _v=30\epsilon _r2\pi \mathrm{}\mathrm{\hspace{0.17em}60}kHz`$. With the tilt angle $`\mathrm{tan}(\varphi )0.1`$, the Raman transition rate is $`U_R0.1\mathrm{}\omega _v`$, providing the figure of merit $`U_R/\gamma _s121`$. The stray magnetic field must be controlled to the level of a few $`mG`$ or less. In the case of $`Cs`$ the pumping field should be applied to the $`F=4F^{\prime \prime }=4`$ transition of the $`D_1`$ line. The repumping field should be tuned to resonance on the $`F=3F^{}=4`$ transition of the $`D_2`$ line, and have an intensity $`10mW/cm^2`$ in order to saturate the transitions from all Zeeman sublevels. The optimal pumping field detuning $`\mathrm{\Delta }_p0.2\mathrm{\Gamma }\omega _v/U_R2\mathrm{\Gamma }`$ and intensity $`_p8mW/cm^2`$ give the cooling rate $`\gamma _{cool}0.4U_R2\pi \mathrm{\hspace{0.17em}2.2}kHz`$, and time of $`\tau \gamma _{cool}^110^4s`$. For these conditions our analysis predict approximately $`95\%`$ of the population will accumulate in the target state $`|F=4,m=0,n=0`$ with lower levels having vibrational numbers up to $`n^{}0.1\mathrm{}\omega _v/\epsilon _r3`$.
## IV Coherence between Vibrational Levels
In the case of a symmetric field configuration for 2D and 3D lattices a degeneracy of the vibrational energy level structure occurs. For a 2D lattice (for example the field configuration of Ref. and our configuration) in the harmonic approximation, the $`n`$-th vibrational level contains $`n+1`$ sublevels $`\{|m,n_x+n_y=n\}`$. We find that the coherence induced between the degenerate or near-degenerate vibrational levels can play an important role, and significantly change the efficiency of the resonance Raman coupling. For instance, consider a field configuration that differs from the scheme presented in Fig. 1.a by the direction and polarization of the pumping beam. Let this circularly polarized beam propagate along the $`z`$-axis as is shown in Fig. 4.a. In this case atoms are optically pumped to the outermost Zeeman substate $`|m=F`$ (see in Fig. 4.b). If we consider two degenerate vibrational levels, for example $`|m=F,n_x=1,n_y=0`$ and $`|m=F,n_x=0,n_y=1`$ coupled by Raman transitions with the unique state $`|m=F1,n_x=n_y=0`$ (Fig. 5.a), we find that there exists a superposition of the states of the degenrate vibrational level
$$|\mathrm{\Psi }_{NC}=|m=F,n_x=0,n_y=1+i|m=F,n_x=1,n_y=0$$
that is not coupled with $`|m=F1,n_x=n_y=0`$ by the operator $`\widehat{U}^{(1)}`$ (6): $`\widehat{U}^{(1)}|\mathrm{\Psi }_{NC}=0`$. Hence, part of the population will be trapped in this superposition state, in an analogy with well-known coherent population trapping in the $`\mathrm{\Lambda }`$-scheme . The result of such trapping is that the cooling efficiency is reduced, i.e. the part of atoms accumulated in the target state $`|m=F,n_x=n_y=0`$ will be significantly less than unity. In the case of coupling between higher levels $`|m=F,n_x+n_y=n`$ and $`|m=F1,n_x+n_y=n1`$, there always exists a coherent superposition of the sublevels $`|m=F,n_x+n_y=n`$, for which the operator of the Raman transitions is equal to zero, as it is for the light-induced $`\mathrm{\Lambda }`$-chains . However, for higher vibrational levels the anharmonicity has to be taken into account and the degeneracy is partially broken. If the vibrational energy levels are only nearly degenerate, due to either the anharmonicity or the optical lattice asymmetry, dark superpositions are not completely stable due to small energy splitting and atoms pass eventually from $`|m=F,n_x+n_y=n`$ to $`|m=F1,n_x+n_y=n1`$. But the corresponding transfer time will be greater than the normal $`U_R^1`$, leading to an increase of cooling time.
In the scheme presented in Fig. 1.a, this unwanted coherence effect is avoided by the simultaneous Raman coupling of the two states $`|m=0,n_x+n_y=1`$ of the degenerate vibrational level, with the two other states $`|m=\pm 1,n_x=n_y=0`$ with different amplitudes, as is shown in Fig. 5.b. It can be seen from this figure that the dark superposition for the transition with $`\mathrm{\Delta }m=1`$ is
$$(X+iY)\widehat{F}_{}|\mathrm{\Psi }_{NC}^{()}=0,|\mathrm{\Psi }_{NC}^{()}=|m=0,n_x=0,n_y=1+i|m=0,n_x=1,n_y=0.$$
While for the transition with $`\mathrm{\Delta }m=+1`$ the uncoupled state is given by
$$(XiY)\widehat{F}_+|\mathrm{\Psi }_{NC}^{(+)}=0,|\mathrm{\Psi }_{NC}^{(+)}=|m=0,n_x=0,n_y=1i|m=0,n_x=1,n_y=0.$$
The states $`|\mathrm{\Psi }_{NC}^{()}`$ and $`|\mathrm{\Psi }_{NC}^{(+)}`$ are orthogonal, so no a superposition can be made which nullifies the Raman coupling operator $`\widehat{U}^{(1)}`$ (6). We also note that the coherence within the vibrational structure might be very useful for other purposes, for instance in quantum state preparation.
## V Conclusion
Concluding, we have proposed a new scheme for 2D Raman-sideband-cooling to the zero-point energy in a far-off-resonance optical lattice and simultaneously pumping atoms to the $`m=0`$ dark state. The main distinguishing features of our proposal is that the method uses a pumping field to Stark-shift the Raman coupling to the red sideband and thus accumulation cold atoms in the $`m=0`$ Zeeman sublevel. An elementary theoretical consideration allowed us to calculate the dependence for the cooling efficiency and for the cooling dynamics. Our estimates for a cold $`Cs`$ show that for easily realizable experimental parameters up to $`95\%`$ of atoms can be accumulated in the $`|F=4,m=0,n=0`$ state in a millisecond time scale. This corresponds to a kinetic temperature on the order of $`100nK`$ after adiabatic release from the lattice . Besides, coherent population trapping between degenerate vibrational levels has been predicted for the sideband Raman transitions.
The authors thank Dr. J. Kitching and Prof. P. Jessen for helpful discussions. This work was supported in part by the Russian Fund for Basic Research (Grant No. 98-02-17794). AVT and VIYu acknowledge the hospitality of NIST, Boulder. |
no-problem/9911/math9911224.html | ar5iv | text | # Lattices in 𝐑² and finite subsets of a circle
## Spaces of finite subsets
For $`X`$ a topological space let $`\mathrm{exp}_kX`$ be the set of all finite subsets of $`X`$ of cardinality at most $`k`$. There is a map from the Cartesian product of $`k`$ copies of $`X`$ with itself to $`\mathrm{exp}_kX`$ which sends $`(x_1,\mathrm{},x_k)`$ to $`\{x_1\}\mathrm{}\{x_k\}`$. The quotient topology gives $`\mathrm{exp}_kX`$ the structure of a topological space. Clearly, $`\mathrm{exp}_1X=X`$ for any $`X`$.
The simplest non-trivial example is provided by the space $`\mathrm{exp}_2S^1`$ which is easily seen to be homeomorphic to the Möbius band. The space $`\mathrm{exp}_3S^1`$ is described by a well-known theorem of R. Bott:
###### Theorem 1.
(Bott, ) The space $`\mathrm{exp}_3S^1`$ is homeomorphic to a 3-sphere $`S^3`$.
The homeomorphism between $`\mathrm{exp}_3S^1`$ and $`S^3`$ is rather non-obvious. This is illustrated by the following result which is, apparently, due to E. Shchepin. Consider the canonical embedding $`\mathrm{\Delta }:S^1\mathrm{exp}_3S^1=S^3`$ which sends a point $`xS^1`$ to the subset $`\{x\}\mathrm{exp}_3S^1`$.
###### Theorem 2.
(Shchepin, \[unpublished\]) The map $`\mathrm{\Delta }:S^1S^3`$ is a trefoil knot.
In his proof of Theorem 1 Bott used a “cut-and-paste” argument. Shchepin’s proof of Theorem 2 was based on a direct calculation of the fundamental group of $`S^3\backslash \mathrm{\Delta }(S^1)`$. The purpose of this note is to show that the above theorems can be deduced from well-known facts about lattices in a plane.
## Spaces of lattices.
A lattice in $`𝐑^2`$ is a subgroup of the vector space $`𝐑^2`$ generated by two linearly independent vectors. The set $``$ of all lattices considered up to multiplication by a non-zero real number can be identified with $`SL(2,𝐑)/SL(2,𝐙)`$ and will be considered with this topology.
The space $``$ can be compactified by adding points which correspond to degenerate lattices, that is, subgroups of $`𝐑^2`$ generated by one non-zero vector. Degenerate lattices modulo multiplication by a non-zero real number form a circle $`S^1`$. The topology on the union $`\widehat{}=S^1`$ can be described as follows. Denote by $`\widehat{T}`$ the space of triangles of perimeter 1 in $`𝐑^2`$ each of whose angles does not exceed $`\pi /2`$ and is possibly equal to zero. The space $`\widehat{T}`$ contains degenerate triangles which have one side equal to 0 and both angles adjacent to it equal to $`\pi /2`$. Let $`T\widehat{T}`$ be the subspace of non-degenerate triangles. There is a map $`p:\widehat{T}\widehat{}`$ which sends a triangle to the lattice generated by any two of its sides considered as vectors. Notice that the restriction of $`p`$ to $`T`$ is a continuous map from $`T`$ onto $``$. We define the topology on $`\widehat{}`$ as the quotient topology with respect to the map $`p`$.
###### Theorem 3.
The compactification $`\widehat{}`$ of the space of lattices $``$ is homeomorphic to a 3-sphere $`S^3`$. The inclusion $`\widehat{}\backslash \widehat{}`$ is a trefoil knot.
The proof of the above theorem (due to D.Quillen) can be found in on page 84 where a homeomorphism between $``$ and the complement of a trefoil is constructed. This map can be easily seen to extend to a homeomorphism between $`\widehat{}`$ and $`S^3`$.
## Main theorem
Our main result relates Theorem 3 to Theorems 1 and 2.
###### Theorem 4.
There is a homeomorphism $`\mathrm{\Phi }:\widehat{}\mathrm{exp}_3S^1`$ which identifies the circle $`\widehat{}\backslash `$ with $`\mathrm{\Delta }(S^1)`$.
###### Remark.
The circle $`S^1`$ acts as a subgroup of $`PSL(2,𝐑)`$ on $``$ (and, in fact, on $`\widehat{}`$) rotating the lattices. It will be clear from the construction below that $`\mathrm{\Phi }`$ sends the action of $`S^1`$ on $`\widehat{}`$ to the action of $`SO(2)`$ by rotations on $`\mathrm{exp}_3S^1`$.
###### Proof.
For any non-degenerate lattice $`L`$ there is a finite number of triangles $`V_i^L`$ with vertices in $`L`$ whose sides are generators of $`L`$ and whose angles do not exceed $`\pi /2`$. There are 12 such triangles for a rectangular lattice and 6 for any other lattice, as shown in the figure.
For each non-degenerate lattice $`L`$ choose one of the triangles $`V_i^L`$ and trace 3 lines which connect the vertices of the triangle with the centre of its circumscribed circle. These lines are distinct unless $`L`$ is rectangular in which case two of them coincide. Each line determines a point in $`\mathrm{𝐑𝐏}^1=S^1`$ so we obtain a point $`\mathrm{\Phi }(L)`$ in $`\mathrm{exp}_3S^1\backslash \mathrm{\Delta }(S^1)`$ for each $`L`$. It is clear that $`\mathrm{\Phi }(L)`$ depends only on $`L`$ and not on a particular choice of a triangle. Now, for a degenerate lattice $`L`$ define $`\mathrm{\Phi }(L)`$ as the point of $`\mathrm{𝐑𝐏}^1`$ which corresponds to the line containing $`L`$. A straightforward check shows that $`\mathrm{\Phi }`$ is a homeomorphism. ∎
## Acknowledgments
I would like to thank Evgenii Shchepin, Abdelghani Mouhallil and Alberto Verjovsky for useful discussions. |
no-problem/9911/astro-ph9911212.html | ar5iv | text | # ACCRETION DISCS AROUND BLACK HOLES: DEVELOPEMENT OF THEORY
ABSTRACT. Standard accretion disk theory is formulated which is based on the local heat balance. The energy produced by a turbulent viscous heating is supposed to be emitted to the sides of the disc. Sources of turbulence in the accretion disc are connected with nonlinear hydrodynamic instability, convection, and magnetic field. In standard theory there are two branches of solution, optically thick, and optically thin. Advection in accretion disks is described by the differential equations what makes the theory nonlocal. Low-luminous optically thin accretion disc model with advection at some suggestions may become advectively dominated, carrying almost all the energy inside the black hole. The proper account of magnetic filed in the process of accretion limits the energy advected into a black hole, efficiency of accretion should exceed $`1/4`$ of the standard accretion disk model efficiency.
Key words: Stars: accretion discs; black holes.
1. Introduction
Accretion is a main source of energy in binary X-ray sources, quasars and active galactic nuclei (AGN). The intensive development of accretion theory began after discovery of X-ray sources and quasars. Accretion into stars is ended by a collision with an inner boundary, which may be a stellar surface, or outer boundary of a magnetosphere for strongly magnetized stars. All gravitational energy of the falling matter is transformed into heat and radiated outward.
In black holes matter is falling to the horizon, from where no radiation arrives. All luminosity is formed on the way to it. The efficiency of accretion is not known from the beginning, and depends on angular momentum of the falling matter, and magnetic field embedded into it. It was first shown by Schwartsman (1971) that during spherical accretion of nonmagnetized gas the efficiency may be as small as $`10^8`$ for sufficiently low mass fluxes. He had shown that presence of magnetic field in the accreting matter increase the efficiency up to about $`10\%`$, and account of heating of matter due to magnetic field annihilation rises the efficiency up to about $`30\%`$ (Bisnovatyi-Kogan and Ruzmaikin, 1974) In the case of a thin disc accretion, when matter has large angular momentum, the efficiency is about $`1/2`$ of the efficiency of accretion into a star with a radius equal to the radius of the last stable orbit. In the case of geometrically thick and optically thin accretion discs the situation is approaching the case of spherical symmetry, where presence of a magnetic field plays a critical role.
Advection dominated accretion flow (ADAF) was suggested by Narayan and Yu (1995), and used as a solution for some astrophysical problems. The suggestions underlying ADAF: ignorance of the magnetic field annihilation in heating of a plasma flow, electron heating only due to binary collisions with protons (ions) had been critically analyzed in papers of Bisnovatyi-Kogan and Lovelace (1997, 1999), Bisnovatyi-Kogan (1999), Quataert (1997). There are contradictions between ADAF model and observational data in radioemission of elliptical galaxies (Di Matteo et al., 1999), and X-ray emission of Seyfert galaxy NGC4258 (Cannizzo, 1998). Account of processes connected with a small-scale magnetic field in accretion flow, strongly restricts solution. Namely, the efficiency of the accretion flow cannot become less then about 1/4 of the standard accretion disc value.
2. Development of the standard model
Matter falling into a black hole forms a disc when its angular momentum is sufficiently high. It happens when matter comes from the star companion in the binary, or from a tidal disruption of the star which trajectory approaches close to the black hole. The first situation is observed in many galactic X-ray sources (Cherepashchuk, 1996). A tidal disruption happens in quasars and active galactic nuclei (AGN) in the model of supermassive black hole surrounded by a dense stellar cluster.
Equations of a standard accretion disk theory had been first formulated by (Shakura, 1972); some corrections and generalization to general relativity (GR) had been done by Novikov and Thorne (1973). Observational aspects of accretion disks have been analyzed by Shakura and Sunyaev (1973). Note, that all authors of the accretion disc theory from USSR were students (N.I.Shakura) or collaborators (I.D.Novikov and R.A.Sunyaev) of academician Ya.B.Zeldovich, who was not among the authors, but whose influence on them hardly could be overestimated. The main idea of this theory is to describe a geometrically thin non-self-gravitating disc of a mass $`M_d`$, much smaller then the mass of a black hole $`M`$, by hydrodynamic equations averaged over the disc thickness $`2h`$.
2.1. Equations
The small thickness of the disc in comparison with its radius $`hr`$ means small importance of the pressure gradient $`P`$ in comparison with gravity and inertia forces. Radial equilibrium equation in a disc is a balance between the last two forces with an angular velocity equals to the keplerian one $`\mathrm{\Omega }=\mathrm{\Omega }_K=\left(\frac{GM}{r^3}\right)^{1/2}`$. Note, that just before a last stable orbit around a black hole this suggestion fails, but in the ”standard” accretion disc model this relation is supposed to hold all over the disc, with an inner boundary at the last stable orbit. The equilibrium equation in the vertical $`z`$-direction is determined by a balance between the gravitational force and pressure gradient $`\frac{dP}{dz}=\rho \frac{GMz}{r^3}`$. For a thin disc this differential equation is substituted by an algebraic one, determining the half-thickness of the disc in the form
$$h\frac{1}{\mathrm{\Omega }_K}\left(2\frac{P}{\rho }\right)^{1/2}.$$
(1)
The balance of angular momentum, related to the $`\varphi `$ component of the Euler equation has an integral in a stationary case, written as
$$\dot{M}(jj_{in})=2\pi r^2\mathrm{\hspace{0.17em}2}ht_{r\varphi },t_{r\varphi }=\eta r\frac{d\mathrm{\Omega }}{dr}.$$
(2)
Here $`j=v_\varphi r=\mathrm{\Omega }r^2`$ is a specific angular momentum, $`t_{r\varphi }`$ is a component of the viscous stress tensor, $`\dot{M}>0`$ is a mass flux per unit time into a black hole, $`j_{in}`$ is equal to the specific angular momentum of matter falling into a black hole. In the standard theory the value of $`j_{in}`$ is determined from physical considerations. For accretion into a black hole it is suggested, that on the last stable orbit the gradient of the angular velocity is zero, corresponding to zero viscous momentum flux. In that case $`j_{in}=\mathrm{\Omega }_Kr_{in}^2,`$ corresponding to the Keplerian angular momentum of the matter on the last stable orbit.
The choice of the viscosity coefficient is the most speculative problem of the theory. In the laminar case at low microscopic (atomic or plasma) viscosity the stationary accretion disc must be very massive and thick, and before its formation the matter is collected by disc leading to a small flux inside. It contradicts to observations of X-ray binaries, where a considerable matter flux along the accretion disc may be explained only when viscosity coefficient is much larger. It was suggested by Shakura (1972), that matter in the disc is turbulent, what determines a turbulent viscous stress tensor, parametrized by a pressure
$$t_{r\varphi }=\alpha \rho v_s^2=\alpha P,$$
(3)
where $`v_s`$ is a sound speed in the matter. This simple presentation comes out from a relation for a turbulent viscosity coefficient $`\eta _t\rho v_tl`$ with an average turbulent velocity $`v_t`$ and mean free path of the turbulent element $`l`$. It follows from the definition of $`t_{r\varphi }`$ in (2), when we take $`lh`$ from (1)
$$t_{r\varphi }=\rho v_thr\frac{d\mathrm{\Omega }}{dr}\rho v_tv_s=\alpha \rho v_s^2,$$
(4)
where a coefficient $`\alpha <1`$ is connecting the turbulent and sound speeds $`v_t=\alpha v_s`$. Presentations of $`t_{r\varphi }`$ in (3) and (4) are equivalent, and only when the angular velocity differs considerably from the Keplerian one the first relation to the right in (4) is more preferable. That does not appear in the standard theory, but may happen when advective terms are included.
Development of a turbulence in the accretion disc cannot be justified simply, because a Keplerian disc is stable in linear approximation to the development of perturbations. It was suggested by Ya.B.Zeldovich, that in presence of very large Reynolds number $`\mathrm{Re}=\frac{\rho vl}{\eta }`$ the amplitude of perturbations at which nonlinear effects become important is very low, so a turbulence may be developed due to nonlinear instability even when the disc is stable in linear approximation. Viscous stresses may arise from a magnetic field, it was suggested by (Shakura, 1972), that magnetic stresses cannot exceed the turbulent ones. It was shown by Bisnovatyi-Kogan and Blinnikov (1976), that inner regions of a highly luminous accretion discs where pressure is dominated by radiation, are unstable to vertical convection. Development of this convection produce a turbulence, needed for a high viscosity.
With alpha- prescription of viscosity the equation of angular momentum conservation is written in the plane of a disc as
$$\dot{M}(jj_{in})=4\pi r^2\alpha P_0h.$$
(5)
When angular velocity is far from Keplerian one the relation (2) is valid with a coefficient of a turbulent viscosity $`\eta =\alpha \rho _0v_{s0}h`$, where values with the index ”0” denote the plane of the disc.
In the standard theory a heat balance is local, all heat produced by viscosity in the ring between $`r`$ and $`r+dr`$ is radiated through the sides of disc at the same $`r`$. The heat production rate $`Q_+`$ related to the surface unit of the disc is written as
$$Q_+=ht_{r\varphi }r\frac{d\mathrm{\Omega }}{dr}=\frac{3}{8\pi }\dot{M}\frac{GM}{r^3}\left(1\frac{j_{in}}{j}\right).$$
(6)
Heat losses by a disc depend on its optical depth. The standard disc model (Shakura, 1972) considered a geometrically thin disc as an optically thick in a vertical direction. That implies energy losses $`Q_{}`$ from the disc due to a radiative conductivity, after a substitution of the differential equation of a heat transfer by an algebraic relation
$$Q_{}\frac{4}{3}\frac{acT^4}{\kappa \mathrm{\Sigma }}.$$
(7)
Here $`a`$ is a constant of a radiation energy density, $`c`$ is a speed of light, $`T`$ is a temperature in the disc plane, $`\kappa `$ is a matter opacity, and a surface density $`\mathrm{\Sigma }=2\rho h`$. Here and below $`\rho ,T,P`$ without the index ”0” are related to the disc plane. The heat balance equation is represented by a relation $`Q_+=Q_{}`$. A continuity equation in the standard model is used for finding of a radial velocity $`v_r`$
$$v_r=\frac{\dot{M}}{4\pi rh\rho }=\frac{\dot{M}}{2\pi r\mathrm{\Sigma }}.$$
(8)
Completing these equations by an equation of state $`P(\rho ,T)`$ and relation for the opacity $`\kappa =\kappa (\rho ,T)`$ we get a full set of equations for a standard disc model. For power low equations of state of an ideal gas $`P=P_g=\rho T`$ ($``$ is a gas constant), or radiation pressure $`P=P_r=\frac{aT^4}{3}`$, and opacity in the form of electron scattering $`\kappa _e`$, or Karammers formulae $`\kappa _k`$, the solution of a standard disc accretion theory is obtained analytically. Checking the suggestion of a large optical thickness confirms a self-consistency of the model. Note that solutions for different regions of the disc with different equation of states and opacities are not matched to each other.
2.2. Optically thin solution
It was found by Shapiro et al. (1976) that there is another branch of the solution for a disc structure with the same input parameters $`M,\dot{M},\alpha `$ which is also self-consistent but has a small optical thickness. That implies another equation of energy losses, determined by a volume emission $`Q_{}q\rho h`$. The emissivity of the unit of a volume $`q`$ is connected with a Planckian averaged opacity $`\kappa _p`$ by a relation $`qacT_0^4\kappa _p`$. In the optically thin limit the pressure is determined by a gas $`P=P_g`$. In the optically thin solution the thickness of the disc is larger then in the optically thick one, and density is lower.
While heating by viscosity is determined mainly by heavy ions, and cooling is determined by electrons, the rate of the energy exchange between them is important for a structure of the disc. The energy balance equations are written separately for ions and electrons. For small accretion rates and lower matter density the rate of energy exchange due to binary collisions is so slow, that in the thermal balance the ions are much hotter then the electrons. That also implies a high disc thickness. In the highly turbulent plasma the energy exchange between ions and electrons may be strongly enhanced due to presence of fluctuating electrical fields, where electrons and ions gain the same energy. In such conditions difference of temperatures between ions and electrons may be negligible. The theory of relaxation in the turbulent plasma is not completed, but there are indications to a large enhancement of the relaxation in presence of plasma turbulence, in comparison with the binary collisions (Galeev and Sagdeev, 1983; Quataert, 1997).
2.3. Accretion disc structure from equations describing continuously optically thin and thick regions
Equations of the disc structure smoothly describing transition between optically thick and optically thin disc, had been obtained using Eddington approximation. The expressions had been obtained (Artemova et al., 1996) for the vertical energy flux from the disk $`F_0`$, and radiation pressure in the symmetry plane
$$F_0=\frac{2acT_0^4}{3\tau _0\mathrm{\Phi }},P_{rad,0}=\frac{aT_0^4}{3\mathrm{\Phi }}\left(1+\frac{4}{3\tau _0}\right),$$
(9)
where $`\tau _{\alpha 0}=\kappa _p\rho h=\frac{1}{2}\kappa _p\mathrm{\Sigma }_0`$, $`\tau _{}=\left(\tau _0\tau _{\alpha 0}\right)^{1/2}`$, $`\mathrm{\Phi }=1+\frac{4}{3\tau _0}+\frac{2}{3\tau _{}^2}`$. At $`\tau _0\tau _{}1`$ we have (7) from (9). In the optically thin limit $`\tau _{}\tau _01`$ we get
$$F_0=acT_0^4\tau _{\alpha 0},P_{rad,0}=\frac{2}{3}acT_0^4\tau _{\alpha 0}.$$
(10)
Using $`F_0`$ instead of $`Q_{}`$ and equation of state $`P=\rho T+P_{rad,0}`$, the equations of accretion disc structure together with equation $`Q_+=F_0`$, with $`Q_+`$ from (6), have been solved numerically by Artemova et al. (1996). Two solutions, optically thick and thin, exist separately when luminosity is not very large. They intersect at $`\dot{m}=\dot{m}_b`$ and there is no global solution for accretion disc at $`\dot{m}>\dot{m}_b`$. It was concluded by Artemova et al. (1996) that in order to obtain a global physically meaningful solution at $`\dot{m}>\dot{m}_b`$, account of advectivion is needed. For the calculated case $`M_{BH}=10^8M_{}`$, $`\alpha =1.0`$ at $`\dot{m}=\dot{m}_b`$ luminosity of the accretion disk is less than the critical Eddington one.
3. Accretion discs with advection
Standard model gives somewhat nonphysical behavior near the inner edge of the accretion disc around a black hole, with a zero heat production at the inner edge of the disk. It is clear from physical ground, that in this case the heat brought by radial motion of matter along the accretion disc becomes important. In presence of this advective heating (or cooling) term, depending on the radial entropy $`S`$ gradient, written as $`Q_{adv}=\frac{\dot{M}}{2\pi r}T\frac{dS}{dr}`$, the equation of a heat balance is modified to
$$Q_++Q_{adv}=Q_{}.$$
(11)
In order to describe self-consistently the structure of the accretion disc we should also modify the radial disc equilibrium, including pressure and inertia terms
$$r(\mathrm{\Omega }^2\mathrm{\Omega }_K^2)=\frac{1}{\rho }\frac{dP}{dr}v_r\frac{dv_r}{dr}.$$
(12)
Appearance of inertia term leads to transonic radial flow with a singular point. Conditions of a continuous passing of the solution through a critical point choose a unique value of the integration constant $`j_{in}`$. First approximate solution for the advective disc structure have been obtained by Paczyński and Bisnovatyi-Kogan (1981). Attempts to find a solution for advective disc had been done by Abramovicz et al. (1988), Matsumoto et al. (1984). For moderate values of $`\dot{M}`$ a unique continuous transonic solution was found, passing through singular points, and corresponding to a unique value of $`j_{in}`$. The number of critical point in the radial flow with the gravitational potential $`\varphi _g`$ (Paczyński and Wiita, 1980) $`\varphi _g=\frac{GM}{rr_g},r_g=\frac{2GM}{c^2}`$. may exceed unity. Appearance of two critical points for a radial flow in this potential was analyzed by Chakrabarti and Molteni (1993). Using of equations averaged over a thickness of the disc changes a structure of hydrodynamic equations, leading to a position of singular points not coinciding with a unit Mach number point.
4. Amplification of the magnetic field at a spherical accretion
A matter flowing into a black hole is usually magnetized. Due to more rapid increase of a magnetic energy the role of the magnetic field increases when matter flows inside. It was shown by Schwartsman (1971), that magnetic energy density $`E_M`$ approaches a density of a kinetic energy $`E_k`$, and he proposed a hypothesis of equipartition $`E_ME_k`$, supported by continuous annihilation of the magnetic field in a region of the main energy production. This hypothesis is usually accepted in the modern picture of accretion (Narayan and Yu, 1995). Account of the heating of matter by magnetic field annihilation was done by Bisnovatyi-Kogan and Ruzmaikin (1974). For a spherical accretion with $`𝐯=(v_r,\mathrm{\hspace{0.17em}0},\mathrm{\hspace{0.17em}0})`$ the equations describing a field amplification in the ideally conducting plasma reduce to (Bisnovatyi-Kogan, Ruzmaikin, 1974)
$$\frac{d(r^2B_r)}{dt}=0,\frac{d(rv_rB_\theta )}{dt}=0,\frac{d(rv_rB_\varphi )}{dt}=0,$$
(13)
where $`\frac{d}{dt}=\frac{}{t}+v_r\frac{}{r}`$ is a full Lagrangian derivative. Consider a free fall case with $`v_r=\sqrt{\frac{2GM}{r}}`$. The initial condition problem is solved separately for poloidal and toroidal fields. For initially uniform field $`B_{r0}=B_0\mathrm{cos}\theta ,B_{\theta 0}=B_0\mathrm{sin}\theta `$ we get the solution (Bisnovatyi-Kogan, Ruzmaikin, 1974)
$$B_r=\frac{B_0\mathrm{cos}\theta }{r^2}\mathrm{\Phi }_1^{4/3},B_\theta =\frac{B_0\mathrm{sin}\theta }{\sqrt{r}}\mathrm{\Phi }_1^{1/3},$$
(14)
where $`\mathrm{\Phi }_1=r^{3/2}+\frac{3}{2}t\sqrt{2GM}`$. The radial component of the field is growing most rapidly. It is $`r^2`$ for large times, $`t^{4/3}`$ at given small radius, and is growing with time everywhere. For initially dipole magnetic field
$$B_{r0}=\frac{B_0\mathrm{cos}\theta }{r^3},B_{\theta 0}=\frac{B_0\mathrm{sin}\theta }{2r^3}$$
we obtain the following solution
$$B_r=\frac{B_0\mathrm{cos}\theta }{r^2}\mathrm{\Phi }_1^{2/3},B_\theta =\frac{B_0\mathrm{sin}\theta }{2\sqrt{r}}\mathrm{\Phi }_1^{5/3}.$$
(15)
Here the magnetic field is decreasing everywhere with time, tending to zero. That describes a pressing of a dipole magnetic field to a stellar surface. The azimuthal stellar magnetic field if confined inside the star. When outer layers of the star are compressing with a free-fall speed, then for initial field distribution $`B_{\varphi 0}=B_0r^nf(\theta )`$ the change of $`B_\varphi `$ with time is described by a relation $`B_\varphi =\frac{B_0f(\theta )}{\sqrt{r}}\mathrm{\Phi }_1^{n+1/3}`$.
5. Two-temperature advective discs
In the optically thin accretion discs at low mass fluxes the density of the matter is low and energy exchange between electrons and ions due to binary collisions is slow. In this situation, due to different mechanisms of heating and cooling for electrons and ions, they may have different temperatures. First it was realized by Shapiro et al. (1976), where advection was not included. It was noticed by Narayan and Yu (1995), that advection in this case is becoming extremely important. It may carry the main energy flux into a black hole, leaving rather low efficiency of the accretion up to $`10^3\mathrm{\hspace{0.17em}10}^4`$ (advective dominated accretion flows - ADAF). This conclusion is valid only when the effects, connected with heating of matter by magnetic field annihilation are neglected.
In the ADAF solution the ion temperature is about a virial one $`kT_iGMm_i/r`$, what means that even at high initial angular momentum the disc becomes thick, forming a quasi-spherical accretion flow. When energy losses by ions are low, some kind of a ”thermo-viscous” instability is developed, because heating increases a viscosity, and viscosity increases a heating. Development of this instability leads to formation of ADAF.
A full account of the processes, connected with a presence of magnetic field in the flow, is changing considerably the picture of ADAF. It was shown by Schwarzman (1971), that in the region of the main energy production, the condition of equipartition takes place, and efficiency of a radiation increase enormously from $`10^8`$ up to $`0.1`$ due to magneto-bremstrahlung. To support the condition of equipartition a continuous magnetic field reconnection and heating of matter due to Ohmic dissipation takes place. It was shown by Bisnovatyi-Kogan and Ruzmaikin (1974), that due to Ohmic heating the efficiency of a radial accretion into a black hole may become as high as $`30\%`$. The rate of the Ohmic heating in the condition of equipartition was obtained in the form
$$T\frac{dS}{dr}=\frac{3}{2}\frac{B^2}{8\pi \rho r}.$$
(16)
In the supersonic flow of the radial accretion equipartition was suggested in a form (Schwartsman, 1971) $`\frac{B^2}{8\pi }\frac{\rho v_r^2}{2}=\frac{\rho GM}{r}`$. For the disc accretion equipartition between magnetic and turbulent energy was suggested by Shakura (1972), what reduces with account of ”alpha” prescription of viscosity to a relation $`\frac{B^2}{8\pi }\frac{\rho v_t^2}{2}\alpha _m^2P`$, where $`\alpha _m`$ characterizes a magnetic viscosity in a way similar to the turbulent $`\alpha `$ viscosity. It was suggested by Bisnovatyi-Kogan and Ruzmaikin (1976) the similarity between viscous and magnetic Reynolds numbers, or between turbulent and magnetic viscosity coefficients $`Re=\frac{\rho vl}{\eta },Re_m=\frac{\rho vl}{\eta _m}`$, where the turbulent magnetic viscosity $`\eta _m`$ is connected with a turbulent conductivity $`\sigma =\frac{\rho c^2}{4\pi \eta _m}`$. Taking $`\eta _m=\frac{\alpha _m}{\alpha }\eta `$, we get a turbulent conductivity
$$\sigma =\frac{c^2}{4\pi \alpha _mhv_s},v_s^2=\frac{P_g}{\rho }$$
(17)
in the optically thin discs. For the radial accretion the turbulent conductivity may contain mean free path of a turbulent element $`l_t`$ in (17) instead of $`h`$. In ADAF solutions, where ionic temperature is of the order of the virial one two above suggestions for magnetic equipartition almost coincide at $`\alpha _m1`$.
The heating of the matter due to an Ohmic dissipation may be obtained from the Ohm’s law in radial accretion
$$T\frac{dS}{dr}=\frac{\sigma ^2}{\rho v_r}\sigma \frac{v_E^2B^2}{\rho v_rc^2}=\frac{B^2v_E^2}{4\pi \rho \alpha _mv_rv_sl_t},$$
(18)
what coincides with (16) when $`\alpha _m=\frac{4rv_E^2}{3v_rv_sl_t}`$, or $`l_t=\frac{4rv_E^2}{3v_rv_s\alpha _m}`$. Here a local electrical field strength in a highly conducting plasma is of the order of $`\frac{v_EB}{c}`$, $`v_Ev_t\alpha v_s`$ for a radial accretion.
Equations for a radial temperature dependence in the accretion disc, separately for the ions and electrons are written as
$$\frac{dE_i}{dt}\frac{P_i}{\rho ^2}\frac{d\rho }{dt}=_{\eta i}+_{Bi}Q_{ie},$$
(19)
$$\frac{dE_e}{dt}\frac{P_e}{\rho ^2}\frac{d\rho }{dt}=_{\eta e}+_{Be}+Q_{ie}𝒞_{ff}𝒞_{cyc},$$
(20)
A rate of a viscous heating of ions $`_{\eta i}`$ is obtained from (6) as
$$_{\eta i}=\frac{2\pi r}{\dot{M}}Q_+=\frac{3}{2}\alpha \frac{v_Kv_s^2}{r},_{\eta e}\sqrt{\frac{m_e}{m_i}}_{\eta i}.$$
(21)
Combining (1),(8),(5), we get
$$v_r=\alpha \frac{v_s^2}{v_K𝒥},h=\sqrt{2}\frac{v_s}{v_K}r,\rho =\frac{\dot{M}}{4\pi \alpha \sqrt{2}}\frac{v_K^2𝒥}{r^2v_s^3},$$
(22)
where $`v_K=r\mathrm{\Omega }_K`$, $`𝒥=1\frac{j_{in}}{j}`$. The rate of the energy exchange between ions and electrons due to the binary collisions was obtained by Landau (1937). Neglecting pair formation for a low density accretion disc, we may write an exact expression for a pressure $`P_g=P_i+P_e=n_ikT_e+n_ekT_p=n_ek(T_e+T_i)`$, and an approximate expression for an energy, containing a smooth interpolation between nonrelativistic and relativistic electrons. The bremstrahlung $`𝒞_{ff}`$ and magneto-bremstrahlung $`𝒞_{cyc}`$ cooling of maxwellian semi-relativistic electrons, with account of free-bound radiation in nonrelativistic limit, may be written as interpolation of limiting cases (Bisnovatyi-Kogan and Ruzmaikin, 1976).
In the case of a disk accretion there are several characteristic velocities, $`v_K`$, $`v_r`$, $`v_s`$, and $`v_t=\alpha v_s`$, all of which may be used for determining ”equipartition” magnetic energy, and one characteristic length $`h`$. Consider three possible choices of $`v_B^2`$=$`v_K^2`$, $`v_r^2`$, and $`v_t^2`$ for scaling $`B^2=4\pi \rho v_B^2`$. The expression for an Ohmic heating in the turbulent accretion disc also may be written in different ways, using different velocities $`v_E`$ in the expression for an effective electrical field $`=\frac{v_EB}{c}`$. A self-consistency of the model requires, that expressions for a magnetic heating of the matter $`_B`$, obtained from the condition of stationarity of the flow (16), and from the Ohm’s law (18), should be identical. It happens at $`\frac{\alpha }{𝒥\alpha _m}\frac{v_E^2}{v_r^2}=\frac{3\sqrt{2}}{4}`$. That implies $`v_Ev_r\frac{\alpha v_s^2}{v_K\sqrt{𝒥}}\frac{v_tv_s}{v_K\sqrt{𝒥}}<v_t`$. In the advective models $`𝒥`$ is substituted by a function which is not zero at the inner edge of the disc. The heating due to magnetic field reconnection $`_B`$ in the equations (19), (20) may be written with account of (21) as $`_B=\frac{3}{16\pi }\frac{B^2}{r\rho }v_r=\frac{1}{2𝒥}_{\eta i}\left(\frac{v_B}{v_K}\right)^2`$. At $`v_B=v_K`$ the expressions for viscous and magnetic heating are almost identical. Observations of the magnetic field reconnection in the solar flares show (Tsuneta, 1996), that electronic heating prevails over the ionc one. Transformation of the magnetic energy into a heat is connected with the change of the magnetic flux, generation of the vortex electrical field, accelerating the particles.
The equations (19), (20) have been solved by Bisnovatyi-Kogan and Lovelace (1997) for nonrelativistic electrons, at $`v_B`$=$`v_K`$. The combined heating of the electrons and ions were taken as $`_e=(2g)_{\eta i},_e=g_{\eta i}`$. The results of calculations for $`g=0.5÷1`$ show that almost all energy of the electrons is radiated, so the relative efficiency of the two-temperature, optically thin disc accretion cannot become lower then 0.25. Increase of the term $`Q_{ie}`$ due to plasma turbulence may restore the relative efficiency to a value, corresponding to the optically thick disc.
6. Discussion
Observational evidences for existence of black holes inside our Galaxy and in the active galactic nuclei (Cherepashchuk, 1996; Ho, 1999) make necessary to revise theoretical models of the disc accretion. The improvements of a model are connected with account of advective terms and more accurate treatment of the magnetic field effects. Account of the effects connected with magnetic field annihilation does not permit to make a relative efficiency of the accretion lower then $`0.25`$ from the standard value. Strong relaxation connected with the plasma turbulence may increase the efficiency, making it close to unity. For explanation of underluminous galactic nuclei two possible ways may be suggested. One is based on a more accurate estimations of the accretion mass flow into the black hole, which could be overestimated. The second is based on existence of another mechanisms of the energy losses in the form of accelerated particles, like in the radio-pulsars, where these losses exceed strongly a radiation. This is very probable to happen in a presence of a large scale magnetic field which may be also responsible for a formation of the observed jets (Bisnovatyi-Kogan, 1999; Blandford and Begelman, 1999). We may suggest, that underlumilnous AGN loose main part of their energy to the formation of jets, like in SS 433. The search of the correlation between existence of jets and lack of the luminosity could be very informative.
Acknowledgements. The author is grateful for partial support to RFBR, grant 99-02-18180, and GPNT ”Astronomy”, grant 1.2.6.5.
References
Abramovicz M.A., Czerny B., Lasota J.P., Szuszkiewicz E.: 1988, ApJ, 332, 646.
Artemova I.V., Bisnovatyi-Kogan G.S., Björnsson G., Novikov I.D.: 1996, ApJ, 456, 119.
Bisnovatyi-Kogan G.S.: 1999, in ”Observational evidences for black holes in the universe”, ed. S.Chakrabarti, Kluwer, p.1.
Bisnovatyi-Kogan G.S., Blinnikov S.I.: 1976, Pisma Astron. Zh., 2, 489.
Bisnovatyi-Kogan G.S., Lovelace R.V.L.: 1997, ApJL, 486, L43.
Bisnovatyi-Kogan G.S., Lovelace R.V.L.: 1999, ApJ (accepted), astro-ph/9902344.
Bisnovatyi-Kogan G.S., Ruzmaikin A.A.: 1974, Ap. and Space Sci., 28, 45.
Bisnovatyi-Kogan, G.S., Ruzmaikin, A.A., 1976, Ap. and Space Sci., 42, 401.
Blandford R.D., Begelman M.C.: 1999, Month. Not. R.A.S., 303, 1.
Cannizzo J.K. et al.: 1998, AAS Abstracts, 192, 4103C.
Chakrabarti S.K., Molteni D.: 1993, ApJ, 417, 671.
Cherepashchuk A.M.: 1996, Uspekhi Fiz. Nauk., 166, 809.
Di Matteo T., Fabian A.C., Rees, M.J. et al.: 1999, Month. Not. R.A.S., 305, 492.
Galeev A.A., Sagdeev R.Z.: 1983, Chap. 6, in Handbook of Plasma Physics, Vol. 1, eds. Rosenbluth M.N., Sagdeev R.Z. (Amsterdam: North-Holland Pub.), Ch. 6.1
Ho L.: 1999, in ”Observ. evidences for b. h. in the Universe”, ed. S.Chakrabarti, Kluwer, p.157.
Landau L.D.: 1937, Zh. Exp. Theor. Phys. 7, 203.
Matsumoto R., Sato Sh., Fukue J., Okazaki A.T.: 1984, Publ. Astron. Soc. Japan, 36, 71.
Narayan R., Yu I.: 1995, ApJ, 452, 710.
Novikov I.D., Thorne K.S.: 1973, in Black Holes eds. C.DeWitt, B.DeWitt (New York: Gordon & Breach), p.345
Paczyński B., Bisnovatyi-Kogan G.S.: 1981, Acta Astron., 31, 283.
Paczyński B., Wiita P.J.: 1980, A&A, 88, 23.
Pringle J.E., Rees, M.J.: 1972, A&A, 21, 1.
Quataert E.: 1997, astro-ph/9710127.
Schwartsman V.F.: 1971, Soviet Astron., 15, 377.
Shakura N.I.: 1972, Astron. Zh., 49, 921; (1973, Sov. Astron., 16, 756)
Shakura N.I., Sunyaev R.A.: 1973, A&A, 24, 337.
Shapiro S.L., Lightman A.P., Eardley D.M.: 1976, ApJ, 204, 187.
Tsuneta S.: 1996, ApJ, 456, 840. |
no-problem/9911/hep-ph9911439.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Spontaneously broken global supersymmetric theories, generate a positive vacuum energy making a flat space solution inconsistent. As is well known the proper framework for discussing spontaneous breaking of supersymmetry is supergravity. The natural scale for such a theory is however the Planck scale $`M_{Pl}`$, which is many orders of magnitude (at least 15) larger than the expected SUSY breaking splitting between standard model particles and their superpartners. The question then arises as to how this wide separation of scales can be generated. In gravity mediated supersymmetry breaking schemes an intermediate scale is typically generated by the condensation of gauginos in some hidden sector gauge group. If $`S`$ is the dilaton, a possible superpotential for the hidden sector could thus take the form
$$W=M_{Pl}^3e^{\frac{S}{b}}+kM_{Pl}^3$$
(1)
with a constant $`k`$ in the superpotential whose origin is not understood at this point and $`b`$ a model dependent parameter. <sup>3</sup><sup>3</sup>3Another popular model, the so-called race track one where two or more gauge groups condense does not need a constant but appears to be ruled out on other grounds . Such a super potential can be used to stabilize the dilaton and give SUSY breaking at an intermediate scale $`\mathrm{\Lambda }=\sqrt{m_{3/2}M_{Pl}}`$ with zero cosmological constant. The latter in fact involves cancellation between two terms both of which are of order $`\mathrm{\Lambda }^4`$ and thus can be accomplished by fine tuning a number of order one. Generically however such gravity mediated SUSY breaking models will lead to large flavor changing neutral currents, unless the $`F`$ term of the dilaton dominates the SUSY breaking <sup>4</sup><sup>4</sup>4If the SUSY breaking is dilaton dominated this could be avoided but it is fair to say that there is no natural framework coming from string theory in which this can be realized..
There are two other scenarios that have been discussed in the literature that lead to SUSY breaking, such that no large flavor changing neutral currents are generated, both of which communicate the SUSY breaking from a hidden sector to the visible by means of gauge interactions rather than by gravity. One, is when the information of supersymmetry breaking is communicated to the visible sector by the usual gauge (non-Abelian) interactions . This mechanism is called Gauge Mediated Supersymmetry Breaking (GMSB), an old idea recently revived in . The quantity that sets the scale for the soft masses is $`\frac{\alpha }{4\pi }\frac{F_X}{<X>}`$, where $`F_X`$ and $`<X>`$ are the highest and lowest components of a chiral superfield $`X`$ in the hidden sector. Phenomenological reasons <sup>5</sup><sup>5</sup>5One of the constraints on $`<X>`$ comes from standard Big Bang nucleosynthesis, giving an upper bound of $`<X>10^{12}`$ $`GeV`$ . require $`<X>`$ to be rather low compared to the cut-off scale $`MM_{Pl}`$.
In contrast to the gravity mediated case though, in these models the cancellation of the cosmological constant is rather artificial. In fact, at the minimum the scalar potential we need,
$$V_0|F|^23\frac{|W|^2}{M_{Pl}^2}=0.$$
(2)
If the supersymmetry breaking scale is $`\sqrt{F}\mathrm{\Lambda }_{SUSY}`$ then to satisfy this we need a constant in the superpotential $`WM_{Pl}\mathrm{\Lambda }_{SUSY}^2`$ which is much larger than the SUSY breaking scale. Thus for instance for $`\sqrt{F}10^5GeV`$ as is typically the case in this scenario, then a constant at a much larger scale $`10^9GeV`$ needs to be added.
One may generate such a constant by adding yet another gauge sector to the model . Thus one possiblity is to add a SUSY preserving SQCD sector, which will generate a non-perturbative contribution to the superpotential giving for instance an effective additional piece,
$$W_{eff}=\mathrm{\Lambda }^{\frac{3N_c1}{N_c1}}(Q\overline{Q})^{\frac{1}{N_c1}}+\lambda Q\overline{Q}\stackrel{~}{S}\frac{g}{3}\stackrel{~}{S}^3$$
(3)
where the gauge theory is $`SU(N_c)`$ and $`Q,\overline{Q}`$ are quark, anti quark superfields. The total superpotential is then $`W_{tot}=W(X,\mathrm{})+W_{eff}`$. The SUSY preserving conditions $`F_{\stackrel{~}{S}}=F_Q=F_{\overline{Q}}=0`$ then give $`\stackrel{~}{S}\sqrt{(Q\overline{Q})}\mathrm{\Lambda }`$ giving effectively a constant of the order of $`\mathrm{\Lambda }^3`$ in the superpotential. Now the question arises as to how this scale is related to the fundamental scale of supergravity, namely $`M_{Pl}`$. In fact one should expect that $`\mathrm{\Lambda }=e^{\frac{S}{3}}M_{Pl}`$ where $`S`$ is the four dimensional dilaton as in the above discussion of gravity mediated SUSY breaking. But now it is clear that without adding a constant to the super potential, $`S`$ will have runaway behaviour and we will not be able to generate an intermediate scale $`\mathrm{\Lambda }`$.
In this paper we will not worry about the origin of such a constant and will leave that to some unknown (stringy) mechanism as in the case of the gravity mediated model. We will simply add such a constant but ask whether whithin the assumptions of the gauge mediated suspersymmetry breaking model, one could stabilize the dilaton in the same way that one is able to do in the supergravity mediated models. The answer will turn out to be negative. This does not imply that this mechanism is not viable. But it suggests (given the remarks in the previuos paragraph) that the dilaton (and presumbaly all moduli) of string theory are stabilized by some nonperturbative stringy mechanism, and that it is frozen out of the low energy field theory dynamics. Indeed given the difficulties of finding viable mechanisms for the gravity mediated case this may be required in that case also (see for example the third paper of ).
The other scenario, is when supersymmetry breaking is communicated to the visible sector by a pseudoanomalous $`U(1)`$ gauge interaction (U1MSB) . The scale of the soft masses in this case is set either by the $`D`$-term corresponding to the $`U(1)`$ or, in the absence of $`D`$-term contributions, it is set by an $`F`$-term. Since the virtue of the pseudoanomalous $`U(1)`$ is mainly the generation of fermion masses, it is assumed to be flavor non universal over the visible sector. Then, however, in order to avoid conflict with data on flavor changing neutral currents (fcnc), one has to require that the $`D`$-terms essentially do not contribute to supersymmetry breaking. <sup>6</sup><sup>6</sup>6This can be achieved in some models with anomalous $`U(1)`$, under additional assumptions. Furthermore, to have universal masses at the scale $`M`$, (again for fcnc reasons) one would like, in addition, supersymmetry breaking to be dominated by the dilaton $`F`$-term, so that the superpartners would get universal masses $`F_S/M`$. We will analyze prototype models for each case and we will try to answer the question under what circumstances supersymmetry is unbroken or broken with vanishing cosmological constant and what are the main features of the corresponding unbroken or broken vacua.
For GMSB, we will take the tree level superpotential to be simply $`w_0=\lambda X\mathrm{\Phi }\overline{\mathrm{\Phi }}`$, with $`\lambda `$ a Yukawa coupling. $`\mathrm{\Phi }`$ and $`\overline{\mathrm{\Phi }}`$ are the messenger fields, vector-like with respect to the gauge group that contains the standard model gauge group. $`X`$ is a field, singlet of the visible and hidden sector gauge groups. For concreteness, we will take a specific, $`SU(N_c)`$ hidden sector that contains one family ($`N_f=1`$) of fields $`Z`$ and $`\overline{Z}`$ transforming respectively as $`𝐍`$ and $`\overline{𝐍}`$ of $`SU(N_c)`$. If the hidden sector is asymptotically free, then at the confining scale the formation of gaugino condensates becomes possible. Below this scale, the relevant physical hidden sector field becomes the hidden sector singlet “meson” $`X\sqrt{Z\overline{Z}}`$, and the corresponding contribution to the superpotential, for $`N_c>N_f+1`$, will be $`w(S,X)=cX^pe^{rS}`$. Here $`c`$, $`p`$ and $`r`$ are model dependent parameters and $`S`$ is the dilaton field. Given that we assumed one hidden sector flavor, $`p=2/(N_cN_f)`$ and $`r=8\pi ^2/(N_cN_f)`$ are both strictly positive quantities. This specific choice of the hidden sector, will not spoil the generality of our final conclusions concerning supersymmetry breaking and the vanishing of the cosmological constant. The full superpotential is then $`𝒲=w_0(X,\mathrm{\Phi },\overline{\mathrm{\Phi }})+w(S,X)+k`$. In $`𝒲`$ we allow for a constant term $`k`$ which can be either the vacuum value of a combination of fields that have been set to constants upon minimization of the scalar potential with respect to the corresponding fields, or a new, non perturbative contribution to $`𝒲`$. In either case what is important for us is that $`k`$ is independent of $`S`$, $`X`$, $`\mathrm{\Phi }`$ and $`\overline{\mathrm{\Phi }}`$.
For U1MSB, we assume a hidden sector which has exactly the same structure as the hidden sector of the GMSB model and we couple it to a visible sector singlet field $`\mathrm{\Theta }`$. Below the condensation scale, the physical degree of freedom is $`\sqrt{Z\overline{Z}}X`$. Both $`\mathrm{\Theta }`$ and $`X`$ are assumed to be charged under the $`U(1)`$. The full superpotential in this case is taken to be $`𝒲=w_0+w+k=\lambda \mathrm{\Theta }X^2+cX^pe^{rS}+k`$, where $`k`$ is independent of $`\mathrm{\Theta }`$ and $`X`$
Due to the dilaton dependence of $`w`$, it is naturally implied that these superpotentials arise from some string compactification which requires supergravity as the correct low energy effective theory. In supergravity, the scalar potential (ignoring $`D`$-terms), is given by :
$$V=e^{K/M^2}\left[\underset{\varphi _i}{}|F_{\varphi _i}|^23\frac{|𝒲|^2}{M^2}\right]\mathrm{with}F_{\varphi _i}𝒲^{(\varphi _i)}+\frac{K^{(\varphi _i)}}{M^3}𝒲,$$
(4)
where $`K`$ is the K$`\ddot{\mathrm{a}}`$hler potential, and $`\varphi _i`$ denotes any of the physical fields. Superscripts in parentheses denote differentiation with respect to the corresponding fields. We will assume for simplicity that all the other moduli besides the dilaton (such as $`T`$ or $`U`$) can be neglected. The K$`\ddot{\mathrm{a}}`$hler potential $`K`$ then will simply be:
$$K=M^2\mathrm{log}(S+\overline{S})\underset{\varphi _i}{}\varphi _i\varphi _i^{}.$$
(5)
The K$`\ddot{\mathrm{a}}`$hler potential has been assumed to have its usual tree level value. One could imagine adding other, weak coupling or strong coupling corrections to it, but it seems unlikely that such terms would significantly affect our analysis. In fact it should be noted that although we are assuming some stringy effect which will generate a constant in the superpotential this does not mean that the four dimensional dilaton $`S`$ is not in the weak coupling region. Indeed, the dilaton being related to the gauge coupling, ought to be fixed at some weak coupling value. The ten dimensional dilaton which governs string perturbation theory on the other hand may be at the S-dual point so that string nonperturbative effects are important. Thus, we believe that we are justified in using the weak coupling form of the K$`\ddot{\mathrm{a}}`$hler potential for $`S`$ and ignoring perturbative and nonperturbative corrections.
It is also convenient to set $`M=1`$ and to define $`𝒢K\mathrm{log}|𝒲|^2`$. The task is to minimize the potential (4) for the GMSB and U1MSB models and see if supersymmetry breaking vacua with desired phenomenological properties exist.
In section $`2`$, we investigate supersymmetric vacuum configurations with vanishing tree level cosmological constant. A necessary and sufficient condition for this is:
$$F_{\varphi _i}=0,D=0\mathrm{and}𝒲=0.$$
(6)
The $`D=0`$ condition is necessary in the $`U(1)`$ case. In fact, we will show that there are no supersymmetric vacua for either case with a nonzero gauge coupling. In section $`3`$, we investigate supersymmetry breaking. To look for supersymmetry breaking vacua, we have to minimize the potential. The scalar potential $`V=e^K[\mathrm{}]`$ upon minimization, gives $`V^{(\varphi _i)}=K^{(\varphi _i)}e^K[\mathrm{}]+e^K[\mathrm{}]^{(\varphi _i)}=0`$, which implies that in a vacuum with vanishing tree level cosmological constant it is sufficient to solve for $`[\mathrm{}]^{(\varphi _i)}=0`$, provided that $`K\mathrm{}`$ at minimum. In section $`4`$, we discuss vanishing of the cosmological constant at one loop.
In section $`6`$, we give our conclusions.
## 2 Supersymmetric Vacua
GMSB model
Let us start by writing out our superpotential again as :
$$𝒲=w_0+w+k=\lambda X\mathrm{\Phi }\overline{\mathrm{\Phi }}+cX^pe^{rS}+k.$$
(7)
Using the K$`\ddot{\mathrm{a}}`$hler potential (5), the $`F`$-terms are computed to be:
$$F_X=\frac{1}{X}(w_0pw)X^{}𝒲,F_S=rw\sigma 𝒲,$$
(8)
$$F_\mathrm{\Phi }=\frac{1}{\mathrm{\Phi }}w_0\mathrm{\Phi }^{}𝒲,F_{\overline{\mathrm{\Phi }}}=\frac{1}{\overline{\mathrm{\Phi }}}w_0\overline{\mathrm{\Phi }}^{}𝒲,$$
(9)
where we have defined for convenience $`\sigma 1/(S+\overline{S})`$. Since the supergravity framework in which we work is valid only up to the scale $`M`$, we will exclude from our analysis cases with the vevs of $`X`$, $`\mathrm{\Phi }`$, $`\overline{\mathrm{\Phi }}`$ or $`Y=\mathrm{}`$. We will denote a field and its vacuum value (vev) by the same symbol, since there is no possibility of confusion. We can distinguish the following possibilities:
* $`w0`$: The constraints (6) allow us to write the dilaton $`F`$-term condition as $`F_S=w(r)=0`$, which requires $`r=0`$. But this is not possible, since we saw that $`r`$ is a strictly positive quantity.
* $`w=0`$.
1. $`S\mathrm{}`$ ($`\sigma 0`$): $`w`$ can be zero only if $`X=0`$ provided $`p<0`$. However, since $`p`$ is a strictly positive quantity, this is not an allowed solution.
2. $`S=\mathrm{}`$ ($`\sigma =0`$): For $`p>0`$, there is a supersymmetric vacuum configuration with $`X0`$ and $`\mathrm{\Phi }=\overline{\mathrm{\Phi }}=0`$.
We conclude that there is no supersymmetric minimum with a vanishing tree level cosmological constant and a finite value for the dilaton and therefore a nonzero gauge coupling.
U1MSB model
The superpotential in this model is:
$$𝒲=w_0+w+k=\lambda \mathrm{\Theta }X^2+cX^pe^{rS}+k.$$
(10)
With the K$`\ddot{\mathrm{a}}`$hler potential (5) the $`F`$-terms are:
$$F_\mathrm{\Theta }=\frac{w_0}{\mathrm{\Theta }}\mathrm{\Theta }^{}𝒲,F_S=rw\sigma 𝒲,F_X=\frac{1}{X}(w_0pw)X^{}𝒲.$$
(11)
Again, we distinguish two possibilities:
* $`w0`$: Using $`𝒲=0`$ we can write $`F_S=w(r)=0`$ and since $`r>0`$, there is no such supersymmetric vacuum.
* $`w`$=0:
1. $`S\mathrm{}`$ ($`\sigma 0`$): $`w`$ can be zero only if $`\mathrm{\Theta }=0`$, which requires $`p<0`$. Since $`p`$ is strictly positive, this is not an allowed solution.
2. $`S=\mathrm{}`$ ($`\sigma =0`$): The $`F_\mathrm{\Theta }=0`$ equation implies $`X=0`$. There is a supersymmetric vacuum with $`p>0`$ and $`\mathrm{\Theta }`$ undetermined.
We conclude that in the U1MSB model too, there is no supersymmetric vacuum except with zero gauge coupling.
## 3 Supersymmetry Breaking Vacua
GMSB model
In the GMSB model, there are four types of fields, namely the hidden sector field $`X`$, the messengers $`\mathrm{\Phi }`$ and $`\overline{\mathrm{\Phi }}`$ and the dilaton $`S`$. The physically relevant case is when the standard model nonsinglet messenger fields do not take vacuum expectation values, i.e. $`\mathrm{\Phi }=\overline{\mathrm{\Phi }}=0`$, which implies through equations (9) that $`F_\mathrm{\Phi }=F_{\overline{\mathrm{\Phi }}}=0`$. The minimization conditions simplify considerably if we notice that the values of the fields $`X`$ and $`\sigma `$ that minimize the potential with $`V=0`$, will minimize $`\stackrel{~}{V}\frac{V}{3|𝒲|^2}`$ as well, since
$$\stackrel{~}{V}^{(\varphi _i)}=\frac{V^{(\varphi _i)}}{3|𝒲|^2}\frac{V}{3|𝒲|^4}(|𝒲|^2)^{(\varphi _i)},$$
(12)
and that $`w_0`$ is essentially absent from the picture due to the vanishing of $`\mathrm{\Phi }`$ and $`\overline{\mathrm{\Phi }}`$. Define then
$$\stackrel{~}{F}_a\frac{F_X}{\sqrt{3}𝒲}=\frac{1}{\sqrt{3}}\left(\frac{p(1\stackrel{~}{k})}{X}+X^{}\right)=\frac{1}{\sqrt{3}}\left(\frac{p(1\stackrel{~}{k})}{a}+a\right)e^{i\alpha },$$
(13)
where $`\stackrel{~}{k}k/𝒲`$. We derived the second part of the above equation using the definition of the $`F`$-term and the expression for $`𝒲`$ and in the third part of the equation we have defined $`Xae^{i\alpha }`$. Similarly, we define
$$\stackrel{~}{F}_S\frac{F_S}{\sqrt{3}𝒲}=\frac{1}{\sqrt{3}}\left(r(1\stackrel{~}{k})+\sigma \right).$$
(14)
The phase in the above is zero if $`k=0`$. These definitions allow us to write $`\stackrel{~}{V}=|\stackrel{~}{F}_a|^2+|\stackrel{~}{F}_S|^21`$. Now let us look for solutions to the minimization conditions with $`V=0`$ and try to find out if it is possible to generate the scale hierarchy necessary for a low energy gauge mediated supersymmetry breaking scenario. For simplicity, we will ignore all phases. The vacuum conditions become:
$$\stackrel{~}{V}^{(a)}:\stackrel{~}{F}_a\left[1\frac{p}{a^2}(1\stackrel{~}{k})(1+p\stackrel{~}{k})\right]+\stackrel{~}{F}_S\left[\frac{pr}{a}\stackrel{~}{k}(1\stackrel{~}{k})\right]=0,$$
(15)
$$\stackrel{~}{V}^{(S)}:\stackrel{~}{F}_a\left[\frac{pr}{a}\stackrel{~}{k}(1\stackrel{~}{k})\right]+\stackrel{~}{F}_S\left[r^2\stackrel{~}{k}(1\stackrel{~}{k})\sigma ^2\right]=0,$$
(16)
$$\stackrel{~}{V}=0:\left[\frac{p}{a}(1\stackrel{~}{k})+a\right]^2+\left[r(1\stackrel{~}{k})+\sigma \right]^2=3.$$
(17)
We can rewrite (15) and (16) as
$$\frac{\stackrel{~}{F}_a}{\stackrel{~}{F}_S}=\frac{ar}{p}\frac{a\sigma ^2}{pr\stackrel{~}{k}(1\stackrel{~}{k})}=\frac{\frac{p}{a}(1\stackrel{~}{k})+a}{r(1\stackrel{~}{k})+\sigma },$$
(18)
$$\left[1\frac{p}{a^2}(1\stackrel{~}{k})(1+p\stackrel{~}{k})\right]\left[r^2\stackrel{~}{k}(1\stackrel{~}{k})\sigma ^2\right]+\left[\frac{pr}{a}\stackrel{~}{k}(1\stackrel{~}{k})\right]\left[\frac{pr}{a}\stackrel{~}{k}(1\stackrel{~}{k})\right]=0.$$
(19)
The first part of the equation (18) comes from the vacuum condition $`\stackrel{~}{V}^{(S)}=0`$ and the second from the expressions (13) and (14) for the $`F`$-terms. Having in mind the constraint $`a𝒪(\eta )10^6`$ mentioned in the introduction and that $`r𝒪(10)`$, $`p𝒪(1)`$, we first notice that (17) can be satisfied only if $`(1\stackrel{~}{k})`$ is also small, say $`(1\stackrel{~}{k})𝒪(\eta ϵ)`$, with $`ϵ`$ at most of order one. Then, (18) can be written as
$$𝒪\left(\eta r+\frac{\sigma ^2}{rϵ}\right)=𝒪\left(\frac{ϵ+\eta }{r\eta ϵ+\sigma }\right).$$
(20)
We distinguish two possibilities. First, assume that $`\sigma 𝒪(1)`$. Then, (20) can be solved if $`ϵ𝒪(\frac{1}{\sqrt{r}})`$. Using, however, this value for $`ϵ`$, (19) becomes $`𝒪(\frac{1}{\eta \sqrt{r}})𝒪(r)`$, which can not be satisfied since $`\eta <<1`$. If on the other hand, $`\sigma <𝒪(1)`$, then from (17) we see that $`ϵ𝒪(1)`$ and (18) becomes
$$𝒪\left(\eta r+\frac{\sigma ^2}{r}\right)=𝒪\left(\frac{1}{r\eta +\sigma }\right).$$
(21)
We now turn to (19), which can be solved only if $`\sigma ^2𝒪(\eta r^2)`$. Substituting this back into (21), we get the condition $`𝒪(\eta r)𝒪(\frac{1}{r\sqrt{\eta }})`$, which is not satisfied for $`\eta <<1`$. <sup>7</sup><sup>7</sup>7If we relax the constraint $`a\eta `$, we can solve the vacuum equations with $`\sigma 1`$, and $`a`$, $`(1\stackrel{~}{k})1/r`$. This seems to be the only possibility to avoid the negative conclusions of our analysis. We conclude that we can not satisfy the vacuum equations with $`a𝒪(\eta )`$ and a vanishing tree level cosmological constant.
It is natural to ask how general this conclusion is, given that we have considered only a certain class of superpotentials. More specifically, we have not considered additional fields and couplings possible in the superpotential involving the field $`X`$. Such a coupling, can be of the form of either $`XAB`$ or $`XXA`$ or $`XXX`$, where $`A`$ and $`B`$ are chiral superfields of the hidden sector. Clearly, the most important contribution comes from $`XAB`$, when $`A,B𝒪(1)`$. Then, defining $`l<AB>`$, $`\stackrel{~}{l}l/𝒲`$ and ignoring phases, we get the leading order modification to the $`F_a`$-term:
$$\sqrt{3}\stackrel{~}{F}_a\sqrt{3}\stackrel{~}{F}_a+\frac{p}{a}\stackrel{~}{l},$$
(22)
which amounts to $`\stackrel{~}{k}\stackrel{~}{k}^{}`$, with $`\stackrel{~}{k}^{}=\stackrel{~}{k}+\stackrel{~}{l}`$. By similar arguments as before, we can see that our previous conclusion remains.
U1MSB model
The first question to address is if the $`U(1)`$ symmetry is flavor universal in the visible sector or not. The main argument for the existence of the pseudoanomalous $`U(1)`$ is that, provided that it is flavor dependent, it is an excellent candidate to explain fermion mass hierarchies . But then, after supersymmetry breaking, its $`D`$-term will contribute to the soft masses which in turn tend to give large flavor changing neutral current (fcnc) contributions. Without any uneven mass splittings between squark generations, the U1MSB scenario is therefore consistent with data only if the dilaton $`F`$-term dominates over the $`D`$-terms. Here, we will assume that the $`D`$-term is negligible (under some additional assumptions it can indeed be small ) and argue that there exists a supersymmetry breaking vacuum with vanishing tree level cosmological constant.
The most important feature in the U1MSB model is that we do not have to make any special assumptions about the vevs of any of the fields. Each minimization condition can in principle determine its corresponding field vev. Indeed, a solution to the minimization conditions, as an expansion in the small parameter $`ϵ\mathrm{\Theta }^2/Y^2`$, has been presented in . Finally, cancelation of the tree level cosmological constant can be achieved by an appropriate choice of $`k`$. Therefore, in this model, the vacuum conditions can be satisfied and at the same time a supersymmetry breaking scale of $`𝒪(100)`$ $`GeV`$ can be generated. This scenario of supersymmetry breaking, in the dilaton dominance limit, qualitatively is very similar to gravity mediation. The difference comes from the field $`\mathrm{\Theta }`$ which when taking a vev, provides us with an additional supersymmetry breaking parameter.
## 4 One loop cosmological constant
To ensure the vanishing of the cosmological constant at one loop, one has to look at the one loop effective potential :
$$V_1=V+\frac{1}{64\pi ^2}Str^0M^4\mathrm{log}\frac{M^2}{\mu ^2}+\frac{1}{32\pi ^2}Str^2M^2+\frac{1}{64\pi ^2}Str^4\mathrm{log}\frac{^2}{M^2},$$
(23)
with
$$Str^n=\underset{i}{}(1)^{2J_i}(2J_i+1)m_i^n$$
(24)
the well known supertrace formula of supergravity. Clearly, to have a zero cosmological constant, it is not sufficient to set $`V=0`$ alone. The second term is zero in theories with equal number of bosons and fermions and in particular in all supersymmetric models, but the third and the last term have to be taken into account. The last term is one that does not destabilize the hierarchy, but its presence is important for the vanishing of the cosmological constant. It can be taken into account by modifying the constraint on the classical potential $`V=0`$, to $`V+1/(64\pi ^2)Str^4\mathrm{log}\frac{^2}{M^2}=0`$. The additional term, however, is expected to be negligible compared to $`V`$, so we could safely ignore it in the previous section. In fact, we used the constant $`k`$ introduced in the superpotential to carry out this “continuous” ($`V=0`$) fine tuning. The vanishing of the third term on the other hand is required in order to have a stable hierarchy, and it involves a “discrete” fine tuning. Given that $`Str^2=2Qm_{3/2}^2`$, with $`m_{3/2}`$ the gravitino mass and
$$Q=N1𝒢^iH_{i\overline{j}}𝒢^{\overline{j}},$$
(25)
with $`H_{i\overline{j}}=_i_{\overline{j}}\mathrm{log}det𝒢_{m\overline{n}}_i_{\overline{j}}\mathrm{log}detRe[f_{ab}]`$, a necessary condition for the vanishing of the cosmological constant is $`Q=0`$ . In the above, $`𝒢_i=\frac{𝒢}{\varphi ^i}`$, $`N`$ is the total number of chiral multiplets, $`f_{ab}=S\delta _{ab}`$ is the gauge kinetic function and $`𝒢_{m\overline{n}}`$ is the K$`\ddot{\mathrm{a}}`$hler metric of the K$`\ddot{\mathrm{a}}`$hler manifold for the $`N`$ chiral superfields. We can readily calculate $`Q`$ for both cases. $`𝒢_S=\frac{F_S}{𝒲}`$, $`𝒢_{\overline{S}}=\overline{𝒢}_S`$ and the K$`\ddot{\mathrm{a}}`$hler metric is $`𝒢_{i\overline{j}}=diag(1/(S+\overline{S})^2,1)`$, where the $`1`$ is to be understood as multiplied by the $`(N1)\times (N1)`$ unit matrix, where $`N`$ is the total number of chiral superfields. This is true only if the only modulus in the model is the dilaton and the K$`\ddot{\mathrm{a}}`$hler potential is as we have chosen it. In more complicated situations the above formulas have to be modified accordingly. Also, $`𝒢^{i\overline{j}}=diag((S+\overline{S})^2,1)`$ and $`𝒢^i=𝒢^{i\overline{j}}𝒢_{\overline{j}}`$, $`𝒢^{\overline{i}}=𝒢_j𝒢^{j\overline{i}}`$. A simple computation then yields:
$$Q=N13(S+\overline{S})^2|\frac{F_S}{𝒲}|^2.$$
(26)
Remembering that $`S+\overline{S}=\frac{2}{g^2}𝒪(1)`$ and that in realistic models $`N𝒪(150)`$ (in the second reference in $`N=143`$), we conclude that if $`|\frac{F_S}{𝒲}|𝒪(1)`$ (this does not exclude the case $`|\frac{F_X}{𝒲}|𝒪(1)`$), using the condition for the vanishing of the tree level cosmological constant $`|\frac{F_S}{𝒲}|3`$, we get the following condition for $`Q=0`$: $`N1\frac{36}{g^4}`$ <sup>8</sup><sup>8</sup>8This relation is rather stable. By including in $`K`$ the $`3\mathrm{log}(T+\overline{T})`$ term, it changes to $`N\frac{36}{g^4}`$.. It is amusing to notice that $`N=150`$ implies $`\alpha 1/26`$. In other words, in dilaton dominated models with $`V=0`$ at the minimum (the same goes through for other moduli dominated models), the one loop cosmological constant vanishes for values of the coupling constant remarkably close to its unification value. The dilaton $`F`$-term, therefore, has to be of at least equal order of magnitude as the $`F_X`$ term.
## 5 Conclusions
We analyzed two simple models of gauge mediated supersymmetry breaking in the context of supergravity and in particular we looked for supersymmetry breaking vacua with zero cosmological constant. To be able to cancel the tree level cosmolological constant, we allowed for a constant $`k`$ in the superpotential. With a K$`\ddot{\mathrm{a}}`$hler potential of the form (5) and $`k=0`$ it is not only impossible to do the latter but also it does not seem possible to get a weak coupling minimum .
For the GMSB model, we showed that it is not possible to have a supersymmetry breaking vacuum state with zero tree level cosmological constant, with $`X`$ in the desired range for low energy gauge mediation. We can speculate what would happen if we relaxed some of our simplifying assumptions. We could generalize the K$`\ddot{\mathrm{a}}`$hler potential by including $`T`$ or $`U`$ moduli but our conclusions regarding the vanishing of the cosmological constant are unlikely to be changed, except that some other modulus $`F`$-term or a linear combination of them would take the place of $`F_S`$. Corrections to the tree level K$`\ddot{\mathrm{a}}`$hler potential would be small in the weak coupling limit and thus unlikely to change these conclusions. Our conclusion then is not that these models are ruled out but that they require a dilaton stabilization at the string scale so that it is frozen out of the low energy field theory dynamics.
For the U1MSB model, on the other hand, we had to assume that the $`D`$-terms were negligible in order to avoid conflict with fcnc data. Provided that this was the case, we argued that there might exist (dilaton dominated) supersymmetry breaking minima with vanishing tree level and one loop cosmological constant with a field theoretic mechanism for dilaton stabilization.
## 6 Acknowledgments
This work was supported in part by the United States Department of Energy under grant DE-FG02-91-ER-40672. |
no-problem/9911/cond-mat9911408.html | ar5iv | text | # Spatial evolutionary prisoner’s dilemma game with three strategies and external constraints
## I INTRODUCTION
Evolutionary game theory has attracted a lot attention during the past years in human sciences, political sciences, biology and economics. In particular, the so-called evolutionary prisoner’s dilemma game (PDG), which is a metaphor for the evolution of cooperation in populations of selfish individuals, has been minutely investigated . In the original form of the PDG, only uniform populations with given strategies were considered. However, it was realized that new interesting phenomena can occur when the PDG was expanded in such a way that local contests in a d-dimensional space could take place (we shall use the abbreviation SPDG for such systems). It turns out that these spatially extended models are similar to the ones studied in nonequilibrium statistical physics. They may exhibit cooperative behavior resulting in phase transitions in the stationary state. Accordingly, it is very fruitful to study SPDG like models using the tools developed in the framework of nonequilibrium statistical physics.
In its simpler form, the PDG is a version of matrix games where the symmetric incomes of the two players depend on their simultaneous decisions, whether they wish to cooperate with the others or to defect. Each player wants to maximize his individual income. The highest individual payoff (the temptation to defect) can be reached by the defector against the cooperator receiving the lowest reward (sucker’s payoff). The mutual cooperation results in the highest total payoff divided equally between the players. For mutual defections the players get a lower payoff exceeding the value of sucker’s payoff. Two rational players will both defect because this choice provides the larger income independently of the partner’s choice.
On the contrary, mutual cooperation dominates in economic and biological systems where the contestants interact frequently. In the iterated round-robin PDG, the players, knowing the previous decisions, have to choose between two options (defection and cooperation). For a given round the contestants can be classified according to the total individual payoffs they have obtained. Following the Darwinian selection principle, at each round the worst player will adopt the winner’s strategy.
Extended numerical simulations have been performed to select the “best strategy” among many . Best does not means that this strategy will always win a fight against another strategy, but means that it will obtain the highest payoff during a tournament during which it will have to fight against many opponents having different strategies. In the highest payoff was obtained by the so-called ”Tit for Tat” ($`T`$) strategy which cooperates in the first round and then always repeats his co-player’s previous decision. The main characteristics of this strategy (never defect first, react to the defection of the opponent and forgiveness) are crucial ingredients to sustain the mutual cooperation against the defectors. In particular, extensive simulations (see for a summary) for the case in which the players adopt one of the two following strategies: $`C`$ (cooperate unconditionally) or $`D`$ (defect unconditionally) have shown that the cooperators were disappearing in the stationary state. The introduction of some $`T`$’s strategies has an important effect. For short times, the D’s population increases while the C’s one decreases, leading to the decrease of D’s payoffs. As a consequence, the T’s can invade the D’s population.
In order to study the spatial effects Nowak and May have introduced an SPDG consisting of a two-state cellular automaton. The players are located on a regular lattice in a d-dimensional space and can adopt the C’s or D’s strategy. Each player is fighting with the individuals belonging to a given neighborhood. The player’s strategies are upgraded simultaneously in discrete time steps according to the following rule: each player adopts the best strategy found in his neighborhood. This model exhibits a rich variety of spatial and temporal patterns as a function of the payoff $`b`$ characterizing the temptation to defect. Other SPDG models have also been investigated . In particular Killingback and Doebeli have shown that “Pavlov” like strategies can be even more efficient than “Tit for Tat” in some circumstances.
Nowak et al. have also extended the above analysis by allowing stochasticity (irrational choices) during the evolutionary process. The degree of stochasticity is governed by a parameter called $`m`$, and in the limit $`m\mathrm{}`$, one recovers the deterministic case. According to the value of $`b`$, several stationary states are possible.
In some related models , Szabó and Tőke have shown that the different stationary state phases were separated by second order nonequilibrium phase transitions lines. The associated critical exponents belongs to the directed percolation (DP) universality class.
Both the importance of the presence of the T’s in the PDG and the richness associated with the spatially extended aspect motivated us to study a new class of SPDG. In the present work, we study a novel aspect of the SPDG, namely what is happening when the cooperators are enforced by some external constraints. More explicitly, we consider a SPDG with the three strategies $`D`$, $`C`$, and $`T`$ and investigate the effects of random adoption (or forcing) of $`C`$ strategies. This new effect can be interpreted as an attempt by a government or by any other organization to enforce the cooperation among individuals by forcing some of them, chosen randomly with a given probability $`p`$, to cooperate. A different interpretation can also be given. Among the player’s community, some old players resign and are replaced by younger ones having a different educational background making them more open to collaborate. As we shall see the effects of this external constraint are rather surprising. Indeed, in dimensions 1,2 or 3, the presence of the constraint reduces the cooperation if $`p`$ is less than a given threshold value $`p_c`$ depending upon the dimensionality of the system. The cooperation is enforced only if the constraint is strong enough, i.e. if $`p>p_c`$.
The nonequilibrium second order phase transitions describing the extinction of one of the possible strategies are found to belong to the directed percolation universality class.
The paper is organized as follows. The model is defined in section II. Its properties are analyzed in mean-field approximation in section III. The properties of the model without constraint are discussed in section IV. The model with constraint is studied in section V, both in mean-field approximation and by Monte Carlo (MC) simulations, for respectively 1, 2 and 3 dimensions. Finally, conclusions are drawn in section VI.
## II THE MODEL
We consider a SPDG model in which the players are located on the sites of a $`d`$-dimensional cubic lattice of linear size $`L`$ where periodic boundary conditions are assumed. Each player can adopt one of the three following strategies $`D`$ (defects), $`C`$ (cooperates) or $`T`$ (Tit for Tat) and interacts with its $`2d`$ nearest neighbors. The total payoff of a given player is the sum of the payoffs coming from the interaction with all its nearest neighbors. We use an extension (including the Tit for Tat strategy) of the payoff matrix used by Nowak and May . The individual payoffs for the players $`P_1`$ and $`P_2`$ as a function of their strategies are given in Table 1. The only free parameter $`b`$ ($`1<b<2`$) measures the temptation to defect. Note that the above payoff matrix does not take into account that the $`T`$ players always try to cooperate with D’s during the first round. Thus these values are considered as the averaged (or in this present case stabilized) payoffs.
| $`P_1`$ $`\backslash `$ $`P_2`$ | $`D`$ | $`C`$ | $`T`$ |
| --- | --- | --- | --- |
| $`D`$ | 0 $`\backslash `$ 0 | $`b`$ $`\backslash `$ 0 | 0 $`\backslash `$ 0 |
| $`C`$ | 0 $`\backslash `$ $`b`$ | 1 $`\backslash `$ 1 | 1 $`\backslash `$ 1 |
| $`T`$ | 0 $`\backslash `$ 0 | 1 $`\backslash `$ 1 | 1 $`\backslash `$ 1 |
Table 1: Payoff matrix of the model
It is legitimate to use this payoff matrix providing that the strategy adoption is rare comparing to the frequency of the game. It makes the simulation simpler and more efficient. Notice that similar payoff matrices can be obtained when substituting some other “nice” strategy (as Pavlov which cooperates in the first round for example, a strategy which is nevertheless less efficient than T in this case) for $`T`$.
For the non constrained case ($`p=0`$), the system evolves in discrete time according to the following MC process. Starting from a random initial state, a site is chosen randomly. This site updates its strategy by first selecting randomly one of its nearest neighbor and second, by adopting the strategy of this nearest neighbor only if it is having a higher payoff. A MC steps consists in updating each lattice site once, on the average.
In the constrained case, at each time step, each player is forced to adopt the cooperative strategy $`C`$ with a probability $`p`$.
Thus the dynamics of the constrained model is the following.
* One chooses randomly one player.
* With probability $`p`$ this player adopts the $`C`$ strategy.
* With probability $`(1p)`$ the player searches for a better strategy according to the procedure described above.
* The players update their payoffs.
The model is characterized by three free parameters: $`b,p,d`$.
It is hopeless to find exact analytical solutions for such models. Accordingly, we shall first study them in the framework of mean-field like approximations and then investigate them by numerical simulations.
## III MEAN-FIELD LIKE APPROXIMATIONS
The simplest mean-field approximation consists in neglecting all the spatial correlations in the systems. This amount to consider a model in which for each player, the partners are chosen randomly in the system instead of being restricted to a particular neighborhood. Each player interacts with the same number of counterparts. The dimensionality of the system plays no role. The simplest mean-field like approximations have been successfully used previously for similar problems and more details concerning this approximative scheme can be found in textbooks .
Within this approximation, the dynamics of the system is completely described by the time dependent concentrations:
$$c_\alpha (t)=N_\alpha (t)/L^d,(\alpha =C,D,T)$$
(1)
where $`N_\alpha (t)`$ is the number of players with strategy $`\alpha `$ at time $`t`$. These concentrations satisfy the normalization condition $`c_D+c_C+c_T=1`$.
According to Table 1 the average payoffs for each strategy are:
$$m_D=bc_C,m_C=c_C+c_T,m_T=c_C+c_T.$$
(2)
Notice that the $`C`$ and $`T`$ strategies have the same payoffs, therefore no strategy exchange will occur among them.
Following the evolutionary rules given in Sec. II, the concentrations $`c_\alpha (t)`$ obey the following equations of motion:
$`\dot{c}_D`$ $`=`$ $`pc_D(1p)c_D(c_C+c_T),`$ (3)
$`\dot{c}_C`$ $`=`$ $`+p(c_D+c_T)\pm (1p)c_Dc_C,`$ (4)
$`\dot{c}_T`$ $`=`$ $`pc_T\pm (1p)c_Dc_T,`$ (5)
where the upper and lower signs refer respectively to the cases when strategy $`D`$ is dominated by $`C`$ and $`T`$ ( $`m_D<m_C=m_T`$), and when $`D`$ dominates $`C`$ and $`T`$ ($`m_D>m_C=m_T`$).
The numerical integration of the above equations of motions leads, for several values of $`p`$, to the flow diagrams shown in Fig. 1.
The quantities represented on the vertical and horizontal axes are respectively $`c_D`$ and $`c_Cc_T`$. The upper corner of the triangle corresponds to the state of $`c_D=1`$, while the lower left and right corners describe respectively homogeneous states with $`c_T=1`$ and $`c_C=1`$.
For $`0<p<1/2`$ the stationary state solution of the above equations is:
$$c_D=\frac{12p}{1p};c_C=\frac{p}{1p};c_T=0$$
(6)
while for $`p>1/2`$ the system goes to the adsorbing state ($`c_C=1`$ and $`c_D=c_T=0`$). Surprising the $`T`$ strategy extincts if $`p>0`$.
Without constraints, i.e. for $`p=0`$, the system tends either to a homogeneous $`D`$ adsorbing state ($`c_D=1`$) or to a mixed state of $`C`$ and $`T`$ strategies ($`c_D=0`$) depending on the initial conditions. Note that the above stationary states are independent of the value of $`b`$. Strikingly different behaviors will be observed beyond this simplest mean-field approximation when the local fluctuations and short-range correlations are taken into account as we shall see in the following sections.
More elaborated mean-field like approximations can be devised. The basic idea is to take explicitly into account some of the spatial correlations by computing the probability of appearance of all the possible configurations of a small cluster containing $`n`$ sites. In one dimension, one considers a cluster of $`n`$ consecutive lattice sites. The equations of motion for these probabilities follows from the evolution rules of the system. Details concerning the one-dimensional case are given in . Note that already the one site mean-field approximation ($`n=1`$) differs from the simple mean-field solution given above. In higher dimensions pair or square mean-field like approximations have been used, and a detailed description of these methods can be found in . Note that now, the predictions of these approximations depends upon the dimension of the system. Accordingly, the cases $`d=1,2`$ and $`3`$ will be discussed in different subsections.
## IV MODEL WITHOUT EXTERNAL CONSTRAINT
Let us start by considering the case with no external constraint, i.e. $`p=0`$. In this case, the dynamic is simple: a given player adopts the strategy of one of his randomly chosen neighbor providing that this neighbor has a higher payoff. As we shall see, the only stationary states are either always trivial cooperator like ($`C`$ or $`C`$-$`T`$) or a pure defector one ($`D`$).
When the random initial state is made of only cooperators and defectors, one founds that, both in $`d=1,2`$ and $`3`$ dimensions, the stationary state of the system is a pure defector one.
The reason in $`d=1`$ is simply that a $`D`$ player has always a higher payoff than a neighboring $`C`$ player.
In $`d=2`$ several configurations should be analyzed. The most favorable situation for $`C`$ to win is when it is adjacent to a flat $`C`$-$`D`$ interface. However, if this interface has an irregularity then the $`D`$’s can invade the sea of $`C`$ players. Indeed, a new born $`C`$ in the $`D`$ half-plane is always weaker than the $`D`$’s at the interface and thus, such $`C`$’s disappears sooner or later. However, at a given time and with some finite probability, two next nearest neighbor $`C`$ players could be present in the $`D`$ half-plane. It results a payoff $`3b`$ for the $`D`$ player squeezed between three $`C`$ and this $`D`$ can now invade the $`C`$’s and sweep off one layer of them independently of the value of $`b`$.
The $`d=3`$ case is very similar, but the process is slower. Indeed, four nearest neighbors of a $`D`$ player (called $`D_1`$) are necessary to be invaded by the $`C`$’s if the value of $`b`$ is close to 1. Once it has occured, $`D_1`$ is strong enough to cross the interface and then destroys all the $`C`$’s.
For a large system with initially a finite proportion of $`T`$ players, the stationary state is always a cooperator like state ($`T`$-$`C`$). This asymptotic behavior can be easily understood in the one dimensional case. Let us suppose that there are four neighboring $`T`$ ($`TTTT`$) in the initial state of a 1d system, (this is practically always the case in a sufficiently large system with a finite probability to have some $`T`$’s in the initial state), then they kill all of the $`D`$’s. Even in the worst case two $`D`$’s could invade into to the $`T`$ population, ($`CDDTTDDC`$), but then the central $`T`$’s become stronger (payoff=1) then the $`D`$’s (payoff=0). The $`D`$’s kill all of the neighboring $`C`$’s, and even these two $`T`$’s can invade the whole $`D`$ area.
This argument can be extended to $`d`$ dimensions. A cube of linear size 4, made of $`T`$ players is enough to guarantee the cooperation in the stationary state. The reason is that the $`D`$’s surrounding this cube cannot destroy the central $`T`$ players inside a cube of linear size 2. Indeed, the central $`T`$’s have always $`d`$ $`T`$’s neighbors to cooperate with (payoff=$`d`$), but the neighboring $`D`$’s can only have one $`C`$ neighbor (payoff$`b<d`$ if $`d2`$).
## V MODEL WITH EXTERNAL CONSTRAINT
We now consider the case in which the players are located on the sites of a d-dimensional cubic lattice in the presence of an external constraint imposing to each player to choose the $`C`$ strategy with probability $`p`$. The cases $`d=1,2`$ and $`3`$ will be considered.
### A One-dimensional system
In the one-dimensional model the players located on the sites of a chain interact with their two nearest neighbors. It is easy to see that the dynamics is independent of the value of the parameter $`b`$ in its domain of definition.
A systematic MC analysis of the stationary states was performed by varying the value of $`p`$ for different system sizes between $`L=32`$ to $`L=16384`$. Our simulations show that the T’s strategy extincts for all values of $`p`$. However, as a function of $`p`$, the stationary state can be either a symbiosis of $`C`$ and $`D`$ strategies or a pure $`C`$ state as shown in Fig. 2. For very small values of $`p`$, one has a pure $`C`$ stationary state ($`c_C=1`$) but, when $`p`$ reaches the value $`p_1`$, a first transition occurs to a stationary state in which the two strategies $`D`$ and $`C`$ coexist. At $`p=p_{c2}>p_1`$ the system undergoes a second continuous transition from the $`C`$-$`D`$ stationary state to the pure absorbing $`C`$ state.
The transition at $`p_{c2}`$ is easy to understand. In one dimension and when only $`C`$ and $`D`$ strategies are present our model is equivalent to the contact process (CP) which was originally introduced as a simple model for epidemics . In the CP a particle (sick person) can disappear at rate 1, and an empty site (healthy person) can become occupied with rate $`\lambda z/2`$, where $`\lambda `$ is the control parameter and $`z`$ is the number of particles in the neighborhood of the empty site. In our model the $`D`$ strategy, which in $`d=1`$ is always better than the $`C`$ one, plays the role of the sick persons and the $`C`$ strategy, which can only be created in the system through the constraint, corresponds to the healthy individuals.
Starting from a $`C`$-$`D`$-$`T`$ initial state and after the extinction of $`T`$’s the possible stationary states are the pure $`C`$ or a $`C`$-$`D`$ states. The $`C`$-$`D`$ state corresponds to the steady state of the CP as well. For $`p`$ around $`p_{c2}`$, the $`T`$’s disappear rapidly and thus do not affect the extinction of the $`D`$’s. Hence the second transition to the absorbing state corresponds to the CP’s one. This transition occurs at $`p_{c2}=1/(1+\lambda _c)=0.23267`$ and is believed to belong to the DP universality class .
The behavior of the first transition at small $`p`$’s is much less clear. It turns out that for our model two characteristic parameters $`p_\alpha (L)`$ and $`p_\beta (L)`$, depending on the system size $`L`$, can be introduced. For $`p<p_\alpha (L)`$ the stationary state is always the pure $`C`$ state while for $`p>p_\beta (L)`$ the system evolves to a $`C`$-$`D`$ coexistence phase, which is a steady state of the CP. For $`p_1(L)<p<p_2(L)`$, the system can evolves towards one of the two possible stationary states, depending on the particular realization of the random numbers and on the initial state.
The probability, $`\rho _L(p)`$, of reaching the $`C`$-$`D`$ state has been investigated numerically. The data are plotted in Fig. 3 and can be well fitted by a function of the type $`1\mathrm{exp}(c_1(pp_0)^{c_2})`$, where $`c_1`$, $`c_2`$ and $`p_0`$ are fitting parameters. The limiting function $`\rho _{\mathrm{}}(p)`$ was obtained by using usual finite size scaling methods. As shown in Fig. 4, the functions $`\rho _L(p)`$ collapse on a single curve if one plots $`\rho `$ as a function of $`\{p\rho _{\mathrm{}}^1[\rho _L(p)]\}/L^{1/3}`$. This shows that $`\rho _{\mathrm{}}(p)`$ do not collapse to zero, hence the phase transition takes place at a finite value of $`p0.025`$.
The respective times $`t_C`$ and $`t_{CD}`$ needed to reach the pure $`C`$ and the $`C`$-$`D`$ stationary states have also been investigated. Both times show a singular behavior (see Fig. 5). Unfortunately, we were not able to find a reasonable scaling fit for those data. However, it is obvious from the simulations, that $`t_{CD}`$ exhibits a much stronger singularity than $`t_C`$. Hence, for a large system, the time needed to reach the $`C`$-$`D`$ state is much larger than the one needed to reach the $`C`$ state, even if the system evolves to the $`C`$-$`D`$ state more frequently.
Beside the MC simulations, the properties of the system were also investigated using generalized mean-field approximations. Assuming the coexistence of all the three strategies these calculations can be performed numerically on clusters of sizes as large as $`n=4`$. Contrary to the MC results (see Fig. 2) these approximations predicted the existence of a $`C`$-$`D`$-$`T`$ state at small $`p`$ values. As far as the T’s are concerned, we observed that the maximum value of $`c_T`$ is decreasing when the cluster sizes $`n`$ were increasing, providing us a trend for the extinction of the T’s.
The contradiction between the present mean-field and MC results refers to the importance of (interfacial) invasion phenomena detailed later on.
Allowing only two strategies ($`C`$ and $`D`$) the above mean-field analysis can be performed for larger cluster sizes up to $`n=7`$. The results are compared with the MC ones in Fig. 2. As expected, the accuracy of the generalized mean-field method increases with the cluster size $`n`$. The extrapolation of the results obtained for finite values of $`n`$ ($`n=1,\mathrm{},7`$) leads to a critical value $`p_{c2}^{(MF)}0.235`$. The quality of this approximative scheme can be estimated by comparing the value of $`p_{c2}^{(MF)}`$ with the best known numerical value $`p_{c2}=0.23267`$ .
In order to understand the behavior of the model around the first transition it is interesting to examine the time evolution of the system in its transient regime. As illustrated in Fig. 6, one can observe a domain growth process controlled by cyclic invasions . This picture suggests that the most relevant aspect of the dynamics is the collision between the fronts separating different strategies.
Let us first investigate the motion of a front separating a cluster of $`C`$ from a $`D`$ one. The $`D`$ domains contain some $`C`$ sites coming from the external constraint. However, the life times of the $`C`$’s are very short \[$`(1p)`$\]. Within a $`D`$ domain the average $`C`$ concentration is equal to that of the CP and can be well approximated by the simple mean-field approach as we have seen above. The $`D`$ strategy invades the territory of the $`C`$ one (absorbing state). Due to the reflection symmetry the invasion front can move to the left or to the right with the same absolute value of the velocity. To first order in $`p`$ in the limit $`p0`$, the absolute value of the average invasion velocity can be estimated by using the configuration probabilities given for $`n=1`$. This calculation yields
$$v=\frac{13p}{2}.$$
(7)
Let us now consider the invasion of the $`T`$ strategy into the $`D`$ one. Inside a territory which has been invaded by the $`T`$ strategy, the $`C`$ strategy is setting up with a probability $`p`$. As both $`C`$’s and $`T`$’s have the same payoff, no adaptation of strategies occur between them. Assuming that this invading front travels with a constant velocity $`u`$, the probability of having a $`C`$ player at a site $`k`$ steps behind the front is $`1e^{p\tau (k)}`$. $`\tau (k)=(k+1/2)/u`$ is the averaged time it takes to the front to move over a distance of $`k`$. Thus, it follows that, to leading order in $`p`$,
$$u=\frac{111p}{2}.$$
(8)
We note that above approximations lead to a velocity for the $`DC`$ front which is larger than the one for the $`TD`$. This prediction can be compared with the results of the MC simulations (see Fig. 7). For the $`DC`$ front, the agreement is very good if $`p<0.1`$, while for the $`TD`$ case, the approximation reproduces well only the linear part near the origin.
Another consequence of the external constraint concerns the life-time of the $`T`$ clusters. According to the above picture, the probability that a $`T`$ cluster dyes out during the time evolution can be approximated by:
$$P_T=\underset{k=0}{\overset{\mathrm{}}{}}\left[1\mathrm{exp}(p(k+1/2)/u)\right].$$
(9)
The inverse of $`P_T`$ can be interpreted as the average life time $`\tau _T`$ of an invading $`T`$ cluster. Substituting an integral for the infinite sum appearing in the logarithm of the above expression leads to:
$$\tau _T\mathrm{exp}(\frac{\pi ^2u}{6p})=\mathrm{exp}(\pi ^2\frac{111p}{12p}).$$
(10)
This expression shows a fast increase of the life-time when $`p`$ is decreasing. For example, a $`T`$ cluster dyes out in about $`10^5`$ MC steps (MCS) if $`p=0.04`$. This estimated life-time is significantly larger than those found by MC simulations \[$`\tau _T^{(MC)}(p=0.04)800`$ MCS\] as indicated in Fig. 6. The $`p`$-dependence of the MC data can be well approximated by the function $`\tau _T=3.26\mathrm{exp}(0.226/p)`$ within the region $`0.025<p<0.08`$ we could study. The large discrepancy between the mean-field and MC results refers to the enhanced role of the velocity fluctuations.
Starting from a random initial state spatially separated, domains are rapidly formed. Then two different situations can occur. First, $`T`$ clusters are present one both ends of each $`D`$ clusters leading to a fast extinction of $`D`$’s. Accordingly, the system evolves to a pure $`C`$ state. Second, after some time the system reaches a state characterized by the presence of only one $`D`$ cluster, having a $`T`$ island at only one of its ends, in a sea of $`C`$. As the $`DC`$ invasion is faster than the $`TD`$, the $`T`$’s can never destroy the $`D`$’s and due to the finite life-time of the $`T`$ cluster the stationary state is a $`C`$-$`D`$ coexisting one. The exponential increase of the life-time of this $`T`$ cluster as $`p0`$ explains the singular behavior observed in $`t_{CD}`$.
For $`p<p_1(L)`$ the life-time of the $`T`$ players are so long that all the $`D`$ clusters are surrounded by $`T`$’s. Accordingly, the first scenario described above is always present, leading to a pure $`C`$ state. In the contrary, for $`p>p_2(L)`$ the short lifetime of $`T`$’s insures that the $`T`$’s disappear rapidly, allowing for the growth of the $`D`$ clusters. As a consequence the stationary state is a $`C`$-$`D`$ one.
### B Two-dimensional system
The players are located on the sites of a two-dimensional square lattice. According to the payoff matrix (see Table 1), two ranges of values of $`b`$ have now to be distinguished, namely, $`1<b<3/2`$ and $`3/2<b<2`$. For any value of $`b`$ in one of those ranges the dynamics is the same.
Let us first consider the simulations performed for the value $`3/2<b<2`$. The situation is summarized in Fig. 8, in which the stationary values of the strategy concentrations are plotted as a function of $`p`$. For $`p<p_{c1}^{(MC)}=0.1329(1)`$, the system reaches a stationary state in which the three strategies $`C,D`$ and $`T`$ coexist. Here it is worth mentioning, that for small $`p`$ values ($`p0.03`$) the finite system can reach the absorbing state if the initial state has been chosen randomly. Below this size-dependent threshold value the three-strategy state can be formed and sustained by slow decreasing of $`p`$ during the simulation. In this case the extinction of $`T`$’s and $`D`$’s is a consequence of fluctuations (detailed later on) and the coexistence of the three strategies is considered as the real stationary state in the limit $`L\mathrm{}`$. The simulations performed for $`p0.005`$ show that $`c_D`$ decreases linearly with $`p`$ when $`p0`$.
When $`p_{c1}^{(MC)}<p<p_{c2}^{(MC)}=0.3671(1)`$ only the strategies $`C`$ and $`D`$ survive. Finally, when $`p>p_{c2}^{MC}`$ the system reaches a pure $`C`$ absorbing state.
The phase diagram obtained by numerical simulations can be compared with the ones obtained using the extended mean-field approximation described in Sec. III. At the level of pair approximation, one finds $`p_{c1}^{(pair)}=0.1704`$ and $`p_{c2}^{(pair)}=0.4236`$, while for the square mean-field approximation one finds $`p_{c1}^{(sq)}=0.1482`$ and $`p_{c2}^{(sq)}=0.3980`$. These latest results are plotted in Fig. 8. They compare well with the MC values given above.
Some complementary information can be obtained by studying the concentration fluctuations defined as:
$$\chi _\alpha =L^d(N_\alpha /L^dc_\alpha )^2,(\alpha =C,D,T).$$
(11)
When $`p0`$, the concentration fluctuations $`\chi _C`$ and $`\chi _T`$ are diverging while $`\chi _D`$ remains regular as shown in Fig. 9. However, the sizes of the systems investigated were not large enough to conclude that $`\chi _C`$ and $`\chi _T`$ are diverging as power laws of $`p`$ in the limit $`p0`$.
Moreover, the simulations suggest that for the stationary state in which the three strategies coexist, the typical size of $`T`$ domains (as well as their persistence time) is proportional to $`1/p`$. Similar behavior was found in the forest fire models introduced by Bak et al. to study the self-organized criticality.
The points $`p_{c1}`$ and $`p_{c2}`$ are critical points where a second order nonequilibrium phase transition takes place. Indeed, the $`T`$ concentration vanishes at $`p_{c1}`$ as $`c_T(p_{c1}p)^{\beta _1}`$, while the $`D`$’s concentration vanishes at $`p_{c2}`$ as $`c_D(p_{c2}p)^{\beta _2}`$. In order to justify this behavior a very careful numerical analysis was performed, using longer sampling times in the vicinity of critical points. Fitting the numerical data leads to the above mentioned values of the critical points and $`\beta _1=\beta _2=0.57(3)`$ which is compatible with the directed percolation exponent as expected on general ground. This fact is confirmed by the study of concentration fluctuations defined by Eq. (11). Sharp increases of the concentration fluctuations are expected at second order phase transitions. Fig. 9 illustrates this point. The concentration fluctuations of $`C`$, $`D`$ and $`T`$ strategies behave, near the transition point $`p_{c1}`$, as $`\chi (p_{c1}p)^{\gamma _1}`$ with $`\gamma _1=0.37(6)`$. A similar behavior was found at the transition point $`p_{c2}`$ with an exponent $`\gamma _2=0.37(9)`$. These values are very close to the one of directed percolation: $`\gamma _{DP}0.35`$ in two dimensions.
The above data suggest that $`p=0`$ is another critical point. However, for the reasons explained previously, it was not possible to extract reliable exponents.
It is interesting to analyze how the three strategies coexist for small values of $`p`$. As an example, let us consider the snapshot of the stationary state of a system with $`p=0.04`$ (see Fig. 10).
One can observe that ”dark areas” (made of defectors) are invading the territory (”white areas”) of cooperators, simultaneously, the ”dark areas” are invaded by the $`T`$ actors (open squares). However, the domination of the $`T`$’s is prevented by the external constraint which is leading to the growing of $`C`$ areas within the $`T`$ territory. When $`p`$ decreases, the $`T`$ territory expands while the growth of white areas slows down. The dark islands become sparse. On the contrary, an increase of the value of $`p`$ accelerates the spreading of the $`C`$ areas as well as their occupation by the defectors. Consequently, the $`D`$ population increases with $`p`$ and the number of $`T`$ competitors decreases and vanishes at $`p=p_{c1}`$.
As a result of these cyclic dominant processes a self-organized domain structure is maintained in the system. Analogous spatio-temporal structures have already been observed by Satulowsky and Tomé in a two-dimensional predator-prey system and by Tainaka and Itoh when studying competing species . Both models belong to the family of spatially-extended Lotka-Volterra models predicting oscillatory behavior of concentrations in the simple mean-field approximation.
One can also analyze the evolution of the typical strategy configurations in the vicinity of the phase transition taking place at ($`p=p_{c1}`$). One recognizes isolated colonies of the $`T`$ strategies whose motions remind us of the branching annihilating random walk models. It is known that this model belongs to the DP universality class too .
The rate of mutual cooperation is related to the average payoff per site. The maximum average payoff (which value is 4) is reached when all the players cooperate with their neighbors. On the other hand, the minimum average payoff per sites coincides with the maximum of $`D`$’s concentration.
In Fig. 11 the average payoff per site is compared with the quantity $`4(c_C+c_T)`$. One can see that the agreement between these two quantities is generally reasonable and is even quite good for $`p<p_{c1}`$. The differences between the two quantities are only coming from the fights taking place at the boundaries $`D`$-$`C`$ or $`D`$-$`T`$. However, for this range of values of $`p`$, the $`D`$ players form clusters due to the presence of the $`T`$’s and accordingly these fights are not frequent.
Finally, let us briefly consider the case for which the parameter $`b`$ belongs to the second possible range, $`1<b<3/2`$. The results of the simulations are qualitatively similar to the case $`3/2<b<2`$. The critical values of $`p`$ are lower, namely, $`p_{c1}=0.112(1)`$ and $`p_{c2}=0.297(1)`$. The most relevant differences can be observed in the limit $`p0`$, where the maximum concentration of the $`T`$ strategy is strikingly lower ($`c_T<0.15`$) than those reported in the previous case (see Fig. 8).
### C Three-dimensional system
We now consider the case in which the players are located on the sites of a three-dimensional cubic lattice. According to the payoff matrix (see Table 1), five ranges of values of $`b:1<b<2`$ have now to be distinguished, separated by the following values: $`b=5/4`$, $`4/3`$, $`3/2`$, and $`5/3`$. Any value of $`b`$ in one of those ranges will lead strictly to the same behavior. Moreover it turns out that all the values of $`b:1b2`$ leads qualitatively to the same behavior.
Figure 12 shows a typical phase diagram which is qualitatively similar to the one found for the two-dimensional model. MC simulations have been performed for systems of sizes $`32^3`$, $`64^3`$ and $`80^3`$ for five representative values of $`b`$ exploring all the different ranges.
When the value of $`p`$ increases, the system undergoes two subsequent phase transitions. The critical values of $`p`$ vary weakly with $`b`$. For example, for $`3/2<b<2`$, we have obtained $`p_{c1}^{(MC)}=0.1441(1)`$ and $`p_{c2}^{(MC)}=0.4292(5)`$, while for $`1<b<3/2`$, we found $`p_{c1}^{(MC)}=0.1512(2)`$ and $`p_{c2}^{(MC)}=0.4130(3)`$. As expected, the vanishing concentrations behave near the critical points as a power law and both transitions belong to the directed percolation universality class. For $`b=1.6`$, the numerical analysis of the MC data in the vicinity of the second transition gives $`p_{c2}^{(MC)}=0.4165(2)`$ and $`\beta _2=0.79(4)`$ in good agreement with the corresponding directed percolation value of $`\beta =0.81(2)`$ .
The analysis of the concentration fluctuations $`\chi _T`$ and $`\chi _D`$ in the vicinity of the points $`p=p_{c1}`$ and $`p=p_{c2}`$ leads to the critical exponent $`\gamma =0.18(8)`$. The large uncertainty is due to the small extension of the critical regime (typically $`|pp_c|10^3`$). The above value of $`\gamma `$ is compatible with the scaling law for the DP exponents.
As far as mean-field like approximations are concerned, they are supposed to be better when increasing the number of nearest neighbors (dimensions). The algebra becomes soon very cumbersome therefore our analysis is restricted to the pair approximation. The results are given in Fig. 12. One can see that the approximate results are in good agreement with the MC data but in the close vicinity of the critical points. Note finally that the $`p`$-dependence of the average payoff is qualitatively similar to those found for the two-dimensional model (see Fig. 11).
## VI CONCLUSIONS
We have studied quantitatively the emergence of cooperation in a spatially extended version of the prisoner’s dilemma game with three possible strategies (cooperation, defection and Tit for Tat) in the absence and presence of externally enforced cooperation. The players are distributed on a $`d`$-dimensional simple cubic lattice and their interactions are restricted to nearest neighbors.
In the absence of external constraint the time evolution is controlled by a local adoption of a neighboring strategy whose introduction is motivated by the Darwinian selection rule. When starting the simulations with a random initial state made of only cooperators and defectors, one founds that, both in $`d=1,2`$ and $`3`$ dimensions, the stationary state of the system is a pure defector one. However, if the $`T`$ strategy is also present in the initial state, then the stationary state is dominated by mutual cooperation.
The external constraint, forcing the players to adopt the strategy $`C`$ with probability $`p`$, has the following consequences. If only the $`C`$ and $`D`$ strategies are present in the initial state, then the external constraint enforces the cooperation for all value of $`p`$. However, if the three strategies $`C,D`$ and $`T`$ are initially present and if the dimensionality of the system is larger than one, then the external constraint reduces the cooperation maintained by $`T`$ for small values of $`p`$. The cooperation will be enforced only if the constraint is large enough ($`p>p_{c1}`$). These conclusions are reached both from the extended mean-field analysis and the MC simulations for $`2`$ and $`3`$ dimensional systems. The general features are not affected by the value of $`b`$ characterizing the temptation to defect.
Our study confirms the crucial role of the $`T`$ strategy which is able to prevent the spreading of defection. The $`T`$ strategy, however, dyes out in the one-dimensional system as well as in the models for which the simple mean-field theory is exact.
The above conclusions are in agreement with several historical facts coming both from the political or economical world. For example, it shows that forcing a fraction of the population to cooperate in a naive way ($`C`$ strategy) does not improve the overall cooperation in the society. It is better to educate more individuals in such a way that they will be able to play the more sophisticated Tit for Tat strategy if one desires to improve the cooperation.
From a nonequilibrium phase transition point of view, the above investigations have confirmed that the two second order phase transitions associated with the extinction processes belong to the robust directed percolation universality class.
Finally we emphasize that similar behavior is expected for spatially extended Lotka-Volterra like systems with three (or more) species providing that where one of the species is externally favored.
## ACKNOWLEDGMENTS
We thank Gunter Schütz for a very helpful discussion. This work was supported by the Hungarian National Research Fund under Grant No. T-23552 and by the Swiss National Foundation. |
no-problem/9911/quant-ph9911056.html | ar5iv | text | # References
Construction of quantum states with bound entanglement
Dagmar Bruß<sup>1</sup> and Asher Peres<sup>2</sup>
<sup>1</sup>Institut für Theoretische Physik, Universität Hannover, D-30167 Hannover, Germany
<sup>2</sup>Department of Physics, Technion—Israel Institute of Technology, 32000 Haifa, Israel
Abstract
> We present a new family of bound-entangled quantum states in $`3\times 3`$ dimensions. Their density matrix $`\rho `$ depends on 7 independent parameters and has 4 different non-vanishing eigenvalues.
PACS 03.67.Hk, 03.65.Bz, 89.70.+c
Entangled quantum states have been used since the very early days of quantum mechanics for computing the properties of atomic and molecular systems . However, it is only in recent years that the existence of a hierarchy of entangled density matrices became apparent, and it is not yet fully understood.
There are classical correlations, but there is no quantum entanglement, in separable density matrices that can be written as convex sums,
$$\rho _{m\mu ,n\nu }=\underset{K}{}w_K(\rho _K^{})_{mn}(\rho _K^{\prime \prime })_{\mu \nu },$$
(1)
where the density matrices $`(\rho _K^{})_{mn}`$ and $`(\rho _K^{\prime \prime })_{\mu \nu }`$ refer to two subsystems, possibly with different dimensions, and the coefficients $`w_K`$ are positive and sum up to unity. If Eq. (1) holds, it readily follows that the partial transpose $`\sigma _{m\mu ,n\nu }=\rho _{n\mu ,m\nu }`$ is another separable density matrix and in particular has no negative eigenvalue. This property gives a very simple necessary condition for separability . It also is a sufficient condition for systems of dimensions $`2\times 2`$ and $`2\times 3`$, but not for higher dimensions . The first counterexamples for dimensions $`2\times 4`$ and $`3\times 3`$ contained one free parameter . Such states are called “bound-entangled” because it is impossible to distill from them pure singlets by means of local operations and classical communication.
Recently, a new class of bound-entangled states was produced by means of unextendible product bases (UPB) . In these states, the density matrix $`\rho `$ depends on 6 parameters and is of rank 4 with equal eigenvalues 0.25. If all the matrix elements are real, there are only 4 free parameters, and $`\sigma \rho `$. In the complex case, $`\sigma `$ and $`\rho `$ have similar structures, but they correspond to different UPBs.
Here, we present a more general construction of bound-entangled states in $`3\times 3`$ dimensions, depending on 7 parameters (only 5 if $`\rho `$ is real and $`\sigma =\rho `$), with 4 different nonvanishing eigenvalues. We hope that this explicit construction will be useful for elucidating properties of bound-entangled states, in particular for proving (or disproving) the conjecture that they satisfy all the Bell inequalities, and therefore are compatible with a local hidden variable description . Other open problems are mentioned in ref. .
We write $`\rho `$ in terms of four unnormalized eigenvectors,
$$\rho =N\underset{j=1}{\overset{4}{}}|V_jV_j|,$$
(2)
where the coefficient $`N=1/_jV_j,V_j`$ normalizes $`\rho `$ to unit trace. The four eigenvalues, $`NV_j,V_j`$, are in general different. Explicitly, we take
$$\begin{array}{c}|V_1=|m,0,s;\mathrm{\hspace{0.33em}0},n,0;\mathrm{\hspace{0.33em}0},0,0,\hfill \\ |V_2=|0,a,0;b,0,c;\mathrm{\hspace{0.33em}0},0,0,\hfill \\ |V_3=|n^{},0,0;\mathrm{\hspace{0.33em}0},m^{},0;t,0,0,\hfill \\ |V_4=|0,b^{},0;a^{},0,0;\mathrm{\hspace{0.33em}0},d,0,\hfill \end{array}$$
(3)
where the components of $`|V_j`$ are listed in the order 00, 10, 20; 01,… It is easily seen that the 9th row and column of $`\rho `$ vanish. The remaining $`8\times 8`$ matrix is like a chessboard, with the odd-odd components depending on $`m,n,s,t`$, and the even-even components on $`a,b,c,d`$ (still, some of these components are zeros).
In principle, all 8 parameters on the right hand side of Eqs. (3) can be complex (special values of these parameters yield results equivalent to those obtained by using UPBs). However, we still have the freedom of choosing the overall phase of each $`|V_j`$ (this obviously does not change $`\rho `$). Furthermore, we can define new phases for the basis vectors $`|e_k^{}`$ and $`|e_\lambda ^{\prime \prime }`$ used to describe the two subsystems. This corresponds to rewriting $`\rho `$ in a different basis without changing its chessboard structure, nor the absolute values of the components of $`|V_j`$. This freedom can be used to make many of these components real, but not all of them, because the two combinations $`cm/bs`$ and $`b^{}t/n^{}d`$ are not affected by these changes of phase. This can be seen as follows: the parameters in $`cm/bs`$ appear in $`|V_1=|m,0,s;\mathrm{}`$ and $`|V_2=|\mathrm{};b,0,c;\mathrm{}`$. The ratios $`m/s`$ and $`b/c`$ are affected only by changes of the relative phase of $`|e_1^{}`$ and $`|e_3^{}`$, and $`cm/bs`$ is not affected at all. Likewise, the parameters in $`b^{}t/n^{}d`$ appear in $`|V_3=|n^{},0,0;\mathrm{};t,0,0`$ and $`|V_4=|0,b^{},0;\mathrm{};0,d,0`$. The ratios $`n^{}/t`$ and $`b^{}/d`$ are affected, both in the same way, only by changes of the relative phase of $`|e_1^{\prime \prime }`$ and $`|e_3^{\prime \prime }`$. There are no other invariants of this type, and we can assume, without loss of generality, that $`s`$ and $`t`$ are complex, while the 6 other parameters are real.
We now prove that in the generic case (random parameters) $`\rho `$ is inseparable: as shown in , a state $`\rho `$ is inseparable if the range of $`\rho `$ contains no product state. This is the case for our $`\rho `$, unless the parameters are chosen in a specific way. Indeed, assume that there is a product state such that
$$|p,q,r|x,y,z=A_j|V_j.$$
(4)
Since the 9th components of all the $`|V_j`$ vanish, we have $`rz=0`$. Assume that $`z=0`$ (the same proof is valid for $`r=0`$, mutatis mutandis). Then the 7th and 8th components vanish, so that $`A_3=A_4=0`$. We then have $`(px)(ry)=(A_1m)(A_2c)`$ while $`(py)(rx)=(A_2b)(A_1s)`$, whence $`mc=bs`$, which does not hold in general for randomly chosen parameters.
Finally, we have to verify that $`\sigma `$ is a positive matrix, so that the entanglement is bound. Namely, all the diagonal subdeterminants of $`\sigma `$ have to be positive or zero. This gives a large number of inequalities. Here, we shall restrict ourselves to the study of two simple cases.
The simplest one is to assume $`\sigma =\rho `$. Owing to the chessboard structure of $`\rho `$, this leads to three nontrivial conditions, which can be written, with the parametrization of $`|V_j`$ in Eqs. (3):
$$\begin{array}{c}\rho _{13}=\rho _{31}\text{or}ms^{}=m^{}s,\hfill \\ \rho _{26}=\rho _{35}\text{or}ac^{}=sn^{},\hfill \\ \rho _{48}=\rho _{57}\text{or}ad=mt.\hfill \end{array}$$
(5)
With our choice of phases, these conditions mean that $`s=ac/n`$ and $`t=ad/m`$ are real. We thus have 6 free parameters in the vectors $`|V_j`$. These parameters can still be scaled by an arbitrary factor (that will be compensated by $`N`$), so that there are 5 independent parameters in our construction.
Another, more general way of constructing bound entangled states is to assume that $`\sigma `$ is not the same as $`\rho `$, but still is spanned by two pairs of eigenvectors with a structure similar to those in Eqs. (3), with new parameters that will be called $`a^{},b^{},..`$. This assumption is obviously compatible with (but not required by) the chessboard structure of $`\sigma `$, which is the partial transpose of $`\rho `$. It is then easily seen that the various parameters in the eigenvectors of $`\sigma `$ may differ only in their phases from those in the eigenvectors of $`\rho `$. We therefore write them as $`a^{}=ae^{i\alpha }`$, and so on. This gives 8 additional arbitrary phases, besides those of $`s`$ and $`t`$. The requirement that $`\sigma `$ is the partial transpose of $`\rho `$ imposes 6 conditions on these phases (apart from those on the absolute values). These are fewer conditions than phases at our disposal, so that the parameters $`s`$ and $`t`$ can now remain complex. Only their absolute values are restricted by
$$|s|=ac/n\text{and}|t|=ad/m.$$
(6)
We thus have 7 independent free parameters for this case.
A natural question is whether this construction can be generalized to higher dimensional spaces, with a larger number of pairs of eigenvectors $`|V_j`$, suitably structured. We have no definite answer: in such a generalization, the number of conditions grows much faster than the number of free parameters, and we think it unlikely that such a generalization is possible, but we have no formal proof.
DB acknowledges support from Deutsche Forschungsgemeinschaft under SFB 407. AP was supported by the Gerard Swope Fund and the Fund for Encouragement of Research. |
no-problem/9911/astro-ph9911361.html | ar5iv | text | # Obtaining Galaxy Masses Using Stellar Absorption and [O II] Emission-Line Diagnostics in Late-Type Galaxies
## 1. Introduction
Kinematic studies of distant galaxies reveal their structure at early times and help trace their evolution to the present day. Measurements of the most important kinematic properties, the velocity widths and rotation curves, are now becoming routine for galaxies at cosmologically significant distances. Conclusions are mixed regarding whether distant field galaxies are consistent with local Tully-Fisher (T-F) relations out to $`z=1`$. The results are consistent with modest evolution in the T-F relation, in the sense that galaxies of a given rotational amplitude are 0.2-1.5 magnitudes brighter at earlier epochs (Forbes et al. 1995; Vogt et al. 1996, 1997; Simard & Pritchet 1998; Rix et al. 1997). In these studies, the bluest galaxies deviate most strongly from the T-F relation suggesting that the conclusions are sensitive to target selection effects (Bershady et al. 1998). Since current studies rely on the brightest and largest galaxies which have well-defined rotation curves, the samples are small and may be biased by including only the most massive spirals. For early-type galaxies, the velocity dispersions and luminosities provide an estimate of evolution in the fundamental plane. Recent work indicates luminosity evolution for both cluster and field ellipticals. Cluster ellipticals show modest evolution—about 0.8 magnitudes of brightening for a fixed kinematic width at z$`0.8`$ compared to $`z=0`$ (Kelson et al. 1997; van Dokkum & Franx 1996; van Dokkum et al. 1998; Treu et al. 1999). This evolution in the fundamental plane is consistent with very early, single-burst, formation scenarios. Field ellipticals show more brightening, $``$1.5 magnitudes (Gebhardt et al. 2000), suggesting a fairly recent formation epoch.
Measuring the dynamics of galaxies at cosmologically significant redshifts is observationally challenging because traditional kinematic indicators are redshifted to wavelengths inaccessible from the ground for many redshift regimes of interest. For the early-type galaxies at intermediate–to–high redshift, traditional stellar dynamical tracers like H$`\alpha `$ emission, Mg II $`\lambda 5167,5173,5814`$ and G-band absorption features at $`\lambda 4300`$, and the Ca II $`\lambda 8498,8542,8662`$ triplet shift out of the optical window. They fall within near-IR atmospheric windows or between night sky lines only for specific redshifts. At redshifts of $`z1`$, the stellar Ca H&K and Balmer lines become the only absorption lines accessible with optical spectrographs. For late type galaxies, the \[O II\] 3727 Å doublet is the most promising emission line for kinematic measurements. It is important to investigate the utility of both spectral features for measuring galaxy masses by comparing the \[O II\] velocity widths with those of other trusted emission lines (H$`\alpha `$ or 21-cm). Early results are encouraging for the \[O II\] widths. There is an excellent correlation between optical H$`\alpha `$ rotation curves and 21-cm profiles (e.g., Mathewson, Ford, & Buchhorn 1992; Raychaudhury et al. 1997; Courteau 1997 and references therein). Courteau & Faber (2000) show a tight correlation between \[O II\], H$`\alpha `$, and HI kinematics in a sample of Sc galaxies.
Another obstacle for mass estimates at intermediate $`z`$ is that most galaxies are smaller than a typical 1<sup>′′</sup> ground-based seeing disk, resulting in unresolved rotation curves. For this reason, it is important to investigate the effects of degraded spatial resolution on kinematic measurements in distant galaxies by comparing global 1-dimensional (i.e., spatially integrated) galaxy spectra with spatially-resolved spectra.
In this paper we present kinematic measurements for 22 local galaxies covering a range of morphological types, from blue compact and irregular, to spiral and active galaxies. We explore whether kinematics derived from \[O II\] $`\lambda \lambda `$3726,3729 doublet profiles yield the same results as H$`\alpha `$ and 21-cm neutral hydrogen data. We also compare the velocity widths derived from global emission line and 21-cm profiles with those derived from Balmer and Ca H&K lines to explore the utility of stellar absorption features for measuring masses in late type galaxies where absorption features have not traditionally been used. We find good agreement between different dynamical tracers. In a small fraction of extreme objects, or in data with low S/N, the \[O II\] line widths may underestimate the kinematic widths by up to 50%.
## 2. Spectroscopic Observations of Local Galaxies
We obtained optical longslit spectroscopy of 22 local ($`V<4000`$ km s<sup>-1</sup>) galaxies over the wavelength range 3650 Å - 4600 Å with the Kitt Peak National Observatory 2.1 m telescope and GoldCam spectrograph during 1999 February 15–19. Table 1 summarizes the targets, their morphological types, and their basic physical properties. Targets were selected to include a diverse assortment of morphological types from blue compact to Sb spiral galaxies. We include 8 objects from the nuclear starburst study of Lehnert & Heckman (1995): NGC 2146, NGC 2798, NGC2966, NGC 3044, NGC 3593, NGC 4527, NGC 4666, and NGC 4818. Most targets have absolute magnitudes in the range $`18<M_B<20.5`$ with inclinations $`30<i<75`$. Thus, the sample includes objects similar to galaxies used for kinematic studies at intermediate redshifts (Forbes et al. 1995; Vogt et al. 1996, 1997; Simard & Pritchet 1998; Rix et al. 1997).
The observed spectral range includes the \[O II\] $`\lambda `$3727 emission line to measure the kinematics of the ionized gas, and the stellar Balmer and the Ca H&K $`\lambda \lambda `$3970, 3932 lines to measure the kinematics of the underlying stellar population. Other prominent features in this spectral range include H$`\delta `$ and H$`\gamma `$ which may be seen either in emission from ionized gas or in absorption from stellar atmospheres. The KPNO grating #47, used in second order with a slitwidth of 2.5<sup>′′</sup> and a CUSO<sub>4</sub> filter to block first order, yielded a spectral dispersion of 0.47 Å pix<sup>-1</sup> and a mean spectral resolution of 1.57 Å (126 km s<sup>-1</sup>) FWHM at 3727 Å (2.35 Å $``$156 km s<sup>-1</sup> at 4500 Å). The spatial scale of the CCD was 0.78<sup>′′</sup> pixel<sup>-1</sup>, and seeing averaged 2<sup>′′</sup>. Calibration included corrections for bias, flatfield, slit illumination and cosmic ray rejection. Frequent exposures of an HeNeAr arc lamp provided a wavelength calibration accurate to 0.05 Å RMS ($``$4 km/s). The spatial and spectral characteristics of the spectrograph remained constant from night-to-night and from position to position on the sky, varying by $`<1`$ pixel (0.47 Å) throughout the run. Furthermore, the dispersion vector was aligned nearly perfectly with the CCD rows so that the wavelength solution (dispersion and zero point) changed by less than 1 pixel over the entire usable detector area. Based on the frequency and stability of the arc lamp exposures, we estimate that relative radial velocities from exposure to exposure are accurate to 0.25 pixels (10 km s<sup>-1</sup>). All velocities reported here correspond to the heliocentric reference frame. Occasional cirrus clouds prohibited photometric calibration, but relative fluxes based on spectrophotometric standard stars and radial velocity calibrations based on frequent arc lamp exposure were not affected.
For each galaxy we obtained total integration times of 40 to 60 minutes, broken into multiple exposures to aid cosmic ray rejection. We oriented the 5 slit along the major axis of each galaxy at the position angle listed in Table 1 and as illustrated in Figure 1. For most of the objects, we also obtained several exposures while the telescope drifted across the galaxy in a direction perpendicular to the slit. These exposures produce a “global” spectrum most closely approximating the spectra being acquired of high-redshift galaxies with typical ground-based resolution elements (several kpc). Generally the drift scans have lower signal-to-noise than the fixed exposures since a smaller fraction of the integration is spent on the high surface brightness portion of the galaxy. Table 1 records the spatial amplitude of the drift for applicable galaxies. Images of the target galaxies appear in Figure 1 along with markers showing the location of the fixed slit exposures and the extent of the drift scan exposures.
We observed 12 stars with well-known spectral types from Leitherer et al. (1996) and Lick IDS spectra from Worthey et al. (1994) in order to construct templates for the measurement of stellar absorption features in the program galaxies. The template stars span a range of spectral types, A2 through K2, and a range of luminosity classes, dwarf through supergiant. Table 2 lists these template stars.
Comparable exposures of the same galaxy were averaged to reject cosmic rays and then background subtracted to remove night sky emission lines. Background subtraction was also necessary to remove the solar spectrum which contamination from a young moon. To simulate most accurately the types of kinematic data being acquired for high-redshift galaxies, we summed the CCD rows over the observable spatial extent of each galaxy to produce a 1-dimensional spectrum. Several extraction apertures of varying size were tested, and produced essentially identical results except in the rare cases where bright H II regions at large radii generate much of the emission line luminosity. In these cases, care was taken to include the entire galaxy in the 1-D spectrum for both the fixed-slit exposures and the drift scan exposures, even though the larger extraction apertures produce 1-D spectra with lower signal-to-noise. We ultimately used the 1-D spectra with smaller extraction windows for analysis. Since the 1-D spectra are effectively intensity-weighted global spectra, the high surface brightness regions at small radii produce most of the kinematic signature. Our results for the small and large apertures yield comparable kinematic measurements, but the smaller extraction window yields more robust results due to the higher signal-to-noise.
In the analysis that follows, we convert all velocities to the heliocentric frame of reference. All velocities and velocity widths discussed in this work refer to projected velocities (in the plane of the sky). In order to compare observable quantities most directly, no assumptions or corrections for galaxy inclination have been made.
## 3. Analysis
### 3.1. Two-D \[O II\] Emission Line Rotation Curves
Courteau (1997) discusses the merits and shortcomings of several methods for extracting rotation curves from spectral data. Here we adopt a straightforward approach, constructing a rotation curve by fitting the centroids of the \[O II\] emission features along the spatial dimension. The size of each spatial bin was varied to ensure approximately equal number of counts in each bin. At each spatial bin we fit the \[O II\] doublet using a Gaussian broadening function convolved with the instrumental profile. Calibration lamp exposures provided a measurement of the instrumental spectral profile. Based on Gauss-Hermite polynomial fitting, the instrumental profile has very little skewness ($`|h|_3<0.02`$) and tail weight ($`|h|_4<0.04`$), allowing us to model it as a single Gaussian shape with $`FWHM=1.57`$ Å (126 km s<sup>-1</sup> at 3727 Å). Since the rotation curve estimates require only knowledge of the velocity centroid, a slight misreprentation of the instrumental or broadening function does not significantly affect the results. Allowing the relative strengths of the \[O II\] doublet to vary over reasonable values ($`0.5<I_{3729}/I_{3727}<1.4`$) produced the same rotation curve to within the formal uncertainties.
The resulting rotation curves appear in Figure 2. NGC 3044 is an example of a classical spiral galaxy rotation curve that is well-defined along many spatial resolution elements. He 2-10 is an example of a compact galaxy and shows little detectable rotation at this spatial and spectral resolution. Table 3 lists the (projected) velocity of maximum rotation, $`V_{max}(OII)`$, as measured from the data in Figure 2. This estimate, and its uncertainty reflects the mean and standard deviation of the last $`n`$ points in the rotation profile, where the number n ranges from 1 to 5. As discussed in Courteau (1997), the best estimate results from parameter fitting, but the differences from our approach are not significant for the goals of this paper.
### 3.2. 1-D \[O II\] Emission Line Profiles
The integrated 1-D spectra result from a convolution of the \[O II\] doublet profile with the line-of-sight Doppler velocity profile. A deconvolution analysis is required to reconstruct the intrinsic velocity profile. Since the velocity profile shapes range from the classic double-horn to a unimodal shape, the deconvolution must accommodate this large variety. We use maximum-penalized likelihood techniques to determine a non-parametric estimate of the velocity profile. The approach taken here is similar to that used in Saha & Williams (1994), Merritt (1997), and Gebhardt et al. (2000). As a first step, we construct the un-broadened \[O II\] doublet profile based on the instrumental spectral profile which is well-characterized by a Gaussian function with $`\sigma =0.67`$ Å (53 km s<sup>-1</sup> for \[O II\]). We begin the deconvolution with a trial velocity profile composed of 54 km s<sup>-1</sup>–wide velocity bins. We then convolve that profile with the measured \[O II\] instrumental doublet profile, and compare to the galaxy spectrum. The program varies the bin heights of the trial velocity profile to converge upon the best convolved galaxy spectrum, assessed using the $`\chi ^2`$ statistic. The addition of a penalty function to the $`\chi ^2`$ ensures smoothness in the velocity profile. The integrated squared second derivative represents the penalty function (as in Merritt 1997). We choose a smoothness parameter that weights the penalty function relative to the $`\chi ^2`$. This choice is partly subjective, however, a broad range of smoothing values exists over which there is little change in the derived velocity profile. More rigorous techniques for optimizing the smoothing parameter involve generalized cross validation (Wahba 1990; Silverman 1986), but are not important for our needs. The ratio of the \[O II\] doublet, $`I_{3729}/I_{3727}`$ was allowed to vary from the low density limit of 1.4 to a high density limit of 0.5 in order to achieve a best fit. In most cases, ratios less than 1.2 (electron densities greater than $`100`$ $`cm^3`$) produced unacceptable fits, consistent with observations that most HII regions exhibit line ratios indicating low electron densities (Osterbrock 1989).
Table 3 summarizes the derived velocity parameters, including the widths at 20% of the peak flux, $`W_{20}(OII)`$, and systemic velocities, $`V_0([OII])`$ (defined as the midpoint between the 20% velocities). Where the profile is Gaussian, $`W_{20}(OII)`$ can be compared to other width measurements such as the full width at half maximum (FWHM) and velocity dispersion, $`\sigma `$, using
$$\sigma =\frac{FWHM}{2.35}=\frac{W_{20}}{3.62}$$
(1)
Figure 3 displays the deconvolved \[O II\] velocity profiles for each galaxy. Profiles from the fixed slit spectra appear in dotted lines, and the drift scan spectra appear in long dashed lines. Although the lower signal-to-noise of the drift scan spectra results in slight differences between the line profiles, the correspondence between the fixed slit and drift scan spectra is generally excellent. Exceptions are NGC 1741 where the fixed slit seems to have been offset from the dominant emission line region. More anomalies and exceptions are discussed in § 4 below.
We estimate uncertainties on the emission line velocity centers and widths using a Monte-Carlo code which adds Gaussian noise to the best-fit velocity profile convolved with the instrumental profile and re-fits the profile. We generate 100 such realizations to characterize the distribution of probable values and estimate confidence limits. Table 3 includes 1 $`\sigma `$ uncertainties on the line centers and widths based on the Monte Carlo simulations.
### 3.3. 1-D Stellar Absorption Line Profiles
As another probe of galaxy kinematics, we use the stellar Ca H&K and Balmer absorption lines to measure the stellar velocity profile of each galaxy. The procedure uses the same maximum-penalized likelihood approach as the emission-line analysis. The only difference is that we must use stellar templates instead of an instrumental profile since the intrinsic widths of the absorption features are broader than the instrumental profile. Furthermore, galaxies consist of a range of stellar populations that have dramatically different spectral features and shapes, making absorption line work more complicated than emission line studies. To obtain reliable kinematic measurements, one must separate the effects of a mixed stellar population from the effects of broadening due to bulk stellar motions within the galaxy. This separation is particularly difficult in the Ca H&K region, and is why these features have not been widely used for kinematic measurements.
Since the targets encompass a broad range of morphological types, we must adequately sample the full range of stellar types in the host galaxies. The spectra of 12 Galactic A main–sequence through K giant stars served as templates (see Table 2). The maximum-likelihood technique simultaneously determines the best velocity profile and the relative contribution each stellar type. As before, we use the integrated squared second derivatives as the penalty function. The analysis software adjusts the velocity profile bin heights and the relative weight to minimize the chi-squared fit to the data. Due to the prominent, but narrow, emission lines in some galaxies in the Ca H and H$`ϵ`$ vicinity, the emission line regions are excised from the spectrum before the fit. Figure 4 shows the 3850 Å – 4040 Å region of each galaxy spectrum with prominent absorption features. Three strong emission lines are present in this range including the blended \[Ne III\] $`\lambda `$3868/He I $`\lambda `$3867, He I $`\lambda `$3889 bended with H8, and the H$`ϵ`$ blended with \[Ne III\] $`\lambda `$3967. Figure 4 shows the raw spectrum, along with with the best fit synthesis of template stars convolved with the estimated velocity profile overplotted in solid lines. Dashed lines show excised regions contaminated by emission features. Some galaxies with strong emission lines such as NGC 4214 require that extensive regions be ignored in the fit. This leads to larger uncertainties on the derived kinematic parameters. The choice of excised regions often depends on the particulars of the spectrum. In general we can fix the excised regions: if emission lines are present, we exclude them (3864 Å – 3872 Å, 3885 Å – 3892 Å, 3965 Å – 3973 Å).
One spectral region was consistently difficult to fit with any combination of template star spectra. The region from 3900 Å – 3940 Å showed significant residuals in the best fit synthesis spectra. Two different problems occur. The galaxies He 2-10, NGC 2276, NGC 2798, NGC 4214, and NGC 5248 show strong positive residuals near 3910 Å resembling emission features. The origin of these weak emission features in some galaxies is not clear. Such stellar features do not appear in any of our template spectra, nor are there any known nebular emission lines in this region. One possible source of extraneous lines in this portion of the spectrum is chromospheric emission from giants and binary systems (Stencel 1975; Strassmeier et al. 1993). However, these features are usually weak and would probably not appear at a significant level in the integrated spectrum of an entire galaxy. Another possibility is CN molecular bands in giant stars, but this effect should produce a larger signature in late type galaxies, contrary to the observed trend. We have not been able to find a satisfactory explanation for the origin of these features. We decided on a case-by case basis whether to exclude this region. When no emission lines are present, we use the full spectral range (e.g., NGC 4527, NGC 4666). Another difficulty involves fitting template spectra to galaxies like NGC 2798, NGC 3044, NGC 4218, and NGC 4818, where the Ca K line is narrower and deeper than any combination of template stars can match. Either of these difficulties are plausibly due to metallicity mismatches between the template stars and the galaxies studied. The stellar templates in Table 2 do not include metal-poor stars which are probably common in some of the smaller star-forming galaxies like NGC 4214.
Figure 3 plots the resulting absorption line velocity profiles for each galaxy. We record in Table 3 the velocity width at 20% of the maximum profile depth, and the systemic velocity defined here as the midpoint between the 20% velocities. We estimate uncertainties on the velocities and widths using the same Monte-Carlo technique described in § 3.2.
Sixteen of the 22 targets had sufficient signal-to-noise and measurable absorption line features for analysis. Five of the six galaxies without detectable absorption features (NGC 1741, Mrk 1089, NGC 5253, UM 439, and UM 462) exhibit strong nebular emission which dominates the spectrum. The low continuum signal-to-noise combined with a strong background solar spectrum from the crescent moon precluded a meaningful measurement for NGC 925.
### 3.4. 21-cm Neutral Hydrogen Spectra
We compiled from the literature the single-dish 21-cm neutral hydrogen spectrum for each galaxy. When more than one profile was available, we selected the one with the best combination of signal-to-noise and velocity resolution, (typically $`11`$ km s<sup>-1</sup>). Most single-dish observations are based on data with the 92 m NRAO Greenbank radio-telescope, which has a FWHP beamsize of 10, large enough to adequately cover the angular extent of galaxies in our sample. In order to compare the 21-cm profiles with the newer optical data, we digitized the original published spectra to generate a spectrum in electronic form.<sup>1</sup><sup>1</sup>1In most cases, the 21-cm spectra date from the early 1980’s, and the data are not recoverable on electronic media from the original authors. From these digitized spectra we measured the full-widths at 20% of the peak intensity, $`W_{20}(HI)`$, a standard parameter commonly tabulated in 21-cm studies. We also computed the systemic velocity, $`V_0(HI)`$, defined here as the midpoint between the 20% velocities. Table 3 lists the original publications for each 21-cm spectrum, along with 21-cm systemic velocities and $`W_{20}(HI)`$. Our re-measurements of the systemic velocities and velocity widths from the digitized data are in good agreement with the published values. We estimate that the digitization process introduced errors of $`<5`$ km s<sup>-1</sup> in velocity and $`<`$5% in amplitude compared to the original data. The resulting HI spectra appear in Figure 3. There are no published HI spectra for NGC 4818 but only marginal detections.
To complement the 1-D 21-cm profiles, we searched the literature for spatially resolved rotation curve measurements from aperture synthesis observations. Table 3 lists the maximum HI rotational velocity, $`V_{max}(HI)`$, for all galaxies with suitable data, as found in the original references. If more than one aperture synthesis measurement was available, we adopted the most sensitive study, usually from recent Very Large Array (VLA) programs.
## 4. Results
### 4.1. Comparison of Velocity Widths Measured from HI \[O II\], and Stellar Absorption Features
One of the goals of this program was to search for systematic differences between morphological types or pathological systems of any type which exhibit markedly different kinematic signatures in one or more of the measured tracers. Figure 3 shows graphically the one-dimensional velocity profiles measured from 21-cm, \[O II\], and stellar absorption line spectroscopy. The 21-cm profiles appear in solid lines, the fixed slit \[O II\] profiles in dotted lines, the drift scan \[O II\] spectra in long dashes, and the absorption spectra in short dashes. A cursory comparison by eye reveals that the overall systemic velocities and velocity widths measured from the different tracers are generally in good agreement. We discuss exceptional systems in more detail below. Henize 2-10 and Mrk 33 are example of low-mass systems with Gaussian velocity profiles where the three dynamical indicators are in excellent agreement. NGC 2798 and NGC 3044 are examples of larger galaxies where all three measurements yield comparable results, even though the 21-cm and \[O II\] profiles are double-peaked and the absorption profile from the stellar component is single-peaked.
Figure 5 compares the derived velocity widths from each of the three kinematic tracers. In the upper right panel we show the the \[O II\] full width at 20% max, $`W_{20}([OII])`$, versus $`W_{20}(HI)`$. Filled symbols distinguish the fixed-position spectra from drift scan spectra (open symbols). Dashed lines denote 20% deviations from the 1:1 correspondence (solid line). NGC 4818 and Mrk 1089 could not be plotted since they do not have published neutral hydrogen spectra. Figure 5 shows a strong correlation between $`W_{20}([OII])`$ versus $`W_{20}(HI)`$, most galaxies falling within 20% of the 1:1 relation. The drift scans yield results consistent with the fixed slit exposures, within the uncertainties. NGC 4449 and NGC 4666 stand out as having broad HI widths compared to the \[O II\] profile both here and in Figure 3. Inspection of their spectra in Figure 3 reveals that NGC 4449 exhibits a small kinematic width in \[O II\] perhaps because the HII regions are concentrated near the center of an extended neutral gas distribution and trace only a small fraction of the gravitational potential. NGC 4666 is a starburst galaxy with low-level \[O II\] emission across the entire HI velocity width, and a strong emission line concentration on one velocity wing. This emission is confined to one portion of the galaxy, and does not trace the full gravitational potential, giving the false signature of a less massive galaxy. If the spectra had been of lower S/N, the weak \[O II\] emission across the entire 300 km s<sup>-1</sup> velocity width seen in HI would have gone unrecognized. NGC 2966 exhibits slightly smaller HI widths compared to the \[O II\] profile and stellar profile. Unfortunately, there is only one published HI profile for this galaxy, so we are unable to explore the possibility that the neutral hydrogen width is somehow underestimated.
The upper left panel of Figure 5 compares the HI line widths with the absorption line widths. There is again a good correlation between the two indicators. Within the often substantial uncertainties on the absorption line widths, all the data are consistent with the 1:1 correspondence except the irregular galaxy NGC 4214. The Seyfert galaxy NGC 1068 is also discrepant, but since we are unable to spatially separate the high-velocity nuclear regions probed by stellar absorption from the rest of the galaxy, agreement is not expected. We do not consider NGC 1068 further, except as an example of an AGN where the nuclear kinematics are distinct from the overall gravitational potential of the galaxy.
The lower right panel of Figure 5 compares the \[O II\] emission versus stellar absorption velocity widths. Once again NGC 4666 and NGC 4449 and NGC 4214 stand out as deviant points for the reasons discussed above. The rest of the data show good agreement between the two kinematic indicators.
As an alternative measure of the velocity width, we computed the difference between velocities where the integrated area under the profile on either side reaches 10% of the total area (see Courteau 1997 for discussion of this and other methods). This approach is more robust when measuring objects like NGC 4666 because it is not as sensitive to instrumental resolution or asymmetric profiles. However, this approach is prone to larger uncertainties for noisy data. The \[O II\] width of NGC 4666 changes from 170 km s<sup>-1</sup> to 392 km s<sup>-1</sup> when measured in this manner, bringing the \[O II\] width into accord with the other dynamical indicators. In general, the measured linewidths of all target galaxies become 15% to 25% smaller using this approach. The correlation between the various dynamical indicators explored here remains consistent with the above discussion.
### 4.2. Comparison of Rotation Curves Measured from HI \[O II\], and Stellar Absorption
When spatially-resolved velocity data are available, rotation curves provide more information about a galaxy’s kinematics than an integrated profile alone. In each of the target galaxies we measured \[O II\] rotation curves and the maximum rotational velocity, $`V_{max}([OII])`$. Table 3 lists these values. Figure 6 shows a comparison of the projected $`V_{max}([OII])`$ with the 21-cm rotation curve maximum, $`V_{max}(HI)`$ where such data were available in the literature. We also compare the $`V_{max}`$ measurements to estimates of $`V_{max}`$ based on $`W_{20}(HI)`$ from the prescription of Tully & Fouqué (1985),
$$W_R^2=W_{20}^2+W_t^22W_{20}W_t[1e^{(W_{20}/W_c)^2}]2W_t^2e^{(W_{20}/W_c)^2}.$$
(2)
Here, $`W_R`$ is the rotation full amplitude which is 2$`\times V_{max}`$. $`W_t=38`$ km s<sup>-1</sup> is the width due to turbulent motions and $`W_c=120`$ km s<sup>-1</sup> is the transition point between galaxies having Gaussian and those having double-horned HI profiles. We see in Figure 6 that the maximum points on the rotation curves measured either from neutral hydrogen aperture synthesis data or \[O II\] are systematically less than those predicted by simply taking half of $`W_{20}(HI)`$. There is better agreement with the analytic expression of Tully & Fouqué which takes into account the effects of turbulence that become more important in the smallest galaxies. The most discrepant points are the compact low-mass galaxies like He 2-10, Mrk 33, and NGC 925 where measuring a rotation curve becomes difficult due to limited spatial resolution, or because the intrinsic line width due to turbulent motions becomes comparable to the (projected) rotational velocity. The \[O II\] rotation curve may easily be underestimated due to poor sensitivity or spatial resolution, especially at large distances.
### 4.3. Comparison of Systemic Velocities Measured from HI \[O II\], and Stellar Absorption
As a measure of the systemic velocity, $`V_0`$, for each system, we record in Table 3 the velocity midway between the 20% intensity points used to define $`W_{20}`$. This formulation of the systemic velocity is a well-defined quantity in the case of massive galaxies with two-horned velocity profiles. It is also less sensitive to asymmetries in the line profile in the case of low-mass galaxies with Gaussian profiles where most of the nebular emission may come from a single HII region. Figure 7 illustrates the differences between systemic velocities measured with each technique. We show, as a function of the (projected) neutral hydrogen width, the difference between the $`V_0(HI)V_0(Abs)`$ (upper left panel), $`V_0(HI)V_0([OII])`$ (upper right panel), and $`V_0([OII])V_0(Abs)`$ (lower right panel). Open symbols distinguish drift scan exposures from fixed-slit integrations (solid symbols). Overall, the three methods yield similar velocities for most galaxies to within the uncertainties on the fits. In nearly all cases the drift scans show better agreement than the fixed slits. In the upper left panel, there is good agreement between the systemic velocities measured with HI and stellar features.
Turning to the upper right panel of Figure 7, a few galaxies show real differences discrepancies as large as 50 km s<sup>-1</sup> between the HI and \[O II\] velocities. In the case of NGC 4527, $`V_0([OII])`$ exceeds $`V_0(HI)`$ by 70 km s<sup>-1</sup>. Since we do not have a drift scan of this object, we suspect that the fixed slit position missed much of the emission line gas. The line width derived from the \[O II\] spectrum is narrower ($`W_{20}=409`$ km s<sup>-1</sup>) than the Lehnert & Heckman \[N II\] width of $`W_{20}=500`$ km s<sup>-1</sup>. Their rotation curve amplitude of 339 km s<sup>-1</sup> is significantly larger than our 260 km s<sup>-1</sup>, consistent with the probability that our choice of slit placement did not cover the entire emission-line extent of the galaxy. Other galaxies show only marginal evidence for systematic offsets between the neutral and ionized gas at the level of $`<10`$ km s<sup>-1</sup>.
### 4.4. Comments on Individual Objects
Henize 2-10: Henize 2-10 is a blue compact galaxy with a solid body rotation curve typical of dwarf and compact objects (Kobulnicky et al. 1995). Figure 3 shows that the HI, and \[O II\] features indicate a similar width of $`W_{20}=156\pm 15`$ km s<sup>-1</sup> and $`W_{20}=158\pm 10`$ km s<sup>-1</sup>. The measured width of stellar features is slightly greater but consistent within the uncertainties $`W_{20}=205_{70}^{+132}`$ km s<sup>-1</sup>. The systemic velocity as measured by the HI spectrum appears consistent with the optical data. Given the strong star formation and high optical depth, it is likely that only the nebular emission on the near side of the galaxy is seen in the optical spectra.
Markarian 33: Another blue compact galaxy, Mrk 33 shows an HI line width ($`W_{20}=208\pm 25`$ km s<sup>-1</sup>) which is 30 km s<sup>-1</sup> larger than, but consistent with, the nebular emission ($`W_{20}=168\pm 5`$ km s<sup>-1</sup>) and the stellar line width ($`W_{20}=187_{48}^{+58}`$ km s<sup>-1</sup>). The emission line gas is evidently confined to only a fraction of the rotation curve traced by the neutral hydrogen and stellar components as suggested by the 1-D profiles in Figure 3. The \[O II\] rotation curve has a very small measurable amplitude compared to the intrinsic line width measured in the optical and neutral gas, causing Mrk 33 to stand out in Figure 6. This small rotation amplitude is most probably a result of limited spatial resolution and the relatively larger distance of Mrk 33.
NGC 925: This is a moderately inclined late-type spiral for which we can only measure \[O II\] in the optical spectrum. The stellar absorption features are weak, and at the low redshift of 550 km s<sup>-1</sup>, its Ca and Balmer features are strongly contaminated by the solar spectrum produced by a nearby 15% illuminated moon. The emission line profile is narrow ($`W_{20}=176\pm 7`$ km s<sup>-1</sup>) compared to the HI profile ($`W_{20}=222\pm 6`$ km s<sup>-1</sup>). The \[O II\] and HI systemic velocities agree well near $`V_0=555\pm 3`$ km s<sup>-1</sup>.
NGC 1068: As the only AGN (Seyfert 2 class) in our sample, NGC 1068 shows a broad nuclear emission component which produces the unusual rotation curve in Figure 2. The drift scan exposure yields a lower velocity width than the fixed slit which is dominated by the nuclear kinematics, so we consider only the former. On the basis of the integrated 1-D optical spectrum, it is not possible to deduce a velocity width that accurately reflects the rotational velocity of the outer disk. The measured emission line width is $`W_{20}([OII])=1129\pm 200`$ km s<sup>-1</sup> compared to the much lower stellar ($`W_{20}(Abs)=564_{59}^{+38}`$ km s<sup>-1</sup>) and HI kinematics ($`W_{20}(H\mathrm{I})=298\pm 30`$ km s<sup>-1</sup>). Because of the unique nature of the this object, we do not consider it further for analysis, except to note this as an instance where the nebular lines would clearly lead to an erroneous estimate of the dynamical mass of the system.
NGC 1741/Mrk 1089: These galaxies are members of a Hickson Compact Group (HCG 31) components a and c respectively. Both objects exhibit strong emission lines observed simultaneously with a single slit placement. A comparison of the drift scan exposure with the fixed slit shows that, in the case of NCG 1741, a small displacement of the slit from the dominant emission line region of the galaxy can yield a discrepant systemic velocity. The available HI profiles encompass both galaxies within the large beam of typical radio telescopes. VLA aperture synthesis maps (Williams et al. 1991) reveal a complex HI distribution indicating that these galaxies are interacting and are possibly connected by a common pool of neutral hydrogen. Most of the neutral gas is affiliated with NGC 1741, with secondary peaks in Mrk 1089 and the other two members of the group. This configuration explains why the HI spectrum in Figure 3 reflects more closely the \[O II\] kinematics of NGC 1741 than Mrk 1089 (not plotted). Comparisons between the nebular and HI kinematics of Mrk 1089 are thus inappropriate until a separate HI profile of Mrk 1089 is available. It was not possible to discern stellar absorption features in either of these objects.
NGC 2146: This early type spiral has the largest velocity amplitude in our sample. Figure 5 shows that the HI width is in good agreement the nebular and stellar tracers. The systemic velocities measured by each tracer are in good agreement for the drift scan measurements, as are the maximum rotation speeds measured from the HI and \[O II\] rotation curves. Systemic velocities measured from the fixed slit only yield \[O II\] centroids which are 60 km s<sup>-1</sup> smaller than the stellar and 21-cm systemic velocities. This result underscores potential difficulties with using single-position slits rather than drift scans to measure recessional velocities of spatially-resolved galaxies.
NGC 2276: A moderate-inclination late-type spiral, NGC 2276 shows good agreement between all three kinematic width tracers in Figure 5. In Figure 7, and from the profiles plotted in Figure 3, we note a 20 km s<sup>-1</sup> offset between the neutral hydrogen profile and the optical \[O II\] and stellar profiles. The $`H\alpha `$ image in Young et al. (1996) reveals a potential reason for this offset. In NGC 2276, HII regions and stellar light preferentially occupy the western half of the galaxy, so that the (intensity weighted) 1-D \[O II\] and stellar absorption profiles are biased away from the systemic velocity. This object serves to illustrate the magnitude of scatter introduced by inhomogeneous distributions of stellar and gaseous components.
NGC 2798: HI and stellar velocity widths in this low-inclination Sa galaxy are consistent near $`310\pm 30`$ km s<sup>-1</sup>, but the optical emission line velocities are somewhat broader at 387$`\pm 21`$ km s<sup>-1</sup>. The measured systemic velocities agree well in Figure 7 within the uncertainties. These substantial errors are due to the relatively low-S/N. The H$`\alpha `$ profile of the nuclear regions which dominates the spectrum is (Lehnert & Heckman 1995)<sup>2</sup><sup>2</sup>2Lehnert & Heckman report \[N II\] profiles in terms of the full width at half maximum (FWHM) and the rotation curve parameters in terms of the rotation full amplitude, $`A_{rot}`$. For comparison, we convert these parameters to 20% widths and rotation curve maxima using $`W_{20}=1.54\times FWHM`$ and $`V_{max}=0.5\times A_{rot}`$. $`W_{20}(H\alpha )=362`$ km s<sup>-1</sup>, consistent with our measurement of $`W_{20}([OII])=387\pm 21`$ km s<sup>-1</sup>.
NGC 2966: This low-inclination spiral exhibits the classic double-horned velocity profile in both HI and in \[O II\]. Both the stellar absorption and \[O II\] profiles are 20-40% broader than the HI profile $`W_{20}(HI)=258\pm 25`$ km s<sup>-1</sup>. Unfortunately, there is only one published 21-cm spectrum of NGC 2966, so we are unable to verify the HI parameters. The drift scan \[O II\] data also yields an unusually broad profile compared to the fixed slit. Lehnert & Heckman (1995) present an H$`\alpha `$ image which reveals a nuclear emission region with additional HII regions in the spiral arms at large radius. Their H$`\alpha `$ longslit spectroscopy yields a FWHM of 257 km s<sup>-1</sup> for their largest measurement, which would correspond to $`W_{20}=395`$ km s<sup>-1</sup>, marginally consistent with our $`W_{20}([OII])=362\pm 15`$ km s<sup>-1</sup>. Their maximum projected rotational velocity, is 116 km s<sup>-1</sup> which is smaller than our 150$`\pm `$10 km s<sup>-1</sup>. Given the agreement between our optical results and those of Lehnert & Heckman, we are suspicious of the low HI width and would urge a new single-dish measurement of this galaxy.
NGC 3044: As one of the largest and brightest galaxies in our sample, NGC 3044 has a well defined rotation curve and velocity profile in every tracer. The profiles in Figure 3, and the kinematic summaries in Figures 5 and 7 shows that the shows that the HI and \[O II\] velocity widths are in good agreement near $`W_{20}=360\pm 10`$ km s<sup>-1</sup>. However, the velocity profile measured from the absorption lines is substantially more narrow, $`W_{20}(Abs)=279_{66}^{+79}`$ km s<sup>-1</sup>. This appears to be an instance where the light-weighted stellar spectrum samples a smaller portion of the rotation curve than neutral hydrogen and ionized gas at larger radius. The systemic velocities are in excellent agreement.
NGC 3593: In this Sa type spiral, linewidths of the HI and stellar absorption agree well in NGC 3593, $`W_{20}300`$ km s<sup>-1</sup>, while the optical emission lines are much broader, $`W_{20}([OII])=411\pm 25`$ km s<sup>-1</sup>. This spectrum has low S/N and fairly large uncertainties. Our optically-measured line widths of $`W_{20}=300400`$ km s<sup>-1</sup> are much larger than the 230 km s<sup>-1</sup> widths implied by the \[N II\] spectroscopy of Lehnert & Heckman (1995). However, a direct comparison of the velocity widths is not possible since Lehnert & Heckman subdivide their spectra in to several different radial bins whereas we consider only the entire spatially integrated profile. Despite this possible difference line widths, the maximum \[O II\] circular velocity of $`V_{max}=103\pm 10`$ km s<sup>-1</sup> is roughly consistent with the \[N II\] circular velocity of 118 km s<sup>-1</sup> reported by Lehnert & Heckman. The systemic velocities measured from all three indicators are in excellent agreement.
NGC 4214: This Magellanic irregular has one of the narrowest velocity profiles in our sample, $`<100`$ km s<sup>-1</sup>, at the limit of our resolution. The HI width is only 82 km s<sup>-1</sup>, comparable to the optical results which suggest $`W_{20}([OII])60`$ km s<sup>-1</sup>. Multiple results from the literature confirm this low neutral hydrogen width. The systemic velocities in Figure 7 show good agreement. Curiously, the stellar velocity profile is much broader, $`W_{20}(Abs)=132_{27}^{+55}`$ km s<sup>-1</sup>, making NGC 4214 one of the most outstanding points in Figure 5.
NGC 4218: The velocity centroids and widths of all measurements are in good agreement for this moderately inclined Sa type spiral, ranging between $`W_{20}=160190`$ km s<sup>-1</sup>. The HI profile is centered centered 20-30 km s<sup>-1</sup> lower than the optical results. The velocity widths of the \[O II\] emission lines are broad, $`W_{20}([OII])=195\pm 6`$ km s<sup>-1</sup> compared to the absorption 21-cm width of $`163\pm 20`$ km s<sup>-1</sup>. Aperture synthesis measurements of the 21-cm profile yield a surprisingly narrow $`W_{20}=138`$ km s<sup>-1</sup> (Verheijen 1997) while single dish profiles range between 157 km s<sup>-1</sup> and 201 km s<sup>-1</sup>, consistent with the optical results.
NGC 4449: Similar in many ways to NGC 4214, this irregular galaxy has an unusually large HI halo extending to 14 times the optical radius (Bajaja et al. 1994). The HI and stellar absorption profiles agree well, near $`W_{20}190`$ km s<sup>-1</sup>, while the \[O II\] fixed and drift exposures are considerably more narrow, $`W_{20}=90\pm 21`$ km s<sup>-1</sup>. Since the signal-to-noise is high in both the optical and 21-cm spectra, this difference s a robust result. NGC 4449 is an example of an object where the global 1-D (intensity-weighted) emission line profile traces only a portion of the rotation curve. Rotational kinematics based only on an \[O II\] profile would underestimate the rotational velocity by 50%. NGC 4449 is the most significant outlier in Figure 5. Figure 7 shows that the systemic velocities measured from each method are in good agreement.
NGC 4527: One of the most massive galaxies in our sample, NGC 4527 shows good agreement between the absorption line and HI kinematics, $`W_{20}=370`$ km s<sup>-1</sup>. Lehnert & Heckman report an \[N II\] width of $`W_{20}=500`$ km s<sup>-1</sup> in the nuclear region. Their rotation curve amplitude of 339 km s<sup>-1</sup> is significantly larger than our 232$`\pm `$13 km s<sup>-1</sup>, suggesting that our choice of fixed slit placement did not cover the entire emission-line extent of the galaxy. The \[O II\] systemic velocity is larger by 60 km s<sup>-1</sup> but carries large uncertainties due to the presence of low-level emission/noise at extreme velocities (see Figure 3).
NGC 4666: Another large, edge-on galaxy, NGC 4666 exhibits good agreement between the velocity widths measured in HI and stellar features near $`W_{20}=400`$ km s<sup>-1</sup>. The double-horned HI profile contrasts with the single-peaked stellar absorption profile as shown in Figure 3. The \[O II\] profile shows one dominant emission peak and superimposed on low-level emission. The dominant emission peak centered near 1700 km s<sup>-1</sup> gives the misleading appearance of a narrow velocity profile with $`W_{20}([OII])=170\pm 13`$ km s<sup>-1</sup>. NGC 4666 serves as a warning that in a low S/N spectrum of a typical high-redshift galaxy, this object could appear to have a small kinematic width. Figure 5 shows that the measured \[O II\] profile is less than half as large as the HI profile. Unfortunately we do not have a drift scan observation of this target to verify that an integrated spectrum would produce the same result. This unusual emission line profile leads to an artificially aberrant emission line centroid, displaced toward higher velocities.
NGC 4818: There is no published HI spectrum for this galaxy, and only recently is a detection reported by Theureau et al. (1998). The absorption line profile is slightly more narrow than the emission line profile ($`231\pm 80`$ km s<sup>-1</sup> versus $`276\pm 45`$ km s<sup>-1</sup>) in this early type spiral.
NGC 5248: Here the HI, \[O II\] and stellar profiles give very similar results, $`W_{20}=293301`$ km s<sup>-1</sup>, for this Sbc galaxy. Likewise the systemic velocities agree well. Like NGC 3044, the stellar profile is single peaked, while the HI and \[O II\] profiles are double-horned.
NGC 5253: NGC 5253 is an amorphous galaxy with a strong central starburst and nebular emission. No stellar features are measurable. The HI and \[O II\] profiles are Gaussian and agree well in their velocity widths. Lehnert & Heckman (1995) find a maximum rotational amplitude of 3 km s<sup>-1</sup> amidst the distorted and irregular velocity field. Martin & Kennicutt (1995) show ordered velocity variations of 25 km s<sup>-1</sup> along the minor axis, consistent with the unusual minor-axis HI kinematics seen in VLA maps (Kobulnicky & Skillman 1995).
UM 462: A blue compact dwarf galaxy, the spectrum of UM 439 is dominated by emission lines and no stellar features are measurable. The \[O II\] rotation curve is consistent with solid body rotation and is measurable over 25 pixels (1.4 kpc). The velocity widths in HI and \[O II\] are in good agreement.
UM 439 : As with UM 462, the velocity profiles are Gaussian and well-matched in both HI and \[O II\] with $`W_{20}100`$ km s<sup>-1</sup>. No stellar features are measurable.
## 5. Discussion and Conclusions
We have examined the rotation curves, systemic velocities, and 1-D velocity widths for a wide variety of galaxy types and inclinations to search for systematic deviations between measurements made with 21-cm, nebular \[O II\], and stellar absorption spectra. The three methods provide kinematic line widths which agree to within 20% for most (19/22) objects, and within 50% in all galaxies studied. NGC 4449, an irregular galaxy with an extremely extended neutral hydrogen halo, is the most discordant: the \[O II\] line width under predicts the 21-cm line width by a factor of 1.6. NGC 4666 also shows broad extended nebular emission and a strong narrow \[O II\] peak which might be interpreted as a signature of a small dynamical mass had the signal-to-noise been lower. NGC 4666 serves as a warning that in a low S/N spectrum of a typical high-redshift galaxy, this object could appear to have a small kinematic width.
In general, however, the stellar kinematics as traced by the absorption features yield line widths which are nearly as reliable as the \[O II\] doublet, even in many late-type galaxies with strong nebular emission lines. Due to the patchy distribution of HII regions within some galaxies, the stellar features yield more reliable estimates of the systemic velocities (as measured by HI 21-cm spectra) than the \[O II\] data. The spatially-integrated drift exposures yield better estimates of the velocity widths and systemic velocities. The intrinsic dispersion in the difference between systemic velocities measured from the gas and stars appears to be less than 10 km s<sup>-1</sup>, comparable to our measurement uncertainty, in most systems.
Although our sample does not include early-type galaxies, we expect that the Balmer and Ca H&K region will prove suitable for absorption-line kinematic analysis in those systems as well. For the galaxies that show no visible emission lines, we were able to use the complete spectrum without needing to excise any spectral region. In these cases, the agreement with the HI profiles was excellent. Since early-type galaxies generally do not show emission features, the full optical spectrum may be used in a kinematic analysis. However, we advise caution when emission lines are present. Spectral regions contaminated by emission lines must be excised before measuring the absorption width. Furthermore, in some galaxies, the region from 3900 Å – 3940 Å is affected by a series of weak unidentified lines which can cause difficulty at discussed in §3 above. This region should be excised if adequate template stars are not available.
One goal of this program was to determine whether galaxy masses derived from \[O II\] emission and stellar absorption features are intrinsically noisier or biased compared to more traditional dynamical indicators. In order to estimate the intrinsic scatter between HI, \[O II\], and absorption-feature velocity widths in Figure 5, we compare the observed RMS to the expected RMS from the formal uncertainties computed by Monte Carlo methods. The measured RMS between the HI width and \[O II\] emission is 47 km s<sup>-1</sup>, and for the absorption width it is 54 km s<sup>-1</sup>. Given the uncertainties, the expected RMS is 40 and 33 km s<sup>-1</sup>, respectively. Thus, the data require an additional intrinsic scatter beyond the experimental measurement errors. This additional scatter is 18 km s<sup>-1</sup> added in quadrature to the \[O II\] width uncertainties, and 20 km s<sup>-1</sup> for the absorption widths. Since a typical galaxy in our sample has a velocity width greater than 200 km s<sup>-1</sup>, either kinematic tracer provides a suitable measure of the width, better than 10% in velocity or 20% in the mass (given that mass, $`M\mathrm{\Delta }V^2`$). Surprisingly, this result is true using absorption line kinematics in some of the very late-type galaxies where the mix of stellar population causes large spatial variations in the spectrum. For low mass galaxies with narrow ($`<150`$ km s<sup>-1</sup>) velocity widths, the measurement uncertainties can approach 50% in velocity and, thus, factors of 2 in mass. While these kinematic indicators should still be useful for measuring the gravitational potential of low-mass galaxies, the relative uncertainties will be significant.
For the most massive systems, however, uncertainties due to the measured intrinsic scatter are smaller than, or comparable to, the dispersion observed in local Tully-Fisher and Fundamental Plane relations. For Tully-Fisher, the scatter has been estimated to be 0.2 magnitudes (Pierce & Tully 1992; Tully & Pierce 1999) to 0.25-0.4 magnitudes (Courteau 1997). This translates into a 12.5% to 20% error in distance or a 10% – 16% error in velocity width (assuming typical slopes, $`a`$, of 6–7 in the Tully-Fisher relation; $`M=a[logV_c2.5]+b`$ ). Thus, the 18 km s<sup>-1</sup> uncertainty, which may be introduced by adopting \[O II\] emission line widths or stellar absorption line widths, only begins to add to the distance scatter for galaxies with W<sub>20</sub> below 200 km s<sup>-1</sup>. Given that the Tully-Fisher galaxies are generally the most massive since they are the easiest to observe and have well-defined rotation curves (as used in Vogt et al. 1997), this additional scatter will have at most a small effect for distant Tully-Fisher analysis.
For the Fundamental Plane, the scatter is around 8% (Jørgensen et al. 1998) in logarithmic effective radius, implying a scatter of 0.25 magnitudes in surface brightness or a 15% error in the velocity dispersion. Since most of our additional uncertainties are below 10% for the absorption line widths, it will have little effect on either the Fundamental Plane scatter or in the estimate of the difference in surface brightness. Since both \[O II\] emission and stellar Balmer/Ca H&K absorption linewidths correlate well with HI 21-cm linewidths we anticipate the increased use of these kinematic indicators to derive galaxy masses in the distant universe.
We are grateful for helpful conversations and suggestions from Matt Bershady, Sandy Faber, Luc Simard, Drew Phillips, Stéphane Courteau, and the referee, Chris Pritchet. We thank Liese van Zee for providing the 21-cm spectra of UM 439 and UM 462, D. J. Pisano for the 21-cm spectra of NGC 925, Siow-Wang Lee for 21 cm spectra of NGC 3044 in electronic form, and Benjamin Weiner for the image of NGC 1068. We thank Evan Skillman and David E. Hogg for their help in tracking down older HI spectra which were surprisingly difficult to locate by electronic means. Greame Smith and Jean Brodie offered helpful advice on the nature of the troublesome stellar features in the 3900 Å – 3930 Å region. H. A. K and K. G. are grateful for support from Hubble Fellowship grants #HF-01094.01-97A and #HF-01090.01-97A awarded by the Space Telescope Science Institute which is operated by the Association of Universities for Research in Astronomy, Inc. for NASA under contract NAS 5-26555. |
no-problem/9911/nucl-ex9911007.html | ar5iv | text | # Nuclear and Coulomb Interaction in the 8B → 7Be + p Breakup Reaction at sub-Coulomb Energies
## Abstract
The angular distribution for the breakup of <sup>8</sup>$``$ <sup>7</sup>Be + p on a <sup>58</sup>Ni target has been measured at an incident energy of 25.75 MeV. The data are inconsistent with first-order theories but are remarkably well described by calculations including higher-order effects. The comparison with theory illustrates the importance of the exotic proton halo structure of <sup>8</sup>B in accounting for the observed breakup angular distribution.
Coulomb dissociation reactions have been used in recent years as a means to obtain information on capture reactions of astrophysical interest. An example is the experiment of Motobayashi, et al. who studied the breakup of <sup>8</sup>$``$ <sup>7</sup>Be + p on a Pb target and related their result to radiative proton capture at solar energies. This reaction corresponds to the projectile breaking up into a core and a valence nucleon due to interactions with virtual photons in the strong Coulomb field of a high-Z nucleus. Although this mechanism is in principle the time reversal of a capture reaction, $`E2`$ photons can contribute significantly to Coulomb dissociation while radiative capture at solar energies proceeds almost exclusively by $`E1`$ transitions. Thus, in extracting information on astrophysical proton capture reactions from the measured dissociation cross section it is crucial to determine the relative contribution of photons having different multipolarity.
The relative importance of $`E1`$ and $`E2`$ contributions to the Coulomb dissociation of <sup>8</sup>B has been investigated both experimentally and theoretically . The earliest experiments suggested that the $`E2`$ strength was much smaller than all published theoretical estimates. Davids, et al. measured the asymmetry in the longitudinal momentum distribution of <sup>7</sup>Be fragments from the dissociation of <sup>8</sup>B on Pb at 44 and 81 MeV per nucleon. Their high-energy data were quite ambiguous, but the 44 MeV/nucleon results gave a clear signal corresponding to an $`E2`$ strength that was 70$`\%`$ of that predicted by the model of Esbensen and Bertsch . This model prediction itself is a factor of two smaller than that of Kim, et al. . Nevertheless the extracted $`E2`$ strength, though considerably quenched, is still larger than implied in Refs. and . Most recently, Iwasa, et al. report a limit on the $`E2`$ strength that is at least an order of magnitude smaller than that of Davids, et al..
It was noted in Ref. that the description of the data by the model of Esbensen and Bertsch is not precise, and that the best-fit values for the $`E1`$ and $`E2`$ strengths differ by 20-30$`\%`$ from the model predictions. The $`E2/E1`$ interference term is, of course, model dependent. The earlier experiment of von Schwarzenberg, et al. had attempted to avoid model dependence by measuring the breakup at sub-Coulomb energies for a low-Z target (<sup>58</sup>Ni) for which multiple Coulomb excitation was expected to be minimal. At these energies, the E2 component is enhanced relative to E1. The very small cross section reported in that work, which was less than that predicted by any reasonable structure model for <sup>8</sup>B , has generated considerable interest. Nunes and Thompson and Dasso, Lenzi, and Vitturi independently suggested that the explanation for this result might be strong destructive nuclear-Coulomb interference effects, despite the fact that at the angle where the measurement was made the classical distance of closest approach is nearly 20 fm, i.e., far outside the range of the nuclear force for a “normal” nuclear system. A strong nuclear-dominated peak in the differential cross section at a center-of-mass angle of $`70^\mathrm{o}90^\mathrm{o}`$ (well inside the expected $`100^\mathrm{o}110^\mathrm{o}`$ for the onset of nuclear breakup of a normal nucleus) was predicted in Refs. and , although it was pointed out that the corresponding calculations are only first-order in the nuclear and Coulomb fields and might be modified by multi-step excitations. Furthermore, it was suggested in Ref. that even pure Coulomb excitation would be considerably modified from that expected in the normal “point-Coulomb” approximation which ignores the extended size of the valence proton wave-function in <sup>8</sup>B (see Ref. for a more complete discussion of this approximation). This leads to a further reduction in the calculated breakup cross section. Since both these effects are directly attributable to the exotic “halo” structure of <sup>8</sup>B, it is important to verify, if possible, the implications of these theoretical calculations. The present work was carried out in an attempt to test these predictions by obtaining angular distribution for the breakup of <sup>8</sup>B on <sup>58</sup>Ni at the same incident energy as the previous experiment , but over as much of the critical angular range as possible.
The experiment was carried out at the Nuclear Structure Laboratory of the University of Notre Dame. To produce the low-energy secondary radioactive <sup>8</sup>B beam, we used the TwinSol radioactive ion beam (RIB) facility and the <sup>6</sup>Li(<sup>3</sup>He,n)<sup>8</sup>B direct transfer reaction. A gas target containing 1 atm of <sup>3</sup>He was bombarded by a high-intensity (up to 300 particle-nA), nanosecond-bunched primary <sup>6</sup>Li beam at an energy of 36 MeV. The entrance and exit windows of the gas cell consisted of 2.0 $`\mu `$m Havar foils. The secondary beam was selected and transported through the solenoids and then focused onto a 924 $`\mu `$g/cm<sup>2</sup> thick, isotopically-enriched <sup>58</sup>Ni secondary target. The laboratory energy of the <sup>8</sup>B beam at the center of this target was 25.75 MeV, with an overall resolution of 0.75 MeV full width at half maximum (FWHM) and an intensity of 2.5 $`\times `$ 10<sup>4</sup> particles per second. The spread in energy was mainly due to a combination of the kinematic shift in the production reaction and the energy-loss straggling in the gas cell windows and the <sup>58</sup>Ni target. The beam had a maximum angular divergence of $`\pm `$ $`4^\mathrm{o}`$ and a spot size of approximately 4.0 mm FWHM. Although the count rate in the detectors was modest (typically less than 2$`\times `$ 10<sup>3</sup> s<sup>-1</sup>), the expected breakup yield is low so pulse-pileup tagging with a resolving time of 50 ns was used to eliminate pileup events.
The <sup>8</sup>B breakup events, and also elastically-scattered particles, were detected with two telescopes consisting of 25 and 30 $`\mu `$m Si $`\mathrm{\Delta }`$E detectors, backed by thick Si E detectors. These were placed on either side of the beam at $`\mathrm{\Theta }_{\mathrm{LAB}}=`$ $`20^\mathrm{o},30^\mathrm{o},40^\mathrm{o},45^\mathrm{o},50^\mathrm{o}`$ and $`60^\mathrm{o}`$. Each telescope had a circular collimator that subtended a solid angle of 41 msr, corresponding to a overall effective angular resolution of 10.9<sup>o</sup> (FWHM), computed by folding in the acceptance of the collimator with the spot size and angular divergence of the beam.
Unambiguous separation of the <sup>7</sup>Be fragments resulting from <sup>8</sup>B breakup from <sup>7</sup>Be contamination in the direct beam elastically scattered by the <sup>58</sup>Ni target was crucial to the success of this experiment. Although contaminants were present in the beam, they could be identified using time-of-flight (TOF) techniques. The TOF of the particles was obtained from the time difference between the occurrence of an E signal in a telescope and the RF timing pulse from the beam buncher. The time resolution of better than 3 ns (FWHM) was adequate to separate <sup>7</sup>Be from <sup>8</sup>B, as illustrated in Fig. 1. At sub-Coulomb energies, it is not easy to carry out a coincidence measurement between the <sup>7</sup>Be fragment and the proton, as was done at higher energies , due to the much reduced kinematic focusing of the protons. Such an experiment requires 4$`\pi `$ geometry and the ability to detect very low-energy protons to avoid biasing the correlation. Thus, we determined only the integrated <sup>7</sup>Be yield from the dissociation reaction <sup>8</sup>B $``$ <sup>7</sup>Be + p. Although the contaminants in the secondary <sup>8</sup>B beam are well separated in the $`\mathrm{\Delta }`$E vs. E<sub>TOTAL</sub> spectrum, as illustrated in Fig. 1(a), it would not have been possible to separate the <sup>7</sup>Be products coming from breakup events from the scattered contaminant <sup>7</sup>Be beam using only this information. However, by also considering the TOF information, particles of different origins could be completely separated as shown in Fig. 1(b), since the <sup>7</sup>Be from <sup>8</sup>B breakup has the same TOF as the <sup>8</sup>B beam.
The experimental angular distribution deduced for the dissociation of <sup>8</sup>B into <sup>7</sup>Be + p on a <sup>58</sup>Ni target is presented in Figs. 2 and 3, as a function of the center-of-mass angle of the detected <sup>7</sup>Be (we used the <sup>8</sup>B elastic-scattering Jacobian to transform the laboratory angles to the center-of-mass frame). The differential cross sections were obtained by integrating <sup>7</sup>Be breakup events over the solid angle subtended by the two telescopes. The number of <sup>8</sup>B ions per integrated charge of the primary beam was determined in a separate run. The normalization was obtained using the information on solid angle, target thickness and the measured integrated charge of the primary beam for each run, and verified by a measurement of <sup>8</sup>B elastic scattering, which is expected to be purely Rutherford at forward angles. The systematic error in the absolute normalization is estimated to be approximately 10$`\%`$, mainly due to the uncertainty in the intensity of the secondary beam. The <sup>8</sup>B beam had a $`1^\mathrm{o}`$ angular offset from the center axis set for the telescopes. This shift, evaluated by analyzing the observed asymmetry in the elastic scattering of <sup>8</sup>B, had a strong effect on the differential cross section at forward angles. Thus, at the most-forward angle setting of the telescopes we display the differential cross sections obtained at $`\mathrm{\Theta }_{\mathrm{LAB}}`$ = 19<sup>o</sup> and 21<sup>o</sup> separately. At backward angles, where the cross section does not change so rapidly as a function of angle, we have taken the average of the yield measured in the two telescopes.
It is obvious from inspection of the experimental angular distribution (see Fig. 3) that our data are completely inconsistent with the large amplitude peak in the vicinity of 70<sup>o</sup> \- 90<sup>o</sup> which was a prominent feature of both first-order theories . Very recently, however, two calculations that incorporate higher-order effects have been published, and they display a much different large-angle behavior. Esbensen and Bertsch performed a dynamical calculation that followed the time evolution of the valence proton wave function to all orders in the Coulomb and nuclear fields of the target. Their results are compared with our data in Fig. 2. The dotted curve corresponds to pure Coulomb breakup while the dashed curve, which includes nuclear effects, can be directly compared with the calculations presented in Refs. and . It can be seen that the higher-order couplings have completely eliminated the large-angle peak predicted by these first-order theories. Nevertheless, it is also clear that Coulomb-nuclear interference at very large distances, due to the extended nature of the “proton halo” in <sup>8</sup>B, still plays an important role in accounting for the experimental data.
Two other curves also appear in Fig. 2. The thin solid line illustrates pure Coulomb excitation under the usual “point-like” assumption. The dotted curve, which is much closer to the experimental data, is the correct pure-Coulomb-excitation calculation which takes account of the extended size of the valence proton orbital of the projectile. This result emphasizes the importance of incorporating the unusual structure of <sup>8</sup>B in all aspects of the reaction dynamics as first discussed in Ref. . (Note that the “point-like” approximation is still valid for neutron-halo nuclei since the relevant distance in this case is that between the core and the center-of-mass of the halo nucleus, which is still small). The thick solid curve includes, in addition to breakup, the effect of nucleon transfer from the projectile to the target. This is the calculation that is most appropriate for comparison with our data since we do not distinguish transfer from breakup. The large-angle peak is partially restored (but transfer was not included in the calculations presented in Refs. and so the computed transfer yield should be added to the angular distributions presented there). The present data suggest that proton transfer may have been somewhat overestimated in Ref. . Nevertheless, the overall agreement between theory and experiment is remarkable, especially considering that there has been no renormalization of the predicted absolute cross section.
Nunes and Thompson have also included higher-order effects, using the coupled discretized continuum channels (CDCC) method combined with the structure model of Esbensen and Bertsch . The advantage of this approach was that they were able to explicitly show that the vanishing of the large-angle peak results directly from the coupling among continuum states. Nunes has added proton transfer to this calculation and the result appears as the thick solid line in Fig. 3. She has also repeated the calculation using the structure model of Kim, et al. which, as mentioned above, has both a larger $`E1`$ and $`E2`$ component. The result is shows as the thin solid curve in Fig. 3. In general, the data favor the CDCC calculation using the wave function of Ref. , but the differences are small.
In conclusion, the angular distribution of the breakup of <sup>8</sup>B into <sup>7</sup>Be + p on a <sup>58</sup>Ni target was measured over a wide range of angles at a laboratory energy of 25.75 MeV. Time-of-flight information allowed us to unambiguously separate the <sup>7</sup>Be fragments coming from the breakup process, considerably improving on a previous measurement . The data are completely inconsistent with first-order reaction theories which predict a large amplitude nuclear dominated peak in the cross section at a center-of-mass angle of $`70^\mathrm{o}90^\mathrm{o}`$. However, recent calculations incorporating higher-order effects are in excellent agreement with experiment. In these calculations, the spurious peak is eliminated by continuum-continuum couplings. Coulomb-nuclear interference at very large distances, and the need to account for the extended size of the valence proton wave function in computing Coulomb breakup, are important features of both calculations. Thus, the present data may well be the best evidence yet of an exotic “proton halo” structure for <sup>8</sup>B. This has been a matter of some controversy, since reaction cross section measurements at relativistic energies by Tanihata, et al. displayed little or no enhancement, while similar measurements at intermediate energies by Warner, et al. and Negoita, et al. showed a rather substantial enhancement. (Enhancements in the reaction cross sections were the first signature of the neutron halo). The present data illustrate that finite-size effects and nuclear-Coulomb interference at very large distances, well outside the “normal” range of the nuclear force, are crucial features for the understanding of <sup>8</sup>B reactions at near-barrier and sub-barrier energies.
The original goal of the experiment described in Ref. was to obtain a model-independent measure of the $`E2`$ component in <sup>8</sup>B breakup and the astrophysical S-factor S<sub>17</sub> for proton capture on <sup>7</sup>Be at solar energies. In light of the discussion above, it appears that this will be very difficult. Even at the farthest forward angles measured in this experiment, corresponding to a distance of closest approach greater than 30 fm, substantial Coulomb-nuclear (and multiple-Coulomb-excitation) interference effects are important. While our data are consistent with the results of Davids, et al. , in the sense that the same structure model provides good predictions for both data sets, this conclusion is model dependent. On the other hand, the results from Ref. are also model dependent and the applicability of first-order perturbation theory and the “point-Coulomb” approximation, used there and in the analysis of breakup data at intermediate energies , should be re-investigated.
There does appear to be some sensitivity to the various structure models in our data. The wave function of Kim, et al. , which has the larger S<sub>17</sub> and $`E2`$ components, does not fit the data as well as other models, but the differences are too small to allow us to make any definitive statements about either quantity at this time.
Finally, the interactions of exotic, weakly-bound nuclei at near- and sub-barrier energies will increasingly be investigated as the next generation of radioactive ion beam facilities using the ISOL technique become available. It is comforting that there exist at least two successful theoretical approaches to the difficult problem of understanding low-energy reaction dynamics of weakly bound nuclei. We have shown that the information obtained from these reactions is complementary to that obtained from studies at intermediate and relativistic energies.
One of us (V.G.) was financially supported by FAPESP (Fundação de Amparo a Pesquisa do Estado de São Paulo - Brazil) while on leave from the UNIP (Universidade Paulista). This work was supported by the National Science Foundation under Grants No. PHY94-01761, PHY95-12199, PHY97-22604, PHY98-04869 and PHY99-01133. |
no-problem/9911/cond-mat9911444.html | ar5iv | text | # New Evidence of Earthquake Precursory Phenomena in the 17 Jan. 1995 Kobe Earthquake, Japan
## Abstract
Significant advances, both in the theoretical understanding of rupture processes in heterogeneous media and in the methodology for characterizing critical behavior, allows us to reanalyze the evidence for criticality and especially log-periodicity in the previously reported chemical anomalies that preceded the Kobe earthquake. The ion ($`Cl^{}`$, $`K^+`$, $`Mg^{++}`$, $`NO_3^{}`$ and $`SO_4^{}`$) concentrations of ground-water issued from deep wells located near the epicenter of the 1995 Kobe earthquake are taken as proxies for the cumulative damage preceding the earthquake. Using both a parametric and non-parametric analysis, the five data sets are compared extensively to synthetic time series. The null-hypothesis that the patterns documented on these times series result from noise decorating a simple power law is rejected with a very high confidence level.
Since the inception of seismology, the search for reliable precursory phenomena of earthquakes has shown that the seismic process is preceded by a complex set of physical precursors. In addition to the direct seismic foreshocks and seismic precursors , many other anomalous variations of various geophysical variables such as electric and magnetic fields and conductivity , as well as chemical concentration have been documented. However, there is no consensus on the statistical significance of these precursors and their reliability , due to a lack of reproducibility and of understanding of the underlying physical mechanisms.
One of the most striking reported seismic precursory phenomena are the time dependence of ion concentrations of ground-water issuing from deep wells located near the epicenter and the ground water radon anomaly preceding the earthquake of magnitude $`6.9`$ near Kobe, Japan, on January 17, 1995. Since the first quantitative analysis of this data, which suggested a discrete scale-invariant time-to-failure behavior , significant advances, both in the theoretical understanding of rupture processes in heterogeneous media and in the methodology needed to characterize critical behavior, permits a reassessment of the data.
Within the critical earthquake model , a large earthquake is viewed as the culmination of a cooperative organization of the stress and of the damage of the crust in a large area extending over distances several times the size of the seismic rupture . Most of the recently developed mechanical models and experiments on rupture in strongly heterogeneous media (which is the relevant regime for the application to the earth) view rupture as a singular ‘critical’ point : the cumulative damage $`D`$, which can be measured by acoustic emissions, by the total number of broken bonds or by the total surface of new rupture cracks, exhibits a diverging rate as the critical stress $`\sigma _c`$ is approached, such that $`D`$ can be written as an ‘integrated susceptibility’
$$DA+B(\sigma _c\sigma )^z,$$
(1)
The critical exponent $`0<z<1`$ is equal to $`1/2`$ in mean field theory and can vary depending on, e.g., the coupling corrosion and healing processes. In addition, it has been shown that log-periodic corrections decorate the leading power law behavior (1), as a result of intermittent amplification processes during the rupture. They have also been suggested for seismic precursors . This log-periodicity introduces a hierarchy of characteristic time and length scales with a prefered scaling ratio $`\lambda `$ . As a result, expression (1) is modified into
$$DA+B\left(\sigma _c\sigma \right)^z+C\left(\sigma _c\sigma \right)^z\mathrm{cos}\left(2\pi f\mathrm{ln}\left(\sigma _c\sigma \right)+\varphi \right),$$
(2)
where $`f=1/\mathrm{ln}(\lambda )`$. Empirical , numerical as well as theoretical analyses point to a prefered value $`\lambda 2.4\pm 0.4`$, corresponding to a frequency $`f1.2\pm 0.25`$ or radial frequency $`\omega =2\pi f7.5\pm 1.5`$. A value for $`\lambda `$ close to $`2`$ is suggested on general grounds from a mean field calculation for an Ising and, more generally, a Potts model on a hierarchical lattice in the limit of an infinite number of neighbors . It also derives from the mechanisms of a cascade of instabilities in competing sub-critical crack growth . Empirically, we see that there is not rarely a bias towards a value twice this, corresponding to a better of signal-to-noise ratio for rescalings of $`\lambda ^2`$ in the underlying renormalization group equation
The physical model underlying our analysis posits that the measured ion ($`Cl^{}`$, $`K^+`$, $`Mg^{++}`$, $`NO_3^{}`$ and $`SO_4^{}`$) concentrations of ground-water issued from deep wells located near the epicenter of the 1995 Kobe earthquake are proxies for the cumulative damage preceding the earthquake. In this reaction-limited model, any fresh rock-water interface created by the increasing damage leads to the dissolution of ions in the carrying fluid that can be detected in the wells. We thus expect that the time evolution of measured ion concentration should follow closely the equation (2). Due to the large heterogeneity of rocks, this ‘deterministic’ signal should be decorated by noise with different realizations for each ion originating from different rock types. To test our hypothesis, we analyze the five ion data sets, thus increasing the statistical significance over our previous report . However, it is possible that their noise realizations are not completely independent as some of the anions and cations are necessarily coupled pairwise. As we shall see, the poor result obtained from the analysis of $`K^+`$ might be due to the fact that it is generally coupled to both $`Cl^{}`$ and $`NO_3^{}`$.
Figure 1 shows the five data sets on which a moving average using three points has been applied. In this moving average, the middle point as usual carries double weight except for the two endpoints. Each of the five data sets is fitted with equation (2). The time intervals used in the fit were determined consistently for all five data sets by identifying the date of the lowest value of the concentration and using that date as the first data point. The last data point was taken as the last measurement before the earthquake. This gave $`53`$ points for the first three data sets, 52 for the fourth and 58 points for the fifth data set. For three of the data sets ($`Cl^{}`$, $`K^+`$, and $`NO_3^{}`$), the fit shown is also the best fit. In the case of $`SO_4^{}`$, the best fit had an angular log-frequency $`\omega 19`$ well above the expected range $`6\omega 9`$, while the second best fit has $`\omega 7`$ within the expected range and is shown instead. This is also the case for $`Mg^{++}`$, where the best fit has a very low angular log-frequency $`\omega 2`$ which capture nothing but a slowly varying trend. Here, the angular log-frequency of the second best fit $`\omega 16`$ is approximately double of what is found for the other four data sets corresponding to a squaring of $`\lambda `$ as previously discussed. We note that the ranking of the fits is done using the variance between the data and the fit with equation (2), where the stress $`\sigma `$ has been replaced with time $`t`$, $`t_c`$ being the critical date of the earthquake. This assumes a Gaussian white noise-distribution, which constitutes a reasonable null-hypothesis for the noise but presumably does not fully reflect the reality.
We observe that the values of the critical time $`t_c`$ predicted for the earthquake from the various fits are not only very stable but also remarkably consistent with the true date $`95.053`$ of the earthquake. And except for the case of $`Mg^{++}`$, where it has approximately doubled, the same is true for the angular log-frequency $`\omega `$ (although in 2 cases, we took the second best fit as explained above). This robustness with respect to the date of the earthquake and the preferred scale ratio $`\lambda =e^{2\pi /\omega }2.5`$ is quite remarkable, considering that the value obtained for the exponent $`z`$ varies by approximately a factor of $`4`$ from the smallest to the largest value.
We complement these fits by a direct ‘non-parametric’ analysis of the log-periodic component, taken as a crucial test of the critical earthquake model captured by (1), (2). After de-trending each of the five chemical time series $`c(t)`$ using
$$c\left(t\right)\frac{c\left(t\right)\left[A+B(t_ct)^z\right]}{C(t_ct)^z},$$
(3)
which should leave a pure log-periodic cosine if no other effects were present, we apply a Lomb periodogram to the de-trended data as a function of $`\mathrm{log}\left(\left(t_ct\right)/t_c\right)`$. Lower fight figure of 1 shows the five spectra superimposed. In all case except for $`K^+`$, we observe significant peaks with a log-frequency $`f\frac{\omega }{2\pi }`$ between $`1`$ and $`2`$, i.e. $`\omega `$ is approximately between $`6`$ and $`12.5`$, and in two cases ($`Cl^{}`$ and $`NO_3^{}`$), we have two very clear peaks at $`f1.2`$ ($`\omega 7.5`$) standing out with a Lomb weight of $`11`$ and $`13`$, respectively, in agreement with our prediction.
In order to assess the significance of these results, we present rather exhaustive statistical tests performed by constructing synthetic time series that differ from the real data only by the log-periodic pattern and we of course follow the same testing procedure.
There are several reasons why the results from the analysis above should be compared with those of synthetic tests. In particular, the real time series have been sampled non-uniformly in time because the water from the wells was collected from commercial bottles of which the production dates could be identified. We observe that the depletion process of bottles in stores implies that the number of bottles with a production date $`t`$ prior to the date of the earthquake $`t_c`$ is proportional to $`\mathrm{log}\left(t_ct\right)`$, leading to a uniform sampling in $`\mathrm{log}\left(t_ct\right)`$ instead of time $`t_ct`$ to the earthquake. In order to investigate the effect of this sampling of the Kobe data with respect to log-periodic signatures, we have as a first step generated twenty synthetic data sets each with $`56`$ points as in the original data analyzed in ref.. The synthetic data was generated using a noisy power law with the parameter values of the leading power law of the fit presented in ref.. Specifically, the equation $`c\left(t_i\right)=15.45+\left(1.877+k\left(0.5\text{ran}\right)\right)\left(95.053t_i\right)^{0.44}`$ was used with different noise-levels $`k`$. Here, ran is a uniform random number generator with values in the interval $`\left[0:1\right[`$ from . In order to obtain a sampling similar to the one found in the original data, the sampling times $`t_i`$ of the synthetic data was chosen as $`t_i=95.053(95.05393.847)e^{5\text{ran}}`$, where $`93.847`$ was the date of the first point in the original analysis. This then gives us a distribution of sampling times, which are uniform in log-time to $`t_c`$, where $`t_c=95.053`$ is the time at which the Kobe earthquake occurred. As noise-level, we used $`k=0`$, $`k=0.37`$ and $`k=0.73`$, corresponding to no-noise, a noise level of the same amplitude as the estimated log-periodic oscillations and a noise level twice that.
In the case of no noise $`k=0`$, the Lomb Periodograms exhibit in general a forest of peaks over a large log-frequency interval which average out to give a flat spectrum. Adding noise of amplitude $`k=0.37`$ and $`k=0.73`$ enhances the peaks but the averaging over the periodograms of the different synthetic data sets again produces an essentially flat spectrum. By “flat”, we mean that, in each individual synthetic time series, about $`7`$ “peaks” are found above $`70`$% of the largest one over the entire log-frequency interval $`0<f<6`$. This suggests a qualifier for the significance of a Lomb peak, corresponding to counting the number of crossings between the periodogram and a $`70`$%-level or a $`50`$%-level, taking the highest peak as the $`100`$% reference (a single peak corresponds to two crossings). For all cases, we obtain at least $`10`$ crossings depending on how one defines a peak. We also see that the noise has the effect that the largest peak is no longer to be found for the lowest frequencies.
We now prodeed with synthetic tests. As a first null-hypothesis, we take white noisy power law. An indication of the statistical significance of the results obtained from the chemical data can be estimated by comparing the periodogram of the de-trended data (3) with the periodograms of a sequence of random numbers uniformly distributed between $`0`$ and $`1`$ after a moving average over three points has been applied. Hence, we have generated $`1000`$ sequences with approximately the same number of points ($`53`$) and the time interval of the true data : none had a Lomb peak above 10. Out of $`10.000`$ synthetic data sets, 9 had Lomb peaks above 10, but none above 12. An additional $`100.000`$ data sets were generated providing three Lomb peaks above 12, but none of them had a log-frequency between $`1`$ and $`2`$. For peak values above 10, the $`100.000`$ data set had six such peaks in that log-frequency range. If we use a Gaussian white noise distribution instead of the uniform one, out of $`1000`$ synthetic data sets, we get $`20`$ peaks with a log-frequency between $`1`$ and $`2`$ and a Lomb weight above $`6`$, but none above $`10`$ with a log-frequency between $`1`$ and $`2`$. Out of $`10.000`$ data sets, only one such peak is observed.
Many other systematic and more elaborate tests have been performed with and without smoothing, as well as by directly generating noisy power laws and then performing the fit with (2) and the de-trending with (3). We find that smoothing after de-trending is not significantly different from smoothing before de-trending. Similarly, a generation of noisy power laws gives results similar to the foregoing white noise hypothesis.
Figure 2 presents a global summary of our tests by showing the bivariate distribution of angular log-frequencies $`\omega `$ and Lomb peak $`h`$ obtained from a series of 100,000 synthetic time series. The five vertical lines correspond to the five real times series and can be obtained from the lower right panel in figure 1 by locating the log-frequency and Lomb peak height of the highest peak for each data set. One can visualize that not only the fundamental frequency but its harmonics are sometimes visible ($`Cl^{}`$ and $`NO_3^{}`$) and, as we said above, may be dominating ($`Mg^{++}`$ and $`SO_4^{}`$).
The interpretation of these results is that there is a confidence level of $`99.99`$ % for a single peak with a log-frequency between $`1`$ and $`2`$ and a Lomb weight above $`10`$. This confidence interval is evaluated with respect to our initial null hypothesis of of uncorrelated (white) noise and would of course change with a different null hypothesis. Nevertheless, the confidence level achieved with the present null-hypothesis is above 99.9% and we hence expect it to remain would remain very high over a range of null-hypothesis. Also, the confidence level will also change depending on whether one assumes that the various ion concentrations are independent or not.
In conclusion, we wish to stress that the presented analysis constitute but a single case-study. Hence, it does not propose a recipe for earthquake prediction. However, we feel that the statistical evidence for this particular analysis is significant enough to encourage further studies along similar lines.
Acknowledgments: We are grateful to H. Wakita for kindly providing the data and for useful correspondence. H.S. thanks Y. Huang for collaboration at an early stage of this work and for many useful discussions. For further discussion of statistical tests and interpretations from a different point of view, see . |
no-problem/9911/gr-qc9911093.html | ar5iv | text | # Slow-roll inflation without fine-tuning preprint
## I Introduction
Inflation avoids the horizon, flatness and monopole problems of the standard cosmological model by just requiring a brief period of accelerated expansion during the very early universe . It also provides an explanation for the origin of the fluctuations associated with the observed cosmic microwave background anisotropies . The latter is a most valued feature at present, since after the detection made by COBE of these anisotropies the prospects for probing the physics of the inflationary models became real. Moreover the coming MAP and Planck missions will push our ability to contrast the theoretical models with observations to the precision required to rule out many of them.
The underlying ideas of inflation are well-kown, but a fully consistent scenario is still unavailable. In most cases the predictions of the models are based on a few features that seem consensual. In the simplest version of inflation one ordinarily considers a self-interacting scalar field $`\phi `$ (the inflaton) being damped by expansion of the universe and eventually yielding the negative pressures required to promote the violation of the strong energy principle, and hence the positive acceleration. The details of the process depend on the form of the potential, but the popular approach has been to associate the bulk of inflation with a stage during which the scalar field slow rolls along a potential well. The slow-roll parameters are defined in this context and a perturbative expansion in terms of these parameters is then used to predict the anisotropies of the cosmic microwave background (CMB), and conversely, to reconstruct from the latter data the shape of the inflaton potential.
A central issue regarding this usual procedure is then to ascertain the legitimacy of the slow-roll approximation. Given an arbitrary potential the slow-roll regime is not a generic dynamical pattern. However the vast majority of inflationary models rely on the slow-roll assumption and then show that it is possible to constrain the parameters of the potential in such a way as to meet the requirements of enough inflation and limited fluctuations in the CMB .
The issues of the genericity of the slow-roll regime and of the possibility of meeting the observational requirements without the slow-roll assumptions have not until now been dealt with satisfactorily. In what concerns the former of these questions, there has been some effort to either avoid it or argue in favour of an affirmative answer in the case of particular models . In particular the results of Ref. have shown that for the pseudo-Nambu-Goldstone-boson of the natural inflation model fine tuning can be avoided if we consider non-zero initial velocities for the field. We recover below this conclusion in an independent context. As for the latter question there has not been to our knowledge any attempt to address it directly. Leaving this question unanswered encourages criticism of the relevance of the theoretical estimates based on the slow-roll approximation. Recently, the reliability of the reconstruction procedure based on the lowest-order slow-roll expansion has been questioned by Wang, Mukhanov and Steinhardt who show that its validity depends on the effective equation of state of the inflaton field being almost stationary during the relevant epoch of horizon exit of the fluctuations. In this reference the authors explicitly consider a potential for which this condition is not met and show that the implications for the CMB anisotropies are significantly different from those predicted by the slow-roll approximation. In Ref. it is shown that the particular potential considered in is such that several requirements, among which the end of inflation, cannot be met.
Given that slow-roll is an atypical dynamical regime for damped oscillations, in the absence of a general negative answer to the second of the questions, the validity of the slow-roll approximation will remain a natural target to criticism such as in .
In the present work our aim has been to determine the consequences of imposing the constraints on the number of e-foldings and on the amplitude of the CMB fluctuations to a general scalar field model giving rise to a limited inflationary period. Our study relies on a numerical analysis of the dynamics, which is both rigorous and general because it stems from the knowledge of the global dynamics of the non-linear dynamical system under consideration. Contrary to what may be thought the numerical approach is more adequate than an alternative analytic approach which is bound here to be either local or too restrictive.
The results of this analysis show that the constraints on the number of e-foldings force the free parameters that control the basic features of the potential to take the values for which the orbits exhibit slow roll dynamics for a relatively long time of their evolution. This is a consequence of the known fact that rapid oscillations in convex potentials do not contribute to inflation . We also show that these values of the parameters correspond to a big region in parameter space. These two conclusions are of great importance on the one hand to avoid the criticism of and, on the other hand, to establish that inflation is likely to happen.
We consider the flat isotropic Friedmann-Robertson-Walker (FRW) universe characterized by the metric
$$ds^2=dt^2+a(t)^2\left[dr^2+r^2(d\theta ^2+\mathrm{sin}\theta ^2d\varphi ^2)\right],$$
(1)
and assume that the matter content of the universe is a homogeneous scalar field with a self-interacting potential. The field equations are
$$H^2=\frac{8\pi }{3m_{pl}^2}\left(\frac{1}{2}\dot{\phi }^2+V(\phi )\right),$$
(2)
$$\ddot{\phi }+3H\dot{\phi }+V^{}(\phi )=0,$$
(3)
where the overdots stand for the derivatives with respect to time, $`m_{pl}^2`$ is the Planck mass, $`H=\dot{a}/a`$ is the Hubble parameter, and $`V^{}=\mathrm{d}V/\mathrm{d}\phi `$. It is also possible to deduce from the latter equations that
$$\frac{\ddot{a}}{a}=\frac{8\pi }{3m_{pl}^2}\left(V(\phi )\dot{\phi }^2\right)$$
(4)
holds and hence that inflation requires $`V(\phi )>\dot{\phi }^2`$.
In our numerical study, we shall consider potentials of the form
$`V_1(A,B,\phi )`$ $`=`$ $`{\displaystyle \frac{A}{B^2}}(\phi ^2B)^2,`$ (5)
$`V_2(A,B,\phi )`$ $`=`$ $`A\left(1+\mathrm{cos}({\displaystyle \frac{\pi \phi }{\sqrt{B}}})\right),`$ (6)
where $`A`$ and $`B`$ are positive constants. However, our results extend to potentials exhibiting several extrema. For these potentials, given any initial conditions outside a potential trap, the damping term in Eq. (3) will make the system evolve to approach the potential well where it will be caught by the stable equilibrium. Provided that this is vanishing (no false vacuum term), the last stage of the evolution of the system corresponds to a potential with one maximum and one vanishing minimum, such as in Eqs. (5,6), together with initial conditions slightly above the energy of the maximum. Futhermore, the behaviour of the solutions of the systems associated with the potentials $`V_1`$ of Eq. (5) or $`V_2`$ of Eq. (6) is typical in this more general class: the qualitative behaviour dos not depend on the detailed form of the potential and is described in some detail below. In this general setting our results apply to any non-degenerate potential with the aforementioned qualitative properties, since the only essential requirement is the non-degeneracy of the consecutive extrema. ¿From a physical viewpoint, our conclusions hold for a non-degenerate potential susceptible of giving rise to a finite inflationary period whose end is followed by the death of the oscillating field, and also to sufficient inflation without fine-tuning the parameters. In fact, the first condition excludes potentials without extrema, such as the exponential potential or the arctan potential of Ref. , as well as potentials with a false vacuum term , while the second discards single minimum potentials as the quadratic potential .
Consider the Eqs. (2,3) with the potentials (5) or (6). In both cases and indeed for any potential with a maximum and a vanishing minimum, the problem reduces to the study of the dynamics of a non-linear, non-linearly damped, oscillator whose damping term tends to zero as the stable equilibrium is approached. In spite of the high non-linearity, the equations are those of an autonomous system in the plane, and therefore the possibility of chaotic behavior is excluded. Indeed, the qualitative description of the behavior of the system in phase-space is quite straightforward. Consider for instance the case of the standard double well potential: there are three equilibrium points, one saddle at $`(\phi ,\dot{\phi })=(0,0)`$ and two stable focus at $`(\phi ,\dot{\phi })=(\pm \sqrt{B},0)`$. The stable manifolds of the saddle point separate the plane into two regions, I and II. One of the stable focuses attracts all the points in I, and the other attracts all of the points in II. Moreover, there is no divergence of nearby orbits, and, apart from the boundary of the two basins of attraction, the future behaviour of the orbits will change smoothly with the initial conditions. Bearing this picture in mind, it is possible to perform a thorough numerical exploration of the system, in order to determine the set of parameters and initial conditions which satisfy the observational constraints.
It is usual to consider the parameters
$$ϵ(t)=\frac{3}{2}\frac{\dot{\phi }^2}{\frac{\dot{\phi }^2}{2}+V(A,B,\phi )}$$
(7)
$$\eta (t)=\frac{\ddot{\phi }}{H\dot{\phi }}.$$
(8)
Inflation is equivalent to having $`ϵ<1`$. The slow roll regime is defined by the following conditions: $`ϵ(t)1`$,which is the condition for neglecting the kinetic energy contribution to the damping term in Eq. (3), and $`\eta (t)1`$, which is the condition for neglecting the acceleration term in Eq. (3).
To succesfully solve the cosmological puzzles of standard cosmology an inflationary model must satisfy the following constraints .
(i) Sufficient inflation \- We demand that the scale factor of the universe inflates by at least 60 e-foldings.
(ii) Energy bounds \- To keep full generality we allow inflation to take place bewteen the Planck energy scale of $`m_{pl}`$ and the electroweak phase transition at $`10^{17}m_{pl}`$. This provides energy bounds in phase-space for the dynamics of the scalar field $`10^{34}m_{pl}^2H^2m_{pl}^2`$.
(iii) Density perturbations \- The observations of the anisotropies of the CMB constraint the amplitude of density fluctuations, $`\delta _H`$, to satisfy the bound $`\delta _H\delta _H^{obs}1.9\times 10^5`$ .
We shall now impose these constraints on the dynamics of Eqs. (2,3) under the potentials $`V_i(A,B,\phi )`$, $`i=1,2`$. First note that both potentials are proportional to $`A`$. Changing the time scale through $`t=\tau /\sqrt{A}`$ and redefining the field’s velocity and acceleration accordingly, the equations of motion (2,3) become independent of the parameter $`A`$. The total number of e-foldings is given by
$$N(A,B)=_0^+\mathrm{}\sqrt{\frac{8\pi }{3m_{pl}^2}\left(\frac{\dot{\phi }^2(t)}{2}+V(A,B,\phi (t))\right)}𝑑t,$$
(9)
where $`\phi (t)`$ is a solution of the Eqs. (2,3). Performing the rescaling of the time variable in the latter integral, we have $`N(A,B)=N(1,B)`$, and, thus, the number of e-foldings is independent of $`A`$. Notice that this scaling property is also independent of the detailed form of $`V`$.
When $`ϵ1`$ and $`\dot{ϵ}0`$ hold, it is known that $`\delta _H=𝒪(1)H^2/\dot{\phi }`$, where the right-hand side is evaluated at horizon crossing . Thus, assuming that $`ϵ(t)`$ satisfies the latter conditions, the amplitude of the density fluctuations scales as $`\delta _H(A)=\sqrt{A}\delta _H(1)`$ and so restriction (iii) can be dealt with separately. Setting $`\phi (0)=0`$, we are thus left with the two parameters $`(B,\phi ^{}(0))`$, where the prime denotes the derivative with respect to the rescaled time, and we have to satisfy the former two constraints, (i) and (ii).
The initial energy bounds define an interval of admissible values for $`\phi ^{}(0)`$. The usual approach is to study the ensemble of initial conditions with energy equal or lower than the symmetry breaking energy $`A`$. Here we consider initial energies that are equal or larger than $`A`$. This is important to avoid fine tuning, since these orbits will have an additional contribution to the number of e-foldings with respect to the lower energy orbits usually considered in the literature.
The question then reduces to find for each $`\phi ^{}(0)`$ in the admissible energy range the interval of values of $`B`$ such that $`N(1,B)[60,+\mathrm{}]`$. The results are shown in Figure 1, and it can be seen that the constraint on the number of e-foldings can be met in a big region of parameter space. More interestingly, it turns out that for all the parameter values that satisfy the constraint on the number of e-foldings, the slow-roll parameters $`ϵ`$ and $`\eta `$, as well as the time derivatives of $`ϵ`$, behave in such a way that the slow-roll assumptions are fulfilled during the whole inflationary period. Thus the slow-roll approximation is valid. In Figure 2 we represent a typical orbit in the admissible region in parameter space, as well as the behaviour of $`ϵ`$ and $`\eta `$. In Figure 3 we show the typical values of $`ϵ`$ and $`\eta `$ for those orbits during the inflationary period. This means that the slow-roll assumption to deal with the constraint on $`\delta _H`$ is fully justified. Moreover, the fact that slow-roll holds during the inflationary period agrees with the observational constraints on the spectral index and on the gravitational waves contribution to the CMB .
More significantly, our analysis provides a negative answer to the question of whether it is possible to meet the basic inflationary requirements without slow-roll in the class of potentials with several extrema yielding a finite inflationary period. In fact, our numerical results amount to a global study of the $`(B,\dot{\phi }(0))`$ parameter plane and show that parameters that give rise to orbits off the slow-roll regime also fail to satisfy the constraint on the number of e-foldings. This finding agrees with the results of regarding the potential considered in , tailor-made to break up the slow-roll approximation, and which they show gives rise to a non-flat perturbation spectrum. The idea that the slow-roll regime is forced by the observational constraints seems to prevail in the general setting of arbitrary potentials exhibiting symmetry breaking, or just several extrema, as well as for other specific models.
It is also clear that the three main constraints, on the number of e-foldings, on the amplitude of the perturbation spectrum, and on the initial energy of the field, can be met in a potential belonging to the broad class considered in this paper without fine tuning of the parameters. The final conclusion is then that in this setting the slow-roll approximation is a general starting point to study further constraints imposed by new observational data.
## Acknowledgements
The authors wish to acknowledge the finantial support from Fundação de Ciência e Tecnologia under the grant PBIC/C/FIS/2215/95 and A.N. thanks the project PRAXIS 2/2.1/MAT/125/94. We are also grateful to Andrew R. Liddle for helpful comments on an preliminary draft of this work. |
no-problem/9911/astro-ph9911084.html | ar5iv | text | # Space-VLBI observations of the twisted jet in 3C 395
## 1 Introduction
Flux density variability at radio wavelengths in the compact cores of radio loud AGNs on time scales of months or even years is generally associated with ejection of components in parsec-scale relativistic jets (e.g. Valtaoja et al. valtaoja (1988)). Such behavior has been observed in many radio sources (e.g. Pauliny-Toth et al. pauliny (1987)) and is successfully reproduced by theoretical work (e.g. Hughes et al. hughes (1991), Gómez et al. gomez (1997)). However, there are objects which, apparently, do not fit into this scenario. One of these, the quasar 3C 395 ($`z=0.635`$), exhibits significant flux density variability which is clearly associated with activity in its compact core (Lara et al. lara1 (1994, 1997)); however, Very Long Baseline Interferometry (VLBI) observations since 1984 show a stationary morphology, with no evidence of the expected correlation between flux density variability and the ejection of new jet components (Lara et al. lara2 (1997)).
In this paper we present new VLBI observations of the quasar 3C 395 at 4.8 GHz, made with the Very Long Baseline Array (VLBA) and the Japanese satellite Halca. These observations not only take profit of the enhanced angular resolution of Space-VLBI, but also of the high sensitivity achieved by the continuous observation of a single source at a single frequency with the VLBA. The new data shed light on the link between flux density variability and the structural properties of 3C 395.
## 2 Observations and data reduction
We made continuum observations of 3C 395 with the VLBA and Halca on May 1st 1998 at a frequency of 4.8 GHz. The observing bandwidth was 32 MHz. Two tracking stations, Robledo (Spain) and Green Bank (USA), participated in the observations providing maser referenced timing tones to the satellite. At the same time, they received the astronomical data from Halca through a Ku-band downlink, recording them on magnetic tapes for later processing in a VLBI correlator. In Fig. 1 we display the uv-coverage achieved in our observations to illustrate the improvement in angular resolution provided by the orbiting antenna. The correlation of the data was done in absentia by the staff of the VLBA correlator in Socorro (NM, USA). After correlation, we used the NRAO AIPS package to determine the bandpass response functions of the antennas, to correct for instrumental phase and delay offsets between the separate baseband converters in each antenna, to determine antenna-based fringe corrections and to apply the a priori amplitude calibration. Data imaging in total intensity was finally performed with the Difmap package (Shepherd et al. shepherd (1994)).
The imaging process consisted of two main steps: initially, we started mapping the VLBA data alone, i.e. without ground-space baselines, at low angular resolution by reducing the weight of the longer VLBA baselines. Once a satisfactory low resolution map of 3C 395 was obtained, the weight of the long baselines was progressively restored to its original values, so that we finally obtained a map with the VLBA alone at its maximum angular resolution. In this process, we also derived accurate self-calibration solutions for the VLBA antennas. We then included data from Halca in a second mapping step, improving the source model at sub-milliarcsecond resolution and obtaining a high resolution map from the whole data set. Finally, we calibrated the absolute flux density scale mapping the compact calibrator source 0133+476, also observed during the experiment and with an assumed flux density of 2.50 Jy at 4.8 GHz (information obtained from the University of Michigan Radio Astronomy Observatory (UMRAO) database).
## 3 Results
Radio maps of 3C 395 are displayed in Fig. 2, with angular resolutions spanning a range from 10 to 0.3 mas.
Fig. 2a shows a low angular resolution map of the jet in 3C 395. The brightness distribution is dominated by a strong double component, but the most remarkable feature at this resolution is the existence of a very sharp bend in the jet at a distance of $``$70 mas from the core. This bend links the compact structure with the extended emission observed at sub-arcsecond resolution (Saikia et al. saikia (1990); Lara et al. lara2 (1997)). We can follow the jet up to a length of nearly 200 mas thanks to the very high sensitivity of the VLBA. The jet at these large scales is not smooth, but knotty. Two of these knots can be associated to components D and E in Lara et al. (lara2 (1997)). The total flux density recovered in our VLBI observations is 1.43 Jy.
Fig. 2b has an angular resolution similar to that of previously published cm-VLBI maps of 3C 395 (Lara et al. lara1 (1994)). Leaving apart the weak component D, the radio structure can be described in terms of three main components named A, B and C, as in previous papers. Component A has been usually identified as the radio core; component B appeared stationary with respect to A and was interpreted as the result of a local bend in the jet towards the observer ; component C, between A and B, was claimed to be moving superluminally after VLBI observations during the 80’s (Waak et al. waak (1985); Simon et al. 1988a ), but no clear evidence of motion was found in later observations (Simon et al. 1988b ; Lara et al. lara2 (1997)).
Figs. 2c-d reveal that a simple description of the compact structure of 3C 395 in terms of only three components is no longer valid when observing with high sensitivity and resolution. Feature A is not a single component: it hosts the core, most plausibly at its western edge, but also a second strong component and a jet-like feature. Such complexity within A was suggested by Lara et al. (lara2 (1997)) after a three-baseline VLBI test observation at 22 GHz. Component B also shows a rather complex structure at high angular resolution. Component C appears resolved (Fig. 2c), and it is difficult to associate it with any single feature, either moving or stationary.
To obtain a quantitative description of the milliarcsecond structure of 3C 395, we have fitted simple elliptical Gaussian components to the visibility data using a least square algorithm within Difmap. We needed a total of 8 components to satisfactorily reproduce the data; we did not attempt to fit the brightness distribution beyond component D along the jet, since this extended emission only marginally affects the data from the shorter baselines. The estimated parameters for each Gaussian component are given in Table 1, where we display the flux density (S), the angular distance from the westernmost component A1 (D), the position angle with respect to A1 (P.A.), the length of the major axis (L), the ratio between the major and minor axis (r) and the orientation of the major axis ($`\mathrm{\Phi }`$), defined in the same sense as the position angle. The elliptical Gaussian components have been plotted on Fig. 4 for direct comparison with the surface brightness distribution.
## 4 Discussion and conclusions
The launch of satellite Halca in February 1997 has probably marked an inflection point in the development of radio astronomy. VLBI, continuously improving its sensitivity since its origins in the sixties, has finally managed to break the hard constraints in angular resolution imposed by the limited size of the Earth. As a sample of the new capabilities of VLBI, our observations show that the milliarcsecond structure of 3C 395 is rather more complex than previously thought, with several features which deserve special attention.
First, there is a sudden decrease in the surface brightness of the jet after component A3, defining a sharp boundary between this component and the rest of the jet. Second, the jet position angle between A2 and A3 is 137, remarkably different from the position angle defined by components C and B (P.A. 118). These two facts argue towards the existence of a bend in the jet soon after component A3 in which the orientation angle of the jet increases significantly with respect to the observer. Third, as mentioned above, the emission between components A and B cannot be easily associated with a single component. Nevertheless, in our Gaussian fit we have used only one component to describe this emission in order to compare our results with previous lower resolution observations and look for any evidence of motion, but the Gaussian component we obtain is very elongated, making any motion estimates very uncertain. Moreover, attempts to model the complex structure of C were not conclusive, suggesting that the emission observed between components A and B might be related more to the underlying jet hydrodynamics, rather than to a traveling shocked component.
Another interesting aspect of 3C 395 is related to its variable flux density. In Fig. 3 we display the time evolution in the flux density of 3C 395 at 4.8 and 14.5 GHz since 1982. There are variations of up to 30% in total flux density at both frequencies. Monitoring the quasar with VLBI at several frequencies and angular resolutions since 1990 shows that the flux density variability is related with the activity within component A (Lara et al. lara2 (1997)).
As noted before, flux density variability in the compact cores of AGNs is usually associated to the ejection of new traveling components along parsec-scale relativistic jets. According to this idea, Fig. 3 suggests that new components should have been ejected from the core of 3C 395 in 1986, 1990 and 1993. However, previous VLBI observations of 3C 395 do not show the expected correlation between flux density variations and the ejection of new VLBI components. The Space-VLBI observations presented here help us to understand this peculiarity: if there exists a large bend in the jet after component A3, it will be very difficult to correlate flux density variability within A with structural variations beyond this component because relativistic time-delay makes the time scales of these two events very different. Moreover, the decrease in the Doppler factor after component A3 produces a large diminution in the flux density of a possible moving component. 3C 395 requires sub-milliarcsecond resolution to study structural variations within A associated to flux density variability. Hence, previous VLBI observations did not have the necessary angular resolution.
The ridge line of the jet of 3C 395 at milliarcsecond scales can be traced from our maps (see Fig. 4) showing the wiggles and curvatures present in this twisted jet. We can conclude that the observed properties of quasar 3C 395 are heavily marked by three main bends which produce changes in the orientation of the jet with respect to the observer, and hence changes in the observed properties due to Doppler factor variations. The first apparent bend, close to the core and producing a departure of the jet direction from the line of sight, might be the reason why flux density flares are not directly followed by ejection of components observable with ground cm-VLBI. The second bend, at 15 mas from the core, orients the jet back towards the observer’s line of sight and may be responsible of the Doppler amplification of the emission at this position resulting in the stationary component B. The third bend, at a distance of 70 mas from the core, produces a sharp deflection in the projected ridge line of the jet, appearing to almost turn back upon itself. While the first two bends could possibly be consequence of a helically twisted geometry in the jet (which might explain also components observed beyond B), the third large curvature would require a global bending of the helix. We note that the intrinsic effects of these bends are most probably highly amplified by a sharp overall orientation of the jet in 3C 395 with respect to the observer’s line of sight.
###### Acknowledgements.
This research is supported in part by the Spanish DGICYT (PB97-1164). It has made use of data from the University of Michigan Radio Astronomy Observatory which is supported by funds from the University of Michigan. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. |
no-problem/9911/astro-ph9911229.html | ar5iv | text | # The Mass Function of Young Star Clusters in the “Antennae” Galaxies
## 1 Introduction
The luminosity functions of the young star clusters in merging galaxies, including the “Antennae” (NGC 4038/39), have roughly power-law form, $`\varphi (L)L^\alpha `$ with $`\alpha 2`$, and no sign of a peak or turnover down to the limiting magnitudes of the observations (Whitmore & Schweizer 1995; Schweizer et al. 1996; Miller et al. 1997; Carlson et al. 1998; Whitmore et al. 1999; Zepf et al. 1999). These are similar to the luminosity functions of open clusters in the Milky Way (van den Bergh & LaFontaine 1984) and M33 (Christian & Schommer 1988) and the populous clusters in the LMC (Elson & Fall 1985) and starburst galaxies (Meurer et al. 1995). In contrast, the luminosity functions of the old globular clusters in the Milky Way, the Andromeda Galaxy, and elliptical galaxies have peaks at $`L_V10^5`$ $`L_{}`$ and decline toward both higher and lower luminosities. These functions are often represented by lognormal distributions of luminosities, corresponding to Gaussian distributions of magnitudes (e.g., Harris 1991).
This immediately raises the question of whether the mass functions of the young star clusters in merging galaxies have power-law or lognormal forms. The mass functions are of greater relevance to our understanding of the physical processes involved in the formation and disruption of the clusters. If the star clusters of any population formed at the same time and with the same stellar initial mass function (IMF), then the luminosity function of the clusters would always have the same shape as the mass function, the two being related by a time-dependent shift along the luminosity axis. However, if the spread in ages within a population is comparable to the mean age, as seems likely for the young star clusters in some recent mergers, one might then worry about the effects of fading, which could cause the shape of the luminosity function to differ substantially from the shape of the mass function (Hogg & Phinney 1997). Based on simulations of this effect, Meurer (1995) concluded that the power-law luminosity function in the Antennae derived by Whitmore & Schweizer (1995) from observations with the WFPC1 before the refurbishment of HST may be consistent with an underlying lognormal mass function. Furthermore, Fritze-von Alvensleben (1999) estimated the mass function of young star clusters in the Antennae directly from these WFPC1 observations and found a lognormal form, similar to that of old globular clusters.
The purpose of this Letter is to determine the mass function of the young star clusters in the Antennae from images taken with the WFPC2 after the refurbishment of HST; these extend about two magnitudes fainter than the earlier images taken with the WFPC1. The data reduction and luminosity functions are described in detail by Whitmore et al. (1999). The novelty of our approach to determine the mass function is that we use reddening-free parameters and a comparison with stellar population synthesis tracks to estimate the intrinsic luminosity and age and hence the mass of each cluster.
## 2 Observations and Selection of Clusters
The Antennae galaxies were observed with the WFPC2 in 1996 January for 2000–4500 s with each of the broad-band filters F336W, F439W, F555W, and F814W. Point-like objects were identified and their magnitudes measured with the IRAF task DAOPHOT; photometry in this system was then transformed to the Johnson $`UBVI`$ system. About 11,000 objects were detected with $`17.4<V<25.4`$, corresponding to $`14<M_V<6`$ at the adopted distance of 19.2 Mpc (for $`H_0`$ = 75 km s<sup>-1</sup> Mpc<sup>-1</sup>). Much of our analysis is based on the reddening-free parameters
$`Q_1=(UB)0.72(BV),`$ (1)
$`Q_2=(BV)0.80(VI).`$ (2)
The first of these is the standard reddening-free parameter in the $`UBV`$ system, while the second is a direct extension to the $`UBVI`$ system. Other reddening-free parameters can be defined analogously, but they are all linear combinations of $`Q_1`$ and $`Q_2`$. Equations (1) and (2) are based on the Galactic extinction curve but would be virtually the same for other familiar extinction curves, such as those of the Magellanic Clouds. For objects with $`M_V=9`$, the photometric errors in $`U,B,V,`$ and $`I`$ (0.04–0.07) lead to $`1\sigma `$ uncertainties of 0.12 in $`Q_1`$ and 0.11 in $`Q_2`$.
We identify cluster candidates and estimate their extinctions and ages by comparing them with population synthesis tracks in the $`Q_1Q_2`$ diagram, shown here as Figure 1. In the presence of reddening by dust, this approach has significant advantages over methods based on the more familiar two-color diagrams. For the population synthesis track, we use the Bruzual & Charlot (1996, hereafter BC96) model with the Salpeter IMF truncated at 0.1 and $`125M_{}`$ and with solar metallicity. We regard all point-like objects brighter than $`M_V=9`$ as cluster candidates, irrespective of their $`Q`$ parameters, because nearly all known stars are fainter than this limit (Humphreys 1983). Furthermore, we regard fainter point-like objects as cluster candidates if they lie within $`\mathrm{\Delta }Q=(\mathrm{\Delta }Q_1^2+\mathrm{\Delta }Q_2^2)^{1/2}=0.3`$ of the population synthesis track (except for a few objects with extremely large photometric errors). This procedure eliminates some but not all stars from our sample (see Figs. 13 and 14 of Whitmore et al. 1999). However, as shown below, stellar contamination has little effect on the mass function we derive because this is based mainly on the brightest objects.
For each cluster candidate, we estimate the intrinsic colors, age $`t`$, and mass-to-light ratio $`M/L`$ from the nearest point on the population synthesis track in the $`Q_1Q_2`$ diagram. A comparison of the intrinsic with the observed colors gives the extinction $`A_V`$, and the intrinsic luminosity $`L_V`$ and mass $`M`$ then follow from $`V`$ and $`A_V`$. The mean value of $`A_V`$ varies from 1.5 for the youngest clusters ($`t<10`$ Myr) to 0.3 for the oldest clusters ($`t>100`$ Myr). Because the BC96 population synthesis track stops evolving in the $`Q_1Q_2`$ diagram at $`\mathrm{log}(t/\mathrm{yr})=6.4`$, no objects can be assigned ages smaller than this by our procedure. Instead, the youngest clusters are all assigned ages in the interval $`6.4<\mathrm{log}(t/\mathrm{yr})<6.8`$. Furthermore, because the track bends sharply at $`\mathrm{log}(t/\mathrm{yr})6.9`$ and has a zigzag at $`\mathrm{log}(t/\mathrm{yr})7.2`$, there is some ambiguity in the assignment of ages in this region of the $`Q_1Q_2`$ diagram. For these reasons, we exclude the intervals $`\mathrm{log}(t/\mathrm{yr})<6.4`$ and $`6.8<\mathrm{log}(t/\mathrm{yr})<7.4`$ from our determination of the mass function.
We have verified by simulations that the scatter of cluster candidates away from the population synthesis track in the $`Q_1Q_2`$ diagram is mostly accounted for by photometric errors. However, for the youngest objects ($`t<10`$ Myr), the scatter is slightly larger, probably because they are affected by nebular emission and residual extinction. (While our approach is designed to correct for extinction, this is never perfect, especially in the dusty regions where the youngest clusters are often found.) The errors in the $`Q`$ parameters lead to typical uncertainties of a factor of 2 in the ages and factors that vary from 1.1 to 1.7 in the luminosities. Fortunately, these uncertainties tend to cancel out in estimates of the masses, because younger clusters have lower mass-to-light ratios but higher extinctions. As a result, the uncertainties in the masses of most cluster candidates are smaller than 50%.
Figure 2 shows the luminosity-age relation for the objects in our sample along with the BC96 population synthesis tracks for $`\mathrm{log}(M/M_{})=`$ 6.0, 5.5, 5.0, 4.5, and 4.0. Evidently, the masses of the cluster candidates range from well below $`10^4M_{}`$ to just above $`10^6M_{}`$. The apparent gaps in the luminosity-age relation at $`\mathrm{log}(t/\mathrm{yr})<6.4`$ and $`7.0<\mathrm{log}(t/\mathrm{yr})<7.2`$ are artifacts caused by little or no evolution of the population synthesis track in the $`Q_1Q_2`$ diagram in these intervals of age (discussed above). Stellar contamination can be neglected above the horizontal dotted line in Figure 2 at $`M_V^c=9`$, an extinction-corrected magnitude that excludes all but the very brightest stars. For reference, the Large Magellanic Cloud, with about 4% of the total blue luminosity of the Antennae, has two stars with $`M_V^c=9`$, three with $`10<M_V^c<9`$, and none with $`M_V^c<10`$ (Humphreys 1983). Thus, we might expect $`100`$ stars (mostly blue supergiants) brighter than $`M_V^c=9`$ in our sample of cluster candidates, corresponding to contamination at the $`5\%`$ level.
## 3 Mass Function of Clusters
We construct the mass function of cluster candidates in two intervals of age: $`6.4<\mathrm{log}(t/\mathrm{yr})<6.8`$ and $`7.4<\mathrm{log}(t/\mathrm{yr})<8.2`$, corresponding to $`2.5<t<6.3`$ Myr and $`25<t<160`$ Myr. These intervals, which are indicated by the bars at the top of Figure 2, were chosen to avoid the problems discussed in the previous section. The first allows us to estimate the mass function down to relatively low masses with a relatively large number of objects, while the second provides a check on the consistency of our procedure. In constructing the mass function, we must correct for incompleteness in our sample. This depends on the brightness of the objects, whether they are on the PC or WF chips, and the local background and/or crowding. For each cluster candidate, we adopt the completeness factor determined by Whitmore et al. (1999) using a false-star method. Our sample as a whole is about 50% complete at $`V=24`$, corresponding to $`\mathrm{log}(M/M_{})3.9`$ for $`6.4<\mathrm{log}(t/\mathrm{yr})<6.8`$ and $`\mathrm{log}(M/M_{})4.4`$ for $`7.4<\mathrm{log}(t/\mathrm{yr})<8.2`$. Above these limits, the younger and older subsamples contain 1140 and 477 objects, respectively.
Figure 3 shows the completeness-corrected mass functions of the cluster candidates in the two intervals of age. The stellar contamination limits are marked by S’s, while the completeness limits are marked by C’s. Stellar contamination is negligible for the younger subsample, but it could begin to affect the older subsample for $`\mathrm{log}(M/M_{})4.7`$. Evidently, the mass functions for the two subsamples are nearly indistinguishable above the completeness limits. They can be represented by a power law, $`\psi (M)M^\beta `$, with $`\beta =1.95\pm 0.03`$ for $`6.4<\mathrm{log}(t/\mathrm{yr})<6.8`$ and $`\beta =2.00\pm 0.08`$ for $`7.4<\mathrm{log}(t/\mathrm{yr})<8.2`$. These are based on weighted least-square fits of the form $`\mathrm{log}\psi =\beta \mathrm{log}M+\mathrm{const}`$.
We have checked that our results are robust with respect to the adopted population synthesis tracks. First, we repeated the entire analysis with the BC96 models but with a different IMF (Scalo vs. Salpeter) and different metallicities ($`0.4Z_{}`$ and $`2.5Z_{}`$ vs. $`1.0Z_{}`$). Second, we repeated the analysis with the Leitherer et al. (1999) models with the Salpeter IMF and solar metallicity. In the first case, the mass function was virtually the same; in the second case, it was slightly steeper, with $`\beta =2.1`$. We have also checked that our results are not biased by observational errors by randomly re-assigning ages to all cluster candidates younger than $`\mathrm{log}(t/\mathrm{yr})=6.8`$. In this case, we obtain $`\beta =1.9`$.
Figure 3 shows a comparison between the mass function of the young star clusters in the Antennae determined here and the mass function of old globular clusters (indicated by the dash curves). The latter was derived from the usual Gaussian distribution of magnitudes and a fixed mass-to-light ratio, $`M/L_V=2`$. Within the observational uncertainties, it appears that the two mass functions may be similar for $`M2\times 10^5M_{}`$. However, for $`M2\times 10^5M_{}`$, they are completely different; that for the young clusters in the Antennae increases rapidly, while that for the old globular clusters decreases rapidly. Whitmore et al. (1999) found that the luminosity function of the star clusters in the Antennae could be be described by two power-laws joined by a weak bend (with $`\alpha _1=1.7`$ and $`\alpha _2=2.6`$), and there may also be hints of curvature in the mass function we have derived. However, any such deviations in the latter from a single power law have low statistical significance ($`<2\sigma `$). Alternatively, the bend in the luminosity function may result from the fading of clusters that formed over a period of a few hundred Myr with a power-law mass function truncated near $`10^6M_{}`$.
Our results differ substantially from those based on earlier observations of the Antennae with the WFPC1 on HST (Meurer 1995; Fritze-von Alvensleben 1999). To understand this difference, we have performed two tests. First, we artificially truncated our sample at $`M_V=9.6`$, the same limit adopted by Fritze-von Alvensleben, and then repeated the entire analysis described above. (Note: after corrections for extinction, $`M_V=9.6`$ corresponds approximately to $`M_V^c=9.9`$ to $`11.1`$). In this case, we obtained a mass function similar to the lognormal one found by Fritze-von Alvensleben. Second, we performed a simulation in which clusters were drawn from a power-law mass function (with $`\beta =2`$) and a uniform age distribution. The luminosities of the clusters were then computed from the BC96 population synthesis models. For simulated clusters brighter than $`M_V=9.6`$, the mass function again resembled the one found by Fritze-von Alvensleben. The reason for this is that, because the clusters fade, those with high masses can be observed over a wide range of ages, whereas those with low masses can be observed only when they are young. As a result, low-mass clusters are underrepresented in the observed mass function, which therefore declines toward both high and low masses.
## 4 Discussion
We have found that the mass function of young star clusters in the Antennae is well represented by a power law, $`\psi (M)M^2`$, over the range $`10^4M10^6M_{}`$. This is similar to the power-law mass function of diffuse and molecular clouds in the Milky Way (Dickey & Garwood 1989; Solomon & Rivolo 1989). However, it differs radically from the lognormal mass function of old globular clusters, which peaks at a $`\mathrm{few}\times 10^5M_{}`$ and declines rapidly toward both higher and lower masses. It is widely believed that galaxies formed hierarchically by the merging of smaller galaxies and/or subgalactic fragments. Thus, our results have potentially important implications concerning the origin of globular clusters during the early phases of galactic evolution. In this connection, it is worth emphasizing that a power-law mass function is “scale free,” whereas a lognormal mass function has a “preferred scale.”
One explanation for the different mass functions is that the conditions in ancient galaxies and protogalaxies were such as to imprint a characteristic mass of a $`\mathrm{few}\times 10^5M_{}`$ but that these conditions no longer prevail in modern galaxies. For example, the minimum mass of newly formed star clusters, set by the Jeans mass of interstellar clouds, will be high when the gas cannot cool efficiently and low when it can, which in turn will depend on the abundances of heavy elements and molecules, the strength of any heat sources, and so forth. These effects favor the formation of clusters in a narrower range of masses in the past than at present (Fall & Rees 1985; Kang et al. 1990). Another explanation for the different mass functions is that populations of stars clusters were born scale free, but later acquired a preferred scale by the selective disruption of low-mass clusters (Fall & Rees 1977; Gnedin & Ostriker 1997, and references therein). In this case, a power-law mass function might evolve into a lognormal mass function (Okazaki & Tosa 1995; Elmegreen & Efremov 1997; Baumgardt 1998; Vesperini 1998). Whether this occurs, however, depends on other aspects of the initial conditions, especially how the mass-radius plane is populated. We plan to address this issue in a future paper.
We thank Claus Leitherer and Brad Whitmore for valuable discussions. Support for this work was provided by NASA through grant number GO-07468 from STScI. |
no-problem/9911/cond-mat9911426.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Electronic transport properties (see also ) as for example the electrical conductivity, the thermopower, as well as the Hall effect are of particular interest due to their sensitivity to the interrelation between atomic structure and electronic system. As described in more detail in the contribution of Häussler , a correspondence between the wave length of the conduction electrons at $`E_F`$ and a frequently occurring distance in the atomic system leads to resonance-like interaction, causing an energy-lowering of the electronic system. This behaviour can, on one hand, stabilize an atomic structure like the quasicrystalline one and, on the other hand, it modifies the electronic system so that transport anomalies arise. Stable quasicrystals can exhibit transport properties which can neither be described as metallic nor as semiconducting (activated behaviour) as shown in fig.9. The anomalies can be strong enough to cause a metal-insulator transition (MIT) in icosahedral Al-Pd-Re .
First investigations on quasicrystalline thin films were performed in order to investigate properties which could be relevant for technical applications, as for example mechanical, tribological and corrosion properties . Additionally, thin films were used to measure transport properties, because of two reasons: first of all, thin films show a much higher electrical resistance than bulk samples and allow a much easier and more precise determination of the resistivity. Films were used to determine the very small magnetoresistance at low magnetic fields, which is useful to interpret some transport anomalies in the framework of quantum corrections . Secondly, some work was performed to investigate the transition from 3 to 2-dimensional behaviour as a function of film thickness .
We want to show here that thin quasicrystalline films do not differ substantially from bulk materials, but offer many advantages with respect to the preparation as well as to measuring electronic transport properties. Additionally, thin films can be prepared in the amorphous phase ($`a`$-phase) and crystallized afterwards into the icosahedral phase ($`i`$-phase) without the intermediate formation of crystalline counterparts. Accordingly, one sample allows the comparison of the quasicrystalline with the amorphous phase as well as the investigation of the transition between both.
Thin-film techniques provide samples which are well suited for systematic investigations as a function of composition and structural quality. We show that for a narrow range of composition large transport anomalies occur, up to insulating films for $`i`$-Al-Pd-Re. We discuss this in the framework of resonant scattering of the conduction electrons with the quasicrystalline structure, causing a reduced electronic mobility and DOS at $`E_F`$.
We should address the question, which films are denoted as thin. The thickness should be compared to the length scales of structural units or relevant physical lengths. The smallest structural units, namely the building clusters of quasicrystals are of order 1 nm . As clusters are selfsimilar, an upper limit for cluster sizes is not easy to specify, but 20 nm is a value of a quite large cluster shown in . Most of the relevant length scales of physical mechanisms are much shorter. The electronic scattering length for example is in the order of some atomic distances in the case of elastic scattering (intrinsic to good quasicrystals), while the inelastic scattering length increases with decreasing temperature and can reach $`l_i=10^2\mathrm{nm}`$ at $`T=1`$ K . From this it can be concluded that the thickness of a typical film is well inside the regime where a 3-dimensional behaviour can be expected. A film may be called ‘thin’ when it is prepared by a deposition process from the vapor phase.
Very thin films with a thickness starting at about $`d=10`$ nm were produced as selective solar absorbers . It is not clear if films of such a thickness are homogeneous or behave more like a percolating network. Therefore, samples for electronic transport measurements cover the thickness range $`30<d<1000`$ nm. The thickness of coatings for technical applications is much higher, up to 200 $`\mu `$m .
## 2 Preparation of thin quasicrystalline films
The films considered here are prepared by evaporation or sputtering from one or more sources onto substrates at different temperatures. Two routes are possible: homogeneous samples may be prepared directly, or crystalline multilayers of the different components may be prepared first, followed by extensive heat treatments. While the first technique can lead directly to quasicrystalline (qc) films (if deposited at a sufficiently high temperature), the second needs interdiffusion between the layers and a solid state reaction to form a qc-film. First films for conductivity measurements were fabricated with the second technique by a sequential sputtering technique of Al, Fe and Cu. As the thickness of each layer could be controlled by a quartz oscillator with high accuracy, the overall composition has been adjusted such that the narrow range of the $`i`$-phase in the phase diagram was reached . Some groups have prepared thin films of Al-Cu-Fe by sputtering from a composite or area- sectional target . This results in samples with homogeneous composition and can lead to amorphous films if the substrate temperature is held below $`T500`$ K and to qc-films for $`T>750`$ K.
The amorphous films could be transformed to the qc state (780 K) via a cubic phase ($`T_{cryst}=580`$ K) by an appropriate heat treatment . A problem of this technique in order to gain samples with systematic varying compositions is the need of an individual target for each composition desired.
The evaporation of a master alloy by an electron beam is another method which has been used . Here the films have compositions which are different from the master alloy and have to be determined afterwards.
In the present paper we report on a sequential flash-evaporation technique (fig.1) in order to prepare amorphous Al-Cu-Fe samples with an exactly defined stochiometry as a precursor for the qc-phase. Grains of a premolten master alloy of the correct composition are feeded to a hot tungsten filament where they flash evaporate. As each grain contributes one monolayer or less to the film thickness, even a complete segregation finally ends up with a homogeneous film of the nominal composition. As a second technique for Al-Pd-Re we use co-sputtering with two magnetron sources (fig.2a). Due to the positions of the two sources in respect to the substrate a defined composition gradient can be achieved along the substrate. With the latter technique in one preparation process a set of amorphous samples can be produced consisting of about 20 samples with a composition, slightly and systematically changing from one sample to the next, cutting the ternary phase diagram at, or close to the optimal composition (fig.2b). A defined shape of the evaporated or sputtered samples can be achieved by applying a mask or a structuring technique using microelectronic technologies, respectively. This allows to determine the absolute resistivity, which is difficult for most bulk samples because of their irregular shape and brittleness.
## 3 Amorphous to quasicrystalline transition
The amorphous state prepared by quenching of the vapor phase onto a cooled substrate is structurally similar to the liquid phase. Therefore, a direct transition from the amorphous to the qc-phase is possible for films at appropriate composition. In contrast to the transition from the liquid to the quasicrystal, the transition from the amorphous to the qc state is irreversible and occurs at much lower temperatures.
Fig.3a shows the resistivity of a thin film of Al-Cu-Fe deposited by sequential flash evaporation onto a substrate cooled down to $`T=4.2`$ K. The resistivity of the amorphous state at this temperature is $`\rho (4.2K)=300\mu \mathrm{\Omega }`$cm and increases to a value of $`\rho (4.2K)=470\mu \mathrm{\Omega }`$cm after an annealing up to $`T=550`$ K. This behaviour is discussed later. Besides the quite small and continuous irreversible increase of the resistivity and its temperature coefficient, a distinct increase related to the direct transition from the $`a`$\- to the $`i`$-phase can be seen at $`T740`$ K. Above, the sample is icosahedral as shown by electron diffraction . The transition temperature is about 350 K lower than the solidus temperature of the same system. This offers new possibilities. First, the transition happens within the solid phase and allows the measurement of electronic transport properties at the transition (fig.4). Second, the transition plotted in fig.3 shows no indication of the formation of crystalline phases like the cubic $`\beta `$\- or $`\tau `$-phases which occur whenever an Al-Cu-Fe melt is cooled down from the liquid state by melt-spinning or conventional casting. Therefore, in films produced via the amorphous route no long-term annealing treatments (typically some hours at $`T=1050`$ K for bulk samples) are necessary for removing crystalline phases. Third, the transition from the $`a`$\- to the qc-phase happens not only at low $`T`$ but also on short time scales in the order of minutes. Fig.4 shows this transition of $`\mathrm{Al}_{63.5}\mathrm{Cu}_{24}\mathrm{Fe}_{12.5}`$films for different heating rates. Together with a low surface roughness achieved by this preparation, this may have consequences for technical applications of qc coatings.
## 4 Comparing the a-with the i-phase - the scattering approach
The preparation of qc thin films via the route of the $`a`$-phase offers the possibility to compare directly the electronic system of the isotropic amorphous with the nearly isotropic quasicrystalline phase in one and the same sample.
The conductivity of very different quasicrystals, stable ones with strong transport anomalies like $`i`$-Al-Cu-Fe, $`i`$-Al-Pd-Mn, or $`i`$-Al-Pd-Re, as well as simple ones (without $`d`$-states at the Fermi level) with much smaller transport anomalies like $`i`$-Al-Mg-Zn have often been described by an inverse Matthiessen rule. That means that their $`T`$-dependencies of the conductivity $`\sigma (T)`$ are claimed to be the sum of the conductivity $`\sigma _0`$ at $`T0`$ K which depends on the particular system, its composition and structural quality, and on a temperature dependent increase $`\mathrm{\Delta }\sigma (T)`$ which is very similar for all qc phases and samples
$$\sigma (T)=\sigma _0+\mathrm{\Delta }\sigma (T).$$
This behaviour is inverse to that of ordinary metals, where the electrical resistivity instead of the conductivity shows a comparable behaviour. Astonishingly, fig.3b (conductivity) shows a parallel behaviour of the curves belonging to the amorphous and the icosahedral phases. Accordingly, also the behaviour of the $`a`$-phase can be described by the inverse Matthiessen rule, suggesting a strong similarity between the amorphous and the quasicrystalline phases.
For quasicrystals theories exist which are based on the particular features of the quasicrystalline structure and its relation to the electronic system. For example a scattering induced electronic hopping between badly propagating electronic states , the localization tendency of electrons in selfsimiliar clusters or a temperature dependent change of the band structure due to the decrease of elastic Bragg scattering at higher temperatures . These models can explain only few aspects of electronic transport as for example the inverse Matthiessen rule.
A different approach to understand the unexpected properties of quasicrystals is motivated by the electronic stabilization of their peculiar structure. This is due to the interaction between the electrons at the Fermi level (characterized by $`2k_F`$) and the structure (characterized by the Jones zone diameter $`k_{pe}`$). Such a behaviour corresponds to the Hume-Rothery mechanism in periodic crystals and was first discussed for quasicrystals by Smith and Ashcroft and Friedel . This description is obvious as the Jones zone of a quasicrystal (fig.6b) consists of many flat areas and hence can interact with a spherical Fermi surface in many directions.
Fig.5 shows that the strongest peaks of the $`i`$-phase are at the position of the electronically induced peak of the $`a`$-phase . This suggests that the size of the Jones zone as well as its shape (see fig. 6) is very similar for the $`a`$\- and the $`i`$-phase. It has been shown that the stability and many other properties of a large class of amorphous systems can be understood by such an interaction. Thin films are well suited to investigate this aspect because they allow the comparison of the amorphous with the quasicrystalline state in one sample. Electronic transport properties of films can be measured in situ up to elevated temperatures due to easy preparation techniques and the high contact stability.
The interaction between electronic and atomic system is based on the elastic scattering of electrons with the Fermi wavenumber ($`k=k_F`$) at a pronounced peak in the structure factor $`S(2k_F)`$ describing the atomic structure. The elastic (Bragg-like) scattering of electrons reduces not only the mean free path, the main effect may be the depression of the DOS at $`E_F`$. If the matching of the $`k`$-vectors of the electronic and the atomic system is good, this results in a resonant-like scattering of electrons. The previously freely propagating electrons change partially to standing waves because of resonant scattering, resulting in a tendency of localization of the electrons at $`E_F`$. As a consequence, the electronic DOS at $`E_F`$ is reduced, the stability of the system increases and electronic transport anomalies arise. The aspect of stability is comprehensively discussed in , while the consequences for electronic transport in amorphous and quasicrystalline films will be briefly discussed in the following. In the case of a weak scattering approach the Ziman formula
$$\rho _z=\frac{1}{\sigma _z}_0^{2k_F}S(K)|v(K)|^2K^3𝑑k$$
has been used to describe the resistivity and its temperature dependence for liquid and amorphous systems. $`K`$ represents the scattering vector, $`v(K)`$ the pseudopotential and $`S(K)`$ the static structure factor. This formula cannot strictly be applied to systems with strong scattering as are some amorphous systems or quasicrystals. But this relatively simple formula can be helpful to follow up the trends to stronger scattering. Due to the weighting with $`K^3`$ the elastic scattering strongly depends on $`S(2k_F)`$, the structural weight at the upper limit of the integral. Thus, a pronounced peak in the structure factor at $`2k_F`$ leads to a strong elastic scattering and a low conductivity. As the temperature increases, the portion of the elastic scattering at $`K=2k_F`$ is reduced in favour of inelastic scattering. The latter is less effective on reducing the conductivity than the diffuse umklapp scattering. Thus, the conductivity increases with increasing inelastic scattering due to rising temperature.
Additionally, the resonant like elastic scattering due to the coincidence of the electronic wavelength with structural distances opens a gap or a pseudogap at $`E_F`$. This is similar to the case of periodic crystals where a matching of the wave length of the electron waves at $`E_F`$ with the wave length of a reciprocal lattice vector causes a gap due to the energy difference for the two solutions of standing electron waves. In quasicrystals the matching may be slightly different. A resonance-like interaction is most effective, if not only the wavelength but also the waveforms of both systems are identical. In quasicrystals and especially in amorphous alloys, it is well known that the atoms are onionshell-like located around ad-atoms. The distance between neighbouring shells $`\lambda _P`$ is in agreement with the Friedel wavelength $`\lambda _F=2\pi /2k_F`$. Each atom is surrounded by ‘mirror spheres’ analog to the equidistant separation of mirror planes in the crystalline case. The electrons should be treated as spherical waves in coincidence with the onionshell-like static structure. Under resonance, any plane wave which once may have been formed will by scattering immediately fall apart into spherical waves which are strongly localized due to the phase-coherent backscattering from all the neighbouring onionshells. The resonance-like coupling causes localization of the electrons at any site and hence a short mean-free path which itself helps to enhance the role of the spherical waves at medium range. In the electronic DOS pseudogaps arise. The onionshell-like position of the neighbouring atoms around any ad-atom may also be seen as a cluster-model with strongly interpenetrating clusters. Whereas in the amorphous phase the location of the atoms on the onionshells is still strongly random and in addition angular correlations are weak or absent, this is no longer the case in the qc-phase. The distances of the individual atoms to the ad-atoms are much better defined and strong angular correlations exist. The phase coherence of the backscattered spherical waves is better fulfilled and therefore the interference effects are even stronger. Accordingly the localization of the electron waves should be stronger and the pseudogap should become deep. As the temperature rises the portion of the resonant elastic scattering is decreased in favour of inelastic scattering reducing the pseudogap both for amorphous and quasicrystalline systems. We mention that this behaviour is simply ruled by the Debye-Waller factor which describes the ratio of the elastic to inelastic scattering for the relevant peaks building up the Jones zone. This can be seen in fig.3b. The conductivity for all reversible parts (marked by $``$ ) of the amorphous states and also for the quasicrystalline state increases with temperature in exactly the same manner. The conductivity irreversibly decreases (marked by $``$) with every annealing step reaching its smallest value in the qc-phase with its strong radial and angular correlation of atoms. The temperature dependence of the conductivity is discussed in detail in the Al-Pd-Re system as this icosahedral system shows even larger transport anomalies (reaching insulating behaviour) and has a much smaller contribution of low temperature effects from spin-orbit scattering.
The effect of resonance in amorphous and quasicrystalline samples can also be seen in the thermopower. For amorphous systems it is well known that a stronger resonance between static structure and electronic system causes a deviation from the free electron value of the thermopower $`S_0`$ to positive values . Fig.7 shows the same behaviour for the icosahedral phase. The samples show a transition from a large negative to a large positive thermopower. This correlates with the increase of the resistance ratio $`\rho (4K)/\rho (300K)`$, which is an indication of the enhanced resonance effect between static structure and electronic system. Note that the $`i`$-phases show thermopowers which are much larger than the free-electron value. This can be attributed to a deep pseudogap which reduces the effective number of charge carriers at the Fermi-level.
## 5 The Al-Pd-Re system
The icosahedral Al-Pd-Re system shows the largest transport anomalies amongst all the quasicrystals, up to a metal-insulator transition (MIT) . Thus it is interesting to check the concept of resonant interaction between the static structure and the conduction electrons on this alloy. Due to the extremely different vapor pressures of the constituents it is difficult to produce homogeneous bulk samples of a defined composition in a reproducible way. With the co-sputtering technique mentioned before, systematic and slightly different samples can be fabricated as thin films.
As an example, Fig.8a shows the conductivity of such a sequence as a function of the Rhenium content for a constant Aluminium to Palladium ratio. There exists a systematic variation of the low $`T`$ conductivity with a minimum around $`\mathrm{Al}_{72.3}\mathrm{Pd}_{20.2}\mathrm{Re}_{7.5}`$. Simultaneously, some of the samples seem to be insulating , that means, the conductivity vanishes for $`T0`$ K (fig. 8b). A magnetic field enhances this behaviour and apparently leads to a vanishing conductivity at a small but finite temperature.
Thereby, the transport properties of $`i`$-Al-Pd-Re match into the common behaviour of stable quasicrystals mentioned before in the $`i`$-Al-Cu-Fe system. Fig.10b shows again the validity of the inverse Matthiessen-rule for the conductivity of the $`a`$\- and $`i`$-phases of different structural qualities. In contrast to all other qc systems, the Al-Pd-Re system offers two exciting aspects. First, the conductivity vanishes at $`T0`$ K, which may be due a stronger resonant scattering because of heavier elements in the case of Al-Pd-Re. Second, at low $`T`$, there is no increase of conductivity, as for high quality $`i`$-Al-Cu-Fe (fig.7b) or for other stable quasicrystals like $`i`$-Al-Pd-Mn. Such a low $`T`$ increase of conductivity (attributed to strong spin-orbit scattering and electron-electron interaction ) makes a MIT unlikely and complicates interpretation.
The model of a resonant-like interaction of the electronic and the static structure is able to explain the transport properties also for the insulating samples. Fig.9 shows a fit to the conductivity data in a quite large temperature range. The fit is indirectly proportional to a function of the Debye-Waller factor $`exp(cP(\frac{T}{\mathrm{\Theta }}))`$, with the Debeye-temperature $`\mathrm{\Theta }`$ and a constant $`c`$ which describes the temperature dependent decrease of the elastic scattering of the electrons at $`2k_F`$ at the Jones zone. There has been only one parameter ($`c_3`$ in the expression written in fig.9) added which is related to the size of the gap between electron- and hole-like charge carriers and is related to the effective pseudopotential. A significant difference between fit and measured data occurs only at temperatures below 100 K. Fig.9b shows that this difference is proportional to $`\sqrt{T}`$ which can be attributed to electron-electron interaction (EEI). In systems of strong
elastic scattering, electrons are less able to screen charge variations and EEI is enhanced, leading to a small dip in the electronic density of states at $`E_F`$ . In the case of amorphous alloys this behaviour can be computed quantitatively as a function of temperature and magnetic field . In weakly insulating icosahedral systems the MIT seems to have two origins. First, the resonance-like interaction between the static structure leads to a deep pseudogap in the electronic DOS at $`E_F`$ and the localization tendency of the conduction electrons due to standing spherical waves in some structural clusters. Second, the EEI lowers the small DOS at $`E_F`$ additionally. The sum of both effects leads to a vanishing conductivity at $`T0`$ K. A magnetic field dephases the electronic wavefunction enlarging the EEI by a further decreased screening. The MIT occurs at a finite temperature.
The thermopower of $`i`$-Al-Pd-Re can be described in the resonance-model mentioned before. As shown in fig.10, the thermopower increases to large positive values as the structural quality increases from the amorphous to the defective and finally the good icosahedral state. The conductivity of the same sample reduces in the same manner as the resonance effect increases.
## 6 Conclusion
Quasicrystalline thin films can be produced with high structural quality. The low temperature electrical conductivity, which is very sensitive to the structure quality, reaches values which are comparable to that of bulk materials. A MIT is achieved in both, the bulk samples as well as in thin films of $`i`$-Al-Pd-Re. Thin films provide the possibility to make systematic investigations as a function of composition and structural quality. Further understanding in the interpretation of the peculiar transport anomalies of quasicrystals could be achieved by the comparison to the isotropic amorphous phase. This has been done on different annealing states of the same sample. Astonishingly, the temperature dependence of the conductivity is qualitatively very similiar for both phases, the inverse Matthiessen rule which is thought to be peculiar to quasicrystals is also valid for amorphous samples of the same composition. We discussed this in the framework of a resonant scattering between the conduction electrons and the static structure. Hereby, the mechanism is the same for amorphous as well as for qc systems. Due to the sharp distance distributions and angular correlations in quasicrystals the quantitiy of the scattering is much larger in the quasicrystalline case, leading to a deeper pseudogap and a stronger localization tendency of the conduction electrons. The temperature dependence of the conductivity could be fitted due to the scattering approach by a simple function of the Debye-Waller factor for the temperature range from 100 to 900 K. At low temperatures the conductivity of $`i`$-Al-Pd-Re is additionally influenced by strong electron-electron interaction. This leads to an additional dip of the electronical DOS at $`E_F`$ and finally to the metal- insulator-transition. The thermopower increases to large positive values as the conductivity vanishes for high-quality qc samples. Again, this is a typical behaviour of electronically stabilized systems, which can be described by a strong interaction between the conduction electrons and the static structure.
## 7 Acknowledgements
Support of this work by DFG under contracts Ha2359/1, Ha1627/8 and Ha1627/9 is gratefully acknowledged. |
no-problem/9911/hep-ph9911229.html | ar5iv | text | # Kaluza-Klein Physics at Muon Colliders 1footnote 11footnote 1To appear in the Proceedings of the Study on Colliders and Collider Physics at the Highest Energies: Muon Colliders at 10 TeV to 100 TeV, Montauk Yacht Club Resort, Montauk, New York, 27 September–1 October 1999
## Introduction
In theories with extra dimensions, $`d1`$, the gauge fields of the Standard Model(SM) will have Kaluza-Klein(KK) excitations if they are allowed to propagate in the bulk of the extra dimensions. If such a scenario is realized then, level by level, the masses of the excited states of the photon, $`Z`$, $`W`$ and gluon would form highly degenerate towers. The possibility that the masses of the lowest lying of these states, of order the inverse size of the compactification radius $`1/R`$, could be as low as $``$ a few TeV or less leads to a very rich and exciting phenomenology at future and, possibly, existing collidersold . For the case of one extra dimension compactified on $`S^1/Z_2`$ the spectrum of the excited states is given by $`M_n=n/R`$ and the couplings of the excited modes relative to the corresponding zero mode to states remaining on the wall at the orbifold fixed points, such as the SM fermions, is simply $`\sqrt{2}`$ for all $`n`$. These masses and couplings are insensitive to the choice of compactification in the case of one extra dimension assuming the metric tensor factorizes, i.e., the elements of the metric tensor on the wall are independent of the compactified co-ordinates.
If such KK states exist what is the lower bound on their mass? We already know from direct $`Z^{}/W^{}`$ and dijet bump searches at the Tevatron from Run I that they must lie above $`0.85`$ TeVtev . A null result for a search made with data from Run II will push this limit to $`1.1`$ TeV or so. To do better than this at present we must rely on the indirect effects associated with KK tower exchange in what essentially involves a set of dimension-six contact interactions. Such limits rely upon a number of additional assumptions, in particular, that the effect of KK exchanges is the only new physics beyond the SM. The strongest and least model-dependent of these bounds arises from an analysis of charged current contact interactions at both HERA and the Tevatron by Cornet, Relano and Ricocornet who, in the case of one extra dimension, obtain a bound of $`R^1>3.4`$ TeV. Similar analyses have been carried out by a number of authorshost ; rw ; the best limit arises from an updated combined fit to the precision electroweak datarw as presented at the 1999 summer conferencesdata and yieldskktest $`R^1>3.9`$ TeV for the case of one extra dimension. From the previous discussion we can also draw a further conclusion for the case $`d=1`$: the lower bound $`M_1>3.9`$ TeV is so strong that the second KK excitations, whose masses must now exceed 7.8 TeV due to the above scaling law, will be beyond the reach of the LHC. This leads to the important result that the LHC will at most only detect the first set of KK excitations for $`d=1`$.
In all analyses that obtain indirect limits on $`M_1`$, one is actually constraining a dimensionless quantity such as
$$V=\underset{𝐧=1}{\overset{\mathrm{}}{}}\frac{g_𝐧^2}{g_0^2}\frac{M_w^2}{M_𝐧^2},$$
(1)
where, generalizing the case to $`d`$ additional dimensions, $`g_𝐧`$ is the coupling and $`M_𝐧`$ the mass of the $`n^{th}`$ KK level labelled by the set of $`d`$ integers n and $`M_w`$ is the $`W`$ boson mass which we employ as a typical weak scale factor. For $`d=1`$ this sum is finite since $`M_n=n/R`$ and $`g_n/g_0=\sqrt{2}`$ for $`n>1`$; one immediately obtains $`V=\frac{\pi ^2}{3}(M_w/M_1)^2`$ with $`M_1`$ being the mass of the first KK excitation. From the precision data one obtains a bound on $`V`$ and then uses the above expression to obtain the corresponding bound on $`M_1`$. For $`d>1`$, however, independently of how the extra dimensions are compactified, the above sum in $`V`$ diverges and so it is not so straightforward to obtain a bound on $`M_1`$. We also recall that for $`d>1`$ the mass spectrum and the relative coupling strength of any particular KK excitation now become dependent upon how the additional dimensions are compactified.
There are several ways one can deal with this divergence: ($`i`$) The simplest approach is to argue that as the states being summed in $`V`$ get heavier they approach the mass of the string scale, $`M_s`$, above which we know little and some new theory presumably takes over. Thus we should just truncate the sum at some fixed maximum value $`n_{max}M_sR`$ so that masses KK masses above $`M_s`$ do not contribute. ($`ii`$) A second possibility is to note that the wall on which the SM fermions reside is not completely rigid having a finite tension. The authors in Ref.wow argue that this wall tension can act like an exponential suppression of the couplings of the higher KK states in the tower thus rendering the summation finite, i.e., $`g_𝐧^2g_𝐧^2e^{(M_n/M_1)^2/n_{max}^2}`$, where $`n_{max}`$ now parameterizes the strength of the exponential cut-off. (Antoniadiskktest has argued that such an exponential suppression can also arise from considerations of string scattering amplitudes at high energies.) For a fixed value of $`n_{max}`$, the exponential approach is found to be more effective and lead to a smaller sum than that obtained by simple truncation and thus to a weaker bound on $`M_1`$. ($`iii`$) A last scenarioschm is to note the possibility that the SM wall fermions may have a finite size in the extra dimensions which smear out and soften the couplings appearing in the sum to yield a finite result. In this case the suppression is also of the Gaussian variety.
We note that in all of the above approaches the value of the sum increases rapidly with $`d`$ for a fixed value of the cut-off parameter $`n_{max}`$. For $`d=2(>2)`$ the sum behaves asymptotically as $`logn_{max}(n_{max}^{d2})`$. This leads to the very important result that, for a fixed bound on $`V`$ from experimental data, the corresponding bound on the mass of the lowest lying KK excitation rapidly strengthens with the number of extra dimension, $`d`$. Table I shows how the $`d=1`$ lower bound of 3.9 TeV for the mass of $`M_1`$ changes as we consider different compactifications for $`d>1`$. We see that in some cases the value of $`M_1`$ is so large it will be beyond the mass range accessible to the LHC as it is for all cases of the $`d=3`$ example.
## SM KK States at the LHC and Linear Colliders
Let us return to the $`d=1`$ case at the LHC where the degenerate KK states $`\gamma ^{(1)}`$, $`Z^{(1)}`$, $`W^{(1)}`$ and $`g^{(1)}`$ are potentially visible. It has been shown kktest that for masses in excess of $`4`$ TeV the $`g^{(1)}`$ resonance in dijets will be washed out due to its rather large width and the experimental jet energy resolution available at the LHC detectors. Furthermore, $`\gamma ^{(1)}`$ and $`Z^{(1)}`$ will appear as a single resonance in Drell-Yan that cannot be resolved and looking very much like a single $`Z^{}`$. Thus if we are lucky the LHC will observe what appears to be a degenerate $`Z^{}/W^{}`$. How can we identify these states as KK excitations when we remember that the rest of the members of the tower are too massive to be produced? We remind the reader that many extended electroweak modelsmodels exist which predict a degenerate $`Z^{}/W^{}`$. Without further information, it would seem likely that this would become the most likely guess of what had been found.
To clarify this situation let us consider the results displayed in Figs. 1 for $`d=1`$ where we show the production cross sections in the $`\mathrm{}^+\mathrm{}^{}`$ channel with inverse compactification radii of 4, 5 and 6 TeV. In calculating these cross sections we have assumed that the KK excitations have their naive couplings and can only decay to the usual fermions of the SM. Additional decay modes can lead to appreciably lower cross sections so that we cannot use the peak heights to determine the degeneracy of the KK state. Note that in the 4 TeV case, which is essentially as small a mass as can be tolerated by the present data on precision measurements, the second KK excitation is visible in the plot. We see several things from these figures. First, we can easily estimate the total number of events in the resonance regions associated with each of the peaks assuming the canonical integrated luminosity of $`100fb^1`$ appropriate for the LHC; we find $`300(32,3,0.02)`$ events corresponding to the 4(5,6,8) TeV resonances if we sum over both electron and muon final states and assume $`100\%`$ leptonic identification efficiencies. Clearly the 6 and 8 TeV resonances will not be visible at the LHC (though a modest increase in luminosity by a factor of a few will allow the 6 TeV resonance to become visible) and we also verify our claim that only the first KK excitations will be observable. In the case of the 4 TeV resonance there is sufficient statistics that the KK mass will be well measured and one can also imagine measuring the forward-backward asymmetry, $`A_{FB}`$, if not the full angular distribution of the outgoing leptons, since the final state muon charges can be signed. Given sufficient statistics, a measurement of the angular distribution would demonstrate that the state is indeed spin-1 and not spin-0 or spin-2. However, for such a heavy resonance it is unlikely that much further information could be obtained about its couplings and other properties. In fact the conclusion of several years of $`Z^{}`$ analysessnow is that coupling information will be essentially impossible to obtain for $`Z^{}`$-like resonances with masses in excess of 1-2 TeV at the LHC due to low statistics. Furthermore, the lineshape of the 4 TeV resonance and the Drell-Yan spectrum anywhere close to the peak will be difficult to measure in detail due to both the limited statistics and energy smearing. Thus we will never know from LHC data alone whether the first KK resonance has been discovered or, instead, some extended gauge model scenario has been realized. To make further progress we need a lepton collider.
It is well-known that future $`e^+e^{}`$ linear colliders(LC) operating in the center of mass energy range $`\sqrt{s}=0.51.5`$ TeV will be sensitive to indirect effects arising from the exchange of new $`Z^{}`$ bosons with masses typically 6-7 times greater than $`\sqrt{s}`$snow . This sensitivity is even greater in the case of KK excitations since towers of both $`\gamma `$ and $`Z`$ exist all of which have couplings larger than their SM zero modes. Furthermore, analyses have shown that with enough statistics the couplings of the new $`Z^{}`$ to the SM fermions can be extractedcoupl in a rather precise manner, especially when the $`Z^{}`$ mass is already approximately known from elsewhere, e.g., the LHC. (If the $`Z^{}`$ mass is not known then measurements at several distinct values of $`\sqrt{s}`$ can be used to extract both the mass as well as the corresponding couplingsme .) In the present situation, we imagine that the LHC has discovered and determined the mass of a $`Z^{}`$-like resonance in the 4-6 TeV range. Can the LC tell us anything about this object?
The obvious step would be to use the LC to extract the couplings of the apparent resonance discovered by the LHC; we find that it is sufficient for our arguments below to do this solely for the leptonic channels. The idea is the following: we measure the deviations in the differential cross sections and angular dependent Left-Right polarization asymmetry, $`A_{LR}^{\mathrm{}}`$, for the three lepton generations and combine those with $`\tau `$ polarization data. Assuming lepton universality(which would be observed in the LHC data anyway), that the resonance mass is well determined, and that the resonance is an ordinary $`Z^{}`$ we perform a fit to the hypothetical $`Z^{}`$ coupling to leptons, $`v_l,a_l`$. To be specific, let us consider the case of only one extra dimension with a 4 TeV KK excitation and employ a $`\sqrt{s}=500`$ GeV collider with an integrated luminosity of 200 $`fb^1`$. The result of performing this fit, including the effects of cuts and initial state radiation, is shown in Fig.2. Here we see that the coupling values are ‘well determined’ (i.e., the size of the $`95\%`$ CL allowed region we find is quite small) by the fitting procedure as we would have expected from previous analyses of $`Z^{}`$ couplings extractions at linear colliderssnow ; coupl ; me .
The only problem with the fit shown in the figure is that the $`\chi ^2`$ is very large leading to a very small confidence level, i.e., $`\chi ^2/d.o.f=95.06/58`$ or CL=$`1.55\times 10^3`$! (We note that this result is not very sensitive to the assumption of $`90\%`$ beam polarization; $`70\%`$ polarization leads to almost identical results.) For an ordinary $`Z^{}`$ it has been shown that fits of much higher quality, based on confidence level values, are obtained by this same procedure. Increasing the integrated luminosity can be seen to only make matters worse. Fig.3 shows the results for the CL following the above approach as we vary both the luminosity and the mass of the first KK excitation at both 500 GeV and 1 TeV $`e^+e^{}`$ linear colliders. From this figure we see that the resulting CL is below $`10^3`$ for a first KK excitation with a mass of 4(5,6) TeV when the integrated luminosity at the 500 GeV collider is 200(500,900)$`fb^1`$ whereas at a 1 TeV for excitation masses of 5(6,7) TeV we require luminosities of 150(300,500)$`fb^1`$ to realize this same CL. Barring some unknown systematic effect the only conclusion that one could draw from such bad fits is that the hypothesis of a single $`Z^{}`$, and the existence of no other new physics, is simply wrong. If no other exotic states are observed below the first KK mass at the LHC, such as $`\stackrel{~}{\nu }`$rp or leptoquarksleptos , this result would give very strong indirect evidence that something more unusual that a conventional $`Z^{}`$ had been found but cannot prove that this is a KK state.
## SM KK States at Muon Colliders
In order to be completely sure of the nature of the first KK excitation, we must produce it directly at a higher energy lepton collider and sit on and near the peak of the KK resonance. To reach this mass range will most likely require a Muon Collider. The first issue to address is the quality of the degeneracy of the $`\gamma ^{(1)}`$ and $`Z^{(1)}`$ states. Based on the analyses in Ref.host ; rw we can get an idea of the maximum possible size of this fractional mass shift and we find it to be of order $`M_Z^4/M_{Z^{(1)}}^4`$, an infinitesimal quantity for KK masses in the several TeV range. Thus even when mixing is included we find that the $`\gamma ^{(1)}`$ and $`Z^{(1)}`$ states remain very highly degenerate so that even detailed lineshape measurements may not be able to distinguish the $`\gamma ^{(1)}/Z^{(1)}`$ composite state from that of a $`Z^{}`$. We thus must turn to other parameters in order to separate these two cases.
Sitting on the resonance there are a very large number of quantities that can be measured: the mass and apparent total width, the peak cross section, various partial widths and asymmetries etc. From the $`Z`$-pole studies at SLC and LEP, we recall a few important tree-level results which we would expect to apply here as well provided our resonance is a simple $`Z^{}`$. First, we know that the value of $`A_{LR}=[A_e=2v_ea_e/(v_e^2+a_e^2)]`$, as measured on the $`Z`$ by SLD, does not depend on the fermion flavor of the final state and second, that the relationship $`A_{LR}A_{FB}^{pol}(f)=A_{FB}^f`$ holds, where $`A_{FB}^{pol}(f)`$ is the polarized Forward-Backward asymmetry as measured for the $`Z`$ at SLC and $`A_{FB}^f`$ is the usual Forward-Backward asymmetry. The above relation is seen to be trivially satisfied on the $`Z`$(or on a $`Z^{}`$) since $`A_{FB}^{pol}(f)=\frac{3}{4}A_f`$ and $`A_{FB}^f=\frac{3}{4}A_eA_f`$. Both of these relations are easily shown to fail in the present case of a ‘dual’ resonance though they will hold if only one particle is resonating.
A short exercise shows that in terms of the couplings to $`\gamma ^{(1)}`$, which we will call $`v_1,a_1`$, and $`Z^{(1)}`$, now called $`v_2,a_2`$, these same observables can be written as
$`A_{FB}^f`$ $`=`$ $`{\displaystyle \frac{3}{4}}{\displaystyle \frac{A_1}{D}}`$
$`A_{FB}^{pol}(f)`$ $`=`$ $`{\displaystyle \frac{3}{4}}{\displaystyle \frac{A_2}{D}}`$
$`A_{LR}^f`$ $`=`$ $`{\displaystyle \frac{A_3}{D}},`$ (2)
where $`f`$ labels the final state fermion and we have defined the coupling combinations
$`D`$ $`=`$ $`(v_1^2+a_1^2)_e(v_1^2+a_1^2)_f+R^2(v_2^2+a_2^2)_e(v_2^2+a_2^2)_f`$ (3)
$`+2R(v_1v_2+a_1a_2)_e(v_1v_2+a_1a_2)_f`$
$`A_1`$ $`=`$ $`(2v_1a_1)_e(2v_1a_1)_f+R^2(2v_2a_2)_e(2v_2a_2)_f+2R(v_1a_2+v_2a_1)_e(v_1a_2+v_2a_1)_e`$
$`A_2`$ $`=`$ $`(2v_1a_1)_f(v_1^2+a_1^2)_e+R^2(2v_2a_2)_f(v_2^2+a_2^2)_e+2R(v_1a_2+v_2a_1)_f(v_1v_2+a_1a_2)_e`$
$`A_3`$ $`=`$ $`(2v_1a_1)_e(v_1^2+a_1^2)_f+R^2(2v_2a_2)_e(v_2^2+a_2^2)_f+2R(v_1a_2+v_2a_1)_e(v_1v_2+a_1a_2)_f,`$
with $`R`$ being the ratio of the widths of the two KK states, $`R=\mathrm{\Gamma }_1/\mathrm{\Gamma }_2`$, and the $`v_{1,2i},a_{1,2i}`$ are the appropriate couplings for electrons and fermions $`f`$. Note that when $`R`$ gets either very large or very small we recover the usual ‘single resonance’ results. Examining these equations we immediately note that $`A_{LR}^f`$ is now flavor dependent and that the relationship between observables is no longer satisfied:
$$A_{LR}^fA_{FB}^{pol}(f)A_{FB}^f,$$
(4)
which clearly tells us that we are actually producing more than one resonance.
Of course we need to verify that these single resonance relations are numerically badly broken before clear experimental signals for more than one resonance can be claimed. Statistics will not be a problem with any reasonable integrated luminosity since we are sitting on a resonance peak and certainly millions of events will be collected. With such large statistics only a small amount of beam polarization will be needed to obtain useful asymmetries. In principle, to be as model independent as possible in a numerical analysis, we should allow the widths $`\mathrm{\Gamma }_i`$ to be greater than or equal to their SM values as such heavy KK states may decay to SM SUSY partners as well as to presently unknown exotic states. Since the expressions above only depend upon the ratio of widths, we let $`R=\lambda R_0`$ where $`R_0`$ is the value obtained assuming that the KK states have only SM decay modes. We then treat $`\lambda `$ as a free parameter in what follows and explore the range $`1/5\lambda 5`$. Note that as we take $`\lambda 0(\mathrm{})`$ we recover the limit corresponding to just a $`\gamma ^{(1)}(Z^{(1)})`$ being present.
In Fig.4 we display the flavor dependence of $`A_{LR}^f`$ as a functions of $`\lambda `$. Note that as $`\lambda 0`$ the asymmetries vanish since the $`\gamma ^{(1)}`$ has only vector-like couplings. In the opposite limit, for extremely large $`\lambda `$, the $`Z^{(1)}`$ couplings dominate and a common value of $`A_{LR}`$ will be obtained. It is quite clear, however, that over the range of reasonable values of $`\lambda `$, $`A_{LR}^f`$ is quite obviously flavor dependent. We also show in Fig.4 the correlations between the observables $`A_{FB}^{pol}(f)`$ and $`A_{FB}(f)`$ which would be flavor independent if only a single resonance were present. From the figure we see that this is clearly not the case. Note that although $`\lambda `$ is an a priori unknown parameter, once any one of the electroweak observables are measured the value of $`\lambda `$ will be directly determined. Once $`\lambda `$ is fixed, then the values of all of the other asymmetries, as well as the ratios of various partial decay widths, are all completely fixed for the KK resonance with uniquely predicted values. This means that we can directly test the couplings of this apparent single resonance against what might be expected for a degenerate pair of KK excitations without any ambiguities.
In Figs. 5a and 5b we show that although on-resonance measurements of the electroweak observables, being quadratic in the $`Z^{(1)}`$ and $`\gamma ^{(1)}`$ couplings, will not distinguish between the usual KK scenario and that of the Arkani-Hamed and Schmaltz(AS) (whose KK couplings to quarks are of opposite sign from the conventional assignments for odd KK levels since quarks and leptons are assumed to be separated by a distance $`D=\pi R`$ in their scenario) the data below the peak in the hadronic channel will easily allow such a separation. The cross section and asymmetries for $`\mu ^+\mu ^{}e^+e^{}`$ (or vice versa) is, of course, the same in both cases. Such data can be collected by using radiative returns if sufficient luminosity is available. The combination of on and near resonance measurements will thus completely determine the nature of the resonance.
We note that all of the above analysis will go through essentially unchanged in any qualitative way when we consider the case of the first KK excitation in a theory with more than one extra dimension as is shown in Fig.6. Here we see that the shape of the excitation curves for the $`d=1`$ case and the $`d>1`$ models listed in Table 1 will clearly allow the number of dimensions and the compactification scheme to be uniquely identified.
## Randall-Sundrum Gravitons at Muon Colliders
The possibility of extra space-like dimensions with accessible physics near the TeV scale has recently opened a new window on the possible solutions to the hierarchy problem. Models designed to address this problem make use of our ignorance about gravity, in particular, the fact that gravity has yet to be probed at energy scales much above $`10^3`$ eV in laboratory experiments. The prototype scenario in this class of theories is due to Arkani-Hamed, Dimopoulos and Dvali(ADD)nima who use the volume associated with large extra dimensions to bring the $`d`$-dimensional Planck scale down to a few TeV. Here the hierarchy problem is recast into trying to understand the rather large ratio of the TeV Planck scale to the size of the extra dimensions which may be as large as a fraction of a millimeter. The phenomenologicalpheno implications of this model have been worked out by a large number of authors. An extrapolation of these analyses to the case of high energy muon colliders shows an enormous reach for this kind of physics.
More recently, Randall and Sundrum(RS)rs have proposed a new scenario wherein the hierarchy is generated by an exponential function of the compactification radius, called a warp factor. Unlike the ADD model, they assume a 5-dimensional non-factorizable geometry, based on a slice of $`AdS_5`$ spacetime. Two 3-branes, one being ‘visible’ with the other being ‘hidden’, with opposite tensions rigidly reside at $`S_1/Z_2`$ orbifold fixed points, taken to be $`\varphi =0,\pi `$, where $`\varphi `$ is the angular coordinate parameterizing the extra dimension. It is assumed that the extra-dimension bulk is only populated by gravity and that the SM lies on the brane with negative tension. The solution to Einstein’s equations for this configuration, maintaining 4-dimensional Poincare invariance, is given by the 5-dimensional metric
$$ds^2=e^{2\sigma (\varphi )}\eta _{\mu \nu }dx^\mu dx^\nu +r_c^2d\varphi ^2,$$
(5)
where the Greek indices run over ordinary 4-dimensional spacetime, $`\sigma (\varphi )=kr_c|\varphi |`$ with $`r_c`$ being the compactification radius of the extra dimension, and $`0|\varphi |\pi `$. Here $`k`$ is a scale of order the Planck mass and relates the 5-dimensional Planck scale $`M`$ to the cosmological constant. Examination of the action in the 4-dimensional effective theory in the RS scenario yields the relationship $`\overline{M}_{Pl}^2=M^3/k`$ for the reduced effective 4-D Planck scale.
Assuming that we live on the 3-brane located at $`|\varphi |=\pi `$, it is found that a field on this brane with the fundamental mass parameter $`m_0`$ will appear to have the physical mass $`m=e^{kr_c\pi }m_0`$. TeV scales are thus generated from fundamental scales of order $`\overline{M}_{Pl}`$ via a geometrical exponential factor and the observed scale hierarchy is reproduced if $`kr_c1112`$. Hence, due to the exponential nature of the warp factor, no additional large hierarchies are generated.
A recent analysisdhr examined the phenomenological implications and constraints on the RS model that arise from the exchange of weak scale towers of gravitons. There it was shown that the masses of the KK graviton states are given by $`m_n=kx_ne^{kr_c\pi }`$ where $`x_n`$ are the roots of $`J_1(x_n)=0`$, the ordinary Bessel function of order 1. It is important to note that these roots are not equally spaced, in contrast to most KK models with one extra dimension, due to the non-factorizable metric. Expanding the graviton field into the KK states one finds the interaction
$$=\frac{1}{\overline{M}_{Pl}}T^{\alpha \beta }(x)h_{\alpha \beta }^{(0)}(x)\frac{1}{\mathrm{\Lambda }_\pi }T^{\alpha \beta }(x)\underset{n=1}{\overset{\mathrm{}}{}}h_{\alpha \beta }^{(n)}(x).$$
(6)
Here, $`T^{\alpha \beta }`$ is the stress energy tensor on the brane and we see that the zero mode separates from the sum and couples with the usual 4-dimensional strength, $`\overline{M}_{Pl}^1`$; however, all the massive KK states are only suppressed by $`\mathrm{\Lambda }_\pi ^1`$, where we find that $`\mathrm{\Lambda }_\pi =e^{kr_c\pi }\overline{M}_{Pl}`$, which is of order the weak scale.
This model has essentially 2 free parameters which we can take to be the mass of the first KK graviton mode and the ratio $`c=k/\overline{M}_{Pl}`$; the later quantity is restricted to be less than unity to maintain the self-consistency of the scenario (to prevent a radius of curvature smaller than the Planck scale in 5 dimensions) and if it is taken too small another hierarchy is formed. Figs.7 and 8 show the cross section and $`A_{FB}`$ for the process $`\mu ^+\mu ^{}e^+e^{}`$ as a function of $`\sqrt{s}`$ in the presence of KK graviton resonances for several values of the parameter $`c`$. For large $`c`$ one does not see the individual resonance structures (since the theory is strongly coupled and they are smeared together by their large widths which grow as $`c^2`$) but only a very large shoulder somewhat similar to a contact interaction. For small $`c`$ one sees the individual resonances with their widths growing rapidly with increasing mass as $`m_n^3`$. Note that for large $`\sqrt{s}`$ where graviton exchange dominates the value of $`A_{FB}`$ is driven to zero. Sitting on any of these KK resonances, in the case of small values of $`c`$, will immediately reveal the unique quartic angular distribution corresponding to spin-2 graviton exchange for the fermions in the final state $`13\mathrm{cos}^2\theta +4\mathrm{cos}^4\theta `$.
## Conclusions
Present data indicates that the masses of KK excitations of the SM gauge bosons must be rather heavy, e.g., $`>3.9`$ TeV if $`d=1`$. We have found that:
* With an integrated luminosity of $`100fb^1`$, the LHC will be able to observe KK excitations in the mass range below $`6`$ TeV but may not see any KK excitations when $`d>1`$ since they are likely to be more massive. The LHC will not see the second set of KK resonances even when $`d=1`$.
* The LHC cannot separate the KK states $`\gamma ^{(1)}`$ from $`Z^{(1)}`$ which will appear together as a single resonance, nor can it obtain significant coupling constant information.
* The LHC cannot see the $`g^{(1)}`$ if its mass is in excess of $`4`$ TeV due to its large width and the energy resolution of the LHC detectors.
* The LHC cannot distinguish an extended electroweak model with a degenerate $`Z^{}/W^{}`$ from a KK scenario. All we will know is the mass of these resonances.
* A LC with $`\sqrt{s}=0.51`$ TeV will be sensitive to the existence of KK states with masses more than an order of magnitude larger than $`\sqrt{s}`$ for reasonable integrated luminosities $`100fb^1`$.
* At a LC, the extraction of the couplings of an apparent $`Z^{}`$, whose mass is known from measurements obtained at the LHC, can be performed in a straightforward manner with reasonable integrated luminosities. However, the $`Z^{}`$ hypothesis will yield a poor fit to the data if the state in question is actually the combined $`\gamma ^{(1)}/Z^{(1)}`$ KK excitation. The LC will not be able to identify this state as such–only prove it is not a $`Z^{}`$.
* A Muon Collider operating at or above the first KK resonance pole will identify it as a KK state provided polarized beams are available.
* Measurements of the KK excitation spectrum at Muon Colliders will be able to tell us both the number of extra dimensions and how they are compactified thus possibly revealing the basic underlying theory upon which the KK scenario is based.
* KK excitations of gravitons in the RS model can be studied in detail at both LC and Muon Colliders with Muon Colliders providing a much larger reach in explorable parameter space. These measurements can completely determine all of the parameters of this model.
Muon Colliders clearly offer a very important window into the physics of Kaluza-Klein excitations. |
no-problem/9911/astro-ph9911424.html | ar5iv | text | # 1 Introduction
## 1 Introduction
The asymptotic giant branch (AGB) stars discussed here include those referred to as Mira variables, OH/IR stars and dusty carbon stars. The classical defining characteristics of Mira variables are: large amplitudes, emission line spectra, and periods in excess of about 100 days. The OH/IR and dusty carbon stars are assumed to be similar to the Miras although they are often too faint optically to determine visual amplitudes or measure spectra. Their periods are generally long, up to 1000 days for the carbon stars and up to about 2000 days for the OH/IR stars. A brief review is given below of red variables in globular clusters, prior to a more detailed description of obscured AGB variables in the LMC.
## 2 AGB Variables in Globular Clusters
Globular clusters provide a well studied and well defined environment in which to consider AGB variables. Oxygen-rich Miras are found only in metal-rich clusters where they are the most luminous stars in the clusters. The cluster semi-regular (SR) variables are less luminous than the Miras, but brighter than the non-variable stars. The luminosities of the Miras are greater than that of core helium flash, clearly putting them on the AGB. Broadly speaking stars evolve up the AGB to higher luminosity. But as is now well known, and as discussed elsewhere in these proceedings an AGB star’s surface luminosity changes quite significantly during the course of the He-shell flash cycle. It seems likely therefore that stars enter and exit the Mira phase, possibly more than once, as their luminosity changes during the course of the shell-flash cycle. There are rather few cluster Miras, as would be expected of an evolutionary phase that lasts no more than $`2\times 10^5`$ years (Renzini & Greggio 1990).
In a period-luminosity (PL) diagram we find the low amplitude SR variables at shorter periods than the Miras. They are on an evolutionary track that leads to, and terminates at, the Mira PL relation (Whitelock 1986; Feast 1988). Thus we can understand the Mira period luminosity relation as the locus of the end-points of the evolution of stars with different mass and/or metallicity. Bedding & Zijlstra (1998) have recently suggested that the luminosities of local SR variables (which are presumably younger than the globular cluster SRs), as determined from Hipparcos parallaxes, follow an evolutionary track parallel to that found for globular cluster SR variables, but about 0.8 mag brighter. The SR variables near LMC clusters, studied by Wood & Sebo (1996), appear to follow roughly the same track as the local SRs. At this stage we have little observational evidence for what happens to stars of even higher initial mass which will evolve at higher luminosity. If, as mounting evidence suggests, the more massive stars are affected by hot bottom burning (HBB) it is quite possible that tracks for AGB stars of several solar masses are quite different.
When the AGB variables drop out of the Mira instability strip, during the low luminosity part of the flash cycle, they will become SR variables for some period of time. This type of SR may show a detached dust shell, from the previous high mass-loss Mira phase, and/or abundance anomalies, such as technetium from the dredge-up which accompanies certain thermal pulses. However, by analogy with the globular clusters<sup>1</sup><sup>1</sup>1In cluster HR diagrams SRs occupy the luminosity range between non-variables and Miras; unless this represents an evolutionary sequence there will be an unexplained luminosity gap between the non-variables and the Miras. the majority of oxygen-rich SRs must be in a pre-Mira evolutionary phase and are not expected to show abundance anomalies.
## 3 AGB stars in the Large Magellanic Cloud
In the following discussion we adopt a distance modulus for the LMC of 18.5 mag, as used in most of the papers to which references are made. The best current distance estimates put it slightly further away (e.g. Feast 1999) so the luminosities discussed here are conservative ones. Our knowledge of AGB variables in the LMC up to a few years ago is nicely illustrated in the PL diagram, Fig. 1, taken from Wood et al. (1992) which was originally produced to show the position of the then newly discovered OH/IR stars. The high luminosity stars were thought to be supergiants, with initial masses in excess of about $`8M_{}`$. The fainter stars are on the AGB, and superimposed are model tracks suggesting their approximate initial masses. The stellar luminosities were derived from single observations of large amplitude variables, so a good deal of the scatter is due to variability. As discussed above, we have good reason to believe that, certainly at low masses, stars only become Miras for a short while, so that only part of any evolutionary track is populated by Miras. The picture has changed somewhat in very recent times with the discovery and detailed investigation of more AGB variables with thick shells, as discussed below.
Feast et al. (1989) established PL relations for carbon- and oxygen-rich Miras in the LMC. These were derived from large amplitude variables monitored over their light cycles to derive mean luminosities. At $`K`$ the carbon and oxygen stars obey the same PL relation with a scatter of only 0.13 mag, up to periods of around 420 days. Bolometric luminosities were determined by fitting blackbodies to $`JHK`$ photometry - a reasonable approximation for these stars which have only very thin dust shells. The bolometric PL relations appeared to be different for the oxygen- and carbon-rich Miras, although it was never clear if this was really so, or if it was an artifact of the way the bolometric luminosities were derived. Groenewegen & Whitelock (1996) derived a bolometric PL relation for LMC carbon Miras using all the available data for spectroscopically confirmed C stars, although there were only single epoch observations for many of these. Their PL relation was essentially indistinguishable from that for the O-rich stars. At the time this work was done no LMC carbon Miras were known with periods significantly longer than 500 days, and the optically visible oxygen-rich stars with periods over 420 days lay clearly above the PL relationship. One of these luminous AGB stars, RCG 69 (0523–6644), was among the sample looked at by Smith et al. (1995) for lithium, and is in fact lithium-rich.
Smith et al. (1995) made a survey for lithium among AGB stars in the SMC and LMC, and found that a very large fraction of those with $`6>M_{\mathrm{𝑏𝑜𝑙}}>7`$ were lithium rich, as were a much smaller number of lower luminosity stars. Our present understanding of lithium enhancements in AGB stars (e.g. Sackmann & Boothroyd 1992) is that they occur principally as a result of HBB. Towards the end of AGB evolution, for stars with an initial mass in the range 4 to 6 $`M_{}`$, the base of the convective envelope can dip into the H-burning shell, with far reaching consequences for the evolution of the star. The transition from O- to C-rich is affected; exactly how seems to depend on the model and in particular on how mass loss is treated. HBB may prevent carbon stars forming at all, or prevent them from happening until the envelope mass is depleted (Frost et al. 1998). In some models C stars do form and then HBB turns them back into O-rich stars (Marigo et al. 1999). Another consequence, for stars in a rather narrow mass range, is the formation of lithium via beryllium. One of the most important results in the present context is a rapid increase in luminosity (Blöcker & Schönberner 1991); again the details depend on mass loss which tends to decrease the effect of HBB. In view of this it is interesting to note that most of the stars illustrated in fig. 6 of Smith at al. (1995) lie above the PL relation and many of these must be undergoing HBB, as their lithium abundance indicates. As it is possible for some stars to experience HBB without showing lithium enhancements (e.g. Mazzitelli et al. 1999) we speculate that all of the stars with luminosities above the PL are there because of HBB. Note that the group of stars discussed by Smith et al. is not representative of AGB stars generally. These are the stars for which it is practical to get optical spectra with sufficient resolution and signal-to-noise to measure lithium line strengths, i.e. the brightest ones with the thinnest dust shells. Furthermore, the luminosities of these stars require more detailed investigation; most are based on single observations of large amplitude variables. The periods are not all well defined and it is possible that some of the stars are SR variables and should not properly be included in this discussion.
Wood (1998) discussed the results of a long term investigation of nine obscured IRAS sources in the LMC, providing periods ($`530<\mathrm{P}<1295`$ days) and luminosities for them. Six of these sources are significantly fainter than an extrapolation of the PL relation and he speculated that these were C stars - available evidence on the subject being somewhat contradictory. IR spectroscopy, either from the ground or from ISO (ISOPHOT or ISOCAM), demonstrates that all but one of Wood’s nine sources are indeed C stars, the presence of the $`3\mu \mathrm{m}`$ $`\mathrm{C}_2\mathrm{H}_2+\mathrm{HCN}`$ feature being a clear diagnostic of carbon-rich chemistry (van Loon et al. 1999b or references therein). The one exception was the star with the longest period which Wood et al. (1992) had already shown to be an OH source, and therefore O rich. Wood’s (1998) conclusion, from the discovery of apparently subluminous C stars, was that his observations “$`\mathrm{}`$ clearly demonstrate that once significant mass loss and dust formation occurs, large amplitude LPVs no longer fall on the tight $`M_{bol}/\mathrm{log}P`$ relation found for Mira variables with $`P<450`$ days”.
The nine stars discussed by Wood are also a subgroup of about 50 LMC sources, originally selected from IRAS data, for which a group of us have been obtaining ground-based and ISO data over the last few years (van Loon et al. 1999b and references therein). We have derived periods for them, which can be further refined by combining our data with Wood’s; these do not differ significantly from those measured by Wood. Van Loon et al. (1999b) combined ISOCAM and/or ISOPHOT photometry and spectroscopy with ground based, $`JHKL`$, photometry and fitted models to derive an independent luminosity for these (and other) stars. These results illustrate the luminosity at the time of the ISO observations, i.e. at random phase, whereas Wood (1998) combined IRAS observations, after applying a correction intended to reproduce mean light, with mean near-infrared photometry. A comparison of our estimates of the C-star luminosities with those from Wood is shown in Fig. 2. Three of the stars agree, while we find brighter luminosities for the other five. The OH/IR star is slightly fainter than Wood found. Our observations were made at random phases, and a rough check suggests that two are close to mean light, two fainter and four brighter than mean light. The reason for the apparent systematic difference between the van Loon at al. and the Wood luminosities is not immediately obvious and requires more detailed investigation, but in view of Fig. 2 and the discussion below it seems premature to conclude that sources with thick dust shells fall systematically below the PL relation.
The analysis of our complete sample of LMC stars is at a preliminary stage so the details of the following discussion may change. Combining preliminary periods with luminosities derived by van Loon et al. (1999b), in the same way as those discussed for the Wood (1998) sources, we obtain the PL data illustrated in Fig. 3. Several of the long period O-rich stars are also OH sources (Wood et al. 1992). Extrapolations of the Feast et al. (1989) PL relations are shown; note that the PL relation determined by Groenewegen & Whitelock (1996) for C stars extrapolates between the two lines illustrated. In interpreting Fig. 3 it is crucial to remember that the luminosities shown here are one-off measurements of large amplitude variables, because we have, in general, only single epoch ISO observations. It is therefore obvious that a good deal of the scatter in this diagram is due to variability and multiple observations of three stars give some impression of this. It is also clear that these data scatter around the extrapolated O-rich PL relation, with both C- and O-rich stars falling close to the same line.
The best sampled light curves for the LMC carbon stars have data spanning 5 to 6 years and show peak-to-peak amplitudes of $`\mathrm{\Delta }K2`$ mag. Several of the C stars exhibit apparently secular variations on top of the regular pulsations, as do many galactic C-stars with high mass-loss rates (Whitelock et al. 1997), probably the results of fluctuations in the mass-loss rate. It is difficult to establish the bolometric amplitudes of these stars as few have been monitored at wavelengths longer than $`3\mu \mathrm{m}`$ where most of the energy is emitted. Nevertheless, the discussion by van Loon at al. (1998) suggests that the C stars may have $`\mathrm{\Delta }m_{bol}1.0`$, while the O-rich stars with $`P>1000`$ days will have $`1.0<\mathrm{\Delta }m_{bol}<1.5`$ mag.
## 4 HBB in Individual Stars
The four separate estimates of the luminosity of IRAS 04496–6958 are shown as connected asterisks in Fig. 3 (van Loon et al. 1998, 1999a, 1999b). The $`3\mu \mathrm{m}`$ spectrum of this star shows a clear $`C_2H_2+HCN`$ absorption feature, indicating it is a C star. Its $`10\mu \mathrm{m}`$ spectrum (Fig. 4), however, is very unusual in that it shows both silicate and silicon carbide features, suggesting a mixed O- and C-rich chemistry (Trams et al. 1999). This is the first example of an extragalactic carbon star with silicate emission. There are several galactic stars which show similar combination features; they are generally understood to be binary systems in which one component is a carbon star and the silicate dust resides in a circum-binary disk (Lloyd Evans 1990). It may be that the same explanation applies here, although in view of its high luminosity, Trams et al. (1999) offered an alternative explanation – that IRAS 04496–6958 was, until recently, a star undergoing HBB, hence the silicate dust. It would thus be an example of a star in which HBB terminated, perhaps due to mass-loss, and which then underwent a thermal pulse and dredge-up turning it into a C star.
The solitary asterisk in Fig. 3 represents a C star which Smith et al. (1995) found to be lithium rich, SHV F4488. The luminosity used here, $`M_{bol}=6.3`$, is considerably brighter than that quoted by Smith et al., $`M_{bol}=5.7`$, which was taken from Hughes & Wood (1990). The difference could be due to variability or possibly Hughes & Wood, who had $`JHK`$ photometry from only one epoch, underestimated the flux. In any case the enhanced lithium tells us that the star is undergoing HBB.
IRAS 04496–6958 and SHV F4488 are more luminous than the other C stars shown in Fig. 3. They lie in the same region of the PL diagram, above the extrapolation of the Mira relation, as do the luminous O-rich AGB stars discussed by Feast et al. (1989) and as do almost all of the lithium-rich stars discussed by Smith et al. (1995).
Finally, it should be noted that one of the stars marked as a short period supergiant in Fig. 1, HV 2572, is among the Smith et al. (1995) lithium-rich sample and must therefore be an AGB star undergoing HBB, not a supergiant.
## 5 Conclusions
Observations of large amplitude variables in the LMC show the following:
1. those which are close to the end of their AGB lifetimes fall close to the Mira PL relation;
2. C- and O-rich Miras obey the same PL as well as we can currently establish;
3. many, perhaps all, of the AGB variables which lie above the PL relation are undergoing HBB.
There is a caveat on item 2 above: we cannot as yet eliminate the possibility that there are stars which lie below the PL, as limitations in the sensitivity of our surveys may have prevented us from detecting them as yet. In this regard we should look at the Galactic Centre where there is evidence for long-period large-amplitude variables with low luminosities (Blommaert et al. 1998; Wood et al. 1998). Finally, if we really want to know the luminosities of these large-amplitude variables we need to monitor them at longer wavelengths than has been done to date.
## Acknowledgements
We are grateful to our colleagues, particularly Jacco van Loon and Albert Zijlstra, for allowing us to discuss data in advance of publication. We also thank Jacco van Loon and John Menzies for a critical reading of a draft of this manuscript. |
no-problem/9911/astro-ph9911464.html | ar5iv | text | # Gravo-thermal properties and formation of elliptical galaxies. Based on observations collected at the Canada France Hawaii Telescope and at the European Southern Observatory, La Silla, Chile
## 1 Introduction
Elliptical galaxies present a striking regularity in their global luminosity distributions. Within a wide range of sizes, the light profile of elliptical galaxies can be described by a non-homologous generalization of the de Vaucouleurs $`R^{1/4}`$ profile, the Sérsic law (e.g. Caon et al. 1993; Graham & Colless 1997; Prugniel & Simien 1997).
This regularity may be understood in terms of a relaxation process: elliptical galaxies seem to be in a quasi-equilibrium state, implying that they should obey the virial theorem. From the second law of thermodynamics, a dynamical system in equilibrium is in a maximum entropy state. Due to their peculiar properties (long range unshielded interactions, equivalence of inertial and gravitational mass, etc.) the thermodynamics of gravitational systems present some difficulties, as well explained in academic books (e.g. Saslaw 1985).
For these systems an equilibrium state is never really reached. Various dynamical time scales can be defined: the natural dynamical time $`t_\mathrm{d}1/\sqrt{4\pi G\rho }`$ (where $`\rho `$ is the mean density of the system), the violent relaxation time scale, $`t_{\mathrm{VR}}`$, with $`t_{\mathrm{VR}}t_\mathrm{d}`$, related to the phase mixing process which leads to a quasi-equilibrium state, and a large secular time scale $`t_{\mathrm{sec}}`$, which is related to the slow effects of two-body gravitational interactions. It is essentially on this scale that one can assert that the equilibrium of a self-gravitating system is never established.
For elliptical galaxies, we have $`t_{\mathrm{VR}}\epsilon t_{\mathrm{sec}}`$ with $`\epsilon (N/\mathrm{log}N)^110^8`$ (where $`N`$ is the number of particles). Therefore even if the entropy $`S`$ of a galaxy is ever growing on the secular time scale, after violent relaxation we have $`\mathrm{d}S/\mathrm{d}t_{\mathrm{VR}}\epsilon `$. Stating that the system is in a quasi-equilibrium stage is equivalent to saying that the entropy is quasi-constant.
However, maximizing the entropy results in an isothermal sphere (Lynden-Bell 1967) which is not valuable either from the point of view of physics (divergent total mass) or from observations (observed density profiles are steeper than the isothermal profile; see also White & Narayan 1987).
It is important to note that although there are no exact stationary entropy states for self-gravitating systems (that is, no absolute maximum entropy states), lowest energy states may exist, as suggested by Wiechen et al. (1988). In order to reach such equilibrium states, the system must necessarily undergo a violent relaxation phase, be it through a collapse or a merger. However the final configuration reached depends, in principle, on how strong the violent relaxation phase was. This raises the interesting question of how these equilibria, based on minimum energy, would relate to the final entropy of the system.
Numerous works have been devoted to the entropy problem (see for instance Merritt 1999). In a previous paper (Lima Neto et al. 1999a, hereafter LGM), a different approach has been adopted: instead of trying to obtain the final expected configuration by maximizing the entropy, LGM admit the existence of a state of quasi-constant entropy and calculate this entropy by deriving it from the observed light (mass) distribution. In order to compare objects of different masses, LGM introduced the specific entropy $`s=S/M`$, that is the entropy normalized by the mass. The specific entropy was then calculated by assuming that the stars obey the equations of state of an ideal gas and using the standard thermodynamical definition of the entropy. LGM showed that the galaxies of two clusters and a group had the same value of $`s`$ and, therefore, that one could derive relative distances between these clusters using the Sérsic profile to model the light distribution. LGM suggested that the galaxies having an unique $`s`$ could explain distance indicators based on the shape of the brightness profile of galaxies, like those proposed by Young & Currie (1994, 1995).
As in LGM, we will describe the light distribution of an elliptical galaxy using a Sérsic profile:
$$\mathrm{\Sigma }(R)=\mathrm{\Sigma }_0\mathrm{exp}((R/a)^\nu )$$
(1)
characterized by three primary parameters: $`\nu `$, the shape parameter (independent of cluster distance), $`a`$, the scale parameter (distance dependent, in arcsec), and $`\mathrm{\Sigma }_0`$, the intensity parameter (in erg s<sup>-1</sup> arcsec<sup>-2</sup>).
In contrast with LGM, who used the thermodynamical definition of the entropy, we will adopt here the microscopic Boltzmann-Gibbs definition, therefore eliminating the assumption based on the equations of state of an ideal gas. Assuming that elliptical galaxies are well described by the Sérsic law, we have derived the specific entropy (see details in Appendix I):
$$s(a,\nu ,\mathrm{\Sigma }_0)=0.5\mathrm{ln}(\mathrm{\Sigma }_0)+2.5\mathrm{ln}(a)+F(\nu ),$$
(2)
with:
$$F(\nu )+0.2\mathrm{ln}(\nu )\frac{1.3}{\nu }+3.9\nu ^{1.3}2.7.$$
(3)
Should the specific entropy of galaxies, $`s(a,\nu ,\mathrm{\Sigma }_0)`$, be a constant or, at least, display a small dispersion around its mean value, then Equation (2) would define a thin surface in the parameter space \[$`\mathrm{\Sigma }_0,a,\nu `$\], or a plane in the space \[$`\mathrm{ln}(\mathrm{\Sigma }_0),\mathrm{ln}(a),F(\nu )`$\]. The results presented in LGM suggest that this is indeed the case.
In the next section we describe the data used in this paper, i.e. the surface brightness of cluster galaxies as well as that of simulated galaxies. We also discuss the fitting techniques used to derive the Sérsic profile parameters appearing in Equation (2); in Sections 3.1 and 3.2 we look for correlations between these parameters; in Section 3.3 we show that besides the uniqueness of the specific entropy of galaxies, another relation is also observed. This question is revisited in the context of the cosmological scenario of hierarchical merging galaxy formation in Section 4. We argue that the observed variations of the specific entropy of galaxies are correlated with their total luminosity (or mass). We then show in Section 5, how this correlation helps to understand the tilt of the Entropic Line defined in LGM and therefore to further refine the profile-shape distance indicator of galaxies based on the shape parameter. We discuss our results in the last section.
## 2 The data and fitting methods
Our goal is to fit the surface brightness of elliptical galaxies with two or three parameters, to search for correlations between these parameters and to look for underlying physical properties.
### 2.1 The data
#### 2.1.1 Real galaxies
We have used data on galaxies belonging to three clusters: Coma, Abell 85 and Abell 496. These galaxies were selected: 1) visually as having an elliptical shape on our CCD images, and 2) spectroscopically as having a redshift within the corresponding cluster range. The photometric data are described in Lobo et al. (1997), Slezak et al. (1998) and Slezak et al. (1999), and the spectroscopic data in Biviano et al. (1995), Durret et al. (1998) and Durret et al. (1999) for these three clusters respectively. We have determined the growth curve of each galaxy using the ellipse task of iraf<sup>1</sup><sup>1</sup>1iraf is the Image Analysis and Reduction Facility made available to the astronomical community by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy (AURA), Inc., under contract with the U.S. National Science Foundation.. The growth curves were determined with and without background subtraction.
The first two of these clusters have already been analyzed for the same purpose in a previous paper (LGM), but with circular apertures. We present here a new analysis of these two clusters, together with a third cluster (Abell 496), based on elliptical fits adapted to the geometry of each galaxy.
#### 2.1.2 Simulated (virtual) galaxies
We have used the merger remnants described in Capelato et al. (1995, 1997). There are three merger generations: (1) the end-products of merging King spheres (with varying impact parameter, relative energy and angular momentum), (2) mergers between first-generation mergers, and (3) mergers between second-generation and between first and second-generation mergers.
### 2.2 The fitting method
#### 2.2.1 Real galaxies
For each galaxy in our cluster sample, we have obtained the growth curve (integrated luminosity within elliptical regions of area $`ϵ\pi A\times B`$, where $`A`$ and $`B`$ are the semi-major and semi-minor axes).
The background contribution was determined individually for each galaxy by fitting the last four points of the growth curve as a function of the surface by a straight line. We checked the robustness of this result by also fitting the last 5, 3 and 2 points of the growth curve.
After determining the background contribution, we have subtracted it from the growth curve. Then we determined the total luminosity, $`L_{\mathrm{tot}}`$, using the last points of the growth curve, the half-luminosity (or effective) radius, $`R_{\mathrm{eff}}`$, and the radius containing 99% of the total luminosity, $`R_{99}`$.
We have then fit the growth curves (corrected for the sky) using the integrated form of the Sérsic law:
$$L(R)=\frac{2\pi a^2}{\nu }\mathrm{\Sigma }_0\gamma (\frac{2}{\nu },\left(\frac{R}{a}\right)^\nu )$$
(4)
where the value of $`R`$ is not a radius but an equivalent radius, $`R=\sqrt{AB}`$, and $`\gamma (c,x)`$ is the standard incomplete Gamma function. The luminosity growth curve fits were done with a standard least square minimization method, using the ‘minuit’ programme from the CERN software library.
In order to avoid effects due to the seeing, we have used only data points from 2.0 arcsec outwards (the seeing was FWHM $`0.9`$ arcsec for Coma and 1.2 arcsec for A85 and A496). The fits were done using data points up to $`R_{99}`$ so that for all galaxies the same amount of light was used for the fits. The results of these fits are given in Tables 3, 4, and 5.
#### 2.2.2 Simulated galaxies
For simulated galaxies we have computed a mean radial mass growth curve (that is, integrated mass instead of integrated luminosity), as follows. Each galaxy was randomly projected in a plane and a growth curve was computed for each projection by simply counting the particles inside iso-density ellipses. These projected growth curves (500 for each galaxy) were then used to compute a mean one. Having determined a growth curve for each simulated galaxy, we have proceeded in the same way as for real galaxies, fitting it to the integrated Sérsic profile.
The effective radius $`R_{\mathrm{eff}}`$ (i.e. the projected radius containing half of the total light) is given by LGM:
$`R_{\mathrm{eff}}`$ $`=`$ $`aR_{\mathrm{eff}}^{};`$
$`\mathrm{ln}(R_{\mathrm{eff}}^{})`$ $`=`$ $`{\displaystyle \frac{0.703480.99625\mathrm{ln}\nu }{\nu }}0.18722.`$ (5)
We notice that the simulated galaxy profiles we obtained were extremely close to Sérsic profiles. This agrees with the results obtained by Capelato et al. (1995) using a different fitting technique. In Fig. 1 we compare the effective radius of simulated galaxies as estimated from relation (5) with that directly measured on the simulations. As it can be seen there is a good agreement between these two quantities.
## 3 Relations and correlations
The last columns of Tables 3, 4, and 5 give the values of the specific entropy, $`s_0`$, for the galaxies of our cluster sample. Notice that we cannot compare directly the values of $`s_0`$ since they depend on $`a`$, which is distance dependent (we use the apparent $`a`$ given in arcsec). Neither can we compare the values of $`s_0`$ for a real cluster and for the simulation, since they have different units for $`a`$ and $`\mathrm{\Sigma }_0`$. However, we can compare different generations of mergers, as well as the relative dispersion around the mean value of $`s_0`$ for all real and simulated galaxies.
Table 1 gives the mean values and dispersions of the specific entropy for the whole sample of galaxies. As expected, the mean specific entropy varies from cluster to cluster, reflecting the different distance of each cluster (cf. LGM). For each real cluster, the dispersion is about 10% around the mean value.
For virtual galaxies, the specific entropy seems to increase with the hierarchy, but with a much smaller dispersion, around 5% of their mean values within each generation. This small dispersion is reminiscent of the results discussed by Capelato et al. (1997), which show that the scatter of the Fundamental Plane defined by virtual galaxies is smaller than for the observed ones by a factor of about 2. The increasing of $`s_0`$ with the hierarchy of the merger will be addressed in Section 4.
It is the relatively small scatter of the specific entropy of galaxies around their mean values which justifies the results discussed in LGM, leading to the definition of a mean specific entropy plane, defined through Equation (2). We will call it the Entropic Plane. However, as we will see in the following, there is another relation linking the observed quantities of galaxies.
### 3.1 Correlations with real galaxies
The correlations of the Sérsic profile parameters taken two by two are displayed in Figures 2, 3 and 4. Notice that we used the “astronomical like” quantity $`2.5\mathrm{log}\mathrm{\Sigma }_0`$ instead of $`\mathrm{\Sigma }_0`$.
These correlations all have the same general aspect as those presented in LGM: the three parameters appear well correlated two by two. This strongly suggests that galaxies are not distributed randomly around their mean Entropic Plane but, instead, are distributed along a thin curve in this plane.
### 3.2 Correlations with virtual galaxies
A similar analysis applied to virtual galaxies is shown in Figure 5. The correlations of the Sérsic parameters taken two by two are similar to those found for real galaxies. However, the scatter for virtual galaxies is much smaller.
### 3.3 Another relation?
The results discussed above suggest that we may go one step further: since the projections of the galaxies belonging to the Entropic Plane are simultaneously three thin curves, the galaxies must in fact be located on a thin curve in the Entropic Plane.
In order to check this hypothesis, we will look at the Entropic Plane edge-on and face-on. This requires a rotation to a new coordinate system \[$`\xi ,\eta ,\zeta `$\], defined by:
$`\xi `$ $`=`$ $`[\mathrm{ln}(a)5\mathrm{ln}(\mathrm{\Sigma }_0)]/\sqrt{26}`$
$`\eta `$ $`=`$ $`[5\mathrm{ln}(a)\mathrm{ln}(\mathrm{\Sigma }_0)+13F(\nu )]/\sqrt{195}`$
$`\zeta `$ $`=`$ $`[5\mathrm{ln}(a)+\mathrm{ln}(\mathrm{\Sigma }_0)+2F(\nu )]/\sqrt{30}.`$ (6)
We apply this rotation to each galaxy and show the result in Fig. 6 for virtual galaxies. This figure suggests several comments:
* The value of the specific entropy depends only on $`\zeta `$, $`s=\sqrt{7.5}\zeta `$. Indeed, as a first approximation, virtual galaxies all have the same specific entropy, its numerical value depending on the choice of units. Deviations from this constant value will be addressed in the next section.
* The galaxies are all effectively located on a curve (in fact very nearly a straight line):
$$L(\xi ,\eta )L(\mathrm{\Sigma }_0,a,\nu )=0.$$
Applying the same rotation to the real galaxies data gives quite similar results as for the virtual ones, as seen in Fig. 7. However:
* The scatter is much larger;
* It is not clear whether the relation $`L(\xi ,\eta )=0`$ found for virtual galaxies can be approximated by a straight line as before.
The curves displayed in Figs. 25 were obtained by assuming that the relation $`L(\xi ,\eta )=0`$ may be approximated by a straight line. Their derivations are given in Apppendix II.
At present we do not know of any physical explanation for this relation. In order to have a curve in a three-dimensional space, we need two surfaces to intersect. If one of them is the Entropic Plane, then the other one must be derived from an independent relation, for instance, a scaling law relating the gravitational potential energy to the total mass, or the depth of the potential well to the mean potential (as also suggested by Lima Neto et al. 1999b). Such an intersection of surfaces is actually a geometrical interpretation of the scaling relations that govern the physics of a given system. This kind of problem is similar, for example, to the one encountered when dealing with the origin of the Tully-Fisher relation (e.g. Mo et al. 1998) or of the luminosity–temperature relation for X-ray clusters of galaxies (e.g. Markevitch 1998).
## 4 Is the specific entropy really unique?
### 4.1 The merging scenario
As explained above, the virtual elliptical galaxies we are using come from the merging of successive generations. In such merging processes the energies and masses of the progenitors will be redistributed in order to generate another elliptical. In this section we analyze the effect of merging on the value of the specific entropy.
When viewing a movie displaying the simulation of a galaxy merger as compared to one displaying a cold self-gravitational collisionless collapse, one is struck by the much more violent matter motions (which are actually the engine for the violent relaxation process) occurring in the collapse simulation. Furthermore, what one observes in such movies is, in a sense, quite similar to an observation of the real universe, that is, a macroscopic observation. This is in contrast with the microscopic description one may obtain from the knowledge of detailed evolution of the phase space positions of each particle provided by the simulation. In other words, the relevant description of the system is done with a coarse-grained distribution function.
In any case, the mixing in phase-space that occurs during violent relaxation is responsible for the increase of the coarse-grained entropy of a dynamical system (Tremaine et al. 1986, Merritt 1999). However, it is not clear how much the entropy or, perhaps more important, the specific entropy, increases during the violent relaxation phase. Does the entropy increase depend on the amplitude of the time-varying potential induced by the violent matter motions occurring during this phase? It could be that the increase of entropy is small compared to the total mass of the system (note that defining the entropy as in Eq. (10), $`S`$ has the unit of a mass). In this case, the specific entropy increment could be insignificant when compared with the change of other quantities like, e.g., the total gravitational energy or the total mass of the system (as indeed will be shown below).
### 4.2 Shift of the specific entropy
A careful inspection of the specific entropy (which is given by $`\sqrt{7.5}\zeta `$) both for virtual and real galaxies indeed shows that it does vary, although slightly. We have zoomed the left panel of Fig. 6 and show the result in Fig. 8: an overall increase of the specific entropy is observed ($`\zeta `$ is not exactly a constant). Three vertical lines have been drawn, corresponding to the mean entropy of the three successive generations of galaxies. The specific entropy is actually different for each generation of galaxies, and seems to increase by quanta of specific entropy from one generation to the next, although the jump of specific entropy is quite small (between 10 and 20 %, see also Table 1).
One obvious difference between galaxies of different generations is the increase of their total mass. In fact, even in a given generation there are galaxies with different masses because of slightly different initial orbital parameters of the progenitors. Therefore, we have plotted the specific entropy as a function of the total light (mass), i.e., the integrated luminosity (mass) given by relation (4) extrapolated to infinity:
$$L_{\mathrm{tot}}=\frac{2\pi a^2}{\nu }\mathrm{\Sigma }_0\mathrm{\Gamma }(\frac{2}{\nu })$$
(7)
The corresponding Fig. 9 shows that the total mass allows to discriminate clearly the entropy of the three generations of galaxies. Notice that it is not the mass by itself which is really responsible for the shift of the specific entropy, but rather the merging process.
Other parameters do not allow such a clear discrimination between generations. For instance we have plotted the specific entropy as a function of the $`\nu `$ parameter for virtual galaxies (Fig. 10). For a given value of $`\nu `$, several values of the specific entropy are possible, implying that the parameter $`\nu `$ is not a good discriminant between generations.
In Fig. 11 we show a plot similar to Fig. 8, but now using the real galaxies data; the same overall trend of an increase of the specific entropy ($`\zeta `$) with $`\eta `$ seems to occur with real galaxies. The striking similarities of these two figures further reinforces the hypothesis that elliptical galaxies have been formed by mergers.
Although we do not know to which generation each real galaxy belongs, we may push the analogy between real and virtual galaxies a little further by searching for correlations between the entropy and luminosity. The corresponding plots are shown in Fig. 12. Again, there is a correlation between $`s_0`$ and $`L_{\mathrm{tot}}`$ for real galaxies, although with lower signal-to-noise ratio. We now investigate the implications of such a correlation in the context of the Entropic Line defined by LGM.
## 5 The Entropic Line revisited
LGM have proposed to rewrite Equation (2) as:
$`Y(\mathrm{\Sigma }_0,a)`$ $`=`$ $`0.5\mathrm{ln}(\mathrm{\Sigma }_0)+2.5\mathrm{ln}(a);`$
$`X(\nu )`$ $`=`$ $`F(\nu );`$
$`Y+X`$ $`=`$ $`s_0.`$ (8)
Equation (8), relating $`Y`$ as a function of $`X`$, is the equation of a straight line with a slope $`1`$, called the Entropic Line in LGM, which is, in fact, the Entropic Plane seen edge-on; in this equation $`s_0`$ is the mean value of the specific entropy of galaxies in a given cluster (as in Table 1). Remember that $`s_0`$ is a distance dependent quantity and thus that values for different clusters are not directly comparable.
In their paper, LGM compared the observed data to the predictions of Eq. (8), showing that there is a small difference between the corresponding slopes (see Table 2, col. 3). The fact that the Entropic Line obtained with the observed data had a slope different from $`1`$ was referred to as the tilt of the Entropic Line. However, as we have seen in the last Section, the specific entropy varies with total mass (or light) and, as we will see below, this may be at the origin of the tilt.
We assume the specific entropy $`s_0`$ to be a function of the total luminosity, $`L_{\mathrm{tot}}`$, as given by Eq. 7:
$$s_0=\alpha \mathrm{ln}(L_{\mathrm{tot}})+s_{0,0}$$
the slope $`\alpha `$ being obtained through an ordinary least square mean fitting of the data (shown as lines in Fig. 12). The intercept $`s_{0,0}`$ may be understood as a corrected specific entropy, that is, taking into account the correlation with $`L_{\mathrm{tot}}`$.
A “corrected” equation can then be written in place of equation (8) as:
$`X_{cor}(\nu )`$ $`=`$ $`F(\nu )\alpha \mathrm{ln}(L_{\mathrm{tot}})`$
$`Y+X_{cor}`$ $`=`$ $`s_{0,0}`$ (9)
In Figs. 13, 14 and 15 we display both the original LGM Entropic Line and the corrected Entropic Line, together with their corresponding residuals, for the clusters of our sample. Notice that the slope of the corrected Entropic Line is close to $`1`$, indicating that the tilt between the data and the predictions based on an unique specific entropy for galaxies, has diminished (See Table 2, col. 4). Moreover, the dispersion around the corrected Entropic Line is improved compared to the dispersion around the uncorrected one (compare cols. 5 and 6 in Table 2).
## 6 Discussion and conclusions
We have shown in this paper that the observational data for galaxies belonging to clusters confirm the hypothesis that their specific entropy is, to first order, unique. As a consequence, the galaxies tend to stay in a thin plane (the Entropic Plane) in the space of the Sérsic light profile parameters. Moreover, we have also shown that the slight observed variations of the specific entropy of galaxies are correlated to their total luminosities. Henceforth, by taking this correlation into account, we were able to apply a correction which resulted in decreasing the scatter of the Entropic Plane.
We have also shown that, besides the Entropic Plane, another relation must exist between the Sérsic profile parameters of galaxies. The intersection of this relationship with the Entropic Plane defines a curve in the 3D Sérsic parameter space which appears very well defined in our data. The physical origin of this new relation is unknown, although it may be related to the gravitational energy of galaxies, as suggested by Lima Neto et al. (1999b).
The existence of this curve explains the tight correlations between the Sérsic parameters, as noted in LGM. It constitutes the theoretical background for the photometrical distance indicator proposed by various authors (Young & Currie 1994, 1995; Binggeli & Jerjen 1998).
Our study is solely based on quantities extracted from photometric data. The dynamics are in fact hidden in the shape parameter of the fitting Sérsic profiles. Usually, both photometric and spectroscopic studies are performed on galaxies, leading for instance to the “Fundamental Plane”. It is interesting to notice that some authors have shown that globular clusters (Bellazzini 1998), galaxies (Guzmán et al. 1993) or even galaxy clusters (Fujita & Takahara 1999) are located on a line in that plane instead of populating the whole area a priori permitted by the natural range of variation of the parameters. The question of how to derive the “Fundamental Plane” (photometry plus spectroscopy) from the Entropic Plane (photometry alone) will be addressed in a forthcoming paper.
Elliptical galaxies are well described by a three parameter profile; since two constraints are acting, only one parameter is free. Because the total luminosity of galaxies correlates with the shape parameter (Prugniel & Simien 1997, Binggeli & Jerjen 1998), their light profile should be completely defined by their total luminosity. Shape and luminosity are probably correlated as a consequence of the relaxation process; this is consistent with the fact that high luminosity galaxies have flat profiles while those of dwarfs are peaked.
We must stress that the calculation of the specific entropy requires a good choice for the density profile. Indeed, besides an intensity and a length scale, which are the basic parameters of all the laws used (for instance the de Vaucouleurs law), structural parameters are also necessary in order to account for the shapes, which reflect the dynamical properties of the galaxies.
In this context, we found that the choice of the Sérsic law is particularly suitable not only for the sake of simplicity, but also because it is the bulk of the light which is affected by the variations of the shape factor, as pointed by Graham & Colless (1997). One can find in the literature other profiles depending on a shape factor, for instance the $`\beta `$-model. However, due to the asymptotic behaviour of this model, the entropy and mass calculated in this case are essentially located at large radial distances. Besides, the values of these two quantities are very sensitive to any computational cut-off, and, moreover, the observational data in the outskirts of clusters are generally too poor to guide the calculation. Therefore, the $`\beta `$-model is not suitable for such calculations.
We do not claim that the Sérsic profile is the ultimate profile of relaxed systems, any more than for instance the de Vaucouleurs profile. However, while the precise analytical expression for a profile may not be fundamental, we understand that taking into account the shape parameters is indeed important. It would then be interesting to see if “Universal” profiles – as for instance the NFW (Navarro et al. 1995) or the Hernquist (1990) profiles – also have shape parameters, allowing the calculation of the specific entropy.
###### Acknowledgements.
This work was partially supported by FAPESP, CNPq, PRONEX-246, USP/COFECUB, CNPq/CNRS Bilateral Cooperation Agreements, and DGCyT grant PB93-0139. We also acknowledge financial support from the French Programme National de Cosmologie, CNRS.
## Appendix I: Entropy and the Sérsic profile
We use the microscopic Boltzmann-Gibbs definition of the specific entropy,
$$sS/M=\frac{1}{M}f(ϵ)\mathrm{ln}f(ϵ)drdv$$
(10)
where $`M`$ is the total mass and we assume that the function $`f`$ (the coarse-grained distribution function) depends only on the energy $`ϵ`$. Then, $`f(ϵ)`$ can be determined by the Abel inversion of the density profile (cf. Binney & Tremaine 1987) and the specific entropy may be computed (see formulae below).
In order to compute $`s`$ we have adopted the following hypotheses:
* Spherical symmetry;
* Isotropy of the velocity distribution;
* $`M/L(r)=`$ constant.
From the 2D mass distribution given by the Sérsic profile (using the hypothesis of constant $`M/L`$ ratio), we have derived a semi-analytical approximation for the 3D mass distribution obtained by deprojecting the Sérsic profile (see LGM):
$$\rho (r)=\rho _0\left(\frac{r}{a}\right)^p\mathrm{exp}([r/a]^\nu );$$
$$p=1.00.6097\nu +0.05563\nu ^2.$$
From the 3D mass distribution we can compute the distribution function and thus the specific entropy. The computation can be done numerically but it is cumbersome. We have therefore found analytical approximations for the specific entropy (see also LGM):
$$s(a,\nu ,\mathrm{\Sigma }_0)=\frac{1}{2}\mathrm{ln}\mathrm{\Sigma }_0+\frac{5}{2}\mathrm{ln}a+F(\nu );$$
$$F(\nu )=0.2\mathrm{ln}(\nu )\frac{1.3}{\nu }+3.9\nu ^{1.3}2.7.$$
If $`s(a,\nu ,\mathrm{\Sigma }_0)=s_0=`$ constant, then the above equations define a surface, the Specific “Entropy Plane” (in the appropriate variables).
Finally, we give for completeness an analytical approximation for the corresponding magnitude for a given set of Sérsic parameters $`(a,\nu ,\mathrm{\Sigma }_0)`$:
$$m=2.5\mathrm{log}L(R\mathrm{})=2.5\mathrm{log}\mathrm{\Sigma }_05\mathrm{log}a+m^{};$$
$$m^{}=0.304\nu 1.708\nu ^{1.44}.$$
## Appendix II: Analytical formulae for the correlations
The projection of the relation $`L(\xi ,\eta )=0`$, introduced in Section 3.3, leads to one dimensional curves in the planes \[$`\mathrm{\Sigma }_0,a`$\], \[$`\mathrm{\Sigma }_0,\nu `$\] and \[$`a,\nu `$\]. We give below semi-theoretical formulæ for these relations:
1. we assume that $`L(\xi ,\eta )=0`$ may be approximated by a straight line:
$$\eta =A\xi +B,$$
(11)
where the constants $`A`$ and $`B`$ are obtained through a fitting of the data;
2. $`\xi `$ and $`\eta `$ depend on $`a`$, $`\nu `$ and $`\mathrm{\Sigma }_0`$ by relations (6);
3. we have postulated that the specific entropy is unique (theoretical aspect) i.e. $`\zeta =s_0/\sqrt{7.5}`$ ($`\zeta `$ also given by Eq. (6)).
Combining (6) and (11) allows us to recover analytical formulae, which are displayed below. We have superimposed these calculated curves on each of the corresponding data in Figs. 2, 3 and 4, for the real galaxies and in Fig. 5 for the simulated ones.
We obtain for the virtual galaxies the formula:
$`a`$ $`=`$ $`\mathrm{exp}[1.771\times (8.2562.5\times \mathrm{log}\mathrm{\Sigma }_0)]`$
$`2.5\mathrm{log}\mathrm{\Sigma }_0`$ $`=`$ $`{\displaystyle \frac{0.328}{\nu }}{\displaystyle \frac{0.9833}{\nu ^{1.3}}}0.0504\mathrm{ln}(\nu )7.5123`$
$`a`$ $`=`$ $`\mathrm{exp}[{\displaystyle \frac{0.5804}{\nu }}{\displaystyle \frac{1.741}{\nu ^{1.3}}}0.089\mathrm{ln}(\nu )+1.317]`$
and successively for Coma:
$`a`$ $`=`$ $`\mathrm{exp}[0.951\times (18.762.5\times \mathrm{log}\mathrm{\Sigma }_0)]`$
$`2.5\mathrm{log}\mathrm{\Sigma }_0`$ $`=`$ $`{\displaystyle \frac{0.6}{\nu }}{\displaystyle \frac{1.798}{\nu ^{1.3}}}0.0922\mathrm{ln}(\nu )+20.42`$
$`a`$ $`=`$ $`\mathrm{exp}[{\displaystyle \frac{0.6304}{\nu }}{\displaystyle \frac{1.89}{\nu ^{1.3}}}0.097\mathrm{ln}(\nu )+1.741]`$
(13)
Abell 85:
$`a`$ $`=`$ $`\mathrm{exp}[0.811\times (19.482.5\times \mathrm{log}\mathrm{\Sigma }_0)]`$
$`2.5\mathrm{log}\mathrm{\Sigma }_0`$ $`=`$ $`{\displaystyle \frac{0.496}{\nu }}{\displaystyle \frac{1.487}{\nu ^{1.3}}}0.0763\mathrm{ln}(\nu )+20.483`$
$`a`$ $`=`$ $`\mathrm{exp}[{\displaystyle \frac{0.611}{\nu }}{\displaystyle \frac{1.834}{\nu ^{1.3}}}0.094\mathrm{ln}(\nu )+1.238]`$
and Abell 496:
$`a`$ $`=`$ $`\mathrm{exp}[1.3131\times (19.802.5\times \mathrm{log}\mathrm{\Sigma }_0)]`$
$`2.5\mathrm{log}\mathrm{\Sigma }_0`$ $`=`$ $`{\displaystyle \frac{0.901}{\nu }}{\displaystyle \frac{2.702}{\nu ^{1.3}}}0.1386\mathrm{ln}(\nu )+21.9`$
$`a`$ $`=`$ $`\mathrm{exp}[{\displaystyle \frac{0.686}{\nu }}{\displaystyle \frac{2.0577}{\nu ^{1.3}}}0.105\mathrm{ln}(\nu )+1.6255]`$
## Appendix III: Growth curve fitting results |
no-problem/9911/astro-ph9911416.html | ar5iv | text | # Properties of the Non-Thermal Emission in Plerions
## 1. Introduction
Not all supernova remnants (SNRs) show the “classical” shell-like structure. Some of them present instead a filled-center structure: the prototype is the Crab Nebula, but other objects with similar properties have been discovered since then. They are called filled-center SNRs, or Crab-like SNRs, or plerions (Weiler & Panagia 1978). Plerions are recognized not just on the basis of their morphology, but also of other properties, like: a flat power-law radio spectrum, with a spectral index ranging from 0.0 to $``$0.3; a high radio polarization, with a well organized pattern (not true for all plerions); a power-law X-ray spectrum, with photon index close to $``$2 (Asaoka & Koyama 1990); the detection of an associated pulsar (not true for all plerions — see Pacini, these Proceedings).
Although the details of the nature and structure of plerions are still unclear, there is a common agreement on the following points: a plerion is an expanding bubble, formed essentially by magnetic fields and relativistic electrons, and the observed synchrotron emission originates from these two components; a continuous supply of magnetic flux and relativistic particles is required, in order to explain the typical synchrotron emissivities, as well as the high frequency emission (from particles with synchrotron lifetimes shorter than the SNR age).
In a simplified approach, one may assume that magnetic fields and particles are uniformly distributed into the plerionic bubble: this approach (Pacini & Salvati 1973; Reynolds & Chevalier 1984; Bandiera et al. 1984) is usually adequate to explain the evolution of the overall nebular spectrum. However the homogeneity assumption is likely to be incorrect. New particles and magnetic flux are released by the associated pulsar, presumably near the pulsar wind termination shock (Rees & Gunn 1974; Kennel & Coroniti 1984a,b). Therefore the degree of homogeneity of the electrons distribution depends on how efficient are the mechanisms (diffusion, or advection) by which particles propagate through the nebula (see Amato, these Proceedings).
Also the structure of the magnetic field can be rather complex. From considerations on the MHD relations it follows that spherical models cannot account adequately for the field structure: it can have at most a cylindrical symmetry (Begelman & Li 1992), but more complicate patterns are suggested by observations. The comparison of high resolution maps at various frequencies may then give important clues on the structure on plerions, and on the processes governing the evolution of the magnetic field as well as of the particle distribution.
The outline of this paper is the following: I begin by reviewing the classical, simplified approach to the evolution of a plerion and of its emission; then I consider the case of the Crab Nebula, the prototype of this class, as well as of some other plerions with characteristics different from the Crab; finally I describe the results and perspectives of a multifrequency study, with high spatial resolution, of the Crab Nebula and ot and other plerionsher plerions.
## 2. Classical models of the evolution of the synchrotron emission
Classical models for the evolution of the synchrotron emission from a plerion as a whole (Pacini & Salvati 1973; Reynolds & Chevalier 1984; Bandiera et al. 1984) are based on the original analysis of Kardashev (1962). The starting points are the two basic synchrotron equations, that for the radiated power from an electron with energy $`E`$ ($`W_s=c_1B^2E^2`$), and that for the typical frequency of radiation ($`\nu _s=c_2BE^2`$): these formulae are averages over pitch angles, under the assumption of isotropy. If $`N(E)`$ is the present distribution of electrons, and the magnetic field $`B`$ is constant throughout the nebula, it immediately follows that the synchrotron spectrum is given by $`L(\nu )=(c_1/2c_2)BEN(E)`$. This relation establishes the well known connection between the particle distribution and the radiated spectrum (if $`L(\nu )\nu ^\alpha `$, then $`N(E)E^{(1+2\alpha )}`$).
Let us consider, for instance, the synchrotron spectrum of the Crab Nebula. One may identify various spectral regions with different spectral indices: the radio, with $`\alpha 0.3`$; the optical, with $`\alpha 0.8`$; the X rays, with $`\alpha 1.0`$; and a further steepening above 100 keV. All the changes of slope in $`L(\nu )`$ correspond to breaks in $`N(E)`$: the issue is to determine which of them are intrinsic to the injected particle distribution, and which result from the evolution. Usually an original power-law distribution is assumed, with the aim of explaining all breaks as originated just from the evolution.
The basic ingredients are: 1. The evolution of the energy input from the spinning down pulsar (this input typically lasts for a time $`\tau _o`$, and then falls down). 2. The fraction by which this power is shared between injected particles and field (usually assumed to be constant). 3. The expansion law of the nebula, $`R(t)`$, that at earlier times may be linear or even accelerated; but later on, with the passage of the reverse shock coming from the outer blast wave, the plerion may shrink, and then re-expand at a lower rate (Reynolds & Chevalier 1984).
The evolution of the magnetic field in the nebula is modelled by including the effects of the adiabatic losses; while for the evolution of particles both adiabatic and synchrotron losses must be taken into account. A special particle energy is that at which the timescales for adiabatic and synchrotron losses are comparable ($`E_b1/c_1B^2t`$): a break in the spectrum occurs at the frequency that corresponds to $`E_b`$. Before the time $`\tau _o`$ this is the only evolutive break present in the spectrum. Kardashev (1962) showed that the change in the spectral index must be $`\mathrm{\Delta }\alpha =0.5`$: this result can be directly tested on the data. After the time $`\tau _o`$, a second break should appear in the distribution, at the energy $`E_c(t)=E_b(\tau _o)R(\tau _o)/R(t)`$: this is the “fossil break”, namely the adiabatic evolution of the break located at $`E_b`$ at the time $`\tau _o`$. This break is at a frequency smaller than that of the truly evolutive break.
The Crab Nebula fits rather well into the scenario described above. In fact: 1. The classical law for the pulsar spin-down seems to be verified: the total energy released by the pulsar since its birth ($`10^{49}`$ erg) is also consistent with the kinetic energy excess in the optical thermal filaments (due to their dynamical coupling with the plerionic bubble). 2. The efficiency in particle production is reasonably high: the present pulsar spin-down power ($`4.5\times 10^{38}\mathrm{erg}\mathrm{s}^1`$) is comparable with the total synchrotron luminosity ($`0.7\times 10^{38}\mathrm{erg}\mathrm{s}^1`$). 3. The efficiency in magnetic field production is rather high: from the position of the break ($`\nu _b10^{13}`$ Hz) one may derive a nebular magnetic field $`B0.4`$ mG, i.e. a magnetic energy $`2\times 10^{49}`$ erg, close to the total energy released by the pulsar. 4. The measured secular variation in radio ($``$0.17% per yr; Aller & Reynolds 1985) agrees with the theoretical estimate (Véron-Cetty & Woltjer 1991). 5. The change in the spectral slope from radio to optical is $`\mathrm{\Delta }\alpha =0.5`$, in agreement with Kardashev; however the further breaks at higher frequencies cannot be explained in this way, unless breaks in the injected distribution are invoked.
It can be noticed that in the Crab Nebula fields and electrons are in near equipartition. Is this equipartition typical for all plerions? Does equipartition hold at injection, or is it the result of a subsequent field-particle coupling?
## 3. Non Crab-like plerions
There is a bunch of plerions characterized by a spectral break at frequencies much lower than in the Crab Nebula: for instance, in 3C58 and in G21.5$``$0.9 the break is at 50 GHz; while in CTB87 it is at 20 GHz. Woltjer et al. (1997) discussed properties and evolutive implications for these objects. Beyond the low-frequency position of the break, they share other features, like a sharp break, with $`\mathrm{\Delta }\alpha `$ larger than the canonical value 0.5. Furthermore, in none of these objects pulsations from the expected neutron star have been detected yet.
It is hard to explain the observed break as the main evolutive break, since it would imply a very large nebular field. The most extreme case is that of CTB87, an extended plerion for which the estimated magnetic energy is greater than $`6\times 10^{51}`$ erg, well above that of the supernova explosion itself. This paradox can be solved if the break observed is the fossil break: in this case also the theoretical limit on $`\mathrm{\Delta }\alpha `$ across the break can be overcome, under the condition that the pulsar slow-down follows a very steep law (even though some difficulties remain, in modelling the break sharpness). This implies that the associated pulsar has slowed down considerably, and therefore has become much fainter than at the origin: this may be a reason why no pulsations are detectable.
Among these plerions, 3C58 is that posing more problems to models. In this object the sharpness of the break requires an abrupt decrease in the rate of injected particles. Beyond that, models must also account for the measured increase in the radio emission (Green 1987, and references therein). Woltjer et al. (1997) show that, in order to match the observations, a sudden change in the relative efficiencies of field and particles production is required: this could actually be associated to a recent “phase change” in the pulsar magnetosphere.
## 4. Plerionic components in composite SNRs
Composite SNRs are those in which a shell-like component (due to the interaction of the supernova ejecta with the ambient medium) co-exists with a plerionic component. Helfand & Becker (1987) introduced the class of composite SNRs and outlined their properties: they are old enough to have developed a shell component, but still young enough to host a detectable plerionic component.
Studying a composite SNR is more interesting than just studying independently a shell-like and a plerionic remnant, since the two components have the same origin (thus the same age and distance), and are interacting. Slane et al. (1998) discuss what kind of diagnostic tools can be used. For instance, the X-ray spectrum of the thermal shell allows one to evaluate its pressure and therefore, by assuming a (rough) pressure equilibrium with the plerion, the plerionic magnetic field. This is an estimate alternative to that based on the position of the evolutive break, and may then represent a test on the nature of that break.
In the cases in which the associated pulsar has not been detected yet, its most likely parameters may be estimated. If a pulsar has been already detected, its spin-down time may be used to estimate the SNR age, that should agree with the age derived from the thermal X-ray spectrum. Moreover one can derive the present pulsar energy output, and then the efficiency by which this energy is converted into magnetic fields and particles. This analysis has been already carried on for various composite SNRs, like G11.2$``$0.3 (Bandiera et al. 1996), CTA1 (Slane et al. 1997), N157B (Cusumano et al. 1998), MSH11$``$62 (Harrus et al. 1998), G327.1$``$1.1 (Sun et al. 1999), G39.2$``$0.3 (Harrus & Slane 1999).
## 5. The Crab Nebula: results of a multifrequency analysis
The above considerations on the Crab Nebula, as well as those on other plerions, are mainly referring to their global properties. But a much deeper insight should follow from a combined study of the spatial-spectral properties of the synchrotron emission from these objects. I present here some results coming from a comparison of optical and X-ray maps of the Crab Nebula, and a preliminary analysis of millimetric observations of this plerion.
The Crab Nebula is a very bright object, over a wide range of frequencies: therefore it is an ideal target for a detailed multifrequency investigation with high spatial resolution. Our project has been inspired by the study of Véron-Cetty & Woltjer (1993) on the Crab synchrotron emission in the optical range: they produced a map of the optical spectral index, and showed that a spectral steepening occurs going outwards; they also noticed that the region with optical spectral index flatter than $``$0.7 matches the shape of the Crab in X rays.
We decided to carry on a detailed, quantitative comparison between optical and X rays: for this reason we have re-analyzed the Véron-Cetty & Woltjer (1993) optical data, while for the X rays we have used public data of ROSAT HRI. In the X-ray data reduction we have stressed, rather than the resolution, the sensitivity to regions with very low surface brightness. For this purpose we have deconvolved the map in order to eliminate the wings of the instrumental Point Spread Function, as well as the halo due to dust scattering. Details on the data analysis and on their interpretation are given by Bandiera et al. (1998).
As already pointed out by Hester et al. (1995) the outer and faint X-ray emission extends almost to the boundary of the optical nebula. We have then performed a quantitative analysis over a large area, by using $`5\mathrm{"}\times 5\mathrm{"}`$ pixels. Fig. 1 gives a plot of the optical-to-X averaged spectral index ($`\alpha _{OX}`$) versus the optical spectral index ($`\alpha _{Opt}`$), where each dot refers to a single pixel: unfortunately no high resolution spectral map in the X rays is available yet.
In the interpretation of this plot our underlining assumption is that, due to the evolution, the spectral indices of the particle distribution, in all energy ranges, tend to steepen: this corresponds to a softening of the emission spectrum at all frequencies, and translates into the requirement that the time evolution of a given bunch of particles produces a drift of the related dot towards the upper right direction of the plot.
The area of the plot can be then subdivided into zones. Most of the dots are in the region confined by the two lines labelled by $`a`$ and $`b`$: let us call it the “Main Region” of the plot. All the pixels corresponding to the dots in the Main Region are located in the main body of the Crab Nebula, namely that obtained by cutting out the N-W and S-E elongations. The line $`a^{}`$, parallel to $`a`$, indicates the cases of no spectral bending between optical and X rays ($`\alpha _{Opt}=\alpha _{OX}`$). Pixels with $`\alpha _{OX}<1`$ outline rather well the region of the X-ray “torus” (see e.g. Hester at al. 1995): in the plot they are located on the lower left side, in agreement with our expectation that this is the main location where to find freshly injected particles.
We introduced the quantity $`m`$, that gives the position of a dot across the strip bounded by lines $`a`$ and $`b`$: this quantity is related to the spectral bending between optical and X rays. A map of $`m`$ is given in Fig. 2-L: points with lower $`m`$ (namely points with a lower bending; brighter pixels in the map) are generally located in a thick “equatorial” belt. In particular, pixels with a very low bending coincide with some prominent thermal filaments. If the absence of bending is a sign of freshly injected, or re-accelerated, particles, then in these regions some secondary acceleration processes could take place.
Let us go back to the plot of Fig. 1. Following the scheme introduced above, the dots above line $`b`$ (“Secondary Region”) cannot be the result of a continuous evolution from dots originally located in the Main Region. They are more likely to represent the evolution of particles originally emitting a spectrum with $`\alpha _{OX}`$ not smaller that 1.75 (line $`c`$), and that afterwards, while moving outwards, have softened considerably their spectrum. Fig. 2-R gives a map of $`\alpha _{OX}`$ for all the points confined to the two lobes (brighter pixels indicate higher values of $`\alpha _{OX}`$): here a softening of the spectrum in the outer zones is apparent. In the lobes, the zones with harder spectra seem to lie on the prolongation of the X-ray “jets” (see Hester et al. 1995). This is very clear for the S-E lobe, and may be the indication that these particles are directly provided by the jets.
All these considerations are largely empirical. Moreover, some conclusions may be partly biased by projection effects: anyway the general results should remain valid. The first result is the great variety of local spectra in the nebula. This may be partly explained in terms of the evolution of electrons injected at the central torus; however, both in the regions with low optical-to-X spectral bending and in the polar lobes further particle components seem to be present, suggesting the presence of secondary acceleration (or re-acceleration) processes.
A similar result follows from recent observations at 230 GHz, with the IRAM 30 m telescope, in collaboration with R.Cesaroni and R.Neri: a comparison with a map at 1.4 GHz shows variations in the spectral index between the two frequencies. If confirmed, this result is very important. In fact the spectral index of the Crab Nebula in the radio shows very little spatial variations (Bietenholz et al. 1997) and, if there is only one channel of injection, the same behaviour is expected up to the evolutive spectral break ($`10^{13}`$ Hz for the Crab). On the contrary, flatter spectra are found, between 1.4 and 230 GHz, in the central regions of the Crab Nebula, similarly as seen in the optical range. Further observations are however required, in order to confirm this result.
## 6. Conclusions
The comparison of high-resolution maps at different frequencies may provide a wealth of information on the plerions, and will help to clarify some open issues on these objects, like: how, where and by how many different mechanisms particles can be accelerated inside plerions? How do they propagate through the nebula? What is the structure of the magnetic field? How the evolution of a plerion and of the associated neutron star affect the synchrotron emission?
A big step towards the understanding of these objects will hopefully come with the arrival of the new generation X-ray telescopes, that will be able to perform spectral mapping with arcsec resolution. X-ray emitting electrons have very short lifetimes, typically of the order of the light crossing time of the nebula, and therefore from an X-ray spectral mapping one could derive information directly on the sites of the present-time injection. But another promising spectral range for effectively probing the physical conditions in the nebula is the millimetric one. In various plerions a spectral break is located near that range: spectral maps may then give indications on the spatial variations of the break position, and therefore on the magnetic structure of the nebula.
Even in the presence of such high quality observations, one may wonder how powerful the synchrotron emission is as a diagnostic tool. In fact it can provide just a mixed information on particles and magnetic field, and only projected along the line of sight. Therefore even the interpretation of high resolution spectral maps is generally not straightforward, unless theoretical models will be developed at a level of detail similar to that of the observations.
### Acknowledgments.
Many of the results presented here come from discussions and collaborations with various persons: E.Amato, F.Bocchino, R.Cesaroni, R.Neri, F.Pacini, M.Salvati, P.Slane, L.Woltjer. This work is partly supported by the Italian Space Agency (ASI) through grant ARS-98-116.
## References
Aller, H.D., Reynolds, S.P. 1985, ApJ, 293, L73
Asaoka, I., Koyama, K. 1990, PASJ, 42, 625
Bandiera, R., Pacini, F., Salvati, M. 1984, ApJ, 285, 134
Bandiera, R., Pacini, F., Salvati, M. 1996, ApJ, 465, L39
Bandiera, R., Amato, E., Woltjer, L. 1998, Mem.SAIt, 69, 901
Begelman, M.C., Li, Z.-Y. 1982, ApJ, 397, 187
Bietenholz, M.F., Kassim, N., Frail, D.A., Perley, R.A., Erickson, W.C., Hajian, A.R. 1997, ApJ, 490, 291
Cusumano, G., Maccarone, M.C., Mineo, T., Sacco, B., Massaro, E., Bandiera, R., Salvati, M. 1998, A&A, 333, L55
Green, D.A. 1987, MNRAS, 225, L11
Harrus, I.M., Slane, P.O. 1999, ApJ, 516, 811
Harrus, I.M., Hughes, J.P., Slane, P.O. 1998, ApJ, 499, 273
Helfand, D.J., Becker, R.H. 1987, ApJ, 314, 203
Hester, J.J., Scowen, P.A., Sankrit, R., et al. 1995, ApJ, 448, 240
Kardashev, N.S. 1962, Soviet Ast., 6, 217
Kennel, C.F., Coroniti, F.V. 1984a, ApJ, 283, 694
Kennel, C.F., Coroniti, F.V. 1984b, ApJ, 283, 710
Pacini, F., Salvati, M. 1973, ApJ, 186, 249
Rees, M.J., Gunn, J.E. 1974, MNRAS, 167, 1
Reynolds, S.P., Chevalier, R.A. 1984, ApJ, 278, 630
Slane, P., Seward, F.D., Bandiera, R., Torii, K., Tsunemi, H. 1997, ApJ, 485, 221
Slane, P., Bandiera, R., Torii, K. 1998, Mem.SAIt, 69, 945
Sun, M., Wang, Z.-R., Chen, Y. 1999, ApJ, 511, 274
Véron-Cetty, M.P., Woltjer, L. 1991, A&A, 251, L31
Véron-Cetty, M.P., Woltjer, L. 1993, A&A, 270, 370
Weiler, K.W., Panagia, N. 1978, A&A, 70, 419
Woltjer, L., Salvati, M., Pacini, F., Bandiera, R. 1997, A&A, 325, 295 |
no-problem/9911/nucl-th9911050.html | ar5iv | text | # Formal comparison of SUSY in the nuclear U(6/2) model and in quantum field theory
## I Introduction
Supersymmetry is a very interesting subject which is widely discussed in different subfields of physics. Supersymmetry has a long and interesting history. It was discovered in the particle physics, however, the ideas of supersymmetry have found in nuclear structure physics a very extensive application. Several good examples of an approximate dynamical supersymmetry have been found there. The investigations of the dynamical supersymmetries in nuclear structure physics has been initiated by F.Iachello and the results obtained are summarized in .
It was found that the combined algebras are particularly useful in a description of the properties of odd–even nuclei. Moreover, an attempt to get a classification in which several nuclei including even and odd ones are described by the same Hamiltonian has been realised and very interesting results have been obtained
However, the formalism to introduce the supersymmetries used is different in particle physics and in nuclear structure physics. This makes a comparison more difficult and also restricts the application of the examples found in nuclear physics. In nuclear structure theory the main attention has been paid to the construction of the Hamiltonians, which exhibit dynamical supersymmetries. Such Hamiltonians are represented by the sums of the Casimir operators for an appropriate subgroup chain of U(6/$`\mathrm{\Omega }`$), where $`\mathrm{\Omega }`$ is a maximum number of the fermions in a system. Taken with the same parameters the Hamiltonians constructed in this way have been applied to the description of different nuclei including even and odd ones. However, previous works considered those subgroup chains in which already at the first step a classical Lie algebra formed by the operators of the Bose sector of U(6/$`\mathrm{\Omega }`$) appears.
For this reason, the operators of the Fermi sector (in the following called super operators), i.e. the operators transforming bosons into fermions and vice versa , do not appear explicitly in the subgroup chains. The super operators of the basic U(6/$`\mathrm{\Omega }`$) graded algebra commute with a Hamiltonian which is a function of only the Casimir operator of U(6/$`\mathrm{\Omega }`$), which is the operator of the total number of bosons plus fermions. Such a Hamiltonian is too simple to describe real nuclei and it does not contain an interaction. In other words, the super operators analogous to those which are included in a construction of the Lagrangians in a QFT do not appear in this consideration. This indicates the difference in the schemes of the realization of the supersymmetry ideas in nuclear and particle physics. The examples of the dynamical supersymmetries in nuclear structure physics where the super operators have been explicitly introduced and discussed are given in . However, in these examples the structure of supergroups is different from that of QFT.
In this paper we consider a well known example of the supersymmetry from the Interacting Boson–Fermion model, which on one side has a good experimental realization in <sup>102</sup>Ru, <sup>103</sup>Rh and <sup>104</sup>Pd and on the other side has an algebraic structure which is similar to the SUSY model in QFT, however. It is the main aim of the present paper to demonstrate this similarity and consider the consequences of it.
## II U(6/2) nuclear structure model
Below we shall assume that fermions occupy a single particle state with angular momentum $`j`$=1/2 (e.g. a $`p_{1/2}`$ state). Thus, the underlying graded algebra is $`U(6/2)`$.
Before starting the consideration let us introduce a Hamiltonian which has a simpler structure and which can serve as an introduction to the more general case considered below.
The simplest Hamiltonian for a system of the noninteracting monopole $`s`$–bosons and a fermion with a spin equal $`\frac{1}{2}`$ and a good helicity quantum number is
$`H=ϵ(a_{\frac{1}{2}\frac{1}{2}}^+a_{\frac{1}{2}\frac{1}{2}}+s^+s)`$ (1)
This is a standard example found in textbooks , by which many features of the supersymmetries can be illustrated. However, this Hamiltonian does not possess rotational symmetry and is not realized in nuclear physics. A simple extension of this Hamiltonian which possesses rotational symmetry is
$`H=ϵ(a_{\frac{1}{2}\frac{1}{2}}^+a_{\frac{1}{2}\frac{1}{2}}+a_{\frac{1}{2}\frac{1}{2}}^+a_{\frac{1}{2}\frac{1}{2}}+s^+s).`$ (2)
The super operators commuting with the Hamiltonian (2) are $`a_{\frac{1}{2}\frac{1}{2}}^+s`$ and $`a_{\frac{1}{2}\frac{1}{2}}^+s`$. They transform eigenstates of this Hamiltonian into another ones with the same energy by changing the number of fermions in the state.
We go one step further and introduce the Hamiltonian which is used in the nuclear physics example below and includes not only monopole $`s`$–bosons but also quadrupole $`d`$–bosons. We shall consider the case of the so called $`U(5)`$ dynamical symmetry limit, for which the numbers of $`s`$– and $`d`$–bosons are conserved separately. In this case the Hamiltonian takes the form
$`H=ϵ(N_F+N_s)+H_d,`$ (3)
where $`N_ss^+s`$ and $`N_F_ma_{1/2m}^+a_{1/2m}`$ are the number operators for the $`s`$–bosons and the fermions, correspondingly, and
$`H_d=ϵ_d{\displaystyle \underset{\mu }{}}d_\mu ^+d_\mu +{\displaystyle \underset{L=0,2,4}{}}C_L(d^+d^+)_{LM}(dd)_{LM}.`$ (4)
The supercharges commuting with the Hamiltonian (3) are
$`P_{1/2m}=a_{1/2m}^+s,P_{1/2m}^+=s^+a_{1/2m},`$ (5)
$`[H,P_{1/2m}]=[H,P_{1/2m}^+]=0`$ (6)
Let us consider a graded algebra which contains $`P_{1/2m}`$ and $`P_{1/2m}^+`$ operators. The anticommutator of the supercharge operators $`P_{1/2m}`$ and $`P_{1/2m}^+`$ is
$`\{P_{1/2m},P_{1/2m^{}}^+\}=\delta _{mm^{}}(N_s+{\displaystyle \frac{1}{2}}N_F)\sqrt{2}(1)^{1/2+m^{}}C_{1/2m1/2m^{}}^{1\eta }S_{1\eta },`$ (7)
where $`S_{1\eta }`$ is the operator of the spin of the fermions
$`S_{1\eta }={\displaystyle \frac{\sqrt{3}}{2}}{\displaystyle \underset{m,m^{}}{}}C_{1/2m1\eta }^{1/2m^{}}a_{1/2m^{}}^+a_{1/2m},`$ (8)
and $`C_{b\beta c\gamma }^{a\alpha }`$ is a Clebsch–Gordan coefficient. The next set of the anticommutators is simple
$`\{P_{1/2m}^+,P_{1/2m^{}}^+\}=\{P_{1/2m},P_{1/2m^{}}\}=0,`$ (9)
The operators $`N_s`$ and $`N_F`$ commute with each other and with $`S_{1\eta }`$. The operators $`S_{1\eta }`$ satisfy the usual commutation relations for spin algebra
$`[S_{1\eta },S_{1\eta ^{}}]=\sqrt{2}C_{1\eta ^{}1\eta }^{1\eta \mathrm{`}\mathrm{`}}S_{1\eta ^{\prime \prime }}`$ (10)
Finally we consider the commutation relations of the operators $`N_s`$, $`N_F`$ and $`S_{1\eta }`$ with the supercharges $`P_{1/2m}`$ and $`P_{1/2m}^+`$. They are
$`[P_{1/2m},N_s]=P_{1/2m},`$ (11)
$`[P_{1/2m},N_F]=P_{1/2m},`$ (12)
$`[P_{1/2m},S_{1\eta }]={\displaystyle \frac{\sqrt{3}}{2}}C_{1/2m1\eta }^{1/2m^{}}P_{1/2m^{}}.`$ (13)
The operators $`P_{1/2m}`$, $`P_{1/2m}^+`$, $`N_s`$, $`N_F`$ and $`S_{1\eta }`$ form a graded algebra, which is $`U(1/2)`$, and the Hamiltonian (3) corresponds to the following reduction chain
$`U(6/2)U(5)U(1/2)`$ (14)
The multiplets of the eigenstates of the Hamiltonian (3) with the same energy combine the states with different number of fermions. For this reason they are called supermultiplets. Applying the super operators $`P_{1/2m}`$, $`P_{1/2m}^+`$ we can transform one member of the supermultiplet into an other one with different number of fermions. All states belonging to the supermultiplet can be created by a repeated application of the supersymmetric operators $`P_{\frac{1}{2}\frac{1}{2}}`$ and $`P_{\frac{1}{2}\frac{1}{2}}`$ to some basic state $`|n_s,n_d>`$ with $`n_s`$ and $`n_d`$ being the numbers of the $`s`$– and $`d`$– bosons in this state, correspondingly. This state does not contain fermions. The complete supermultiplet thus includes the following states
$`|n_s,n_d>(s^+)^{n_s}\{(d^+)^{n_d}\}_{IM}|0>,`$ (15)
$`P_{\frac{1}{2}\frac{1}{2}}|n_s,n_d>,P_{\frac{1}{2}\frac{1}{2}}|n_s,n_d>,`$ (16)
$`P_{\frac{1}{2}\frac{1}{2}}P_{\frac{1}{2}\frac{1}{2}}|n_s,n_d>.`$ (17)
Above the state $`|n_s,n_d>`$ is an analog of the Clifford (fermion) vacuum of the Susy model of the QFT. Examples of the complete supermultiplets are shown in Table 1. Looking at these examples we see that there is an interesting relation between the total number of the bosonic states, which include by definition states without fermions and also states with even numbers of fermions, and fermionic states, i.e. states with odd numbers of fermions, in the supermultiplet. In the supermultiplets, which include the states with the number of $`s`$–bosons equal or exceeding two, the total number of the fermionic (magnetic) substates is equal to the number of the bosonic (magnetic) substates. In the supermultiplets, which include states with the number of $`s`$–bosons equal or not exceeding one, the total number of the fermionic (magnetic) substates is equal to twice the number of the bosonic (magnetic) substates. This relation will be extended in the next section where the formal proof will be presented.
A nice and well known example of the approximate supersymmetry in nuclear physics corresponding to the Hamiltonian (3) is given in . It is a complete supermultiplet containing the three nuclei $`{}_{44}{}^{}{}_{}{}^{102}`$Ru<sub>58</sub> ($`N_s`$ \+ $`N_d`$=7), $`{}_{45}{}^{}{}_{}{}^{103}`$Rh<sub>58</sub> ($`N_s`$ \+ $`N_d`$=6, $`N_F`$=1) and also the two quasiparticle states of $`{}_{46}{}^{}{}_{}{}^{104}`$Pd<sub>58</sub> ($`N_s`$ \+ $`N_d`$=5, $`N_F`$=2 ). However, the experimental information on the two quasiparticle states in <sup>104</sup>Pd is insufficient. A possible candidate for the lowest state in <sup>104</sup>Pd belonging to the supermultiplet is $`0^+`$ (1793 KeV) state seen in $`\beta ^{}`$ decay of <sup>104</sup>Rh with $`logft`$=5.5. Experiments on <sup>103</sup>Rh (<sup>3</sup>He,d) will be important to clarify a situation. The odd proton occupies the $`p_{1/2}`$ single particle state. The experimental spectra are given in Fig.1. We put in Fig.1 the $`0^+`$ two quasiparticle state of <sup>104</sup>Pd at the same energy as the ground states of <sup>102</sup>Ru and <sup>103</sup>Rh because in SUSY approach to nuclear structure we are dealing with the energies of the states relative to the lowest state of the multiplet in the same nucleus. This is also the reason why in (3) the same coefficient is used in front of $`N_F`$ and $`N_s`$ terms. The corresponding full Hamiltonian $`\overline{H}`$ which includes total energies is
$`\overline{H}=ϵ_FN_F+ϵ_sN_s+H_d.`$ (18)
If we consider only relative energies $`\overline{H}`$ is equivalent to $`H`$ of (3), however. This can be seen by rewriting $`\overline{H}`$ as follows
$`\overline{H}=(ϵ_Fϵ_s)N_F+ϵ_s(N_F+N_s)+H_d.`$ (19)
One notes that the first term on the r.h.s. of (19) does not influence the relative energies. This is why this term is omitted which leads to the Hamiltonian $`H`$ of (3).
We see in Fig.1 that although the supersymmetry is realized with a good accuracy in <sup>102</sup>Ru and <sup>103</sup>Rh the symmetry is broken. So, to describe the experimental data in more details it is necessary to add to the Hamiltonian (3) a term which breaks the supersymmetry described above. This term can be choosen as
$`H_{break}=\alpha N_sN_F+\beta \stackrel{}{L_d}\stackrel{}{S},`$ (20)
where $`\stackrel{}{L_d}=\sqrt{10}(d^+\stackrel{~}{d})_1`$ is the operator of the orbital momentum of $`d`$–bosons.
## III Formal comparison of supersymmetry in U(6/2) model and in quantum field theory
In this section we shall compare the commutation and anticommutation relations between the operators of the U(1/2) superalgebra considered in the preceeding section with those of the SUSY model of the quantum field theory. This comparison will demonstrate a close similarity between both sets of relations and also some difference between them. The similarity will be used further to derive a formal relation between the numbers of the bosonic and fermionic states in the supermultiplets. In QFT the supercharges analogous to our operators $`P_{1/2m}`$ and $`P_{1/2m}^+`$ are denoted by $`Q_\alpha ^L`$ and $`(Q_\alpha ^L)^+`$, where $`\alpha `$ is a spinorial index and $`L`$ is an index connected with an intrinsic symmetry group. They commute with the 4–dimensional Lorentz momentum $`P_\mu `$
$`[Q_\alpha ^L,P_\mu ]=[(Q_\alpha ^L)^+,P_\mu ]=0,`$ (21)
and therefore with the Hamiltonian, which is a $`\mu `$=0 component of $`P_\mu `$.
The anticommutator of the supercharge operators $`Q_\alpha ^L`$ and $`(Q_\alpha ^L)^+`$ is
$`\{Q_\alpha ^L,(Q_\beta ^M)^+\}=2\delta ^{LM}{\displaystyle \underset{\mu }{}}\sigma _{\alpha \beta }^\mu P_\mu ,`$ (22)
where $`\sigma ^\mu `$ is a Dirac matrix. The second anticommutator is
$`\{Q_\alpha ^L,Q_\beta ^M\}=0.`$ (23)
In addition to the super operators $`Q_\alpha ^L`$, $`(Q_\alpha ^L)^+`$ and the Lorentz 4–dimensional momentum $`P_\mu `$ the SUSY model of QFT can includes also the boson type operators, which form the intrinsic symmetry Lie algebra in QFT, where they are denoted by $`B_l`$. The operators $`B_l`$ satisfy to the following commutation relations.
$`[B_l,B_m]=i{\displaystyle \underset{k}{}}f_{lmk}B_k,`$ (24)
where $`f_{lmk}`$ are the structure constants. The commutators between the intrinsic operators and super operators are
$`[Q_\alpha ^L,B_l]=i{\displaystyle \underset{M}{}}A_l^{LM}Q_\alpha ^M,`$ (25)
Now these relations can be putted in a correspondence with those relations obtained in the preceeding section. This is done below, where the corresponding relations are connected by the left–right arrows
$`P_{1/2m}Q_\alpha ^L,HP_\mu `$ (26)
$`N_sN_F,S_{1\eta }B_l,`$ (27)
$`[P_{1/2m},H]=0[Q_\alpha ^L,P_\mu ]=0,`$ (28)
$`\{P_{1/2m^{}},P_{1/2m}^+\}=\delta _{mm^{}}\left(N_s+{\displaystyle \frac{1}{2}}N_F\right)+\sqrt{3}{\displaystyle \underset{\eta }{}}C_{1/2m1\eta }^{1/2m^{}}S_{1\eta }`$ (29)
$`\{Q_\alpha ^L,(Q_\beta ^M)^+\}=2\delta ^{LM}{\displaystyle \underset{\mu }{}}\sigma _{\alpha \beta }^\mu P_\mu `$ (30)
$`\{P_{1/2m},P_{1/2m^{}}\}=0\{Q_\alpha ^L,Q_\beta ^M\}=0,`$ (31)
$`[S_{1\eta },S_{1\eta ^{}}]=\sqrt{2}C_{1\eta ^{}1\eta }^{1\eta ^{\prime \prime }}S_{1\eta ^{\prime \prime }}[B_l,B_m]=i{\displaystyle \underset{k}{}}f_{lmk}B_k`$ (32)
$`[P_{1/2m},N_s]=P_{1/2m},[P_{1/2m},N_F]=P_{1/2m},`$ (33)
$`[P_{1/2m},S_{1\eta }]={\displaystyle \frac{\sqrt{3}}{2}}C_{1/2m1\eta }^{1/2m^{}}P_{1/2m^{}}[Q_\alpha ^L,B_l]=i{\displaystyle \underset{M}{}}A_l^{LM}Q_\alpha ^M`$ (34)
It is easy to see a strong similarity between analogous relations in the QFT and in the nuclear case. This similarity is evident for the relations (28,3134). We notice, however, some difference in the structure of the expressions on the right hand sides of the anticommutators in (30). This distinction is due to the difference in the underlying bosonic algebras in QFT and in nuclear case. This difference is reflected in the relation between the numbers of the bosonic and fermionic states in the supermultiplets, which we shall consider below.
As in quantum field theory this relation can be obtained by multiplying both sides of the relation (7) by $`(1)^{N_F}`$ and taking a trace. On the left side we get
$`Tr\left((1)^{N_F}P_{1/2m}P_{1/2m^{}}^++(1)^{N_F}P_{1/2m^{}}^+P_{1/2m}\right)=`$ (35)
$`Tr\left((1)^{N_F}P_{1/2m}P_{1/2m^{}}^++P_{1/2m}(1)^{N_F}P_{1/2m^{}}^+\right)=`$ (36)
$`Tr\left((1)^{N_F}P_{1/2m}P_{1/2m^{}}^+(1)^{N_F}P_{1/2m}P_{1/2m^{}}^+\right)=0`$ (37)
Above we have used the invariance of the trace operation under the cyclical permutations of the operators and the commutation relation (12). Since the relation (7) can be also written as
$`\{P_{1/2m},P_{1/2m^{}}^+\}=\delta _{mm^{}}s^+s+a_{1/2m}^+a_{1/2m^{}}`$ (38)
by taking $`m=m^{}`$ and summing over $`m`$ we get
$`Tr\left((1)^{\widehat{N_F}}(2\widehat{N_s}+\widehat{N_F})\right)=0.`$ (39)
In fact (39) contains the relation between the numbers of the bosonic and fermionic states in the supermultiplets. To show it let us rewrite the trace in (39) in an explicit way as a sum of the diagonal matrix elements of the operator in circular brackets. Since the super operators on the left of (38) have nonzero matrix elements only between the states belonging to the same supermultiplet this sum is restricted to these states
$`{\displaystyle \underset{asupermultiplet}{}}<a|(1)^{\widehat{N_F}}(2\widehat{N_s}+\widehat{N_F})|a>=0`$ (40)
Above the index $`a`$ includes bosonic states, i.e. the states without fermions and with two fermions, and fermionic states, i.e. states with one fermion. Since two fermions can be coupled in our case only to zero angular momentum the number of magnetic substates without fermions is equal to the number of magnetic substates with two fermions. Therefore, both are equal to the half of the total number of the bosonic magnetic substates in the supermultiplet, which we denote as $`n_B`$. Let $`n_F`$ be the total number of the fermionic states in the supermultiplet. Applying (40) to the supermultiplets with $`N_s+N_FN_s^{max}`$ 2 one obtains
$`{\displaystyle \frac{1}{2}}n_B2N_s^{max}n_F\left(2(N_s^{max}1)+1\right)+{\displaystyle \frac{1}{2}}n_B\left(2(N_s^{max}2)+2\right)`$ (41)
$`=(n_Bn_F)(2N_s^{max}1)=0`$ (42)
and thus $`n_B=n_F`$. Above $`N_s^{max}`$ is a maximum possible number of the $`s`$–bosons in the states belonging to a supermultiplet.
For the supermultiplet with $`N_s+N_F`$=1 we get from (40)
$`{\displaystyle \frac{1}{2}}n_B1n_F=0`$ (43)
Thus we can formulate the following rules:
–In the supermultiplets which include the states with $`N_s+N_F2`$ the number of bosonic (magnetic) substates is equal to the number of the fermionic (magnetic) substates.
–In the supermultiplets containing states with $`N_s+N_F`$ = 1 the number of the fermionic (magnetic) substates is equal to twice the number of the bosonic (magnetic) substates.
There is some difference in the relations between the numbers of bosonic and fermionic states in nuclear case and in QFT where for all supermultiplets (excluding ground state) $`n_F=n_B`$. This difference is a consequence of the difference in the anticommutation relations (7) and (22).
Since the relation between the numbers of the bosonic and fermionic (magnetic) substates is a fundamental point in QFT, consider from this point of view the eigenstates of the Hamiltonians (1,2). The supermultiplets of the eigenstates of the Hamiltonian (1) correspond to the supermultiplets of the QFT with the zero rest mass Clifford (fermion) vacuum. In this case helicity is a good quantum number. The one–fermion state is, for instance, an analog of neutrino and a one $`s`$–boson state is an analog of S–neutrino. In the case of the Hamiltonian (2) a one–fermion state is an analog of electron with two possible projections of the spin.
In the QFT the following expression has been derived for the Hamiltonian. If we multiply both sides of (22) by $`\sigma _{\alpha \beta }^0`$ and sum over $`\alpha `$ and $`\beta `$ we get
$`HP_0={\displaystyle \frac{1}{4}}{\displaystyle \underset{\alpha ,L}{}}\{Q_\alpha ^L,(Q_\alpha ^L)^+\}.`$ (44)
In our case the relation, which can be derived in a similar way, is different because of the difference in the anticommutation relations (7) and (22)
$`H={\displaystyle \frac{1}{2}}N_F+{\displaystyle \frac{1}{2}}{\displaystyle \underset{m}{}}\{P_{1/2m},P_{1/2m}^+\}+H_d,`$ (45)
which is equivalent from the excitation energy point of view to
$`H={\displaystyle \frac{1}{2}}{\displaystyle \underset{m}{}}\{P_{1/2m},P_{1/2m}^+\}+H_d,`$ (46)
However, the ground state (i.e. the state with zero numbers of fermions, $`s`$– and $`d`$– bosons) of our Hamiltonian has zero energy as in QFT.
## IV Conclusion
We have considered the comparison of the supersymmetry concept in nuclear structure theory and in quantum field theory. The nuclear structure model based on the U(6/2) graded algebra is considered. It is shown that the U(6/2) algebra has a graded subalgebra whose generators commute with the model Hamiltonian. Thus, the existence of the supersymmetric operators transforming bosons into fermions and vice versa and commuting with the Hamiltonian is demonstrated. We have furthermore obtained explicit expressions of the wavefunctions of the various members of supermultiplets. For supermultiplets with $`N_F+N_s2`$ we find an equal number of boson and fermion type (magnetic) substates in the supermultiplet. This corresponds to the fundamental property of supermultiplets in QFT.
###### Acknowledgements.
The authors would like to express their gratitude to Profs. R.F.Casten, A.Gelberg, T.Otsuka, P. van Isacker and M.Zirnbauer for discussions. The work was supported in part by the DFG under the contract Br 799/8-2. One of the authors (R. V. J) is grateful to the Universität zu Köln for support. |
no-problem/9911/hep-ph9911214.html | ar5iv | text | # Amplification of Isocurvature Perturbations induced by Active-Sterile Neutrino Oscillations
## 1 Introduction
If one assumes the existence of a mixing between an active neutrino flavor $`\alpha `$ and a sterile neutrino flavor, then a lepton number can be generated in the early universe . To this purpose a small vacuum mixing angle ($`\mathrm{sin}2\theta _01`$) is necessary (the mass eigenstates almost coincide with the interaction eigenstates). Moreover the sterile neutrino must be lighter than the active neutrino and the absolute value of $`\delta m^2`$, the difference of squared masses, must be in the range $`(10^5÷10^5)\mathrm{eV}^2`$. In fact the critical temperature at which the generation occurs is given, if the sterile neutrino production is until that time negligible, by the expression:
$$T_c14.5(18.0)\mathrm{MeV}\left(\frac{|\delta m^2|}{\mathrm{eV}^2}\right)^{\frac{1}{6}}\alpha =e(\mu ,\tau )$$
(1)
It follows that if $`|\delta m^2|\stackrel{<}{}10^5\mathrm{eV}^2`$, then $`T_c\stackrel{<}{}3\mathrm{MeV}`$ and the lepton number variation would be dominated by the MSW effect. In this case a lepton number larger than $`10^7`$ cannot be generated .
On the other hand if $`|\delta m^2|\stackrel{>}{}10^5\mathrm{eV}^2`$ then $`T_c\stackrel{>}{}150\mathrm{M}\mathrm{e}\mathrm{V}`$ and one should be able to describe the neutrino oscillations in a quark-gluon plasma, something beyond the present level of matter effects account (moreover it is not interesting in this context, because it would imply active neutrino masses much higher than $`100\mathrm{eV}`$, cosmologically excluded if neutrinos are stable).
A rigorous description of neutrino oscillations in the early universe requires a quantum kinetic approach able to describe the evolution of the statistical density matrix for the two mixed neutrino flavors . Such a description must include the different effects of matter on the neutrino mixing: a coherent effect due to forward scattering , a loss of coherence due to the collisions that change the neutrino momentum and the repopulation of active neutrino states (depleted by the oscillations into the sterile neutrino states) through the collisions .
For a better understanding of the generation of a lepton number, it has been shown to be more convenient to turn to a simpler description that neglects the possibility of a MSW effect at the resonance. This approximation proves to be valid at temperatures $`T\stackrel{>}{}5\mathrm{MeV}`$ because the collisions destroy the coherence at the resonance. In this way the physics underlying the lepton number generation is isolated and can be clearly understood . The role of collisions, from this point of view, is crucial: they can be usefully considered, in a heuristic sense, as a measurement process able to make the two-quantum state collapsing in one of the two interaction eigenstates. Through this effect sterile neutrinos are produced and the presence of a small baryon number induces a tiny asymmetry between the production rates of neutrinos and antineutrinos. If the sterile is lighter than the active neutrino and if the values of mixing parameters are in the intervals previously indicated, this initial asymmetry is amplified through the generation of a lepton number that starts at the critical temperature $`T_c`$.
In this mechanism the momentum dependence plays an important role and a monochromatic approximation provides only a rough description. The interesting feature is that, within this simplified physical picture, the quantum kinetic description through the density matrix collapses into an effective (classic) kinetic description (Pauli-Boltzmann approach). This is possible because in the regime where collisions are rapid enough, the evolution of the off-diagonal terms can be disentangled by the diagonal terms evolution and explicit expressions can be derived for them. In this way, in the description of the evolution of the statistical properties of mixed neutrinos, only the diagonal terms are left, the usual statistical distributions: a simpler physical picture yields a simpler mathematical description. The derivation of the equations for the distributions can be done either directly from the simplified physical model or also, with a more formal procedure, from the quantum kinetic equations themselves, via some approximations, indicated by the authors globally as static approximation, valid under appropriate conditions on the mixing parameters . The differences that one expects from a full quantum kinetic description are at low temperatures ($`T\stackrel{<}{}5\mathrm{MeV}`$), when the MSW effect becomes important and must be taken into account, and at the critical temperature for large enough vacuum mixing angles, when the growth of lepton number is so rapid that an adiabatic condition to describe the process does not hold anymore. This adiabatic condition is the possibility to neglect any change in the effective potentials, and thus in the mixing angle, between the collisions on average. The collisions, if vacuum mixing angles is not too large, are able to average out the coherent effects of the oscillations in the macroscopic quantity (like the neutrino asymmetry).
This analysis on the validity of the approach is confirmed by the numerical calculations performed using the QKE. In it has been shown how the MSW effect at low temperature is able to amplify the growth of lepton number up to values slightly lower than the maximum absolute value obtainable of $`3/8`$, corresponding to a situation when all active antineutrinos (or neutrinos) are converted into sterile neutrinos and to a value $`\xi _\alpha 0.5`$ for the chemical potential <sup>1</sup><sup>1</sup>1 The authors of a recent paper obtain much lower values for the final lepton number . This result has been obtained solving an approximated equation that, compared to the static approximation equation, contains a new term responsible, according to them, for the different values. We do not hide our scepticism toward these results, due to the robust coherent picture of previous ones, obtained both in a physically clear approximated picture and confirmed by numerical calculations using the exact QKE equations . The authors of try to justify this situation claiming a not rigorous numerical procedure in the exact numerical calculations, without being aware of the conspiracy that would exist between a supposed numerical error in the solution of the exact QKE and the effect of a missing term in the static approximation equation. Moreover new numerical solutions of the exact QKE will be soon presented together with clear checks of their validity and an accurate description of the adopted numerical procedure. Although there is no final agreement at the present, we wish to thank A.D. Dolgov, S.H. Hansen, S. Pastor and D.V. Semikoz, for the kind availability to answer our questions. . On the other hand it has been shown how, for temperatures $`T\stackrel{>}{}5\mathrm{MeV}`$, an effective kinetic approach agrees almost perfectly with the quantum kinetic one for small mixing angles. It is also possible to extend the numerical study of QKE to large mixing angles and it has been found that for $`\mathrm{sin}^22\theta _0\stackrel{>}{}10^6`$ and $`\mathrm{sin}^22\theta _0\stackrel{<}{}\mathrm{\hspace{0.17em}3}\times 10^4(\mathrm{eV}^2/|\delta m^2|)`$, in the quantum kinetic approach, at the critical temperature $`T_c`$, rapid oscillations in the lepton number take place. This behaviour was first studied in and later confirmed in (even though in a smaller region of mixing parameters), but in these works the momentum dependence was not taken into account and the rapid oscillations are observed for a much larger region of mixing parameters. In is concluded that, outside the special region of mixing parameters where rapid oscillations are observed at the resonance, at temperatures $`T\stackrel{>}{}5\mathrm{MeV}`$, the effective kinetic approach provides a very good description not only for the evolution of the absolute value of lepton number but also in predicting its final sign. In this paper we will extend this analysis to the case when some tiny inhomogeneities in the baryon number (baryon isocurvature perturbations) are present, dealing with the region of mixing parameters where the effective kinetic approach can be used. This can be done including into the equation a term that describes the diffusion of neutrinos and we will show how this effect can induce the generation of lepton domains <sup>2</sup><sup>2</sup>2 The possibility of a generation of lepton domains was first claimed in due to a sign indetermination in the obtained equation for the lepton number evolution. This results in a generation of lepton domains with sign randomly determined. However in it was shown how the account of a correcting term produces a full sign determination of the solution (see Section 2). The idea of a ”chaotic” generation of lepton domains has been recently re proposed in . This model assumes that the lepton number undergoes at the resonance very rapid and unstable changes of sign for any choice of mixing parameters and this would again result in a sign indetermination in different points of space. The analysis presented in , including a full momentum dependence, excludes this possibility for almost all values of mixing parameters, except in a special region where the numerical calculations cannot be, at the moment, conclusive. In the mechanism we present here, the generation is not chaotic but perfectly determined by the spectral features of the baryon number inhomogeneities and moreover the horizon scale is not a limit to the size of lepton domains. .
The plan of the paper is the following. In section 2 we will show how the problem of sign of lepton number is fully determined in the effective kinetic approach, re analyzing and extending a procedure presented in . In section 3 we will extend the analysis to the case where small inhomogeneities in the baryon number are present, through the introduction of a diffusion term. We will illustrate how this term can induce, in some regions, an inversion in the sign of lepton number growth, bringing to the formation of lepton domains. In section 4 we describe qualitatively the evolution of lepton domains once they have been generated. In section 5 we conclude and discuss the possible applications of the new proposed mechanism.
## 2 Final sign of lepton number in a homogeneous background
Within the Pauli-Boltzmann approach developed in , the evolution of the lepton number carried by an active $`\alpha `$-neutrino flavor that is mixed with a sterile neutrino flavor, is described by the following equation:
$$\frac{dL_{\nu _\alpha }}{dt}=\left[A(T,L)\text{ }LB(T,L)L_{\nu _\alpha }\right]$$
(2)
where we defined <sup>3</sup><sup>3</sup>3 We are actually neglecting a third term that arises only when a sterile neutrino asymmetry is produced, that means when the lepton number of active neutrinos is changing, considering that $`L_{\nu _s}+L_{\nu _\alpha }=`$const. This term can give effects only when an initial $`\alpha `$-lepton number much higher than $`\stackrel{~}{L}`$ is assumed. Here we do not consider this situation and in this case this third term can be neglected. :
$$A\frac{T^3}{2\pi ^2n_\gamma }𝑑yy^2f_{eq}^0(\mathrm{\Gamma }_{\alpha s}\overline{\mathrm{\Gamma }}_{\alpha s})(z_s^+z_\alpha ^+)$$
(3)
$$B\frac{6\zeta (3)T^3}{\pi ^4n_\gamma }𝑑y\frac{y^2e^y}{(1+e^y)^2}(\mathrm{\Gamma }_{\alpha s}+\overline{\mathrm{\Gamma }}_{\alpha s})$$
(4)
The variable $`y`$ in the integrals is the adimensional momentum $`p/T`$. The quantity $`L_{\nu _\alpha }`$ is rigorously defined as the lepton number of the $`\alpha `$-active neutrino in the portion of comoving volume that contains a fixed number of photons $`N_\gamma ^{in}`$ at some initial temperature $`T_{in}\stackrel{<}{}150\mathrm{MeV}`$. Considering that the evolution of lepton number freezes at temperatures around $`T_f1\mathrm{MeV}`$, one can safely consider the number of photons in the element of comoving volume as a constant (neglecting muon annihilations) and write:
$$L_{\nu _\alpha }\frac{N_{\nu _\alpha }N_{\overline{\nu }_\alpha }}{N_\gamma ^{in}}\frac{n_{\nu _\alpha }n_{\overline{\nu }_\alpha }}{n_\gamma }$$
(5)
We indicated with $`n`$ the particle densities and with $`N=nR^3`$ the numbers of particles in the comoving volume $`R^3`$. We also introduced the (effective) total lepton number $`L`$ defined as:
$$LL_{\nu _\alpha }+L_{\nu _e}+L_{\nu _\mu }+L_{\nu _\tau }\pm \frac{1}{2}B_n2L_{\nu _\alpha }+\stackrel{~}{L}$$
(6)
where $`+`$ holds for $`\alpha =e`$ and $``$ for $`\alpha =\mu ,\tau `$. The field $`\stackrel{~}{L}`$ is the total charge number of non oscillating neutrinos plus a contribution from the baryon number carried by neutrons: we will refer to it as the background charge. It is constant while the oscillations occur, and must be considered as a parameter given by some earlier phase of baryo-leptogenesis. In this section we will assume that it is also strictly homogeneous, while in the next section we will study the effect of the presence of inhomogeneities. Both in $`L`$ and in $`\stackrel{~}{L}`$ we dropped an index $`\alpha `$ to simplify the notation.
The quantities $`\mathrm{\Gamma }_{\alpha s}`$, $`\overline{\mathrm{\Gamma }}_{\alpha s}`$ are the production rates for sterile neutrinos and antineutrinos, $`f_{eq}^0`$ is the Fermi-Dirac distribution with zero chemical potential and $`z_{\alpha ,s}^\pm =(f_{\nu _{\alpha ,s}}\pm f_{\overline{\nu }_{\alpha ,s}})/2f_{eq}^0`$ are the sum and difference distributions of neutrinos, relative to the Fermi-Dirac one. The active neutrinos distributions can be safely described by thermal equilibrium distributions because the process of generation of lepton number occurs for temperatures $`T1\mathrm{M}\mathrm{e}\mathrm{V}`$ (the total collision rate is therefore much higher than the expansion rate) and for small mixing angles (the collision rates that refill the quantum states of active neutrinos are much higher than the sterile neutrinos production rates that deplete them). On the other hand the sterile neutrinos distributions must be described by two other rate equations that, together with the expressions for the production rates, can be found in (we do not need them in the present context).
For a qualitative understanding of the evolution of lepton number, with a specific attention to its sign, it is useful to recast the equation (2) in the form:
$$\frac{dL_{\nu _\alpha }}{dt}=\left[2A(T,L)B(T,L)\right](L_{\nu _\alpha }L_{eq})$$
(7)
where we introduced the quantity:
$$L_{eq}=\frac{A\stackrel{~}{L}}{2AB}$$
(8)
While the term $`B`$ is always positive, the term $`A`$ changes sign at a critical temperature $`T_c`$, being negative for higher temperatures and positive for lower temperatures. When the temperature is far from the critical value ($`|TT_c|\mathrm{\Delta }T(23)\mathrm{MeV}`$), the $`B`$ term acts as a correcting term ($`B|A|`$) and the fixed point $`L_{eq}L_{eq}^0(1/2)\stackrel{~}{L}`$. Moreover, until the lepton number is below a threshold value $`L_{}10^6(|\delta m^2|/\mathrm{eV}^2)^{\frac{1}{3}}`$ (see ), its non linear effect inside $`A`$ and $`B`$ can be neglected. In this situation the equation (2) becomes extremely simple:
$$\frac{dL_{\nu _\alpha }}{dt}=A(T)L$$
(9)
While for temperatures above $`T_c`$ the quantity $`A`$ is negative, the fixed point $`L_{eq}`$ is stable and thus the total lepton number is destroyed, for temperatures below $`T_c`$, $`A`$ is positive, the fixed point is unstable and the total lepton number starts to grow. To answer the question toward which direction it grows, one has to take into account the term $`B`$ and study its action in the vicinity of the critical temperature when $`A`$ is small and $`B`$ becomes important in driving the evolution of lepton number. Let us see in different stages what happens when temperature approaches the critical value from the above. When $`B`$ is still small compared to $`A`$, expanding up to the first order in $`B/A`$, one gets:
$$L_{eq}=\frac{1}{2}\stackrel{~}{L}+\frac{1}{4}\frac{B}{|A|}\stackrel{~}{L}$$
(10)
During this period the term $`2AB`$ is negative and the solution tracks the fixed point $`L_{eq}`$. An approximate expression for the quantity $`\delta LL_{\nu _\alpha }L_{eq}`$ can be derived neglecting a term $`d(\delta L)/dt`$ in the Eq. (7):
$$\delta L=\frac{1}{2AB}\frac{dL_{eq}}{dt}$$
(11)
Until $`2ABdL_{eq}/dt`$, the solution tracks the fixed point. From the expression (10) it is clear that the term $`B`$ drives the growth of the lepton number toward the same sign as the background charge $`\stackrel{~}{L}`$. At the critical temperature the term $`A`$ vanishes and simply $`L_{eq}=0`$. In this moment the fixed point is changing very rapidly while the term $`2AB`$, that should force the solution $`L_{\nu _\alpha }`$ to track $`L_{eq}`$, is small and the solution starts to diverge from the fixed point. Immediately below the critical temperature, when $`2A=B`$, the fixed point has a vertical asymptote and changes sign and at still lower temperatures it rapidly approaches zero again. This time however $`A`$ is positive and the fixed point $`L_{eq}`$ does not attract the solution any more and this continues to grow toward the same direction transmitted by the ”derailment” action of the term $`B`$ at the critical temperature.
The behaviour of the solution around the critical temperature is shown in fig. 1 for a particular choice of the mixing parameters.
## 3 Lepton domains formation in presence of inhomogeneities
In this section we will generalize the process of active-sterile neutrino oscillations to the case when spatial inhomogeneities are present. From the observation of CMB we know that small inhomogeneities in the temperature field (adiabatic perturbations) were present in the early universe. However taking into account the presence of these perturbations does not change the basic results of the homogeneous scenario. The lepton number growth starts at slightly different times in different points, but the final result is unchanged: the presence in the end of a final lepton number with the same sign everywhere in the space and with the same absolute value as in the homogeneous case.
More interesting is to consider the possibility that small inhomogeneities are present in the background charge $`\stackrel{~}{L}`$ (isocurvature perturbations). This quantity is the sum of the lepton number carried by the non oscillating neutrinos and a term given by the presence of a baryon number. Whatever is the mechanism that created the inhomogeneities, it is reasonable to think that the inhomogeneities in the lepton numbers carried by the non oscillating neutrinos become soon much smoother than those in the baryon number, due to the much higher diffusion of neutrinos. More precisely we can consider the baryon number field expressed in the comoving coordinates as a constant, neglecting the neutron diffusion. This will make much simpler our following discussion, without altering the main results.
A first trivial effect on the equation (2) is simply that now $`\stackrel{~}{L}=\stackrel{~}{L}(𝐱)`$, where $`𝐱`$ are the comoving coordinates. At temperatures above $`T_c`$ this effect is enough to understand how the process is modified by the presence of the inhomogeneities. The $`\nu _\alpha `$-lepton number will evolve in a way to destroy in any point the total lepton number $`L`$ and a zero order solution, considering only the $`A`$-term, is given by $`L_{\nu _\alpha }=(1/2)\stackrel{~}{L}(𝐱)`$. When temperature drops down, approaching the critical value, again the action of the correcting term $`B`$ must be considered in order to understand toward which direction, positive or negative values, the total lepton number will grow.
This time however a new correcting term must be considered in the equation, a term that arises due to the neutrino diffusion, and the full equation becomes now:
$$\frac{dL_{\nu _\alpha }}{dt}=A(T,L)\left[2L_{\nu _\alpha }+\stackrel{~}{L}(𝐱)\right]B(T,L)L_{\nu _\alpha }+\frac{𝒟}{R^2}^2L_{\nu _\alpha }$$
(12)
where clearly $`L_{\nu _\alpha }=L_{\nu _\alpha }(𝐱,t)`$, $`𝒟`$ is the diffusion coefficient and $`R`$ is the scale factor that appears because we are dealing with comoving coordinates (we normalize $`R`$ in a way that $`R_0=1`$), while the diffusion term must be calculated in physical lengths.
The appearance of a diffusion term is something intuitively clear. For a more rigorous derivation one has simply to write the Liouville operator in the left hand side of the Boltzmann equation in the inhomogeneous case:
$$\frac{d}{dt}f_{\nu _\alpha }(𝐱,𝐩,t)=\frac{}{t}f_{\nu _\alpha }(𝐱,𝐩,t)H𝐩_𝐩f_{\nu _\alpha }(𝐱,𝐩,t)+\frac{c}{R}\widehat{p}_𝐱f_{\nu _\alpha }(𝐱,𝐩,t)$$
(13)
where $`H`$ is the expansion rate $`\dot{R}/R`$. In this way, integrating on the momenta the Boltzmann equation for the distribution difference of active neutrinos $`f_{\nu _\alpha }^{}f_{\nu _\alpha }f_{\overline{\nu }_\alpha }`$ and dividing for the photon number density, one obtains:
$$\frac{dL_{\nu _\alpha }}{dt}=\left[\frac{dL_{\nu _\alpha }}{dt}\right]_{\mathrm{osc}}+\frac{1}{R^2}_𝐱\left(𝒟_𝐱L_{\nu _\alpha }\right)$$
(14)
where the first term in the r.h. side is the contribution from the oscillations into the sterile neutrinos (the r.h. side in the Eq. (2)), while the second term is the contribution due to the diffusion. The diffusion coefficient is defined by the expression:
$$\frac{d^3p}{(2\pi ^3)}f_{\nu _\alpha }^{}(𝐱,𝐩,t)\widehat{p}\frac{𝒟}{R}_𝐱\frac{d^3p}{(2\pi ^3)}f_{\nu _\alpha }^{}(𝐱,𝐩,t)$$
(15)
As we are considering the case of small inhomogeneities in the baryon number, the diffusion coefficient can be safely considered homogeneous and one gets the diffusion term in the form written in the equation (12).
The order of magnitude of $`𝒟`$ is given by $`c\mathrm{}_{int}`$, where $`\mathrm{}_{int}=\mathrm{\Gamma }_{int}^1`$ is the interaction length and $`\mathrm{\Gamma }_{int}=3.15k_\alpha G_F^2T^5`$ is the total collision rate, with $`k_\alpha 1.27\text{ }(0.92)`$ for $`\alpha =e(\mu ,\tau )`$ <sup>4</sup><sup>4</sup>4An explicit calculation of the diffusion coefficient in the simpler case of ”light” particles diffusing in a ”heavy” particles medium can be found in . .
If the diffusion term is negligible compared to the $`B`$ term, then the homogeneous scenario is practically unchanged. In the equation (10) one has simply to consider that now $`\stackrel{~}{L}=\stackrel{~}{L}(𝐱)`$ and, as we are considering small inhomogeneities so that the sign of $`\stackrel{~}{L}`$ is spatially constant, the $`B`$ term pushes the solution toward the same direction in all the points.
On the other hand if we assume that the diffusion term is dominant, the situation can be much different. Let us assume for definiteness that $`\stackrel{~}{L}>0`$. In the regions where the background charge is lower, the $`\alpha `$-neutrino lepton number, before the critical temperature, would be higher (because $`L_{\nu _\alpha }0.5\stackrel{~}{L}`$). In these regions the diffusion term pushes the lepton number to be depleted with the result that at the critical temperature one can have $`L_{\nu _\alpha }<0.5\stackrel{~}{L}`$ and the growth starts toward a negative sign. The vice versa would happen in the regions where the background charge is higher. The final result is the creation of regions with different sign of lepton number with the same size as the scale of baryon number inhomogeneities. The sign of lepton number is in fact determined at any point by the curvature of the background charge field and thus the global properties of the baryon number inhomogeneities are somehow transmitted to the lepton domains.
We have now to study the conditions for which this mechanism can work. The involved parameters are the mixing parameters, the amplitude of the inhomogeneities and the size of the inhomogeneities. We can assume, for simplicity, that inhomogeneities have only one size scale. Let us introduce the field $`\rho (𝐱)`$ of the inhomogeneities, writing the background charge field as:
$$\stackrel{~}{L}(𝐱)=\overline{L}\left[1+\rho (𝐱)\right]$$
(16)
where $`\overline{L}`$ is the mean value, and where we assume that the field $`\rho `$ is a perturbation ($`\rho 1`$ at any point). As we already said, while also non oscillating neutrinos give a non negligible contribution to the mean value, the baryon number gives the dominant contribution to the inhomogeneities.
At temperatures higher than $`T_c`$ the solution will approach again the fixed point that at the zero order will be simply $`L_{eq}^0(𝐱)=(1/2)\stackrel{~}{L}(𝐱)`$. The final sign of lepton number will be again determined by the correcting terms that this time include also the diffusion term. This can be generically estimated through the following expression (we drop the dependence on $`𝐱`$ in $`\rho `$ and $`L_{eq}^0`$):
$$\frac{𝒟}{R^2}^2L_{\nu _\alpha }\mathrm{\Gamma }_{int}\rho \left(\frac{\mathrm{}_{int}^{(0)}}{\lambda ^{(0)}}\right)^2L_{eq}^0DL_{eq}^0$$
(17)
where we indicated with $`\lambda `$ the scale of the inhomogeneities and, given the generic physical length $`\mathrm{}`$, we mean with $`\mathrm{}^{(0)}=(R_0/R)\mathrm{}`$, the comoving length normalized at the present. We also introduced the convenient quantity $`D`$. In this way the expression (8), valid for the fixed point in the homogeneous case, becomes:
$$L_{eq}=\frac{A\stackrel{~}{L}}{2ABD}$$
(18)
or at the first order in $`(B+D)/A`$:
$$L_{eq}=\frac{1}{2}\stackrel{~}{L}+\frac{1}{4}\frac{B+D}{|A|}\stackrel{~}{L}$$
(19)
This expression is now describing what we previously said in words: if the term $`D`$ is negative (it means $`\rho <0`$, the regions where the background charge is lower than the mean value if we assume it is positive) and its absolute value higher than $`B`$, then the lepton number growth is addressed toward the opposite sign than in the regions where the opposite condition holds.
To determine the conditions for which an inversion of sign is possible, we have thus simply to compare the term $`B`$ with the absolute value of $`D`$ at temperatures around the critical temperature. An estimation of the order of magnitude of $`B`$ can be easily done considering that the integral on the momenta receives a dominant contribution only around the resonant value. This procedure has already been employed in (in that paper it was used to estimate the $`A`$-term) and here we only give the result (valid for temperatures $`T\stackrel{>}{}T_c`$, when the lepton number is still small and can be neglected):
$$B10^2s^2\mathrm{\Gamma }_{int}$$
(20)
where we defined $`s\mathrm{sin}2\theta _0`$. A calculation of the comoving interaction length expressed in parsec gives the result:
$$\mathrm{}_{int}^{(0)}100(70)\mathrm{pc}\left(\frac{\mathrm{MeV}}{T}\right)^4$$
(21)
for $`\alpha =\mu ,\tau (e)`$. Using this expression and imposing the condition $`|D|>B`$ for the generation of lepton domains, it is thus straightforward to derive the following condition on the (orders of magnitude) of the involved parameters:
$$\rho \left(\frac{\lambda ^{(0)}}{pc}\right)^2\stackrel{>}{}10^7s^2\left(\frac{|\delta m^2|}{\mathrm{eV}^2}\right)^{\frac{4}{3}}$$
(22)
This analytical expression is confirmed by the numerical calculations. In figure 2 we show the evolution of the terms $`|A|`$, $`B`$, calculated numerically for a particular choice of the mixing parameters and compared with the term $`|D|`$, calculated through the expression (17) with different values of the parameter $`|\rho |/(\lambda ^{(0)})^2`$. The result is in agreement with the one that could be derived using the eq. (22).
This condition can also be expressed as a condition on the order of magnitude of the maximum size of lepton domains that can be generated <sup>5</sup><sup>5</sup>5 There is also a condition on the minimum size of lepton domains that can be generated : $`\lambda ^{(0)}\stackrel{>}{}\mathrm{}_{int}^{(0)}(TT_c)10^3\mathrm{pc}(|\delta m^2|/\mathrm{eV}^2)^{\frac{2}{3}}`$. Below this scale neutrino are free streaming and they can destroy the lepton domains more rapidly than they can be generated, considering that $`\dot{L}\stackrel{<}{}\mathrm{\Gamma }_{int}`$. :
$$\lambda ^{(0)}\stackrel{<}{}10\mathrm{Kpc}\left[\frac{|\delta m^2|}{10^5\mathrm{eV}^2}\right]^{\frac{2}{3}}\left(\frac{\rho }{10^1}\right)^{\frac{1}{2}}\left(\frac{10^{10}}{s^2}\right)^{\frac{1}{2}}\stackrel{<}{}10\mathrm{Kpc}$$
(23)
In this last expression we indicated the extreme possible values of the different quantities that yield the maximum value for $`\lambda ^{(0)}`$. In particular the condition $`s^2\stackrel{>}{}10^{10}`$ must be imposed to have a not negligible final absolute value of lepton number.
In next section we will use this result to sketch a scenario of how lepton domains with different sizes evolve and affect the BBN. This analysis will suggest a way to circumvent the upper limit found on the maximum size of a lepton domain.
In conclusion of this section we notice that because of the presence of the term $`B`$ there are always points where, even though $`\rho <0`$, the condition $`B|D|<0`$ is not verified. Therefore we can say that the $`B`$ term will always favour a dominance of regions where the generation of lepton number occurs with the same sign of $`\stackrel{~}{L}`$: we will refer to it as the dominant sign, compared to regions with inverted sign. There are however two much different extreme situations of dominance. In a first one $`B/|D|1`$ everywhere, except in thin walls around the surfaces where $`\rho =0`$ (weak dominance). In this case the size of lepton domains coincides with the size of baryon number inhomogeneities and in a comoving sphere with a much larger radius than the size of inhomogeneities, the volume of regions with dominant sign lepton number is almost equal to that of regions with inverted sign lepton number. The topology of lepton domains is that one of a cubic lattice with each cube surrounded by first closest neighbours cubes with opposite sign and second closest neighbours cubes with the same sign.
In a second extreme situation the condition $`B|D|<0`$ is verified only in small regions around the points where $`|\rho |`$ is maximum, with a size much smaller than that of the inhomogeneities (strong dominance). In this case one has a structure of lepton domains with inverted sign (3-dim) islands in a dominant sign background.
## 4 Lepton domains evolution
Lepton domains start to be formed at $`T=T_c`$. With the increase of lepton number the diffusion at the border of domains also increase. There is some interplay between the rate of generation (the $`A`$-term) and the rate of diffusion.
Let us assume that the dominant sign is positive. In this case the diffusion can be described as a process that gradually fills at the border the regions with negative sign (the ”holes”). At the same time, in the positive sign regions, the lepton number that has been used to fill the holes, is restored by the oscillations through the growth term. In this situation it is clear that if the diffusion is able to cover the whole hole, the generation of lepton domains is only a transient regime without any practical effect. We can refer to this phase as cannibalization regime, in which the dominant sign regions enlarge at the expense of the inverted sign regions, until eventually their complete disappearance.
However this process can last only until the generation of a lepton number can occur at the border of domains, where there is a change of sign and the lepton number is kept small by the diffusion. When the temperature drops down to $`T0.6T_c`$ <sup>6</sup><sup>6</sup>6 It corresponds to a situation when the resonant neutrinos have a momentum $`p\stackrel{>}{}10T`$, in the tail of the distribution. For small values of $`L`$, neutrinos and antineutrinos are both resonant but if $`L>0(<0)`$ and if $`yy_{\mathrm{peak}}2.2`$, the number of resonant antineutrinos (neutrinos) is a little bit higher than the number of resonant neutrinos (antineutrinos): this explains why the fixed point $`L0`$ is unstable and lepton number starts to grow (see ). or in any case down to $`3`$ MeV, the growth starts to be inhibited. We have to calculate the diffusion length to know how large the lepton domains with inverted sign must be in order to survive to the cannibalization regime. This is given by the following expression that takes into account the Universe expansion:
$$\lambda _d^{(0)}=_{t_c}^t\frac{v_ddt^{}}{R(t^{})}$$
(24)
The diffusion velocity is approximately given by $`v_d\sqrt{𝒟/(tt_c)}`$. Considering that $`tt_c`$ and that $`t\mathrm{}_H`$, it is an easy task to get the following expression for the comoving diffusion length:
$$\lambda _d^{(0)}100(70)\mathrm{pc}\left[\left(\frac{\mathrm{MeV}}{T}\right)^{\frac{5}{2}}\left(\frac{\mathrm{MeV}}{T_c}\right)^{\frac{5}{2}}\right],$$
(25)
for $`\alpha =\mu ,\tau (e)`$. The generation of lepton number can only occur down to temperatures $`T\mathrm{min}(3\mathrm{MeV},0.6T_c)`$, therefore for $`T_c\stackrel{>}{}5\mathrm{MeV}`$, using the expression (1), we get:
$$\lambda _d^{(0)}(T=0.6T_c)0.1\mathrm{pc}\left(\frac{|\delta m^2|}{\mathrm{eV}^2}\right)^{\frac{5}{12}}$$
(26)
For sizes below this value (very small scales) lepton domains with inverted sign are destroyed before they can produce any effect. For $`T_c\stackrel{<}{}5\mathrm{MeV}`$ the diffusion length between the beginning of the generation of lepton number at $`T_c`$ and its end at $`T3\mathrm{M}\mathrm{e}\mathrm{V}`$ tends to zero for $`|\delta m^2|10^5\mathrm{eV}^2`$, but in any case there is a lower limit on the size of lepton domains that can be generated and in the end the same expression (26) can be approximately used.
For larger scales the cannibalization cannot destroy completely the lepton domains before that the generation of lepton number at the border stops. When this happens, the lepton number that diffuses to fill the hole is not generated any more: the result is an effect of dilution of lepton number. The generation of lepton domains leaves a mark that modifies the homogeneous scenario.
In this case, to understand the fate of lepton domains, it is useful to compare the scale of lepton domains with the horizon scale, given by the following expression:
$$\mathrm{}_H^{(0)}=100\mathrm{p}\mathrm{c}\left(\frac{\mathrm{MeV}}{T}\right)$$
(27)
It coincides with the diffusion length at $`1\mathrm{M}\mathrm{e}\mathrm{V}`$ when neutrinos start to free stream. All lepton domains with small scales $`\lambda ^{(0)}\stackrel{<}{}100\mathrm{p}\mathrm{c}`$, do not survive and one gets in the end a homogeneous lepton number field. However, as already stated, this time the generation of lepton domains produced the effect to dilute the final value of lepton number. If the condition for the formation of domains (22) is satisfied only in the peaks of the inhomogeneities (strong dominance), then the contribution from the regions with inverted sign is negligible and the dilution effect too. Otherwise in the other extreme case, when the relation (22) is satisfied in almost all the space (weak dominance), then there would be an almost total reciprocal cancelation of regions with opposite sign, with only a small relic value of lepton number due to the presence of the $`B`$-term. All intermediate values of lepton number are possible.
Scales of inhomogeneities $`\lambda ^{(0)}\stackrel{>}{}100\mathrm{p}\mathrm{c}`$ can produce lepton domains able to survive until the freezing of neutron to proton ratio and therefore the presence of these scales would give rise to an inhomogeneous BBN scenario.
Another important distinction is between scales able to produce primordial nuclear abundance inhomogeneities that survive until the present (large scales) or not (medium scales). In fact, even though nuclear abundances are produced in a inhomogeneous scenario of BBN, subsequent astrophysical mixing mechanisms, such as shock waves induced by supernovae explosions, would be able to homogenize the products of BBN . These processes are effective on scales $`\lambda ^{(0)}\stackrel{<}{}100\mathrm{K}\mathrm{p}\mathrm{c}`$. From this point of view the mechanism we proposed, in its simplest version, is unable to produce visible primordial abundance inhomogeneities.
We can however circumvent this limit if we relax our initial assumption on the presence of only one characteristic scale in the field of the inhomogeneities $`\rho (𝐱)`$. We can in fact imagine a simple extension with two characteristic scales contemporarily present (for examples two Fourier modes): one small scale and one large scale. The amplitude of the first one could be modulated by the presence of the second. In this case one can have the simultaneous existence of regions where the small scales inhomogeneities are able to efficiently dilute the lepton number generated at the critical temperature down to negligible values, and regions where the amplitude of small inhomogeneities is depressed and there is no dilution effect. In this way one can produce islands where lepton number is present in a background with zero lepton number or vice versa. This time the size of the island regions has no intrinsic limit, but it is a feature of the model that produced the baryon inhomogeneities.
More realistically one has to think that the inhomogeneities are described by a spectrum of lengths. The interesting property is that the spectral features of the isocurvature perturbations would be reflected in the lepton domains features and eventually in the BBN products. Another interesting aspect is that very large scale lepton domains would not be constrained by microwave background observations as models where large amplitude isocurvature perturbations are directly present in the baryon number . This because the inhomogeneity in the active neutrino energy density would be compensated by an opposite inhomogeneity in the sterile neutrino energy density. In fact during active-sterile neutrino oscillations one has that $`L_{\nu _s}+L_{\nu _\alpha }=`$const. Even though sterile neutrinos free stream while active neutrinos are still diffusing, for very large scales (surely larger than the diffusion length of active neutrinos) this different behaviour cannot change the conservation of the sum of the asymmetries of active neutrinos and sterile neutrinos at each point.
In the end of this section we want also to show that small amplitudes of inhomogeneities are sufficient for the mechanism to work. From the expression (22), imposing the lower limit (26) on $`\lambda ^{(0)}`$ (in order to have lepton domains with inverted sign large enough not to be cannibalized), one gets the following condition on the mixing parameters:
$$s^2\left(\frac{|\delta m^2|}{\mathrm{eV}^2}\right)^{\frac{1}{2}}<10^5\rho $$
(28)
that is satisfied for large regions of possible values of the mixing parameters, even for very small values of $`\rho `$. These regions are represented in figure 3 for three different values of $`\rho `$. It can be noticed that even for $`\rho =10^5`$, there is a significant region where lepton domains can be generated and produce interesting effects.
## 5 Conclusions
We proposed a mechanism according to which active-sterile neutrino oscillations would amplify small baryon number inhomogeneities generating a structure of lepton domains. The assumption on the presence of baryon number inhomogeneities is a very reasonable one, because very small amplitudes are sufficient to produce lepton domains large enough to give some effect. The structure of the lepton domains and its effect on the BBN would reflect the spectral features of the same perturbations that seeded the formation of the domains. The main attraction of this mechanism is that the size of these domains is not limited by the horizon scale. Moreover large scale lepton domains cannot be ruled out by current CMB observations (but maybe could be constrained or discovered in future experiments).
The mechanism presents different applications. It provides a scenario for a non standard BBN (see for a review and see also the recent paper where the effects of inhomogeneous chemical potentials on BBN are considered). If the active neutrino is an electron neutrino the effects on the nuclear abundances can be remarkable and some of them could help to explain the present observational picture, for example claimed inhomogeneities in the D/H abundances from measurements in quasar absorption systems . From this point of view one possibility is that we live in a background of zero electron lepton number but some observed high redshift absorption systems, for example the one observed with an high abundance, could be included in an island of non zero lepton number <sup>7</sup><sup>7</sup>7 The possibility that neutrino chemical potentials inhomogeneities can be responsible for the primordial deuterium abundance inhomogeneities able to explain the observations from high redshift quasar absorbers, has been proposed in . A recent analysis of the current measurements of primordial deuterium abundance in high redshift quasars absorbers can be found in . .
We also mention that isocurvature perturbations can have interesting effects on large scale structure .
Very generically we can say that the generation of lepton domains enlarge the possibility of the existence of some observational signature of active-sterile neutrino oscillations in the early universe. In the very fortunate case that active-sterile neutrino oscillations, with the right parameters for a lepton domain formation, do really occur in nature, then we would have a powerful probe for baryogenesis models that predict a spectrum of inhomogeneities. The simple non observation of a signature of the generation of lepton domains would put strong constraints. We conclude saying that the planned earth experiments will be able in next future to test the hypothesis of active-sterile neutrino oscillations and thus to rule out or support the proposed mechanism.
Acknowledgements
P. Di Bari acknowledges many interesting and stimulating discussions with R. Foot, P. Lipari, M. Lusignoli and R.R. Volkas and is grateful to them for many suggestions that made this manuscript clearer. He also thanks K. Jedamzik for a nice discussion during COSMO-99.
## Figure Captions
Fig. 1 The behaviour of $`L_{\nu _\alpha }L_{eq}^0`$ (solid line) around the critical temperature when the sign is determined by the action of the correcting term $`B`$. The dotted line is the quantity $`L_{eq}L_{eq}^0`$, with $`L_{eq}`$ defined by the expression (8) and $`L_{eq}^0=(1/2)\stackrel{~}{L}`$. We used $`\stackrel{~}{L}=10^{10}`$, while the values of mixing parameters are those indicated.
Fig. 2 Comparison of the quantities $`A,B,D`$ times $`|dt/dT|=1/HT`$ (for numerical calculations it is more convenient to replace time with temperature). The term $`A`$ vanishes at the critical temperature $`T15\mathrm{M}\mathrm{e}\mathrm{V}`$. The values of mixing parameters are those indicated. The diffusion term is plotted for different values of the parameter $`\rho /(\lambda ^{(0)}\mathrm{pc}^1)^2`$. It can be noticed that the diffusion term $`|D|`$ is dominant on $`B`$, around the critical temperature, for $`\rho /(\lambda ^{(0)}\mathrm{pc}^1)^2\stackrel{>}{}10^1`$, as it could be deduced by the expression (22).
Fig. 3 Represantation of the regions where the condition (28) for the surviving of lepton domains to the canibalization regime is satisfied. The regions are those below the three solid lines and correspond to the three indicated values for the amplitude $`\rho `$ of the baryon number inhomogeneities that seed the lepton domain formation. Above the dotted line the contribution of sterile neutrinos to the number of effective neutrinos during the BBN is higher than $`0.8`$ for $`\alpha =e`$ (see ). Even for very small amplitudes ($`\rho =10^5`$) there is a significant region of parameters where the condition (28) is satisfied. |
no-problem/9911/quant-ph9911099.html | ar5iv | text | # A remark on the isotropic model
( FOM Institute AMOLF, Kruislaan 407, 1098 SJ Amsterdam, The Netherlands
and
I. Institut für Theoretische Physik, Jungiusstrasse 9, Universität Hamburg, D-20355 Hamburg, Germanythanks: present address)
abstract
The applicability of the so-called isotropic complete photonic-band-gap (CPBG) models \[S. John and J. Wang, Phys. Rev. Lett. 64, 2418 (1990)\] to capture essential features of the spontaneous emission (SE) of a fluorescent atom or molecule near a band-gap-edge of a CPBG structure is discussed.
PACS number: 32.80.-t - Photon interactions with atoms
PACS number: 42.70.Qs - Photonic bandgap materials
In their influential article , John and Wang considered the quantum electrodynamics (QED) of an atom, minimally coupled to the radiation filed, in the presence of a complete photonic-band-gap (CPBG). In the latter case, there is a frequency interval in which, independently of the photon direction and polarization, no photon modes can propagate. Such a QED vacuum differs significantly from the conventional QED vacuum. In the vicinity of a CPBG edge $`\omega _c`$ a number of exotic phenomena were predicted, among others, radical changes in the spontaneous emission (SE) and an anomalous Lamb shift. (It is worthwhile to notice that some of the exotic phenomena have been discussed much earlier in less known papers by Bykov .) Atom properties were shown to depend strongly on the exponent $`\eta `$ of the density of states (DOS) asymptotic
$$\rho (\omega )\text{const}|\omega \omega _c|^\eta $$
(1)
near the CPBG edge. Calculations involving photonic crystals in more than one dimension (1D) are notoriously difficult. Therefore, in order to determine $`\eta `$, John and Wang made use of approximations, subsequently employed in a number of recent discussions of the SE near a CPBG edge , and often called isotropic and anisotropic CPBG models. In the first model, the DOS near the band edge $`\omega _c`$ is obtained from an approximated dispersion relation of the CPBG material $`\omega _𝐤\omega _c+A(𝐤𝐤_0)^2,`$ where $`A\omega _c/𝐤_0^2`$ and $`𝐤_0`$ is a vector at the Brillouin zone boundary. The second model is then a slight generalization of the first one.
However, already in the Wigner-Weisskopf approximation the SE of an isolated fluorescent atom (or molecule) in a fixed position $`𝐫`$ within the unit cell is determined by the local DOS (LDOS) and not the DOS . Considerations based on the DOS can only be valid in the two hypothetical cases which are difficult to achieve: (i) when the atom is allowed to freely propagate within a CPBG structure, as in experiments with cavity QED, and (ii) when atoms are distributed homogeneously within the entire unit cell. Only then the averaged and not the local properties of the QED vacuum within the unit cell are probed. The fluorescent atoms are usually not distributed uniformly. If the atoms can be considered as independent and are only radiatively coupled to the crystal, the measured SE is determined by the weighted average (with the atomic position probability distribution) of the LDOS.
In contrast to the DOS, the LDOS behavior near a CPBG edge shows more complex behavior. Within a given frequency band, the LDOS as a function of $`𝐫`$ exhibits minima and maxima, their number depending on the order of the band starting from the lowest one . The LDOS behavior is especially sensitive to $`𝐫`$ in the vicinity of the LDOS minima, where the LDOS approaches zero. If sufficiently close to a minimum of the LDOS, a shift in position by $`10^4`$ of the unit cell length (i.e., a shift by one atom for optical photonic crystals) can cause a change in the LDOS by a factor of $`3`$ or higher. Surprisingly enough, except at the exact position of a minimum, the LDOS near a band edge has asymptotic described by Eq. (1) and the value of $`\eta `$ is still universal. Therefore, at least in 1D, the predictions derived from the isotropic model hold true.
I should like to thank S. John and O. Toader for pointing to an error in my original calculation. This work is part of the research program by the Foundation for Fundamental Research on Matter which was made possible by financial support from the Netherlands Organization for Scientific Research<sup>1</sup><sup>1</sup>1Fortran code to calculate the LDOS asymptotic at band edges is available upon request.. |
no-problem/9911/hep-lat9911023.html | ar5iv | text | # Pion decay constant in quenched QCD with Kogut-Susskind quarksPresented by T. Kaneda
## 1 Introduction
The Kogut-Susskind (KS) quark action has the well-known feature that SU(4) flavor symmetry is broken down to U(1) subgroup at finite lattice spacing. The restoration of full flavor symmetry toward the continuum limit has been previously examined for pion mass. Here we extend the examination to the pion decay constant by comparing results for various KS flavors to that in the U(1) channel for which the renormalization constant equals unity.
This comparison is made both for perturbative and non-perturbative renormalization factors, the latter evaluated with the method of Ref. .
Numerical simulations are carried out in quenched QCD at $`\beta =6.0`$ and 6.2 employing $`32^3\times 64`$ and $`48^3\times 64`$ lattices. Other lattice parameters are summarized in Table 1.
## 2 Formalism and simulation
In the hypercubic notation for the KS fermion fields, the axial vector current in the KS flavor channel $`F`$ has the form
$$A_\mu ^F=\overline{\varphi }(\gamma _\mu \gamma _5\xi _F)\varphi .$$
(1)
The fields $`\overline{\varphi }`$ and $`\varphi `$ in (1) are generally separated over a 4-dimensional hypercube. We use both gauge-invariant and non-invariant operators, inserting the product of link variables between $`\overline{\varphi }`$ and $`\varphi `$ for the former, and employing the Landau gauge fixing to deal with the latter.
The one-loop perturbative results for the renormalization constants $`Z_A^F`$ have been worked out in Ref. . We use tadpole-improved values employing the tadpole-improved $`\overline{\mathrm{MS}}`$ coupling with $`q^{}=1/a`$; they will be denoted as $`Z_A^{(\mathrm{P})F}`$. For non-perturbative renormalization constants $`Z_A^{(\mathrm{N})F}`$, we take results of Ref. at $`(ap)^2=1.0024`$.
Quark propagators are calculated for 16 wall sources, each corresponding to a corner of a hypercube. These propagators are combined to form the correlator $`A_\mu ^F(t)\pi _W^F(0)`$ for each flavor $`F`$, where $`\pi _W^F`$ is the pion field for the wall source, thereby enhancing signals.
We find the quark mass dependence of $`(m_\pi ^F)^2`$, $`f_\pi ^F`$ and vector meson masses to be well described by a linear function. Hence a linear fit is employed for the chiral extrapolation in the quark mass $`m_q`$. The scale is set by the $`\rho `$ meson mass in the spin-flavor channel $`\gamma _k\xi _k`$. Errors are estimated by the single elimination jackknife procedure.
## 3 Pion masses
We show in Fig. 1 values of $`(m_\pi ^F)^2`$ at $`m_q=0`$ as a function of $`a^2`$. The 16 KS flavors are grouped into 8 irreducible representations (irreps) given by $`\xi _5,\xi _4\xi _5,\xi _4,I`$ (1-d irreps) and $`\xi _k\xi _5,\xi _k\xi _4,\xi _{\mathrm{}}\xi _m,\xi _k`$ (3-d irreps). We observe that the 8 irreps form a degeneracy pattern $`\xi _5,(\xi _k\xi _5,\xi _4\xi _5),(\xi _k\xi _4,\xi _{\mathrm{}}\xi _m),(\xi _4,\xi _k),I`$ at finite lattice spacing. This feature was initially observed numerically in Ref. , and recently theoretically explained in Ref. . We also see that non-zero values of $`(m_\pi ^F)^2`$ for the channels other than $`\xi _5`$ vanish quadratically in $`a`$ toward the continuum limit as expected.
## 4 Pion decay constants
In Fig. 2 we illustrate how the bare values of $`f_\pi ^F`$ change under renormalization, taking results for gauge invariant current at $`\beta =6.2`$. The bare values for the 8 irreps again form a degeneracy pattern. The pattern reflects the distance of $`\overline{\varphi }`$ and $`\varphi `$ in (1), and is different from that of $`m_\pi ^F`$.
The perturbative renormalization constants (middle frame in Fig. 2) help to reduce the discrepancy among the bare values of $`f_\pi ^F`$. It is clear, however, that the discrepancy is much more reduced with the non-perturbative renormalization constants (bottom frame).
In Fig. 3 we plot the continuum extrapolation of the pion decay constants obtained with gauge-invariant currents. The top figure shows results with perturbative renormalization factors, and the bottom figure with non-perturbative factors. Lines are quadratic fits in $`a^2`$ following the $`O(a^2)`$ scaling violation expected for the KS quark action. Similar figures for the decay constants obtained with the gauge non-invariant currents are shown in Fig. 4.
The gauge-invariant current for the flavor $`\xi _5`$ does not require renormalization, from which we obtain $`f_\pi =89(6)`$ MeV in the continuum limit. This result is consistent with the experimental value of 92.4(3) MeV; possible deviations due to quenching is not visible within our 7% error.
We observe in Figs. 3 and 4 that the perturbative renormalization factors at one-loop order are not sufficient to ensure restoration of flavor symmetry in the continuum limit. In contrast, the decay constants evaluated with the non-perturbative renormalization factors agree much better already at finite lattice spacings, and their continuum limits are well convergent, the central values coinciding with each other well within the errors.
This work is supported by the Supercomputer Project No.45 (FY1999) of High Energy Accelerator Research Organization (KEK), and also in part by the Grants-in-Aid of the Ministry of Education (Nos. 09304029, 10640246, 10640248, 10740107, 10740125, 11640294, 11740162). K-I.I is supported by the JSPS Research Fellowship. |
no-problem/9911/astro-ph9911263.html | ar5iv | text | # RXTE Observations of the Vela Pulsar: The X-ray-Optical Connection
## 1. Introduction
The Vela pulsar (PSR B0833-45) is the strongest $`\gamma `$-ray source in the sky, but it is one of the most difficult pulsars to detect at X-ray energies. This is in part because it is embedded in a very bright X-ray synchrotron nebula providing a large unpulsed background, but also because its pulsed X-ray emission is comparatively weak. The first detection of pulsed emission at X-ray energies was made by ROSAT in the 0.1 - 2 keV band (Ogelman 1993), and the spectrum is consistent with a blackbody. We detected the pulsar for the first time in hard X-rays (2 - 30 keV) during a 93 ks RXTE Cycle 1 observation (Strickman, Harding & De Jager 1999 \[SHD99\]) and the pulse profile shows two peaks. The first RXTE peak is closely aligned with the first EGRET $`\gamma `$-ray peak, but the second peak has an energy-dependent phase and is aligned with the second EGRET peak only at the highest energy (16 - 30 keV). In our lowest energy RXTE band (2 - 8 keV) the second peak is roughly aligned with the second peak of the optical profile (Gouiffes 1998). The average pulse spectrum joins smoothly with the high-energy spectrum of OSSE, COMPTEL and EGRET, although the spectrum of the first peak is significantly harder than that of the second peak, which appeared to be consistent with an extrapolation to the optical flux points. We (SHD99) suggested that the second RXTE peak was a blend of separate hard and soft components. In this paper, we report preliminary results of our analysis of a 300 ks RXTE Cycle 3 observation which confirms this picture.
## 2. Cycle 3 RXTE Observations and Analysis
The Cycle 3 observations were carried out during April/May and July/August 1999 with a good exposure of 274 ks. For this analysis we epoch-folded the PCA data in GoodXenon event-by-event mode at the pulsar period using the Princeton Pulsar Database (Arzoumian et al. 1992), to obtain the energy-dependent light curves which we have summed into three broad energy bands shown in Figure 1. To separate the various possible components of the light curve and compute individual component spectra, we fit a five peak sinusoid model with peaks of the form
$`C(i)`$ $`=`$ $`A(i)|\mathrm{cos}[{\displaystyle \frac{\pi }{2}}(\varphi \varphi __0(i))]|^{\xi (i)},`$ (1)
$`\xi (i)`$ $`=`$ $`{\displaystyle \frac{0.693}{\mathrm{log}[\mathrm{cos}(\frac{\pi }{2}W(i))]}}`$
to the data in 94 energy channels (roughly spanning 2 - 30 keV), where $`A(i)`$, $`\varphi _0(i)`$ and $`W(i)`$ are the amplitude, center phase and width of peak $`i`$ respectively. The model also includes a constant background level. We found that $`\varphi _0(i)`$ and $`W(i)`$ did not vary significantly with energy, except for the phase of Peak 3 which has a modest variation (see Table 1), so they were fixed while the values of $`A(i)`$ as a function of energy were determined. The counts for each peak were then found by integrating each sinusoid curve over phase. We could then obtain the photon spectrum of each peak separately by fitting a power law with photoelectric absorption of fixed column density $`10^{20}\mathrm{cm}^2`$.
## 3. Results
The RXTE light curves in Figure 1 show a narrow peak (Peak 1) at the phase of the EGRET first peak and a second peak that is now clearly seen to be a blend of two components. The first of these components, which we call Peak 2, becomes dominant in the 2 - 8 keV band and is in phase with the optical second peak. The second component, Peak 3, is harder and is in phase with the EGRET second peak. There are two other statistically significant peaks that appear in the RXTE light curve: a peak (Peak 5) at the radio phase (0.0) and a weaker peak (Peak4) leading the radio peak. Note that there is also a peak in the optical light curve at the radio phase. The energy spectra have been computed using the sinusoid model fits to each peak. The resulting power law spectral indices and goodness of fit for each peak are listed in Table 1. The peaks in phase with the EGRET peaks (Peak 1 and Peak 3) both have hard, but significantly different, spectra. The spectra of Peak 2 and Peak 5 are much softer and both their flux levels and indices are consistent within the uncertainties. Interestingly, an extrapolation of the Peak 2 spectrum falls near the optical flux points (Nasuti et al. 1997).
## 4. Discussion
These Cycle 3 observations have thus independently confirmed the multicomponent nature of pulsed emission from the Vela pulsar in the energy range 2 - 30 keV suggested by the Cycle 1 observations. With the improved statistics of the Cycle 3 data, we have been able to separate the broad second peak in the RXTE X-ray light curve into soft (Peak 2) and hard (Peak3) spectral components which maintain their phase integrity throughout the RXTE energy range. In addition, we have discovered a new feature in the RXTE light curve: a peak (Peak 5) at the phase of the radio pulse with an extremely soft spectrum. There is, in addition, significant emission leading the radio phase (Peak 4). Peaks 1 and 3 make up the hard spectral component whose light curve peaks are in phase with those of the gamma-ray light curve, and whose spectrum smoothly connects to the 100 keV - 5 GeV spectrum. Peaks 2 and 5, whose phases match those of the second optical and radio pulses respectively, make up the soft component. Their spectra, consistent with each other in both flux and spectral index, extrapolate to the optical flux points.
Although the RXTE hard component spectrum connects to the $`\gamma `$-ray spectrum, the X-ray spectrum is harder, requiring a break around 100 keV. Such a break at the local cyclotron energy, blueshifted by the parallel momentum of the pairs, is predicted by the polar cap cascade model (Harding & Daugherty 1999). The RXTE soft component may be either inverse Compton scattering radiation of pairs in polar cap cascades (Zhang & Harding 1999) or synchrotron radiation of backflowing particles from the outer gap (Cheng & Zhang 1999).
## References
Arzoumian, Z., Nice, D. & Taylor, J. H. 1992, GRO/radio timing data base, Princeton Univ.
Cheng, K. S. & Zhang, L. 1999, ApJ, 515, 337.
Gouiffes, C. 1998, in Neutron Stars and Pulsars, ed. N. Shibazaki, N. Kawai, S. Shibata & T. Kifune (Univ. Acad. Press: Toyko), p. 363.
Harding, A. K. & Daugherty, J. K. 1999, Proc. of 3rd Integral Workshop, in press.
Kanbach, G. et al. 1994, A & A, 289, 855.
Nasuti, F. P. et al. 1997, A & A, 323, 839.
Ogelman, H. 1993, in Lives of Neutron Stars, ed. M. A. Alpar et al. (Dordrecht: Kluwer), p. 101.
Strickman, M. S., Harding, A. K. & DeJager, O. C. 1999, ApJ, 524. 373.
Zhang, B. & Harding, A. K. 1999, ApJ, in press. |
no-problem/9911/nucl-th9911040.html | ar5iv | text | # Photoproduction of meson and baryon resonances in a chiral unitary approach
## 1 Introduction
Chiral Perturbation Theory ($`\chi PT`$) has proved very successful in order to describe the physics of mesons at very low energies. The key point of the whole approach is to identify the lightest pseudoscalar mesons $`\pi ,K`$ and $`\eta `$ as the Goldstone bosons associated to the chiral symmetry breaking. These particles will be the only degrees of freedom at low energies and their interactions can be described in terms of the most general effective Lagrangian which respects the chiral symmetry constraints.
So far as this is a low energy approach, the amplitude of a given process is basically given as an expansion in the external momenta over the scale of symmetry breaking $`4\pi f_\pi 1.2`$GeV. The approach is known to provide a good description of meson interactions up to about 500 MeV. However, if one is interested in resonances in particular, as it happens in meson spectroscopy, there is little that one can do with just plain $`\chi PT`$. The method that we expose here naturally leads to low lying resonances and allows one to face many problems so far intractable within $`\chi PT`$.
The method incorporates the following elements: 1) Unitarity is implemented exactly; 2) It can deal with coupled channels allowed with pairs of particles from the lightest octets of pseudoscalar mesons and ($`\frac{1}{2}^+`$) baryons; 3) A chiral expansion in powers of the external four-momentum of the lightest pseudoscalars is done for Re $`T^1`$, instead of the $`T`$ matrix itself as it is done in standard $`\chi PT`$.
We sketch here the steps involved in this expansion for the meson meson interaction. One starts from a $`K`$ matrix approach in coupled channels where unitarity is automatically fulfilled and writes
$$T^1=K^1i\sigma ,$$
(1)
where $`T`$ is the scattering matrix, $`K`$ is a real matrix in the physical region and $`\sigma `$ is a diagonal matrix which measures the phase-space available for the intermediate states
$$\sigma _{nn}(s)=\frac{k_n}{8\pi \sqrt{s}}\theta \left(s(m_{1n}+m_{2n})^2\right),$$
(2)
where $`k_n`$ is the on shell CM momentum of the meson in the intermediate state $`n`$ and $`m_{1n}`$, $`m_{2n}`$ are the masses of the two mesons in the state $`n`$. The meson meson states considered here are $`K\overline{K}`$, $`\pi \pi `$, $`\pi \eta `$, $`\eta \eta `$, $`\pi K`$, $`\pi \overline{K}`$, $`\eta K`$, $`\eta \overline{K}`$. Since $`K`$ is real in the physical region, from eq. (1) one sees that $`K^1`$ = Re $`T^1`$. In non-relativistic Quantum Mechanics, in the scattering of a particle from a potential, it is possible to expand $`K^1`$ in powers of the momentum of the particle at low energies as follows (in the s-wave for simplicity)
$$\text{Re}T^1K^1=\sigma ctg\delta \frac{1}{a}+\frac{1}{2}r_0k^2,$$
(3)
with $`k`$ the particle momentum, $`a`$ the scattering length and $`r_0`$ the effective range.
The ordinary $`\chi `$PT expansion up to $`O(p^4)`$ is given by
$$T=T_2+T_4,$$
(4)
where $`T_2`$ is obtained from the lowest order chiral Lagrangian, $`L^{(2)}`$, and is of $`O(p^2)`$, whereas $`T_4`$ contains one loop diagrams in the s, t, u channels, constructed from the lowest order Lagrangian, tadpoles and the finite contribution from the tree level diagrams of the $`L^{(4)}`$ Lagrangian and is $`O(p^4)`$. This last contribution, after a suitable renormalization, is just a polynomial, $`T^{(p)}`$. Our $`T^1`$ matrix, starting from eq. (4) is given by
$`T^1`$ $`=`$ $`\left[T_2+T_4+\mathrm{}\right]^1=T_2^1[1+T_4T_2^1+\mathrm{}]^1`$ (5)
$`=`$ $`T_2^1[1T_4T_2^1+\mathrm{}]=T_2^1[T_2T_4]T_2^1`$
Due to the fact that $`\text{Im}T_4=T_2\sigma T_2`$, the above equation is nothing but eq. (1), but using eq. (4) to expand $`K^1=\text{Re}T^1`$. Inverting the former result, one obtains:
$$T=T_2[T_2T_4]^1T_2,$$
(6)
which is the coupled channel generalization of the inverse amplitude method of .
Once this point is reached one has several options to proceed:
a) A full calculation of $`T_4`$ within the same renormalization scheme as in $`\chi PT`$ can be done. The eight $`L_i`$ coefficients from $`L^{(4)}`$ are then fitted to the existing meson meson data on phase shifts and inelasticities up to 1.2 GeV, where 4 meson states are still unimportant. This procedure has been carried out in . The resulting $`L_i`$ parameters are compatible with those used in $`\chi PT`$. At low energies the $`O(p^4)`$ expansion for $`T`$ of eq. (6) is identical to that in $`\chi PT`$. However, at higher energies the nonperturbative structure of eq. (6), which implements unitarity exactly, allows one to extend the information contained in the chiral Lagrangians to much higher energy than in ordinary $`\chi `$ PT. Indeed it reproduces the resonances present in the L = 0, 1 partial waves.
b) A technically simpler and equally successful additional approximation is generated by ignoring the crossed channel loops and tadpoles and reabsorbing them in the $`L_i`$ coefficients given the weak structure of these terms in the physical region. The fit to the data with the new $`\widehat{L}_i`$ coefficients reproduces the whole meson meson sector, with the position, widths and partial decay widths of the $`f_0(980)`$, $`a_0(980)`$, $`\kappa (900)`$, $`\rho (770)`$, $`K^{}(900)`$ resonances in good agreement with experiment . A cut off regularization is used in for the loops in the s-channel. By taking the loop function with two intermediate mesons
$$G_{nn}(s)=i\frac{d^4q}{(2\pi )^4}\frac{1}{q^2m_{1n}^2+iϵ}\frac{1}{(Pq)^2m_{2n}^2+iϵ},$$
(7)
where $`P`$ is the total meson meson momentum, one immediately notices that
$$\text{Im}G_{nn}(s)=\sigma _{nn}.$$
(8)
Hence, we can write
$$\text{Re}T_4=T_2\text{Re}GT_2+T_4^{(p)},$$
(9)
where $`\text{Re}G`$ depends on the cut off chosen for $`|\stackrel{}{q}|`$. This means that the $`\widehat{L}_i`$ coefficients of $`T_4^{(p)}`$ depend on the cut off choice, much as the $`L_i`$ coefficients in $`\chi PT`$ depend upon the regularization scale.
c) For the L = 0 sector (also in L = 0, S = $`1`$ in the meson baryon interaction) a further technical simplification is possible. In these cases it is possible to choose the cut off such that, given the relation between $`\text{Re}G`$ and $`T_4^{(p)}`$, this latter term is very well approximated by $`\text{Re}T_4=T_2\text{Re}GT_2`$. This is possible in those cases because of the predominant role played by the unitarization of the lowest order $`\chi PT`$ amplitude, which by itself leads to the low lying resonances, and because other genuine QCD resonances appear at higher energies.
In such a case eq. (5) becomes
$$T=T_2[T_2T_2GT_2]^1T_2=[1T_2G]^1T_2,$$
(10)
or, equivalently,
$$T=T_2+T_2GT,$$
(11)
which is a Bethe-Salpeter equation with $`T_2`$ and $`T`$ factorized on shell outside the loop integral, with $`T_2`$ playing the role of the potential. This option has proved to be successful in the L = 0 meson meson sector in and in the L = 0, S = $`1`$ meson baryon sector in .
In the meson baryon sector with S = 0, given the disparity of the masses in the coupled channels $`\pi N`$, $`\eta N`$, $`K\mathrm{\Sigma }`$, $`K\mathrm{\Lambda }`$, the simple “one cut off approach” is not possible. In higher order Lagrangians are introduced while in different subtraction constants in G are incorporated in each of the former channels leading in both cases to acceptable solutions when compared with the data.
In fig. 1 we show the results done with the method of for some selected phase shifts and inelasticities in the meson meson sector, showing resonances in different channels, (see for an update of the results). The agreement with the meson meson data is quite good up to 1.2 GeV and the parameters $`\widehat{L_i}`$ obtained from the fit are essentially compatible with those of $`\chi PT`$.
## 2 $`\overline{K}N`$ interaction in free space
The meson-baryon interaction Lagrangian at lowest order in momentum is given by
$$L_1^{(B)}=\overline{B}i\gamma ^\mu \frac{1}{4f^2}[(\mathrm{\Phi }_\mu \mathrm{\Phi }_\mu \mathrm{\Phi }\mathrm{\Phi })BB(\mathrm{\Phi }_\mu \mathrm{\Phi }_\mu \mathrm{\Phi }\mathrm{\Phi })],$$
(12)
where $`\mathrm{\Phi }`$ represents the octet of pseudoscalar mesons and $`B`$ the octet of $`1/2^+`$ baryons. The symbol $``$ denotes the trace of SU(3) matrices.
The coupled channel formalism requires to evaluate the transition amplitudes between the different meson-baryon channels. For $`K^{}p`$ scattering there are ten channels, namely $`K^{}p`$, $`\overline{K}^0n`$, $`\pi ^0\mathrm{\Lambda }`$, $`\pi ^0\mathrm{\Sigma }^0`$, $`\pi ^+\mathrm{\Sigma }^{}`$, $`\pi ^{}\mathrm{\Sigma }^+`$, $`\eta \mathrm{\Lambda }`$, $`\eta \mathrm{\Sigma }^0`$, $`K^+\mathrm{\Xi }^{}`$ and $`K^0\mathrm{\Xi }^0`$, while in the case of $`K^{}n`$ scattering there are six: $`K^{}n`$, $`\pi ^0\mathrm{\Sigma }^{}`$, $`\pi ^{}\mathrm{\Sigma }^0`$, $`\pi ^{}\mathrm{\Lambda }`$, $`\eta \mathrm{\Sigma }^{}`$ and $`K^0\mathrm{\Xi }^{}`$. These amplitudes have the form
$$V_{ij}=C_{ij}\frac{1}{4f^2}\overline{u}(p_i)\gamma ^\mu u(p_j)(k_{i\mu }+k_{j\mu }),$$
(13)
where $`p_j,p_i(k_j,k_i)`$ are the initial, final momenta of the baryons (mesons) and $`C_{ij}`$ are SU(3) coefficients that can be found in Ref. . At low energies the spatial components can be neglected and the amplitudes reduce to
$$V_{ij}=C_{ij}\frac{1}{4f^2}(k_j^0+k_i^0).$$
(14)
The coupled-channel BS equations in the center of mass frame read
$$T_{ij}=V_{ij}+\overline{V_{il}G_lT_{lj}},$$
(15)
where the indices $`i,l,j`$ run over all possible channels and $`\overline{V_{il}G_lT_{lj}}`$ corresponds to the loop integral involving $`V`$, $`T`$ and the meson baryon propagators of $`G`$, all functions of the loop variable. However, as was shown in Ref. , the off-shell part of $`V_{il}`$ and $`T_{lj}`$ goes into renormalization of coupling constants and $`V_{il}`$, $`T_{lj}`$ factorize outside the integral with their on-shell values, thus reducing the problem to one of inverting a set of algebraic equations.
The value of the cut-off, $`q_{\mathrm{max}}=630`$ MeV, was chosen to reproduce the $`K^{}p`$ scattering branching ratios at threshold, while the weak decay constant, $`f=1.15f_\pi `$, was taken in between the pion and kaon ones to optimize the position of the $`\mathrm{\Lambda }(1405)`$ resonance. The predictions of the model for several scattering observables are summarized in Table 1. Cross sections for $`K^{}p`$ scattering to different channels are also calculated in and good results are obtained for low energies of the kaons where the s-wave is dominant.
A recent application of these methods in the S=0 sector is done in , where the $`\mathrm{\Delta }(1232)`$ resonance is also nicely reproduced.
## 3 Application to the photoproduction of meson baryon pairs in resonant states
As quoted above, a good description of the interaction of $`K^{}p`$ and its coupled channels is obtained in terms of the lowest order Lagrangians and the Bethe Salpeter equation with a single cut off. One of the interesting features of the approach is the dynamical generation of the $`\mathrm{\Lambda }(1405)`$ resonance just below the $`K^{}p`$ threshold. The threshold behavior of the $`K^{}p`$ amplitude is thus very much tied to the properties of this resonance. Modifications of these properties in a nuclear medium can substantially alter the $`K^{}p`$ and $`K^{}`$ nucleus interaction and experiments looking for these properties are most welcome. Some electromagnetic reactions appear well suited for these studies. Application of the chiral unitary approach to the $`K^{}p\gamma \mathrm{\Lambda }`$, $`\gamma \mathrm{\Sigma }^0`$ reactions at threshold has been carried out in and a fair agreement with experiment is found. In particular one sees there that the coupled channels are essential to get a good description of the data, increasing the $`K^{}p\gamma \mathrm{\Sigma }^0`$ rate by about a factor 16 with respect to the Born approximation.
In a recent paper the $`\gamma pK^+\mathrm{\Lambda }(1405)`$ reaction was proposed as a means to study the properties of the resonance, together with the $`\gamma AK^+\mathrm{\Lambda }(1405)A^{}`$ reaction to see the modification of its properties in nuclei. The resonance $`\mathrm{\Lambda }(1405)`$ is seen in its decay products in the $`\pi \mathrm{\Sigma }`$ channel, but as shown in the sum of the cross sections for $`\pi ^0\mathrm{\Sigma }^0`$, $`\pi ^+\mathrm{\Sigma }^{}`$, $`\pi ^{}\mathrm{\Sigma }^+`$ production has the shape of the resonance $`\mathrm{\Lambda }(1405)`$ in the I = 0 channel. Hence, the detection of the $`K^+`$ in the elementary reaction, looking at $`d\sigma /dM_I`$ ($`M_I`$ the invariant mass of the meson baryon system which can be induced from the $`K^+`$ momentum), is sufficient to get a clear $`\mathrm{\Lambda }(1045)`$ signal. In nuclear targets Fermi motion blurs this simple procedure (just detecting the $`K^+`$), but the resonance properties can be reconstructed by observing the decay products in the $`\pi \mathrm{\Sigma }`$ channel. In fig. 2 we show the cross sections predicted for the $`\gamma pK^+\mathrm{\Lambda }(1405)`$ reaction looking at $`K^+\pi ^0\mathrm{\Sigma }^0`$, $`K^+all`$ and $`K^+\mathrm{\Lambda }(1405)`$ (alone). All of them have approximately the same shape and strength given the fact that the I = 1 contribution is rather small. In the figure the dashed dotted line indicates what one should expect to see in nuclei, just detecting the $`K^+`$, from the effect of Fermi motion.
The energy chosen for the photon is $`E_\gamma `$ = 1.7 GeV which makes it suitable of experimentation at SPring8/RCNP, where the experiment is planned , and TJNAF.
One variant of this reaction is the crossed channel reaction $`K^{}p\mathrm{\Lambda }(1405)\gamma `$. This reaction, for a $`K^{}`$ momentum in the 300 to 500 MeV/c range, shows clearly the $`\mathrm{\Lambda }(1405)`$ resonant production and has the advantage that the analogous reaction in nuclei still allows the observation of the $`\mathrm{\Lambda }(1405)`$ resonance with the mere detection of the photon, the Fermi motion effects being far more moderate than in the case of the $`\gamma AK^+\mathrm{\Lambda }(1405)X`$ reaction which requires larger photon momenta and induces a broad distribution of $`M_I`$ for a given $`K^+`$ momentum.
## 4 Photoproduction of resonant two meson states
Another application which can be done using the same reaction is the photoproduction of resonant two meson states. Particularly the $`f_0`$(980) and $`a_0`$(980) resonances. These states appear in $`L=0`$ in isospin zero and one respectively. The scalar sector of the meson is very controversial and the chiral unitary theory has brought a new perspective on these states. In particular, it has been possible to identify the lightest scalar octet, made of the $`\sigma `$, $`f_0(980)`$, $`a_0(980)`$ and $`\kappa (900)`$ resonances. All of them can be simply generated by unitarization of the lowest order ChPT, with just a cutoff as a free parameter.
The $`O(p^4)`$ chiral parameters can be understood as the residual contact terms that appear when one integrates out heavier states and the resulting Lagrangian is that of ChPT. Hence, the values of the chiral constants can be related to the masses and widths of the preexisting heavier resonances (“Resonance Saturation Hypothesis”). Indeed, most of their values are saturated by vector resonances alone (that is vector meson dominance) but some other parameters still need the existence of scalar states. Recently , using the N/D unitarization method with explicit resonances added to the lowest order ChPT Lagrangian, it has been established that these heavier scalar states should appear with a mass around 1.3 - 1.4 GeV for the octet and 1 GeV for the singlet. In addition, the $`\sigma `$, $`\kappa `$, $`a_0`$ and a strong contribution to the $`f_0`$, were also generated from the unitarization of the ChPT lowest order. These states still survive when the heavier scalars are removed. That agrees with our observation that the $`\sigma `$, $`\kappa `$, $`f_0`$ and $`a_0`$ are generated independently of the chiral parameters, that is, of the preexisting scalar nonet, which is heavier. In addition, it was also stablished in that work that the physical $`f_0(980)`$ resonance is a mixture between the discussed strong $`K\overline{K}`$ scattering contribution and the preexisting singlet resonance with a mass around 1 GeV. Since Chiral Perturbation Theory does not deal with quarks and gluons, it is very hard to make any conclusive statement about the nature of these states ($`q\overline{q}`$, four-quark, molecule, etc…), unless we make additional assumptions. However, any model of the nature of these states should be able to explain the different features of these resonances as they appear in the chiral unitary approach.
In addition, it would be very interesting to obtain further information from other processes. In the present case the reaction suggested is $`\gamma ppM`$. where $`M`$ is either of the resonances $`a_0`$(980) or $`f`$(980). In practice the meson M will decay into two mesons , $`\pi \pi `$ or $`K\overline{K}`$ in the case of the $`f_0`$(980) or $`K\overline{K}`$, $`\pi \eta `$ in the case of the $`a_0`$(980).
In Fig. 3 we show the results for the 5 channels considered. We observe clear peaks for $`\pi ^+\pi ^{}`$, $`\pi ^0\pi ^0`$ and $`\pi ^0\eta `$ production around 980 MeV. The peaks in $`\pi ^+\pi ^{}`$ and $`\pi ^0\pi ^0`$ clearly correspond to the formation of the $`f_0(980)`$ resonance, while the one in $`\pi ^0\eta `$ corresponds to the formation of the $`a_0(980)`$. The $`\pi ^0\pi ^0`$ cross section is $`\frac{1}{2}`$ of the $`\pi ^+\pi ^{}`$ one due to the symmetry factor . The $`K^+K^{}`$ and $`K^0\overline{K}^0`$ production cross section appears at energies higher than that of the resonances and hence do not show the resonance structure. Yet, final state interaction is very important and increases appreciably the $`K^+K^{}`$ production cross section for values close to threshold with respect to the Born approximation.
It is interesting to notice that the $`f_0`$(980) resonance shows up as a peak in the reaction. This is in contrast to the cross section for $`\pi \pi \pi \pi `$ in $`I=0`$ which exhibits a minimum at the $`f_0`$ energy because of the interference between the $`f_0`$ contribution and the $`\sigma (500)`$ broad resonance.
However, we should bear in mind that we have plotted there the contribution of the $`f_0`$ resonance alone. The tree level contact term and Bremsstrahlung diagrams, plus other contributions which would produce a background, are not considered there.
In any case it is interesting to quote in this respect that a related reaction from the dynamical point of view, which also involves the interaction of two mesons in the final state, the $`\varphi \pi ^0\pi ^0\gamma `$ and the $`\varphi \pi ^0\eta \gamma `$ decay, which have been measured recently at Novosibirsk , show clearly the $`f_0`$ and $`a_0`$ excitation, respectively, in the invariant mass spectra of the two mesons. A theoretical study along the lines reported here has been done in where a good description of the experimental spectra as well as the absolute rates is obtained.
## 5 Summary
We have reported on the unitary approach to meson meson and meson baryon interaction using chiral Lagrangians, which has proved to be an efficient method to extend the information contained in these Lagrangians to higher energies where $`\chi PT`$ cannot be used. This new approach has opened the doors to the investigation of many problems so far intractable with $`\chi PT`$ and a few examples have been reported here. We have applied these techniques to the problem of photoproduction of scalar mesons $`f_0`$(980), $`a_0`$(980) and the photoproduction of the $`\mathrm{\Lambda }(1405)`$, a resonant state of meson baryon in the $`S=1`$ sector and have found signals which are well within measuring range in present facilities. The experimental implementation of these experiments confronted with the theoretical predictions will contribute with new tests of these emerging pictures implementing chiral symmetry and unitarity, which for the moment represent the most practical approach to QCD at low energies.
## Acknowledgments.
This work is partly supported by DGICYT, contract number PB 96-0753. |
no-problem/9911/astro-ph9911184.html | ar5iv | text | # Investigations of Positron Annihilation Radiation
## I Introduction
One of the primary objectives of the OSSE instrument on NASA’s COMPTON Observatory has been to understand the nature of galactic positron annihilation radiation. Through 8$`\frac{1}{2}`$ years of observations, 511 keV line emission has been detected, but the emission has never been unambiguously attributable to a given discrete source. The galactic center (GC) region’s 511 keV emission was monitored by the Gamma-Ray Spectrometer on-board the Solar Maximum Mission (SMM) (1980-1988), the Transient Gamma-Ray Spectrometer (TGRS) on-board the WIND mission (1995-1997) and with multiple OSSE observations (1991-present). None of these detections require variable sources in addition to the two component models discussed here to explain the measured fluxes (Share et al. 1990, Harris et al. 1998, Purcell et al. 1997). These results have supported the suggestion that the majority of the emission is diffuse. Reported here are preliminary results of the extension of the OSSE analysis into three new areas: the inclusion of observations in regions with no $`apriori`$ expectations of positron annihilation radiation, the extension of the analyzed region to include a larger fraction of the Galaxy, and the mapping of the positronium continuum component (PCONT) of the total annihilation emission. The derived PCONT flux values have a stronger dependence upon fitting the underlying continuum than do the 511 keV line flux values, and thus detailed analyses (model-fitting & 1D cuts) are only performed upon the 511 keV data-set in this preliminary presentation.<sup>2</sup><sup>2</sup>2Among the potentially important effects not yet addressed are emission from the diffuse cosmic-ray continuum and a correction for scan-angle dependent background.
Previously published OSSE results have focussed upon the 511 keV line emission emanating from the central radian of the inner Galaxy, mapping emission from $`|l|`$ $``$ 33, $`|b|`$ $``$ 17, and model-fitting on a $`|l|`$ $``$ 90,$`|b|`$ $``$ 45, 1x1 grid. In a result first reported at the Fourth Compton Symposium and based on 6+ years of OSSE data, Purcell et al. 1997 (hereafter PURC97), showed evidence for three components to the 511 keV line emission, (1) an intense slightly extended emission centered in the direction of the GC, (2) a fainter planar emission, and (3) an unexpected enhancement of emission from positive latitudes (PLE). To generate that data-set, two basic types of data were used; (1) standard & offset pointing data where the scan angle crossed the GC and/or was perpendicular to the galactic plane, and (2) mapping data which searches for emission by observing a large sky region at regularly spaced intervals along a scan path.<sup>3</sup><sup>3</sup>3See PURC97 for details of the OSSE pointing strategies and background techniques. The live-time from mapping observations is spread over many more pointings (16 or 32 for mapping versus 3-9 for standard), so the sensitivity per source pointing achieved is inferior to the standard or offset observations.
The present work relaxes all selection criteria, initially including all observations whose source and background pointings are within the $`|l|`$ $``$ 90, $`|b|`$ $``$ 45 region. To extract the positronium component from the total emission, three spectral models have been fit to each spectrum (from 60 keV to 700 keV); (1) a single power-law +511 keV line + PCONT (as fit by PURC97, though they fit from 50 keV -4 MeV), (2) a power-law with an exponential fall-off +511 keV +PCONT, (3) a thermal bremsstrahlung model +511 keV +PCONT. If the best-fitting model (of the three) is deemed to provide an acceptable fit, then that result is selected for subsequent mapping analysis. This has been done for the 1153 observations in the data-set. Fewer than 20 observations have been rejected in this preliminary, all-inclusive analysis. The combination of including archival data and the use of new data collected in the 2+ years since the PURC97 paper has increased the total GC exposure from 1.9 x 10<sup>7</sup> det$``$s to 8.6 x 10<sup>7</sup> det$``$s, as seen in Figure 1. As a result of this exposure, the fraction of the GC region ($`|l|`$ $``$ 90, $`|b|`$ $``$ 45) mapped above our exposure threshold has increased from 22% to 83%.
The combined OSSE/SMM data-sets are used rather than the OSSE-only data-set because the OSSE background-subtraction technique leads to differential rather than absolute fluxes. OSSE is insensitive to both isotropic emission and modest intensity gradients. The SMM fluxes are not absolute either, being insensitive to isotropic emission. Assuming isotropic emission to be zero, the SMM data contributes an overall normalization (Share et al. 1988). The SMM FoV is wide ($``$130 FWHM), but it does contribute limited spatial information as the response peak swept through the GC region along the ecliptic. One important difference between the PURC97 data-set and this one is the use of TGRS data. PURC97 used the TGRS data from Teegarden et al. (1996). That data has since been re-analyzed by Harris et al. (1998), with the resultant 511 keV line flux reduced by almost 20%. The Harris et al. (1998) data-set is used for the model-fitting studies shown in Table 1. SVD maps of the combined OSSE/SMM/TGRS data-set were not ready for these proceedings, but a RL map of the OSSE/SMM/TGRS data-sets (not shown) is in general agreement with the OSSE/SMM map.
## II Mapping Positron Annihilation Emission
Two techniques have been employed to map the OSSE data-set: minimizing $`\chi ^2`$ with an adaptation of the Richardson-Lucy Algorithm (RL), and response matrix inversion using truncated Singular Valued Decomposition (SVD). SVD was described in PURC97, RL will be described in detail in an upcoming work.<sup>4</sup><sup>4</sup>4In short, RL adds flux distributed according to the instrument response to the source/background regions to raise/lower individual flux values to match each observation. Through successive iterations, map structure develops and the overall $`\chi ^2`$ lowers. Differences between the two resulting maps suggest the level of the uncertainties involved.
The upper and middle panels of Figure 2 shows the SVD and RL 511 keV maps of the combined OSSE and SMM data-sets. Both maps show three principle features, intense emission centered near the GC (hereafter called bulge emission), a fainter planar emission (hereafter called disk emission), and emission from the negative longitude/positive latitude region (hereafter called a PLE). The PLE is more pronounced in the SVD map than in the RL map. The hatched region does not meet a minimum exposure threshold (or maps negative intensity). We emphasize the level to which the emission is concentrated in these three components. Although a far larger fraction of the inner radian is mapped than in PURC97, there is no indication of intense emission from any of these newly mapped regions. The apparent dominance of the bulge and disk emissions along with a contribution from the PLE drive the use of two and three component models when model-fitting.
Shown in the lower panel of Figure 2 is a RL map of the fitted positronium continuum fluxes of the OSSE data-set. In most astrophysical environments, more PCONT photons are produced in annihilation events than 511 keV line photons. For a positronium fraction of 0.95, the PCONT:511 ratio is 3.7:1.<sup>5</sup><sup>5</sup>5OSSE has measured f<sub>Ps</sub>=0.97$`\pm `$0.03 at the GC (Kinzer et al. 1996). This value is consistent with the TGRS measured value, f<sub>Ps</sub>=0.94$`\pm `$0.04 (Harris et al. 1998). As a result, the PCONT map is more intense. As explained in the introduction, this map is preliminary as a number of potential biases have not yet been addressed. Nonetheless, the dominance of intense bulge and fainter disk components appears to agree with the 511 keV maps. The principal difference is the lack of evidence of PLE emission, as will be discussed in the next section.
## III Longitude and PLE Cuts
A measure of the planar structure in the 511 keV emission can be seen by taking a cut along the galactic plane of the RL and SVD maps. Cuts for the RL and SVD maps, as well as a model combining an R<sup>1/4</sup> bulge and the DIRBE 100 disk are shown in the upper panel of Figure 3. Each is the sum of the $`|b|`$ $``$ 2 pixels. The SVD data is plotted with 1$`\sigma `$ error bars per degree. The SVD and RL 511 maps show rough agreement, except in the +18 to +27 region, where the peaks and valleys are exaggerated in the SVD map relative to the RL map. The model is more centrally peaked and the centroids of the maps are slightly offset towards negative longitude, but general agreement exists between the model and the maps. A systematic survey of the inner galactic plane scheduled for CGRO Cycle 9 will improve the sensitivity of the longitudinal cut.
The lower panel of Figure 3 shows a cut through the PLE at an angle of 60 relative to the negative-longitude galactic plane (an angle suggested by the SVD map). As seen in the longitudinal cut, the R<sup>1/4</sup> shape is more centrally peaked. In the anti-PLE direction, all three maps agree, all falling smoothly. Both the SVD-511 and the RL-511 are brighter in the +3 $``$ +8 region than the corresponding negative region. Beyond +8, the sensitivity becomes poor, and there is little significance to the differences between the SVD and RL maps. The PCONT map (not shown) does not show a corresponding enhancement, though interpretation of this result as being due to annihilation physics, or alternatively being due to fitting systematics is not justified in this preliminary analysis.<sup>6</sup><sup>6</sup>6The “annihilation fountain” model (Dermer & Skibo 1997) would have a low positronium fraction due to the high temperature, as direct annihilation with free electrons dominates over radiative recombination above 10<sup>6</sup>K (Bussard, Ramaty & Drachman 1979). However, the absence of broad emission in the TGRS spectra provides a constraint to this scenario (Harris et al. 1998).
## IV Model-fitting the Galactic Center Region
The maps of 511 keV line emission support the PURC97 representation of the galactic emission as being due to bulge, disk and PLE components. To quantify the contributions by the bulge and disk components, three bulge shapes have been combined with 28 disk shapes and compared with the OSSE, then the OSSE/SMM, then the OSSE/SMM/TGRS 511 keV data-sets. The bulge shapes tested are; a GC point source, a Gaussian, and the projection of a truncated R<sup>1/4</sup> function. Shown in the left panel of Figure 4 are the chi-squared values for best-fitting bulges of each shape paired with 28 disk models and fit to the OSSE data-set. The disks have been ordered by the equivalent latitude FWHM of a Gaussian profile.<sup>7</sup><sup>7</sup>7The map references are: DIRBE maps (Hauser et al. 1998), the CO map (Dame et al. 1987), the Hot Plasma (Koyama et al. 1989), the R<sup>1/4</sup>, HF Light & Disk Light (Higdon & Fowler 1987), the M31 maps (Ciardullo et al. 1987). “90x10”, etc. refer to the FWHM of Gaussian disks. In the fits, the FWHM of the Gaussian bulges have been permitted to vary (best-fit FWHM values range from 3.9 to 5.7). The R<sup>1/4</sup> radial function has been truncated to a constant value inside of a radius (R<sub>min</sub>).<sup>8</sup><sup>8</sup>8The R<sup>1/4</sup> function is $`\rho `$(R $``$ 0.24) = A$``$R<sup>-6/8</sup>$``${exp(-B$``$R<sup>1/4</sup>) }, $`\rho `$(R $``$ 0.24) = 5/4$``$A$``$R<sup>-7/8</sup> {exp(-B$``$R<sup>1/4</sup>) -C$``$R<sup>-1/4</sup>\], where R = radial distance from GC in kpc, and A,B,C are constants. R<sub>min</sub> has been constrained to be between 50 pc and 700 pc. The radial distribution has then been projected onto the line of sight assuming the GC to be 8 kpc distant. For all R<sub>min</sub>, the R<sup>1/4</sup> has extended wings relative to the Gaussian shape; for small R<sub>min</sub>, the R<sup>1/4</sup> is also more centrally peaked. For every disk tested, the extended bulges are strongly favored over the GC point source. The R<sup>1/4</sup> bulge shows a slight preference for thin disks, while the Gaussian bulge shows a stronger preference for thicker disks. This is interpreted as a hint of emission separated from both the galactic plane and the GC. For the R<sup>1/4</sup> bulge shape, the extended bulge wings account for this emission. For the Gaussian, a thicker disk is required.
In the right panel of figure 4, the SMM data is combined with the OSSE data. The fits for the R<sup>1/4</sup> solutions are similar to the OSSE-only plots. However, only the thicker disk-Gaussian bulge solutions are able to approximate the quality of the R<sup>1/4</sup> fits. As seen in Table 1, the B/D ratios are larger for R<sup>1/4</sup> solutions than for Gaussian solutions. The total flux of the R<sup>1/4</sup> solutions change little when SMM and then TGRS data is added, but for the Gaussian solutions the total flux can change considerably. For many Gaussian bulge-disk combinations, the OSSE data would not permit insertion of the SMM or TGRS-required flux without violating OSSE constraints, leading to poor fits. These solutions do not necessarily span the range of possible bulge-disk combinations, nor are they unique, but they do suggest ranges of plausible B/D ratios and total fluxes. The uncertainties of the parameters are not shown in Table 1 due to concern about them being misinterpreted. The standard approach is to fix all but a single parameter, and calculate the degradation to the fit that results from varying that single parameter. The uncertainties that result do not account for the effect of varying the other parameters, nor does it account for other potential model shapes. In this paper, the uncertainties are evident in Table 1, but are not calculated.
The results shown in Table 1 have ignored the existence of a PLE and its potential influence upon the B/D and F<sub>Tot</sub> parameters. PURC97 quantified the PLE flux by two methods; (1) by subtracting the mirror region from the outputs of the mapping techniques, and (2) by fitting the data-set with three components (bulge, disk and PLE), all with Gaussian profiles. The two methods yield very different PLE flux results. The PURC97 maps suggest the PLE flux to be (1.3-1.5) x 10<sup>-4</sup> phot cm<sup>-2</sup> s<sup>-1</sup>. The 3-Gaussian fitting suggests (5.4-8.8) x 10<sup>-4</sup> phot cm<sup>-2</sup> s<sup>-1</sup>. The current maps suggest (1.2$`\pm `$0.5) x 10<sup>-4</sup> phot cm<sup>-2</sup> s<sup>-1</sup>, in agreement with the PURC97 maps. Quantifying the PLE by mirror-region subtraction, the B/D ratios for the RL and SVD maps are 0.63 and 0.49, as shown in Table 1. The B/D values are intermediate to the R<sup>1/4</sup> and Gaussian solutions, suggesting a general agreement between mapping and two component modeling.
The PURC97 3-Gaussian solution, when inserted with quoted parameters, is not an acceptable solution for the current OSSE/SMM/TGRS data-set ($`\chi ^2`$/$`\alpha `$ = 1.34), but the same method can be applied to examples of better-fitting models from Table 1. When a Gaussian representing the PLE is added to a 4.3 Gaussian bulge + 60x15 Gaussian disk, and the B/D re-optimized, the B/D rises from 0.163 to 0.169 ($`\mathrm{\Delta }\chi ^2`$=-16.4). The PLE centroid for this Gaussian fit is determined to be ($`l,b`$) = -2, +8, and the PLE flux is 1.1 x 10<sup>-4</sup> phot cm<sup>-2</sup> s<sup>-1</sup>. When a PLE Gaussian is added to the R<sup>1/4</sup> +CO Smooth model, the B/D lowers slightly from 3.3 to 2.6 ($`\mathrm{\Delta }\chi ^2`$=-12.3), and the parameters become: ($`l,b`$) = -2, +1, PLE flux = 0.7 x 10<sup>-4</sup> phot cm<sup>-2</sup> s<sup>-1</sup>. These values are given not to suggest a refinement of the PLE, but rather to demonstrate that the B/D depends more on the bulge shape than on the PLE (and vise versa). The current modeling of the PLE lowers the flux values to better agreement with the mapping fluxes, but the interpretation must be that the characteristics of any possible PLE are too poorly constrained and too model dependent to claim flux values and emission centroids (and the corresponding uncertainties).
## V Discussion
OSSE investigations of galactic positron annihilation have recently been expanded into three new areas; utilization of archival data near the GC, utilization of archival data away from the GC and mapping the positronium continuum emission. This expansion makes the results more dependent upon the spectral analysis, which is currently in its preliminary stage. The portion of this analysis least dependent upon these new complications are the 511 keV line flux values. Although the spatial coverage of maps from these values has increased, characterization of the emission with three components (bulge, disk and PLE) remains adequate. An extended bulge emission has been shown to be favored over point-like emission from the GC for a collection of 28 disk models. Comparing two approximations of the bulge shape, Gaussian bulges require thick disks to fit the data-sets, while R<sup>1/4</sup> bulges are less dependent upon the disk thickness, slightly favoring thin disks. The bulge shape must be better resolved with future measurements to improve the constraints upon the B/D ratio (currently ranging from 0.2 -3.3 for the combinations tested). The total flux is better constrained, ranging between (20.9-31.4) x 10<sup>-4</sup> phot cm<sup>-2</sup> s<sup>-1</sup>. The effect of a potential PLE upon these parameters is shown to be less than the bulge-shape uncertainty when the PLE is approximated by a Gaussian.
Positron astronomy is 30 years old but remains in its infancy. Of current generation gamma-ray telescopes, OSSE is the best-suited to investigate the sky distribution of this emission. The time allotted to OSSE annihilation studies has increased in recent cycles of the CGRO mission. This time, combined with the expanded OSSE analysis, and the impending launch of the INTEGRAL telescopes should produce significant improvements of our understanding of the nature of galactic positron annihilation. |
no-problem/9911/astro-ph9911123.html | ar5iv | text | # The origin of the highest energy cosmic rays Do all roads lead back to Virgo?
## I Introduction
The origin of the highest energy particles observed in the universe continues to present a major enigma to physics. These particles reach energies as high as $`\mathrm{3\hspace{0.17em}10}^{20}`$ eV. The flux of such nuclei is expected to drop sharply at $`\mathrm{5\hspace{0.17em}10}^{19}`$ eV due to the interaction with the microwave background, commonly referred to as the GZK-cutoff after its discoverers . However, the number of particles known to be beyond $`10^{20}`$ eV continues to increase, with now 14 published, and a further 10 expected from new observations with HIRES and a reanalysis of the Yakutsk data .
There are three basic difficulties: First, we need to find a site that can either accelerate particles to such energies, or produce them outright through the decay of even higher energy particles (topological defects for example, ). While there are many sites which are in principle capable to produce such particles , there is only one class that has been argued to require protons at such energies in the source, namely radio galaxies with powerful jets and/or hot spots, such as M87 or Cyg A. M87 has been under suspicion to be the primary source for ultra high energy cosmic rays for a long time . For both compact and extended radiojets as well as hot spots such as in M87 the emission of synchrotron radiation with a steep cutoff at frequencies about $`\mathrm{3\hspace{0.17em}10}^{14}`$ Hz implies an initial turbulence injection scale of the Larmor radius of protons at $`10^{21}`$ eV. Radio galaxies of sufficient power can provide through acceleration sufficient momentum to particles to overcome losses within the space limitation and the source magnetic and radiation fields . Gamma ray bursts are another possible class.
Secondly, there is the additional difficulty of getting these particles to us, that is, overcoming the losses in the bath of the cosmological microwave background . That implies that the source should not be very much further than 20 Mpc, and that corresponds to the travel distance for the presumably charged particles, not necessarily as short as the light travel time distance. There are only two certain candidates, the radio galaxy M87 at about 20 Mpc, and the radio galaxy NGC315 at about 80 Mpc - this latter distance is hard to accept already. Hence, we require that not much additional travel time should be spent through scattering in large scale magnetic irregularities .
The third difficulty is the explanation of the nearly isotropic distribution of the arrival directions of these events. The anisotropy expressed in a weak correlation with the supergalactic plane, or the detection of multiples mostly along the supergalactic plane is a small effect. The arrival directions are best described as isotropic, at least in the accessible sky in the North. There are two extreme explanations of the isotropy, one is to argue that we have many sources contributing, even at the highest particle energies, and the other is that we have one source dominating at the maximum energies, and the particle orbits are bent. In the second option the bending should not add substantially to the travel time.
There are several ways out of these difficulties. The highest energy particles could be particles generated by topological defects or from extensions of supersymmetry . Alternatively, they could be cosmologically local . If many sources contribute to the observed particle fluxes they have to be fairly common in space, as in the scenarios of accelerating such particles in the environment of quiescent black holes , or gamma ray bursts .
Here<sup>*</sup><sup>*</sup>*inquiries to plbiermann@mpifr-bonn.mpg.de we present the consequences of introducing a magnetic Galactic wind in analogy to the solar wind. The magnetic field of the wind bends the particle orbits without adding substantial travel time.
## II A model for a magnetic galactic wind
It has long been expected that our Galaxy has a wind akin to the solar wind . Recent modelling shows that such winds can be quite fast, and ubiquituous.
It seems plausible that this wind is powered by the combined action of cosmic rays and magnetic fields and starts in the hot phase of the interstellar medium seen in X-rays by ROSAT. This phase has a density of $`\mathrm{3\hspace{0.17em}10}^3`$ particles per cc, a temperature of about $`\mathrm{4\hspace{0.17em}10}^6`$ K, a radial scale of about 5 kpc, and an exponential scale in $`z`$, perpendicular to the disk, of almost 2 kpc . We note that the corresponding equipartition magnetic field strength is 10 microGauss.
Parker has shown that in a spherical wind the poloidal component of the magnetic field becomes negligible rather quickly with radius $`r`$, the radial component decays with $`1/r^2`$, but the azimuthal part of the magnetic field quickly becomes dominant with $`B_\varphi \mathrm{sin}\theta /r`$ in polar coordinates.
An available measure of magnetic field along any line of sight is the Rotation Measure: It is proportional to the line of sight integral of the product of electron density and magnetic field component (including the sign) from us to a distant linearly polarized radio source. We verified that the magnetic field topology of the Parker model is consistent with the data for a base density below that of the hot interstellar medium.
Cosmic ray driving is similar to radiation driving of winds in massive stars, and so magnetic fields can lead to an increase of the momentum of the wind. For a steady wind and Parker topology the Alfvén speed in the wind becomes independent of radius $`r`$ for large distances. The ultimate wind velocity is then a small multiple of the Alfvén velocity, suggesting a rather strong magnetic field. On this basis we have adopted here a magnetic field somewhat higher than in other current models .
The data on the sign of the azimuthal component show that in the disk of the Galaxy there are reversals but in the direction of the anti-center, the part of the sky most relevant for calculating orbits of energetic charged particles, the field points to the direction of galactic longitude about 90 degrees . That means immediately that positively charged particles traced backwards have their origin above us, at high positive galactic latitudes.
There is one additional observational argument in favor of such a topology of the magnetic field: Krause & Beck have found that the dominant magnetic field symmetry in spiral galaxies points approximately along the local spiral arms inwards towards the center. This effect is unexplained at present, but it seems nevertheless to be observationally well established. One might expect that magnetic field sign reversals in the disk should translate into similar reversals in the wind. However, if the dominant directionality is established by the most active spirals arms, then the symmetry as found by Krause & Beck should prevail well outside the disk. We will assume this to be the case. This symmetry implies an overall current in the Galaxy possibly driven by star formation through stellar winds and cosmic rays.
If the entire disk has this symmetry, then there cannot be a sign reversal of the azimuthal component close to the disk along the direction perpendicular to the disk. This implies that the radial magnetic field component $`B_r1/r^2`$ is pointed inwards inside the disk and in its immediate neighborhood below and above. However, considering all of $`4\pi `$ the radial component has to be pointed outwards roughly over half the sky, and pointed inwards in the other half in order to satisfy the condition of a source free magnetic field. Thus, one possible and simple configuration is that $`B_r`$ is pointed inwards within 30 degrees of the disk both above and below the disk, and pointed outwards within 60 degrees of both poles. Such a pattern is different from the Solar wind, where the components change sign in mid-plane.
We adopt the simplest possible model. We assume that the magnetic field in the galactic wind has a dominant azimuthal component, and ignore all other components. We assume that this azimuthal component has the same sign everywhere. This then means that the strength of this azimuthal component at the location of the Sun is the first key parameter, and the distance to which this wind extends is the second one. Since the irregularities in magnetic fields should decay into the halo, the total magnetic field strength near the Sun may be used as an approximation to the local halo wind field, and in fact one might expect a lower limit. Most measures of magnetic field underestimate its strength. Therefore we will consider for reference a model which has a field strength near the Sun of 7 microGauss . The second parameter, the distance to which this wind extends, is more uncertain: Our Galaxy dominates its near environment well past our neighbor, M31, the Andromeda galaxy, and might well extend its sphere of influence to half way to M81. Therefore we will adopt as outer the halo wind radius half the distance to M81, 1.5 Mpc.
## III Tracing the path backwards
To follow the particle trajectories in the Galactic halo we trace protons backwards from their arrival direction at Earth. We use the 14 published cosmic ray events above 10<sup>20</sup> eV, the list from Watson (included in ) and the new list from AGASA . There is a big uncertainty with the energy estimate of the highest energy Yakutsk event, which we therefore exclude from the present analysis, and hence we arrive at a final tally of 13 events used.
Fig. 1 shows the directions of the events at that point when they leave the halo wind of our Galaxy in a polar projection. For reference we show the direction to the active galaxy M87 (Virgo A), the dominant radio galaxy in the Virgo cluster. We show the two highest energy events twice: under the assumption, (i) that they are protons, and (ii) that they are Helium-nuclei (filled black symbols). We also show the supergalactic plane as a shaded band.
The interesting result of these model calculations is that the directions of all tracks point North. With the exception of the two events having the highest energy, all other 11 events can be traced to within less than about 20 degrees from Virgo A. Considering the uncertainty of the actual magnetic field distribution, we find then that all events are consistent with arising originally from Virgo A. Since these particles are assumed to be accelerated out of cosmic gas, about 1/10 of all particles may be Helium nuclei with the same energy per particle. If the two highest energy events are in fact He nuclei, all 13 events point within 20 of Virgo A.
If Virgo A is indeed the acceleration site of the highest energy cosmic ray events, they all require systematic bending at a ten to twenty degree level. Such bending could be easily accomodated within the plausible magnetic field strength within the supergalactic sheet from here to Virgo . Bending by 20 degrees on a pathlength of 20 Mpc implies a regular transverse magnetic field strength of order 2 nanoGauss for 10<sup>20</sup> eV protons. In fact, the expected total magnetic field strength in the supergalactic plane is significantly higher .
How critical are our assumptions for these results?
$``$ The assumption of the symmetry of the magnetic field above and below the Galactic disk is important; if the magnetic field were to have opposite sign above and below the disk, then all those events that arrive below the disk, would in fact go down and point in the general direction of the Galactic South pole.
$``$ The value of the magnetic field, here adopted as 7 microGauss, for the wind near the Sun, is a key parameter. A decrease of this value by up to $``$ 30% would not change essentially the results. If the magnetic field were considerably weaker, however, of order one microGauss, then the effective pointing to a common direction of origin would be largely removed. However, it would still be sufficient to smear arrival directions, especially if its structure is regular.
$``$ The scale of the Galactic wind here 1.5 Mpc, is not a critical parameter, since the calculations show that most of the bending happens within the first few 100 kpc.
## IV Discussion and implications
This particular model can be tested in several ways:
It has been suggested repeatedly in the literature, that radio galaxies may inject a major fraction of the total energy output (of order 0.1) in energetic particles . In order to determine this fraction for M87 we need a model for the magnetic field structure along the supergalactic sheet, and test it with orbit calculations.
Second, if the appearance of pairs and triplets of events along the supergalactic sheet is confirmed, then clearly the magnetic field structure along the sheet may be the reason in that it gives rise to some caustics with an enhancement just along the sheet direction. Again, orbit calculations will be able to test this.
Third, the very concept of a magnetic wind, driven by cosmic rays, but with an initial magnetic field as strong as in the disk, needs to be examined more closely. The conditions to be met are a) the available energetics, b) agreement with the Rotation Measure data, c) the condition that the wind be super-Alfvénic, and d) that the wind does extend to large scales. The model for Wolf Rayet star winds gives us some confidence, that this is indeed reasonable. Since this should be true for any galaxy with a high star formation rate and therefore, presumably with a high rate of cosmic ray production, this latter condition may be the easiest to test with sensitive absorption line data at high redshift, with observations of the shell of gas around these extended winds.
Fourth, the current experimental statistics cover only a fraction of the Southern sky. We would expect some asymmetry between the Northern and Southern hemispheres of our sky. How much asymmetry depends again on the more detailed magnetic field configurations, but the data from the Auger observatory will clearly provide stringent conditions on this as on any other model.
Fifth, if the model proposed here could be confirmed, then it would constitute strong evidence that all powerful radiogalaxies produce high energy cosmic rays, and that they do this at a level of $``$ 0.1 of their total power output. This then implies that compact radio galaxies do provide a good test bed for particle interactions, since they have a large screen of interstellar gas around the radio hot spots and jets as seen in mm-wavelength radio data . There the paradigm is materialized of having a gigantic accelerator, and a beam dump. These radio galaxies may be used for particle interaction experiments in the sky. If a significant correlation in arrival direction between ultra high energy cosmic rays and this specific class of radio quasars could be confirmed , then properties of new particles could be constrained.
In summary, we propose here that a very simple model for a Galactic wind rather analoguous to the Solar wind, may allow particle orbits at $`10^{20}`$ eV to be bent sufficiently to allow “super-GZK” particles to get here from M87, and also explains the apparent isotropy in arrival directions.
Acknowledgments. This work was started at the 1999 astrophysics summer school at the Vatican Observatory; both EJA and PLB are grateful to George Coyne, S.J. and his colleagues for inviting us and for the fruitful and inspiring atmosphere at these schools. PLB, GMT and TST would like to acknowledge also useful discussions with Rainer Beck, Pasquale Blasi, E. Boldt, Glennys Farrar, J.L. Han, Anatoly Ivanov, Randy Jokipii, Frank Jones, Hyesung Kang, Phil Kronberg, Martin Lemoine, Friedrich Meyer, Hinrich Meyer, Biman Nath, Dongsu Ryu, Michael Salomon, Günther Sigl, Alan Watson, Arnold Wolfendale and many others. We thank Phil Kronberg for a careful reading of the manuscript. Work with PLB is partially supported by a DESY-grant, GMT is partially supported by the Brazilian agencies FAPESP and CNPq, TST is supported by NASA grant NAG5–7009, and PLB and TST have jointly a grant from NATO. |
no-problem/9911/astro-ph9911259.html | ar5iv | text | # IAU 177 — A week in review
## 1. Introductory Remarks
The onerous task of summarising meetings such as these falls traditionally on some poor unsuspecting soul. Having accepted this duty (and I still can’t believe I ever agreed to it!) it’s now time to look back on what has been one of the most enjoyable meetings I’ve had the good fortune to attend.
As always on these occasions, the conference summariser is given a free rein to review the meeting as he/she feels appropriate. As we all know, this often results in a significant deviation from the “ideal conference summary” — an unbiased account of the proceedings. This review is no exception and, as is the tradition, I shall make it very clear from the start that the following paragraphs only scratch the surface of the many varied topics discussed at this meeting. I choose to concentrate on the topics I was most interested in, and the new results that really caught my attention. As such, the many excellent presentations on pulsar emission theory are therefore not mentioned here (see however the preceding review by Melrose). My apologies to all such participants who do not get a mention in this review. This merely reflects my personal bias as an observer who takes the “black-box” approach to pulsars and life in general.
In all we listened to 95 contributed talks and pondered over 180 posters. Topics covered included pulsar surveys, the interstellar medium, anomalous X-ray pulsars, magnetars, single pulse studies, radio emission phenomenology, pulsar timing, supernova remnants, interactions with companion stars, high energy observations, neutron star emission theories, fundamental astrometry, general relativity, neutron star demography, plasma physics and the future of the field. Many of us also heard the excellent public lecture given by Jocelyn Bell-Burnell (a tour de force presentation of astronomy to the general public) as well as additional evening sessions discussing orthogonal polarisation modes and magnetars. There is clearly much to talk about and, without further ado, I will begin.
## 2. Meeting Highlights
### 2.1. Pulsar Surveys
The week got off to a fine start with talks by Camilo and Manchester on the latest results from the Parkes Multibeam survey. With 13 $`\lambda `$ 21-cm 25 K receivers on the sky, along with $`13\times 2\times 288`$-MHz filterbanks, the system is presently making major contributions in a number of different pulsar search projects. In its main use for the Galactic plane survey described by Camilo, the system achieves a sensitivity of 0.15 mJy in 35 min and covers about one square degree of sky per hour of observing — far beyond the capabilities of any other system at present.
The staggering present total of 439 new pulsars from an analysis of just under half the total data leads the team to predict that the final body count should be over 600. Such a large haul is resulting in significant numbers of interesting individual objects: Several of the new pulsars are observed to be spinning down at high rates, suggesting that they are young objects with large magnetic fields. The inferred age for the 400-ms pulsar J1119$``$6127, for example, is only 1.6 kyr. Another member of this group is the 4-s pulsar J1814$``$1744, an object that may fuel the ever-present “injection” controversy surrounding the initial spin periods of neutron stars. Further studies of these objects to look for extended radio/optical emission (i.e. supernova remnants/bow shocks) will undoubtably help us to understand the birth properties of pulsars.
A number of the new discoveries from the survey have orbiting companions. Several low-eccentricity systems are known where the likely companion is a white dwarf star. Two probable double neutron star systems are presently known: J1811$``$1736 is in an 18-d highly eccentric orbit, while J1141$``$65 has a lower eccentricity but with an orbital period of only 4.75 hr. The fact that J1141$``$65 may have a characteristic age of just over 1 Myr implies that the likely birth-rate of such objects may be large. Whilst it is premature to start speculating on statistics of one object, it is clear that these binary systems and the many which will undoubtably come from this survey will teach us a lot about the still poorly-understood population of double neutron star systems.
The most massive binary system from the multibeam survey to date is J1740$``$3052, whose orbiting companion must be at least 11 M. As Manchester mentioned, recent optical observations reveal a K-supergiant as being the likely companion star in this system. With such high-mass systems in the Galaxy, not to mention the massive X-ray binary systems, surely it is only a matter of time before a radio pulsar will be found orbiting a stellar-mass black hole. Future searches for these elusive beasts in globular clusters, where long integration times are common, would do well to utilise Ransom’s novel technique to detect short orbital period binary systems.
We also heard during the meeting that the Parkes multibeam system has not only been finding young and distant pulsars along the Galactic plane. Two other search projects were described in the posters by Edwards et al. and Freire et al. Edwards et al. have been using all 13 beams to do a “quick” (5 min per pointing) search at intermediate Galactic latitudes ($`5^{}|b|15^{}`$). The discoveries of 8 “recycled” pulsars during this search, not to mention 50 long-period objects, strongly support a recent suggestion by Toscano et al. that an L-band search of intermediate latitudes is an excellent means of finding millisecond pulsars which, as they and others demonstrated (see Kramer’s review) now appear to have significantly flatter spectra than previously thought.
Freire et al., on the other hand, have used only the central beam to search for new pulsars in the globular cluster 47 Tucanae. This has revealed no less than 10 new binary millisecond pulsars, bringing the total number of pulsars in this cluster to 21! The new discoveries include a 95-min binary system. This is presently the shortest orbital period for a radio pulsar binary. Clearly these are exciting times for the pulsar hunting community and the coming months will undoubtably throw up further surprises.
### 2.2. Gravitational Wave Astronomy
While pulsar searches in the radio and X-ray regimes are enjoying a renaissance due to new instrumentation and advances in sensitivity, our colleagues in the gravitational wave community are about to open a whole new window on the Universe. As Schutz reported during his talk, a number of sensitive detectors are about to come on-line. Specifically, these are the two LIGO sites in the US, the GEO detector in Hannover (a joint German/UK project) and, later, the Italian VIRGO project. Possible events that are being targeted by these detectors include: (1) compact object mergers at cosmological distances; (2) continuous emission from spinning neutron stars (though detecting even nearby pulsars will challenge the limits of present sensitivity); (3) stellar collapse in the Galaxy. Millisecond pulsar timing still remains an attractive means of probing the long-wavelength regime, with space interferometers still some way off.
In anticipation of the data that will soon be pouring off these detectors, astronomers at various institutes around the world have been putting a tremendous effort into understanding the numerous signal processing/detection hurdles they face. As always, the main problem is one of sensitivity… astronomers are trying to measure events that produce a detector “strain” of order $`10^{23}`$ or less! In addition, as we heard, analyses of $`10^7`$ s time series are required with the initial detector systems. The requirement to deal with time series of this length is prompting novel techniques to enable even present state-of-the-art computers analyse the data within a Hubble time. Radio and X-ray pulsar hunters would do well to stay in touch with this exciting field for new tricks in the future.
### 2.3. Radio-Quiet Neutron Stars and Supernova Remnants
The so-called “radio-quiet neutron stars” were the subject of a number of presentations during this meeting. This term includes the soft gamma-ray repeaters (SGRs; thought to be magnetars) and the anomalous X-ray pulsars (AXPs) and the enigmatic Geminga pulsar (see below). One significant advance in the high energy observations has been in the timing analyses that are now possible. The new Rossi X-ray phase-coherent timing observations of two AXPs by Kaspi et al. provide a nice confirmation that these neutron stars are spinning down in a similar manner to most radio pulsars. Whilst there seems to be overwhelming evidence that the AXPs are young neutron stars at the centres of supernova remnants (see e.g. the contributions by Gotthelf and Gaensler) I certainly got the impression during some of the discussions that whilst the evidence associating SGRs with supernova remnants is tantalising, it is presently only circumstantial.
Part of the problem in making the link arises because of uncertainties in independently measuring the distances to the remnants and the SGRs. An interesting development in this regard is the detection of dispersed low-frequency radio pulses from SGR 1900+14 by Shitov and collaborators. The dispersion measure obtained implies a distance of 6 kpc for the most recent electron density model. Future combined radio and high-energy timing analyses to monitor the spin-down of this magnetar will be most interesting, particularly if e.g. regular radio observations reveal magnetar glitches. It presently remains a mystery why this pulsar and Geminga are not detected at higher radio frequencies e.g. 430 MHz, this is despite numerous observing campaigns at Arecibo, Jodrell Bank and the VLA (see Kassim & Lazio’s poster and references therein).
### 2.4. Pulsar Timing
A number of recent and interesting pulsar timing results were presented during the week. High-precision timing of millisecond pulsars continues to demand careful attention to understanding the properties of the telescope and data taking system in order to reap the wealth of astrophysical information that these clocks have to tell us. This was well highlighted during Britton’s study of the systematic effects of polarisation on the Parkes timing model residuals obtained for the bright, nearby millisecond pulsar J0437$``$4715. The clever use of the “invariant profile” (Stokes $`I^2Q^2U^2V^2`$, as opposed to the total intensity) in the timing analysis is an elegant means (for weakly polarised pulsars) to circumvent the many non-trivial steps in a proper polarisation calibration analysis. The quality of the residuals obtained via this relatively simple analysis should inspire other observers to try this technique as a means of improving their timing precision.
Recent results from timing observations that have (finally!) resumed after the Arecibo upgrade include further confirmation of the orbital decay of the binary pulsar B1913+16 predicted by general relativity (see Taylor’s paper), as well as a continuing study of the pulse profile evolution of this pulsar caused by geodetic precession of the orbit (see Weisberg’s paper). Post-upgrade Arecibo observations of the “planets pulsar” B1257+12, when combined with Effelsberg and pre-upgrade Arecibo observations, and a clever perturbation analysis, have permitted the orbital inclination of two of the planets and hence their absolute masses to be determined (see papers by Wolsczcan and Konacki).
On the subject of planets, a long-awaited update on the timing of the pulsar B1828$``$11 was presented by Lyne. The curious behaviour of this system was first reported by Bailes et al. in the proceedings of the “Planets around Pulsars” meeting back in 1992 where the timing residuals showed a strong periodicity indicative of a sum of three sinusoids that could be attributed to the Doppler shifting of the pulsar period by orbiting planetary bodies. The new results presented here show a strong correlation between pulse profile changes and the periodic oscillations seen in the timing residuals. Whilst this would rule out a planetary origin, an interesting twist in the story is that the phases of the sinusoids sum to a constant value — as seen in e.g. the Jovian satellites. Presently, the most plausible explanation seems to be free precession of the neutron star. If correct, this would be the first clear detection of such an effect. It remains a mystery why similar behaviour is not seen in other young pulsars.
The very useful contributions that can be made by smaller radio telescopes, which are being used to perform intense observing campaigns on selected objects, was highlighted several times during the week. Wolszczan and collaborators have now been performing regular timing observations with the 32-m telescope in Torun for several years. Their growing database of observations is being used to search for planets around pulsars, and to investigate pulsar scintillation (see papers by Lewandowski et al). The Crab pulsar still continues to surprise astronomers after almost 30 years of regular monitoring with the remarkable echoing events (ghost pulse profile components) seen in observations made using the 85-foot telescope at Green Bank by Backer and the 42-foot telescope at Jodrell Bank (Smith & Lyne). Current interpretations of the echos, which have now been seen on several occasions, are that they are caused by ionised shells drifting around the nebula, or intrinsic to the pulsar magnetosphere.
### 2.5. Neutron Star Demography
Significant progress in our understanding of the Galactic population of neutron stars is being made, and Cordes’ presentation summarised the present status of an ambitious attempt his group is making to model the population. A big improvement that this model has over previous studies is the use of a self-consistent beaming model. Assuming that the core-cone model of the radio beam is correct, and Desphande and Rankin’s polar cap maps seem to be confirming this, apparent pulse shapes can be simulated and compared to those that we observe. I look forward to seeing the full results of this study in the near future. A useful check of this model would be to simulate the Parkes multibeam survey.
As emphasised by Cordes, even non-detections from pulsar surveys are interesting results and pulsar hunters at all wavelengths should continue to write up their negative results (boring though the task of writing them may be!) for use in likelihood analyses. A good example in this regard is the modelling of the gamma-ray pulsar luminosity function based on OSSE and EGRET results presented by McLaughlin. This study provides a novel means of constraining a number of population parameters, such as the initial spin period of pulsars.
Our understanding of poorly-sampled subsets of the Galactic population, such as the double neutron star systems, is receiving a significant boost from the work presented by Kalogera. This population is presently dominated by two objects: the original binary pulsar B1913+16 and B1534+12. Kalogera and Narayan are presently addressing the burning question “how representative are these pulsars of the underlying sample of objects?” — a novel approach to this is to create a model population (whose underlying parameters are well understood) and estimate the size of the underlying sample based on a few objects. It turns out that there is a significant bias which can be derived from this model, and applied to the true sample of objects to infer the Galactic population. This is an important new result and could also be applied to other small samples.
One such application is the population of long-period pulsars. We heard from Young that the period of PSR J2144$``$3933, originally discovered in the Parkes Southern Sky Survey, is 8.5 s — three times that previously thought. This is presently the longest period for a radio pulsar. Young et al. make the valid point that such pulsars could be very numerous in the Galaxy since they have very narrow emission beams and therefore radiate to only a small fraction of the celestial sphere. Why this pulsar is radiating at all challenges some current neutron star emission theories and equations-of-state.
Theoretical studies concerning globular cluster pulsar demography and dynamics are beginning to become popular again. As we heard during Rasio’s talk, the recent flurry of new discoveries in 47 Tucanae and improved timing solutions are allowing more detailed studies to be carried out. The radial distribution of pulsars in this cluster out to a few core radii seems to be consistent with an isothermal sphere. 47 Tucanae is a cluster that has not yet gone through a core collapse phase and is supporting itself by “burning” binaries (thereby releasing kinetic energy) in the core. Among the questions posed is why there are so many short-period binaries in 47 Tucanae but relatively few low-mass X-ray binaries. In addition, there is a distinct absence of long-period ($`\stackrel{>}{_{}}3`$ day) binaries. Presumably, the latter objects get quickly disrupted during exchange interactions, which may in turn result in short-period binaries and solitary millisecond pulsars. Fossil evidence for such interactions may be the small, but significant, eccentricities now measurable for some of the binaries in 47 Tucanae.
## 3. The Future
This review has (shamelessly!) focused on radio pulsar astronomy — a classic summariser selection bias. Whilst the current status of this field is alive and well, giant leaps are anticipated in the future with the Square Kilometer Array. Funding and radio frequency interference permitting, there should be no shortage of tasks to be carried out in the near and long-term future of this field. High energy astronomy is already embarking on a busy period with the new generation of satellites, in particular CHANDRA, as we heard is already producing fantastic results. As already discussed, gravitational wave astronomy should be increasingly augmenting the science presented at these meetings in future as new detectors come on-line. In short, the future of pulsar astronomy looks well set.
## 4. And finally…
The friendly atmosphere which prevailed during the week inevitably included some lighthearted moments. Three particularly memorable quotes from the speakers are listed below for posterity:
The only good pulsar is a dead pulsar!” — G. Pavlov
We obtained a good fit to the data with only 18 free parameters.” — A. Somer
We have convinced ourselves that this will probably work.” — J. Cordes
Joking and back-slapping aside, I’m sure that these proceedings will serve as a useful testament to the events of this meeting and a snap-shot of the field as it stands. This is particularly important not only for researchers wanting to catch up on the latest results, but also for the continual influx of young researchers to pulsar astronomy, many of whom were present at this meeting.
Und dat, as they say in Rheinland, war es denn. I’d like to close by saying what a pleasure it was to participate in this meeting — not only because of all the hot scientific results being discussed, but also because of the efficient way in which it was organised and run throughout the week. Whilst many would expect nothing less from a conference held in Germany, nevertheless the standard set here will be hard to surpass in future meetings. On behalf of all of the participants, I’d like to thank all the members of the Local Organising Committee who put themselves at our disposal during this week. In particular, hearty thanks go out to Michael Kramer, Norbert Wex, Gabi Breuer, Ute Runkel and Richard Wielebinski, all of whom must have breathed a sigh of relief, perhaps even saying “nie wieder!”, when the meeting was over and the last participant had shuffled out of the building. They can pride themselves on a job well done and be assured that we are all eagerly looking forward to another pulsar IAU meeting in Bonn at some point early in the next millennium! |
no-problem/9911/astro-ph9911512.html | ar5iv | text | # The CANGAROO-III Project
## Introduction
Following the CANGAROO-I (3.8 m) and CANGAROO-II (7 m) telescopes, CANGAROO-III is a project to study celestial gamma-rays in the 100 GeV region utilizing a stereoscopic observation of Cherenkov light flashes with an array of four 10-meter telescopes. The CANGAROO-II telescope (hereafter C-II), which has a 7-meter reflector and has been operational since 1999 May Tanimori99 Mori99 Kubo99 , is going to be expanded in early 2000 by adding more small mirrors and will be the first 10 m telescope of this array.
The CANGAROO-III project started in April 1999 and is planned as a five-year program. The schedule is shown in Figure 1. This year we will expand the 7 m telescope to 10 m, and the second year we will build the second telescope which will be installed in the third year. The other two telescopes will be installed in the fourth and fifth years. Each telescope will be set on a corner of a diamond of about 100 m side in order to have a maximum number of pairs of telescopes of the same baseline length. The first stereoscopic observation will be performed in 2002 and the full four telescope will be in operation in 2004.
## Expansion of CANGAROO-II
Expansion of the 7 m telescope to 10 m is simple. Since C-II is originally designed as a 10 m telescope, all we have to do is add 54 mirrors and tune their attitude (Figure 2). This work will be completed in early 2000.
Additional outer mirrors will worsen the point image at the focus but it is not a serious problem. Simulations show that the concentration of photons in one pixel will be reduced from 56% to 42% at the center, and from 50% to 36% at one-degree off-axis. In any case, the number of collected photons will be almost doubled, reducing the energy threshold by a factor of two.
## CANGAROO-III Telescope Design
At this stage we will use basically the same design for the support structure and the driving mechanism as the C-II telescope, which is originally designed as a 10-meter telescope and has been proved to work well. The reflector will have a parabolic, composite mirror consisting of 114 small mirrors of 80 cm in diameter Kawachi99 . The focal length will be 8 m if we use the same mirrors.
Mirrors made of plastic laminates used for C-II are very light and pose little stress on the support structure. Observation of star images at various zenith angles showed the deformation of mirrors was negligible. But the image quality of these mirrors are not as good as glass-made mirrors since they are made by molding: thus we are still investigating other possibilities. The attitude of each mirror will be controlled by stepping motors as for the present 7 m telescope. Tuning this number of small mirrors to a common focus is not a simple task. For C-II we tuned the mirrors one by one using lids to cover the other mirrors, but this is not easy for larger numbers of mirror segments.
The prime focus camera will be similar to the present CANGAROO-II camera consisting of 512 half-inch photomultipliers and subtending about 3 degrees, but the optimization for stereoscopic observation is underway.
The electronics and data acquisition system will be improved to match higher data rates. In any case, we take timing information of each signal, in addition to pulse height, to utilize the isochronous nature of our parabolic reflector. For stereoscopic observation, we must introduce an inter-telescope trigger to compensate for geometrical delays using programmable delays between telescopes. The local triggers will be as frequent as 1 kHz but the delayed coincidences at the main trigger will be reduced to about 100 Hz, we hope.
## Stereo Simulation
Here we briefly show some results of simulations of stereo observations Hara99 . This work was done before the whole CANGAROO-III project was approved and takes only two telescopes into account, but the result is valid if we use a twofold coincidence in the inter-telescope trigger.
The detection efficiency as a function of baseline length between telescopes is given in Figure 3. If we cut some detected events using the core distance, the energy resolution will be better and an angular resolution less than $`0.1^{}`$ can be achieved if we use the baseline longer than 100 m. Thus we will adopt the baseline length of around 100 m, which agrees with other calculations. Figure 4 is a comparison of effective area and energy resolution between single and stereo observations.
## Expected performance
Expected sensitivity assuming one 10 m telescope shown in Figure 5 is around $`10^{12}`$ cm<sup>-2</sup>s<sup>-1</sup> above threshold energy of $`100200`$ GeV, and we may detect many EGRET sources in tens of hours of observation if their spectra extend to higher energies. Also shown are gamma-ray spectra of 22 X-ray selected BL Lacs which are predicted by Stecker et al. Stecker96 , however only 5 are in the southern hemisphere. We note that observations at other wavelengths, especially ground-based ones, are rather biased to the northern hemisphere sky and there are undoubtedly more candidate XBLs in the southern sky.
## Observation targets
Table 1 shows the list of objects observed by the CANGAROO 3.8m telescope for its 6 years of operation. We had been given preference to Galactic sources because of the rather high threshold energy ($`2`$ TeV) of the 3.8m telescope, but we may spend more time on extragalactic objects taking account of the lower threshold of new telescopes. One can see from the table we have needed more than 50 hours of observation at least as the necessary condition to conclude “positive detection” with sufficient statistics on the number of gamma-rays. In addition, the imaging Cherenkov technique still suffers from systematic errors which are not negligibly small when compared with gamma-ray signal strength from even “strong sources”, and careful estimation on the experimental errors is indispensable by using the data spanning over a long period of observation. We performed survey observations of shorter duration on many sources, which possibly provide a chance of time varying activities of episodic flares, as well as the objects like X-ray binaries which might be “strong sources” if the claims in earlier days are true. The prime efforts of CANGAROO-III will be on those types of objects appearing as top-ranked sources in the table, extending a systematic survey on more sources. In the case of sources of soft spectra, better statistics in the 100 GeV energy region will enable us to detect them in 10 to 20 hours of observation. However, we still have shortage of total observation time available during a year. It is necessary to develop world-wide efforts for more new types of high-energy gamma-ray sources in collaboration with other groups proposing next-generation telescopes.
## Summary
CANGAROO-III will start to explore the southern half of the 100 GeV gamma-ray sky in 2004, complementing projects located in the northern hemisphere to ensure the entire sky is covered at these energies. |
no-problem/9911/astro-ph9911272.html | ar5iv | text | # Be stars in X-ray binary systems
## 1. Introduction
The objectives of this paper are :
* To explain what High Mass X-ray Binaries (HMXBs) are, and, in particular, Be/X-ray binary systems.
* To discuss the general observational characteristics of the group.
* To present examples of how work on such systems has furthered our knowledge of Be stars.
## 2. General properties
The Be/X-ray systems represent the largest sub-class of massive X-ray binaries. A survey of the literature reveals that of the 96 proposed massive X-ray binary pulsar systems, 67% of the identified systems fall within this classes of binary. The orbit of the Be or supergiant star and the compact object, presumably a neutron star, is generally wide and eccentric. X-ray outbursts are normally associated with the passage of the neutron star through the circumstellar disk. The optical star exhibits H$`\alpha `$ line emission and continuum free-free emission (revealed as excess flux in the IR) from a disk of circumstellar gas.
The physics of accretion-powered pulsars has been reviewed previously (e.g. White, Nagase & Parmar 1995, Nagase 1989, Bildsten et al 1997). This paper will concentrate on the optical and IR observational properties of these systems and will address what such observations have contributed to our understanding of Be star behaviour.
X-ray behavioural features of these systems include:
– regular periodic outbursts at periastron called Type I;
– giant outbursts at any phase probably arising from a dramatic expansion of the circumstellar disk described as Type II;
– “missed” outbursts frequently related to low H$`\alpha `$ emission levels (hence a small disk), or other unknown reasons (eg perhaps centrifugal inhibition of accretion (Stella, White & Rosner 1986);
– shifting outburst phases (see below).
Progress towards a better understanding of the physics of these systems depends on a multi-wavelength programme of observations. From observations of the Be star in the optical and IR, the physical conditions under which the neutron star is accreting matter can be determined. In combination with hard X-ray timing and flux observations, this yields a near complete picture of the accretion process. It is thus vital to identify the optical counterparts to these X-ray systems in order to further our understanding.
Neutron stars are found in many astronomical configurations - see Figure 1 - and Be/X-ray binaries contain the second largest group of known neutron stars. The largest group comprises thousands of isolated rotation-powered pulsar systems.
Currently (September 1999) there are about 100 known or suspected HMXBs (see Figure 2). Surprisingly, nearly one third of all of these sytems lie in the Magellanic Clouds. This very large fraction, particularily noticeable in the Small Magellanic Cloud, will be discussed below in Section 7. Currently about one third of all the 100 systems have no known optical counterpart - frequently this is simply due to the inaccuracy of the X-ray observatory in locating the X-ray source.
There are two main sub-groups of HMXBS - the supergiant counterparts (normally of luminosity class I or II), and the Be/X-ray binary systems (normally luminosity class III or V). Both systems involve OB type stars and are commonly found in the galactic plane and the Magellanic Clouds. They differ, however, in accretion modes with the supergiant systems accreting from a radially outflowing stellar wind, and the Be/X-ray binaries accreting directly from the circumstellar disk (maybe with some limited Roche lobe overflow on rare occasions). As a result the supergiants are persistent sources of X-rays, whilst the Be/x-ray systems are very variable and frequently much brighter.
Recently Reig and Roche (1999) have suggested a third sub-group of systems; the X Persei like systems. The main characteristics of this proposed group are: long pulse periods (typically 1000s), persistent low X-ray luminosity ($`10^{34}`$ erg/s) and low variability, and rare uncorrelated weak X-ray outbursts. There are currently 5 such objects in this new group.
A summary of the observational properties of Be/X-ray binaries is :
– they were first discovered in 1974–5 (eg A1118-616 and A0535+26)
– there are currently 38 optically identified systems
– there are a further 23 systems proposed based upon their X-ray characteristics
– there are probably 100 – 1000 Be/X-ray binaries in our galaxy (Bildsten et al 1997), but maybe up to 10,000 (Rappaport & van den Heuvel 1982, Meurs & van den Heuvel 1989)
– binary periods lie in the range 16d – 400d
– pulse periods lie in the range 0.07s – 1413s
– the source names are primarily based upon their coordinates.
## 3. Evolution
The evolutionary history of Be stars in these Be/X-ray binary systems is somewhat different to that of their isolated colleagues. Figure 3 shows the widely accepted evolutionary path based upon conservative mass transfer that has been developed by van den Heuvel (1983) and Verbunt & van den Heuvel (1995). The important consequences of the scenario is that wide binary orbits (200 – 600d) are produced before the final supernova explosion. Hence, any small asymmetries in the subsequent SN explosion will then produce the frequently observed wide eccentric orbits.
Of particular interest is the narrow range of spectral class of the identified optical counterparts (see Negueruela 1998). In contrast to the sample of isolated Be stars to be found in the Bright Star Catalogue, there are no known Be/X-ray objects beyond spectral class B3 - see Figure 4. Most commonly the systems have counterparts in the B0-B2 group.
The explanation offered by Van Bever and Vanbeveren (1997) for this phenomenom is that the wide orbits produced by the evolutionary models are very vulnerable to disruption during the SN explosion. This will be particularily true for the less massive objects that would make up the later spectral classes. Hence the observed distribution is confirming the evolutionary models.
Some evolutionary models predict the existence of other Be star + compact object systems. In particular, the work of Raguzova and Lipunov (1999) discusses the emergence of Be + Black Hole systems and cites GRS 1915+105 as a possible example. Raguzova & Lipunov (2000) also predict that “46% of all Be stars formed in binary evolution should have White Dwarf companions”. To date, however, only two B + WD systems are known (Vennes, Berghofer & Christian 1997 and Burleigh & Barstow 1999), and no Be + WD systems - but it is very hard to detect the WD in the presence of such bright companions.
## 4. Periodicities
One of the most exciting developments in the study of Be/X-ray binaries occurred when Corbet (1984) realised that a strong correlation exists between the spin period of the neutron star and the binary period of the system. Subsequently Corbet et al (1999) developed this line of work and recently produced the more comprehensive diagram reproduced here as Figure 5. From this figure two things may be clearly seen: the intitial correlation for the Be/X-ray systems is strongly confirmed, and different classes of HMXB lie in distinctly different locations on the diagram. Hence not only does the diagram provide a valuable tool for estimating unknown binary periods in Be/X-ray systems, but if both periods are known then the class of object may be deduced.
The explanation for the striking relationship with the Be/X-ray binaries is to be found in the process of accretion on to the neutron star. For accretion to occur the Alfven radius must be less than the co-rotation radius of the accreting material - otherwise the neutron star continues to spin down until this condition is met. However, the size of the Alfven radius depends upon the density of the surrounding medium and larger orbits mean lower stellar wind densities in the environment of the neutron star. Hence a relationship naturally develops between the orbital period and the equilibrium spin period of the neutron star.
## 5. Example 1 : A1118-616
The system A1118-616 provides an excellent example of a classic Be/X-ray binary system. It was discovered in 1991 by chance while the Ariel 5 satellite was observed the nearby source Cen X-3 (Eyles et al 1975). Coe et al (1994) show the time history of its X-ray flux over 2 months starting just before the initial outburst. During the outburst the flux of the source increased by much more than 1 order of magnitude and became the dominant hard X-ray source in the Centaurus region for 7-10 days. The X-ray source then disappeared from sight for 27 years before re-emerging in the same dramatic style in 1991.
The explanation for these sudden massive X-ray outbursts lies in the H$`\alpha `$ observation carried out prior to, during, and after the 1991 outburst. Note, no similar observations exist for the 1974 outburst since the optical counterpart was not identified at that time. The history of the H$`\alpha `$ observations over a 10 year period are shown in Figure 6. From this figure it is immediately apparent why the source went into outburst when it did - the H$`\alpha `$ equivalent width had reached an exceptionally high value of more than 100Å. Since the pulse period is 404s the Corbet diagram suggests a binary period in the range 200-300d, therefore it is unlikely that the two recorded outbursts relate to binary motion (Type I outbursts). Far more likely, we are seeing two Type II outbursts, and the normal Type I outbursts are either very minor or don’t occur at all. This could simply be due to the orbit not normally taking the neutron star through the circumstellar disk. However, during these abnormal levels of H$`\alpha `$ activity, the disk probably expands to include the orbit of the neutron star, and hence accretion immediately begins. Note the strikingly rapid decline of the H$`\alpha `$ immediately after the outburst suggesting that the whole period of activity was probably just due to one major mass ejection event from the Be star.
It is interesting to compare the size of the circumstellar disk seen in A1118-616 with that observed in other systems. Figure 7 shows a plot of H$`\alpha `$ EW against the intrinsic infrared excess (also thought to arise from free-free and free-bound emission in the circumstellar disk). An obvious and unsurprising correlation exists between these two parameters as is clear from the figure. The quiescent location of A1118-616 is shown on the diagram together with two other Be/X-ray systems. For comparison, data from a sample of isolated Be stars (Dachs and Wamsteker 1982) is also presented. It is interesting to note that even in quiescence, A1118-616 is at the extreme edge of the diagram, and in fact, its peak H$`\alpha `$ EW value of $``$110Å may be one of the very largest recorded values for any Be star.
## 6. Example II : EXO2030+375
The system EXO 2030+375 was discovered in the same manner as A1118-616 (a sudden outburst of activity from a previously unknown object), but by EXOSAT in 1985 (Parmar et al 1989). The source was regularily monitored by the BATSE all sky instrument on the CGRO spacecraft (see Stollberg 1997 for an extensive discussion of this system). For long periods of time the source has shown classic Type I outbursts every periastron passage (binary period = 46d). Typical X-ray outbursts are shown in Figure 8 (Norton et al, 1994).
Then for about two years the source disappeared from the X-ray sky and re-merged in 1996. Detailed RXTE observations (Reig & Coe 1998) revealed a significant shift in the binary phase of the outbursts of $``$0.15. This phase has since remained unaltered, so possible precessional models of disks are ruled out. In addition, the H$`\alpha `$ EW had decreased from -18Å to -8Å, thereby eliminating the possibility that the earlier outburst phase arose from to an expanded circumstellar disk which the neutron star encountered earlier in its orbit. It seems more likely that a fundamental change has occurred in the accretion mechanism resulting from the presence of a much smaller disk. Probably, while the disk was large, the accretion was by Roche lobe overflow, but now it is simply by direct stellar wind accretion. This idea is supported by the very rapid pulse period changes seen prior to 1994 which had essentially ceased after 1996 - the more direct coupling of Roche lobe overflow would exert greater torques on the neutron star.
## 7. Neutron star interactions with the circumstellar disk
One of the particularily interesting questions related to studying Be/X-ray binaries is the question of the effect of the neutron star on the circumstellar disk. Initial work in this area by Norton et al (1994) suggested that the dimensions of the circumstellar disk, compared with the amount of material undergoing accretion on to the neutron star, made it very unlikely that the presence of the neutron star would affect the growth of the circumstellar disk.
This view has now been altered by the longer term studies carried out by Reig, Fabregat & Coe (1997). In Figure 9 the maximum observed H$`\alpha `$ EW is plotted against the orbital period of several systems. Though the data are still rather sparse, there is a strong suggestion of a correlation. The conclusion the authors draw from this diagram is that the continually orbiting neutron star gradual erodes away the outer edges of the circumstellar disk and inhibits its growth.
Similarily, the same authors present a comparison of the maximum H$`\alpha `$ EWs of Be stars in Be/X-ray binary systems with isolated Be stars. Again there is a strong suggestion that, on average, the disks are smaller in those systems with neutron star companions.
As yet the statistical sample sizes are small for these studies, but as more binary periods are determined these propositions can be checked. It is certainly important to understand how these two components in the binary system interact.
## 8. Small Magellanic Cloud
It has come as a great surprise to discover that there are a large number of Be/X-ray binaries in the Small Magellanic Cloud. Figure 10 shows the distribution of about half of these superimposed upon an optical image of the SMC.
It is possible to estimate the number of systems one would expect based upon the relative masses of our galaxy and the SMC. This ratio is $``$50, so with 64 known or suspected systems in our galaxy we would only expect 1 or 2 systems in the SMC. However, Maeder, Grebel & Mermilliod (1999) have shown that the fraction of Be stars to B stars is 0.39 in the SMC compared with 0.16 in our galaxy. So this raises the expected number of Be/X-ray systems to $``$3 - but we now know of 20 such systems!
The reason for the large number of Be stars almost certainly lies in the history of the Magellanic Clouds. Detailed H1 mapping by Stavely-Smith et al (1997) and Putman et al (1998) has shown a strong bridge of material between the Magellanic Clouds and between them and our own galaxy. Furthermore, Stavely-Smith et al have demonstrated a the existence of a large number of supernova remnants of a similar age ($``$5 Myr), strongly suggesting enhanced starbirth has taken place as a result of tidal interactions between these component systems. Consequently it seems very likely that the previous closest approach of the SMC to the LMC $``$100 Myrs ago may have triggered the birth of many new massive stars which have given rise to the current population of HMXBs. In fact, other authors (eg Popov et al 1998) claim that the presence of large numbers of HMXBs may be the best indication of starburst activity in a system.
Whatever the reason for the large number, the Small Magellanic Cloud now provides us with an excellent sample of Be/X-ray systems in a relatively compact and easily observed region of the sky.
## 9. Conclusions
Be stars in X-ray binary systems offer a whole new avenue for exploring the Be phenomenon. The evolutionary histories are different from the isolated systems, but their basic characteristics seem identical. The presence of a neutron star companion provides us with an extremely valuable probe of the circumstellar disk, and X-ray monitoring satellites provide one of the most rapid methods for identifying a Be star in outburst. Work in this field should continue to make substantial contributions to the much broader studies of these mysterious objects.
## 10. Acknowledgements
I am very grateful to my many collaborators over the years who have made major contributions to the topics discussed in this paper. In particular, I wish to thank Dave Buckley (SAAO), Simon Clark (Sussex), Juan Fabregat (Valencia), Ignacio Negueruela (Rome), Andy Norton (OU), Pablo Reig (Crete), Paul Roche (Leicester) and Iain Steele (Liverpool).
I also wish to thank the organisers of IAU Colloquium 175 for giving me an opportunity to discuss the work in this field.
## References
Bildsten, L., Chakrabarty, D., Chiu, J., Finger, M.H., Koh, D., Nelson, R.W., Prince, T.A., Rubin, B.C., Scott, D.M., Stollberg, M., Vaughan, B.A., Wilson, C.A. & Wilson, R.B. 1997, ApJS, 113, 367
Burleigh, M.R. & Barstow, M.A. 1999, A & A 341, 795
Coe, M.J., Roche, P., Everall, C., Fishman, G.J., Hagedon, K.S., Finger, M., Wilson, R.B., Buckley, D.A.H., Shrader, C., Fabregat, J., Polcaro, V.F., Giovannelli, F. & Villada, M. 1994, A & A, 289, 784.
Coe, M.J. & Orosz, J.A. 1999, MNRAS, (in press)
Corbet, R.H.D. 1984, A & A, 141, 91
Corbet, R.H.D., Marshall, F.E., Peele, A.G. & Takeshima T. 1999 ApJ, 517, 956
Dachs, J. & Wamsteker, W. 1982, A & A, 107, 240
Eyles, C.J., Skinner, G.K., Wilmore, A.P. & Rosenberg, F.D. 1975, Nature, 254, 577
Maeder, A., Grebel, E.K. & Mermilliod, J.C. 1999, A & A, (in press)
Meurs, E.J.A. & van den Heuvel E.P.J. 1989, A & A, 226, 88
Nagase, F. 1989 PASJ, 41, 1
Negueruela, I. 1998 A & A, 338, 505
Norton, A.J., Chakrabarty, D., Coe, M.J., Everall, C., Finger, M.H., Prince, T.A., Roche, P., Stollberg, M.T. & Wilson, R.B. 1994, MNRAS, 271, 981
Parmar, A.N., White, N.E., Stella L., Izzo C., Ferri, P. 1989, ApJ, 338, 359
Popov, S.B., Lipunov, V.M., Prokhorov, M.E. & Postnov, K.A. 1998, Astron Reports, 42, 29
Putman, M.E. and 25 other authors 1998, Nature, 394, 752
Raguzova, N.V. and Lipunov, V.M. 1999, A & A, 349, 505
Raguzova, N.V. and Lipunov, V.M. 2000, this conference
Rappaort S. & van den Heuvel E.P.J. 1982, in Be Stars, Proceedings of the Symposium, (Dordrecht: Reidel), 327
Reig, P., Fabregat, J. & Coe, M.J. 1997, A & A, 322, 193
Reig, P. & Coe, M.J. 1998, MNRAS, 294, 118
Reig, P. and Roche, P. 1999, MNRAS (in press)
Stavely-Smith, L., Sault, R.J., Hatzidimitriou, D., Kesteven, M.J. & McConnell, D. 1997, MNRAS, 289, 225
Stella, L., White, N.E. & Rosner, R. 1986, ApJ, 308, 669
Stollberg, M.T. 1997, PhD Thesis, University of Alabama in Huntsville
van den Heuvel, E.P.J. 1983, in Accretion driven stellar X-ray sources, eds Lewin, W.H.G. & van den Heuvel, E.P.J., (Cambridge University Press)
Van Bever, J. & Vanbeveren, D. 1997, A & A, 322, 116
Vennes S., Berghofer, T.W. & Christian, D.J. 1997, ApJ, 491, L85
Verbunt, F. & van den Heuvel, E.P.J. 1995, in X-ray Binaries, eds. Lewin, W.H.G., Van Paradijs J. & van den Heuvel, E.P.J., page 457, (Cambridge University Press)
White, N.E., Nagase, F. & Parmar, A.N. 1995, in X-ray Binaries, eds. Lewin, W.H.G., Van Paradijs J. & van den Heuvel, E.P.J., page 1, (Cambridge University Press)
## Discussion
Raguzova: I would like to make some comments about Be stars with white dwarfs. All white dwarfs in such systems must have very high effective temperatures and we can detect such objects in the extreme ultraviolet range.
Coe: Yes, that is correct. The two known B/WD systems have been found with EUVE - though it is proving much more difficult than one might think. Some people believe that the numbers of such systems may not be as large as you suggest.
Hummel: Are the detected Be/X-ray binaries in the SMC field stars, or members of open clusters?
Coe: There is no evidence so far from the identified systems that they are mainly found in clusters. |
no-problem/9911/astro-ph9911159.html | ar5iv | text | # RX J0420.0–5022: An isolated neutron star candidate with evidence for 22.7 s X-ray pulsations∗
## 1 Introduction
To date five candidates for isolated neutron stars (INS) were discovered in ROSAT data (RX J1856.6–3754, Walter et al. 1996; RX J0720.4–3125, Haberl et al. 1997; RX J0806.4–4123, Haberl et al. 1998; 1RXS J130848.6+212708, Schwope et al. 1999 and RX J1605.3+3249, Motch et al. 1999) which can be classified by their similar X-ray properties. The soft X-ray spectra are well represented by pure blackbody emission with temperatures kT between 50 and 120 eV and the X-ray flux is constant on time scales of months to 10 years. The X-ray brightest source, RX J1856.6–3754, could be identified with a faint blue object (V = 25.6) by Walter & Matthews (1997) on HST images and for the second brightest, RX J0720.4–3125, a likely counterpart with B = 26.1-26.6 was found (Motch & Haberl 1998, Kulkarni & van Kerkwijk 1998). The inferred extremely high X-ray to optical flux ratios log(f<sub>x</sub>/f<sub>opt</sub>) of 4.8 and 5.3, respectively, exclude any known object other than an isolated neutron star. Similarly high lower limits for f<sub>x</sub>/f<sub>opt</sub> are derived from the faint brightness limits obtained for the other candidates (see table in Schwope et al. 1999).
The five candidates constitute the bright end of the log N – log S distribution of X-ray detected INS (with no detectable radio mission) of which up to several thousand were expected to be seen in the ROSAT all-sky survey (RASS, Blaes & Madau 1993, Colpi et al. 1993, Madau & Blaes 1994). The estimates were based on models of old neutron stars heated by the accretion of interstellar matter as proposed by Ostriker et al. (1970). From a compilation of available data on the number and space density of INS candidates and upper limits obtained from RASS follow-up identification programs Neuhäuser & Trümper (1999) concluded that the log N – log S curve lies between the theoretical expectations for middle-aged cooling neutron stars and old accreting neutron stars. They suggest that the larger number of expected accreting old neutron stars than observed is mainly caused by the assumed velocity distribution of the neutron stars.
From one of the five INS candidates X-ray pulsations were detected, indicating the spin period of the neutron star. The pulse period of 8.391 s found for RX J0720.4–3125 constrains the magnetic field strength of the neutron star. In the case of an accreting neutron star a weak field of less than 10<sup>10</sup> G is derived for accretion to be possible (Haberl et al. 1997). A higher magnetic field strength in the past is required to spin down the neutron star to the present rate within a Hubble time if the neutron star was born with a typical spin period of 10 ms. Consequently this would indicate magnetic field decay and for a birth field strength of 10<sup>12</sup> G decay time scales of $`>10^7`$ y are derived (Wang 1997) for an old ($`>10^9`$ y) neutron star. Alternatively RX J0720.4–3125 may prove to be a relatively young ($`<10^6`$ y) cooling neutron star for which the argumentation from above would require a very high magnetic field strength of the order of 10<sup>14</sup> G (Kulkarni & Kerkwijk 1998). The population of these ”magnetars” was suggested to explain the properties of soft $`\gamma `$-ray repeaters (Kouveliotou et al. 1999 and references therein). The extremely strong magnetic field can suppress radio emission and supply a significant source of heat, allowing magnetars to remain detectable in X-rays over longer times than ordinary pulsars (Heyl & Kulkarni 1998). However a doubling of the pulse period derivative observed in SGR 1900+14 casts doubt on the magnetar model as it requires a 100% increase of the magnetic field energy if magnetic dipole radiation were the primary cause of the pulsar spin-down (Marsden et al. 1999). Also the different X-ray spectrum of RX J0720.4–3125 in comparison to those of soft $`\gamma `$-ray repeaters would need to be explained by the magnetar model.
As long as the distribution of initial spin period / magnetic field strength is highly uncertain, conclusions on the evolution of these important neutron star parameters are tentative. The nature of RX J0720.4–3125 remains therefore unclear until an accurate measurement of the period derivative may further constrain the models.
In this letter we report on ROSAT X-ray and NTT optical observations of a new INS candidate, RX J0420.0–5022, which shows evidence for 22.7 s X-ray pulsations. Together with RX J0720.4–3125 it is only the second long-period INS candidate and their investigation is crucial for our understanding of their evolutionary status within the whole class of isolated neutron stars.
## 2 Observations
### 2.1 Soft X-rays
RX J0420.0–5022 was discovered as X-ray source in the ROSAT all-sky survey data and originally associated with the galaxy ESO 202-G008. It is included in the ROSAT all-sky survey bright source catalogue (1RXS J042003.1–502300, Voges et al. 1996) with 0.12$`\pm `$0.03 counts s<sup>-1</sup>. Details on the ROSAT mission can be found in Trümper (1982) and the focal plane instruments PSPC (Position Sensitive Proportional Counter) and HRI (High Resolution Imager), both sensitive in the 0.1 – 2.4 keV energy range, are described by Pfeffermann et al. (1986) and David et al. (1993), respectively. From two follow-up HRI observations in June and December 1997 (the observations are summarized in Tab. 1) it became clear that the soft X-ray source is nearly 1′ away from ESO 202-G008 and unrelated. The best position was derived from the June 1997 HRI observation using the maximum likelihood technique of EXSAS (Zimmermann et al. 1994) to RA = 04<sup>h</sup> 20<sup>m</sup> 2$`\stackrel{s}{.}`$2, Dec = -50° 22′ 46″ (J2000.0) with a 90% confidence error of 8″ (dominated by the 7″ systematic error of the attitude reconstruction). No other sources were detected in the HRI image which could be utilized for bore-sight correction.
RX J0420.0–5022 was serendipitously observed in a short PSPC pointing in February 1997 at an off-axis angle of 31′. The count rate is consistent with the survey detection and with the HRI detections in June and December 1997 (see Tab. 1 and below). The average PSPC spectrum (Fig. 1), although of low statistical quality, is very similar to the spectra of the known INS candidates without any significant number of counts above 0.4 keV. A power-law fit to the spectrum is acceptable but results in an unrealistic photon index of $``$10. The soft emission is well described by a blackbody model (kT = 57$`{}_{47}{}^{}{}_{}{}^{+25}`$ eV, column density 1.7$`\times 10^{20}`$ cm<sup>-2</sup> (0 – 8$`\times 10^{20}`$ cm<sup>-2</sup>)) with an observed flux of 6.9$`\times 10^{13}`$ erg cm<sup>-2</sup> s<sup>-1</sup> (0.1 – 2.4 keV). Assuming a distance d for the source a bolometric luminosity of L<sub>bol</sub> = 2.7$`\times 10^{30}`$ (d/100 pc)<sup>2</sup> erg s<sup>-1</sup> is derived. Assuming an emission area with 10 km radius, the derived upper limit for the blackbody temperature (104 eV) yields un upper limit for the distance of 3.9 kpc. For the best fit temperature the distance is 700 pc but is further reduced for a smaller emitting area. Folding the blackbody model with the best fit parameters through the HRI energy response yields for the observed PSPC count rate (pointing) an expected HRI count rate of 1.9$`\times 10^2`$ counts s<sup>-1</sup>. The PSPC position of RA = 04<sup>h</sup> 20<sup>m</sup> 1$`\stackrel{s}{.}`$6, Dec = -50° 22′ 44″ (J2000.0, error of 17″) is 6.1″ away from the HRI position. The survey position is 16.5″ from the HRI position, also consistent within the errors. A formal average of the positions (weighting with errors) yields RA = 04<sup>h</sup> 20<sup>m</sup> 2$`\stackrel{s}{.}`$36, Dec = -50° 22′ 49$`\stackrel{}{.}`$5 (7$`\stackrel{}{.}`$4 error), somewhat south of the HRI position.
A temporal analysis of the PSPC data using Fourier transformation and a Rayleigh Z<sup>2</sup> folding test (Buccheri et al. 1983) reveals evidence for periodic modulation of the soft X-ray flux with 22.69$`\pm `$0.03 s. The probability for a false detection of the period was derived to 7$`\times 10^5`$ corresponding to a 4 $`\sigma `$ detection. The PSPC light curve folded by the pulse period is plotted in Fig. 2 and shows a sinusoidal semi-amplitude modulation of 43$`\pm `$14%. The folding analysis was also applied to the June 1997 HRI observation. The periodogram shows peaks around the expected period, supporting the reality of the modulation. However, the HRI observation was performed in five intervals unevenly spread over 49 hours causing a large number of alias periods spread over $``$0.5 s around the 22.7 s found from the PSPC observation (which was uninterrupted) and does not provide a better estimate for the period.
### 2.2 Optical
We observed the region of RX J0420.0–5022 with the ESO-NTT and SUSI2 on 1999 January 15. The combination of the two EEV CCDs yields a field of view of 5$`\stackrel{}{.}`$5 $`\times `$ 5$`\stackrel{}{.}`$5 with a pixel size of 0$`\stackrel{}{.}`$08. A total of 30 min in B and 30 min in R were accumulated using 10 min long individual exposures. The night was of photometric quality and observations of the standard field PG0942-029 allowed photometric calibration to an accuracy of $``$ 2% . Mean seeing was 1″ FWHM. Raw images were corrected for flat-field, bias and cleaned from cosmic-ray impacts using standard MIDAS procedures.
Fig. 3 shows the ROSAT HRI, PSPC and Survey error circles overlayed on the sum of the B and R images of RX J0420.0–5022 representing a total integration time of 1 hour. We also show in Fig. 4 and Fig. 5 the individual B and R summed images with the HRI error circle overlayed. Optical positions were computed using 8 stars extracted from the USNO A-1.0 catalogue and spread around the X-ray positions. The final astrometric accuracy should be that of the USNO catalogue, i.e. better than 1 arcsec.
Several faint objects labeled A to H on Fig. 3 are detected in or close to the HRI error circle. Table 2 lists the B and R magnitudes with one $`\sigma `$ errors and information on image profile when possible.
The brightest object in the HRI error circle (object A) is stellar like and has a B-R colour index suggesting a remote G5V-K0V late type star. The next two brightest objects (B and D) located at the edge of the error circle appear clearly extended and are probable field galaxies. Among the remaining candidates, the brightest one which lacks reliable B-R colour index and spatial extent is object E with B = 25.25 $`\pm `$ 0.20. Our detection limit estimated from magnitudes of object H is B $``$ 26.5 and R $``$ 26.4.
## 3 Discussion
The soft X-ray source RX J0420.0–5022 is characterized by the same properties as the five previously published isolated neutron star candidates discovered in ROSAT data. A blackbody-like spectrum with kT = 57 eV and little interstellar absorption of 2$`\times 10^{20}`$ cm<sup>-2</sup> are well within the range observed from the known candidates. No significant long-term flux variations were observed between ROSAT survey (1991, January 23-25) and three pointed observations in 1997.
Optical NTT imaging failed to identify the counterpart to the X-ray source. The brightest stellar-like object in the error circle has B = 24.44 $`\pm `$ 0.08 implying log(f<sub>x</sub>/f<sub>opt</sub>) $``$ 3.0 (using the observed X-ray flux of 6.9$`\times 10^{13}`$ erg cm<sup>-2</sup> s<sup>-1</sup> and f<sub>opt</sub> = 10<sup>-0.4(B+13.42)</sup> from Maccacaro et al. 1988). It exhibits a red B-R colour index, compatible with that of a remote late type star and can be ruled out on this basis. The next possible candidate has B = 25.25 $`\pm `$ 0.20 yielding a log(f<sub>x</sub>/f<sub>opt</sub>) ratio of 3.3. This virtually excludes all possible kinds of soft X-ray emitters other than isolated neutron stars. In particular an identification with a hot white dwarf or with a very soft Seyfert-1 AGN (e.g. Grupe et al. 1998) seems to be excluded as this would imply the presence of an optically bright object in the error circle. A very distant white dwarf is also incompatible with the low observed absorption.
With 0.11 counts s<sup>-1</sup> detected in the PSPC RX J0420.0–5022 is the faintest of the known INS candidates so far (see Schwope et al. 1999). It is a factor of 33 fainter than RX J1856.6–3754 which shows a similar blackbody temperature. Scaling the optical B magnitude of RX J1856.6–3754 by this factor yields an expected value of B $``$ 29.6 for RX J0420.0–5022, beyond the reach of our imaging. The serendipitous discovery near an unrelated object is the first of this kind and more INS may be unrecognized because of positional mis-identification. It is remarkable that it shows the same soft spectral properties as the previously found INS candidates like RX J0806.4–4123 and RX J1605.3+3249 which were discovered from dedicated searches of objects with such properties. This suggests that most INS have temperatures kT below $`120`$ eV and that the distribution is not strongly biased by selection effects.
In the ROSAT PSPC observation of RX J0420.0–5022 we detected X-ray pulsations with a significance of 4$`\sigma `$. If this is confirmed RX J0420.0–5022 would share this remarkable feature with RX J0720.4–3125. For none of the four other candidates periodic flux variations were found although it must be noted that small amplitude variations could not be detected in the available data due to insufficient statistics. The pulse period of 22.7 s for RX J0420.0–5022 is even longer than that for RX J0720.4–3125 (8.391 s) and the modulation is deeper ($`>`$29% compared to 12$`\pm `$2%).
As in the case of RX J0720.4–3125 similar conclusions can be drawn for RX J0420.0–5022, based on the different scenarios conceivable for its evolution. If the X-rays are powered by accretion of interstellar matter then the magnetic field strength of RX J0420.0–5022 should be lower than 5.3$`\times 10^9`$ d<sub>100</sub> G where d<sub>100</sub> is the distance in units of 100 pc. This is a similarly low limit as obtained for RX J0720.4–3125 for a distance of a few hundred pc and would indicate magnetic field decay unless the neutron star was born with a relatively long spin period (Wang 1997). The lower temperature derived for RX J0420.0–5022 may indicate a factor of $``$4 lower accretion rate than for RX J0720.4–3125. This in turn increases slightly the allowed velocity of the neutron star relative to the ambient interstellar medium. However as in the case of RX J0720.4–3125 the relative velocity would probably be a few tens of km s<sup>-1</sup>, well below 100 km s<sup>-1</sup>.
In the magnetar model (Heyl & Hernquist 1998, Kulkarni & van Kerkwijk 1998) the factor 2.7 longer pulse period of RX J0420.0–5022 compared to RX J0720.4–3125 implies a different age and/or magnetic field strength of the neutron stars. Assuming that the neutron stars have spun down by magnetic dipole radiation the age is proportional to P<sup>2</sup>/B<sup>2</sup>. A similar magnetic field strength B would imply that RX J0420.0–5022 is a factor of 7.3 older than RX J0720.4–3125 while for similar age the magnetic field of RX J0420.0–5022 is stronger than that of RX J0720.4–3125 by the factor 2.7. A measure of the pulse period derivative RX J0720.4–3125 and RX J0420.0–5022 will be essential to determine the evolutionary status of these isolated neutron stars.
###### Acknowledgements.
The ROSAT project is supported by the German Bundesministerium für Bildung und Forschung (BMBF/DLR) and the Max-Planck-Gesellschaft. |
no-problem/9911/astro-ph9911103.html | ar5iv | text | # The dearth of halo dwarf galaxies: is there power on short scales?
\[
## Abstract
$`N`$-body simulations of structure formation with scale-invariant primordial perturbations show significantly more virialized objects of dwarf-galaxy mass in a typical galactic halo than are observed around the Milky Way. We show that the dearth of observed dwarf galaxies could be explained by a dramatic downturn in the power spectrum at small distance scales. This suppression of small-scale power might also help mitigate the disagreement between cuspy simulated halos and smooth observed halos, while remaining consistent with Lyman-alpha-forest constraints on small-scale power. Such a spectrum could arise in inflationary models with broken scale invariance.
\]
The broad-brush picture painted by the inflation-inspired hierarchical-clustering paradigm accounts for the smoothness of the cosmic microwave background (CMB), its tiny temperature fluctuations, the flatness of the Universe, and the observed distribution of galaxies. However, finer inspection yields some possible—and possibly troubling—discrepancies between the models and observations that still need to be ironed out. One of these is the dearth of substructure in galaxy halos.
Recent high-resolution $`N`$-body simulations have confirmed earlier analytic arguments that suggested that hierarchical-clustering models should produce far more dwarf galaxies around the Milky Way than are observed. There are only 11 dwarf galaxies with internal velocity dispersions greater than 10 km sec<sup>-1</sup> within the virial radius of the Milky Way halo. However, numerical simulations of structure formation in a hierarchical model indicate that a halo of the Milky Way’s mass and circular speed should contain roughly an order of magnitude more dwarf galaxies. These theoretical results are robust to changes in the values of cosmological parameters or in the tilt of the primordial spectrum of perturbations.
One might at first be tempted to dismiss this discrepancy between theory and observations as a consequence of some nasty astrophysics. After all, the simulations consider only gravitational interactions and identify only virialized dark halos, while the real Universe is filled with gas, and dwarf galaxies are identified by their visible matter. So, for example, one might guess that the halos are there but remain invisible because the gas has been expelled by an early generation of supernovae. However, even generous estimates of the efficiency of supernova-driven winds fall short of explaining the absence of luminous dwarf galaxies . Further, even if the baryonic matter could somehow be driven out of the mini-halos or kept dark, we would still have trouble explaining how a spiral disk could have formed in the strongly fluctuating potential of such a clumpy halo. See Refs. for more detailed reviews of such arguments.
In the absence of any prosaic astrophysical mechanism, it is natural to think of more exotic explanations. One strategy is to modify the nature of the dark matter, in order to prevent low-mass halos from forming. One option is that the dark matter is warm, for example a neutrino of mass around a keV, which would suppress the formation of small-scale structure by free-streaming out of potential wells . However, such a neutrino would diminish power on scales $`2`$ to $`12h^1`$ Mpc to a degree that may conflict with the power inferred from the Lyman-alpha forest . Another possibility is that the dark-matter particles interact strongly with each other, but not with ordinary matter . However, the properties required of this particle (elastic-scattering cross-sections of order $`10^{25}\mathrm{cm}^2`$) are almost inconceivable in the predominant paradigms of weakly-interacting massive particles and axions .
Here we consider an alternative strategy: broken scale invariance in the primordial power spectrum, arising from some feature in the inflaton potential. In slow-roll inflation, the amplitude of density perturbations on some comoving scale is proportional to $`V^{3/2}/V^{}`$, where $`V`$ is the inflaton potential, and $`V^{}`$ its derivative, when the comoving scale under consideration exits the horizon. Suppose that $`V^{}`$ is initially very small and that there is then a discontinuity in the second derivative, $`V^{\prime \prime }`$. The slope $`V^{}`$ will then jump, and the density-perturbation amplitude will drop steeply. In this way, the density-perturbation amplitude on some suitably small scale can be suppressed relative to the power on some larger scale. Such an idea has been invoked as a possible explanation of a claimed break in the power spectrum at $`100`$ Mpc scales (and in an attempt produce non-Gaussian features ). Here, we use this idea to account for the lack of halo substructure. Below, we show how a suppression of small-scale power can affect the dwarf-galaxy abundance and illustrate a particular inflationary model. We also argue briefly that the absence of small-scale power may help explain the discrepancy between simulated cuspy halos and observed smooth halos.
For simplicity and clarity of presentation, we restrict our discussion here to an Einstein-de Sitter (EdS) model, but everything can readily be generalized to a low-density Universe. Although we focus on an EdS cosmogony, we use a cold-dark-matter (CDM) power spectrum with $`\mathrm{\Gamma }=0.25`$ (e.g., as in a $`\tau `$CDM model ) to ensure that observational constraints from the shape and amplitude of the galaxy power spectrum, the cluster number density, and the COBE normalization are satisfied . The solid curve in Fig. 1 shows this power spectrum, along with two other power spectra motivated later. The corresponding rms mass fluctuations are also shown.
We need to understand how changes to the power spectrum of density perturbations will affect the abundance of dwarf galaxies. We are interested in the abundance of mini-halos of some given mass that exist within some larger halo of mass $`2\times 10^{12}M_{}`$, comparable to that of the Milky Way and the halo mass used by Moore et al. . The simulations of Moore et al. show that the number of mini-halos in the Galaxy today is very close to the number that existed in the proto-galaxy; i.e., no more than a small fraction are fully disrupted during their subsequent orbital motion in the Milky Way halo. Thus, our task is simplified: we can calculate the abundance of mini-halos in the proto-galaxy that later became the Milky Way halo. To do so, we use the conditional mass function , given by
$$F(>M_{\mathrm{small}})=\mathrm{erfc}\left[\frac{\delta _\mathrm{c}z_\mathrm{f}}{\sqrt{2\left[\sigma ^2(M_{\mathrm{small}})\sigma ^2(M_{\mathrm{big}})\right]}}\right].$$
(1)
This equation gives the fraction of the mass in a galactic halo of total mass $`M_{\mathrm{big}}`$ that was in mini-halos of mass greater than $`M_{\mathrm{small}}`$ at the redshift $`z_\mathrm{f}`$ at which the protogalaxy broke off from the expansion. Here $`\sigma (M)`$ is the linear-theory rms fractional mass fluctuation in spheres of radii that on average enclose a mass $`M`$, and $`\delta _\mathrm{c}=1.7`$ is the critical threshold for collapse.
It is the circular velocity of the dwarf galaxies, not the mass, which is observed. We relate the circular speed $`v_\mathrm{c}`$ to the mass by assuming the mini-halos underwent collapse at the protogalaxy-formation epoch. Then the circular speed is obtained from $`v_\mathrm{c}^3=10MGH(z_\mathrm{f})`$, where $`G`$ is Newton’s constant and $`H(z)`$ is the expansion rate at the collapse redshift $`z`$. To compare with the numerical results of Moore et al. , we consider a Galactic halo of mass $`M=2\times 10^{12}M_{}`$ with a circular speed of 220 km sec<sup>-1</sup>. Using the conditional mass function, we obtain for the CDM power spectrum the cumulative number of halos shown by the solid curve in Fig. 2 (cf. Fig. 1 in Ref. ). Although our calculation of the number of halos has several shortcomings, the good agreement with the numerical results shown in Fig. 1 of Ref. clearly indicates that we are including the essential physics. In particular, we reproduce the order-of-magnitude excess of halos in the theoretical prediction, as compared to observations, for $`v_\mathrm{c}/v_{\mathrm{global}}`$ in the range 0.05 to 0.1.
We now consider how the dwarf-galaxy abundance changes if the power spectrum is modified on short scales. Clearly, if power is reduced on dwarf-galaxy scales ($`10^{10}M_{}`$) relative to that at the Galactic scale ($`10^{12}M_{}`$), the number of dwarf galaxies will be suppressed. To investigate what is required, we first sharply cut off the power spectrum at $`k=4.5h\mathrm{Mpc}^1`$, shown by the dotted curve in Fig. 1; the lower panel shows how the rms mass fluctuations are modified. The corresponding prediction for halo substructure is shown as the dotted curve in Fig. 2, and is in striking agreement with the observations.
However, such a cutoff is purely phenomenological, and much stronger motivation is needed. In particular, the question arises as to whether inflation can produce such a power spectrum. The heuristic arguments given above suggest that the power spectrum can undergo a sharp drop on small scales if there is a discontinuity (or near-discontinuity) in the slope of the potential, and this is known as the broken-scale-invariance (BSI) model. Remarkably, there is an exact solution for the power spectrum in this situation, due to Starobinsky , even though the slow-roll approximation does not apply. Further, this spectrum has a universal form, depending only on the change in the derivative across the discontinuity, in the sense that the form of the spectrum is preserved if the discontinuity is smoothed out, provided the smoothing is over a sufficiently short range of scalar-field values. The underlying reason for this is that the field briefly becomes kinetic-energy dominated as it goes over the discontinuity, and the precise form of the potential becomes irrelevant.
The dashed curve in Fig. 1 shows this power spectrum for a suitably chosen location and amplitude of the discontinuity (the amplitude is $`p=10`$, in the notation of Ref. ). The universal form features a modest rise, followed by a series of oscillations asymptoting to the original spectral shape at a much lower amplitude; in this case the power is reduced by a factor $`p^2=100`$ on short scales. Such a model represents the fastest possible cutoff in power that can be obtained from a single-field inflation model.<sup>*</sup><sup>*</sup>*We suspect that this is true in multi-field models too, as additional fields will supply extra friction, slowing down the fields’ evolution and hence stretching features in $`k`$-space.
This power spectrum gives rise to the dashed curve shown in Fig. 2. This model is almost as successful as the cutoff power spectrum; it produces roughly 20 mini-halos with $`v_\mathrm{c}/v_{\mathrm{global}}0.04`$, in much better agreement than the original power spectrum with the $`𝒪(10)`$ that are observed. We conclude that the BSI model provides a promising possibility for reconciling the predictions with observations.
Our modification to the power spectrum sets in only on very short scales, and so does not affect successes of the standard paradigm on much larger scales, such as the cluster number density or galaxy correlation function. However, we need to check that the loss of short-scale power is not inconsistent with object abundances at high redshift. The most powerful and direct constraint comes from the abundance of damped Lyman-alpha systems . Observations exist at redshifts 3 and 4, giving comparable constraints. Redoing the calculation of Ref. to include updated observations gives a 95% confidence lower limit of
$$\sigma (10^{10}h^1M_{},z=0)>2.75+h,$$
(2)
in a critical-density Universe, which is (just) satisfied by the BSI model (though not by the cutoff power spectrum).
This narrow escape makes it look as if the model is quite marginal. However, we have only analyzed the critical-density case, and the damped-Lyman-alpha constraint is much weaker in a low-density model. The observational constraint on $`\sigma (M)`$ at redshift 4 is almost the same (no more than 10% lower for $`\mathrm{\Omega }_0`$ values of interest ). However at redshift 4 suppression of the growth of perturbations has yet to set in, and the normalization to COBE (and/or cluster abundance) means $`\sigma `$ is higher by a factor $`1/\mathrm{\Omega }_0`$. With this additional factor all our models would easily pass the damped Lyman alpha system test were $`\mathrm{\Omega }_00.3`$.
There are two other key short-scale predictions worth considering. One is the power spectrum of perturbations obtained from the Lyman-alpha forest . However, the power spectrum derived in Ref. does not extend much beyond $`k=5h`$ Mpc<sup>-1</sup>, which is where our model begins to differ from the usual one. It therefore does not appear in conflict with the power in our theoretical spectra; indeed, if anything, their analyses see less power than expected on short scales. The second test is whether the model has early enough structure formation to reionize the Universe by a redshift 5, as required by the Gunn–Peterson test. Estimates using the technique of Ref. would suggest this is marginal in the critical-density case. However, once again in the low-density case we benefit from the much higher early-time normalization of the power spectrum, which increases the predicted redshift by a factor $`1/\mathrm{\Omega }_0`$. We thus conclude that this BSI model is likely to be consistent in the observationally-favored low-density flat model, though a more careful investigation is merited.
The suppression of small-scale power may also help to explain the discrepancy between the cuspy inner halos observed in simulations and the smooth halos inferred from observations. It has been argued that the steepness of the halo density profile at small radii depends on how the characteristic density of the mini-halos that merge to form galactic halos scales with their mass. If small halos are denser—as they would be if they collapsed earlier—then the cusp is steeper . If power is suppressed on sub-galactic scales, then all mini-halos with sub-galactic masses will undergo collapse at about the same time and thus have similar densities . If so, then halos should be less cuspy than prior simulations have shown. However, these heuristic arguments have been tested numerically for the case of galaxy clusters , where no change to the core structure was found when short-scale power was cut off. More work will be needed assess the effects on galactic halos.
In conclusion, the arguments suggesting that the standard structure-formation paradigm predicts too much substructure in galactic halos have now been refined to an extent that requires that the problem be taken very seriously and that radical answers to the discrepancy be considered. We have proposed here that the discrepancy indicates a dramatic lack of short-scale power in the primordial power spectrum. While such a sharp feature in the spectrum would not be expected a priori, it can be obtained from inflation, with the BSI model being the best available example. This model has already been invoked to introduce features on much larger scales of around $`100h^1`$ Mpc , but the observational motivation for a short-scale break appears stronger. If our suggestion is correct, the lack of short-scale power will soon become apparent in new observations of high-redshift phenomena.
We thank Bruno Guiderdoni, Ben Moore, John Peacock, and Paul Steinhardt for discussions, and we acknowledge the hospitality of the Isaac Newton Institute where this work was initiated. MK acknowledges the support of the DoE, NASA, and the Alfred P. Sloan Foundation. |
no-problem/9911/astro-ph9911372.html | ar5iv | text | # Constraints on the Structure of Dark Matter Halos from the Rotation Curves of Low Surface Brightness Galaxies
## 1 Introduction
Cosmological models make specific predictions for the structure of dark matter halos. A generic prediction of the cold dark matter (CDM) model is that dark matter halos should be very dense, with steep central cusps. Simulations by Navarro, Frenk & White (1996, 1997) found a nearly universal density profile (hereafter the NFW-profile), which rises as $`r^1`$ to the center. Higher resolution simulations suggest that the central profiles should be even steeper, rising as $`\rho r^{1.5}`$ (Fukushige & Makino 1997, Moore et al. 1998; Ghigna et al. 1999), although other comparable simulations have found much shallower density cusps ($`r^{0.3}`$; Kravtsov et al. 1998, Bullock et al. 1999).
Observationally, one may hope to test these predictions by measuring the dynamics of disk galaxies. For disks with a central surface brightness close to the Freeman (1970) value, however, the actual insight gained into the structure of dark halos is severely limited by the disk’s contribution to the circular velocity. The unknown mass-to-light ratio of the stellar component results in ambiguity regarding the actual density distribution of the dark matter. Although some limits can be obtained (i.e., the maximum disk), the lack of any obvious transition region from a disk-dominated to a halo-dominated part in the rotation curves (the “disk-halo conspiracy”) severely impedes the ability to obtain a unique decomposition. Many different combinations of mass-to-light ratio and halo density profile yield equally good descriptions of the data (e.g., van Albada et al. 1985; Lake & Feinswog 1989).
The structure of dark matter halos is more directly revealed in systems where the disk contributes little to the dynamics. Several studies have noted that the rotation curves of dark matter dominated dwarf galaxies are inconsistent with steeply cusped dark halos, and instead indicate constant density cores (Flores & Primack 1994; Moore 1994; Burkert 1995; Navarro, Frenk & White 1996; Burkert & Silk 1997; Stil 1999). This presents an important problem for the CDM picture, unless as suggested by Kravtsov et al. (1998) and Bullock et al. (1999), CDM halos are not as steeply cusped as suggested by others. Several solutions to the problem have already been suggested. Some argue against CDM, in favor of either modified Newtonian dynamics (McGaugh & de Blok 1998a,b) or self-interacting dark matter (Spergel & Steinhardt 1999). Others have suggested the presence of dark spheroids of baryons (Burkert & Silk 1997), or violent outflows of baryonic matter from stellar feedback (Navarro, Eke & Frenk 1996; Gelato & Sommer-Larsen 1999).
An important question in this regard is whether all dark matter halos have constant density cores as observed in the dwarfs. On the mass scale of clusters, gravitational lensing provides a promising tool to constrain the density profiles. So far, however, results have been contradictory, and the issue is still under debate (e.g., Tyson, Kochanski & Dell’Antonio 1998; Williams, Navarro & Bartelmann 1999). On the scale of massive galaxies, the most promising constraints on the density distributions of dark halos come from the rotation curves of low surface brightness (LSB) galaxies. It has been shown that the observed rotation curves of LSB galaxies are remarkably similar to those of dwarf galaxies, suggesting similar density profiles for their dark halos (Kravtsov et al. 1998; Stil 1999). Indeed, McGaugh & de Blok (1998a) have argued that, as for nearby dwarf galaxies, the rotation curves of LSB galaxies are inconsistent with the NFW profile (see also Navarro 1998).
These results imply that constant density cores are not restricted to low mass dwarf galaxies. However, while the rotation curves of dwarf galaxies are well resolved (due to their proximity), the data on more massive disk galaxies is typically of much poorer spatial resolution. Therefore, before using low resolution LSB rotation curves to put stringent constraints on the nature of dark matter, it is important that the effects of resolution (i.e. “beam smearing”) are properly taken into account.
In this paper we re-examine a set of HI rotation curves taken from the literature, focusing on LSB disk galaxies with $`V_{\mathrm{max}}80\mathrm{km}\mathrm{s}^1`$, as described in §2. We perform disk-halo decompositions by fitting both the observed HI surface brightness and rotation curve, including the effects of beam smearing and of the adiabatic contraction of the dark halo. In §3 we describe these models, and the method used to fit the data. We present the resulting fits for low resolution and high resolution LSB data in §4 and §5, respectively, and show that all fits are consistent with NFW profiles. We discuss the implications of this result in §6, and summarize in §7.
## 2 Sample
To identify an appropriate sample for measuring the dark matter profiles of massive galaxies, we have selected from the literature all disk galaxies that meet the following criteria: (i) central $`B`$-band surface brightness fainter than $`\mu _0(B)=23.0`$ mag arcsec<sup>-2</sup>; (ii) $`V_{\mathrm{max}}80\mathrm{km}\mathrm{s}^1`$; and (iii) readily available $`B`$-band surface photometry, HI rotation curves and HI surface brightness profiles. The 16 galaxies that meet these criteria are listed in Table 1, in order of increasing distance. In addition we list two dwarf galaxies, DDO 154 and NGC 3109, which we use for comparison in § 5.2.
To quantify the resolution of the HI rotation curves, we introduce the following two parameters: $`_1R_d/S`$ and $`_2R_{\mathrm{max}}/S`$, where $`R_d`$ is the scalelength of the stellar disk (in the $`B`$-band), $`R_{\mathrm{max}}`$ is the radius of the last measured point of the rotation curve, and $`S`$ is the FWHM of the beam of the HI observations along the beam’s major axis. Larger values of $`_1`$ and $`_2`$ imply that a rotation curve is better spatially resolved and has more independent data points. The majority of the galaxies listed in Table 1 are at relatively large distances and have $`_1<1`$. The only galaxy in the sample with $`V_{\mathrm{max}}80\mathrm{km}\mathrm{s}^1`$ that is well resolved is the nearby NGC 247, which has $`_1=7.4`$ and a rotation curve that extends for almost 25 beams.
Throughout this paper we adopt a Hubble constant of $`H_0=70\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$. All data is converted to this value. Where necessary, we list parameter values in units of $`h_{70}`$. All magnitudes and surface brightnesses are corrected for inclination and galactic extinction. Because of the poorly understood dust properties of LSB galaxies, we have not attempted a correction for internal extinction.
## 3 Modeling
### 3.1 Mass components
When fitting the rotation curves described above, we consider models with three mass-components: an infinitesimally thin gas disk, a thick stellar disk, and a spherical dark halo.
Neutral gas disks in external galaxies are typically exponential at large radii, but often contain a central hole. To model this profile, we adopt an HI surface density described by
$$\mathrm{\Sigma }_{\mathrm{HI}}(R)=\{\begin{array}{cc}\mathrm{\Sigma }_0(R/R_{\mathrm{HI}})^\beta \mathrm{e}^{R/R_{\mathrm{HI}}}\hfill & (R<R_c)\text{ ,}\hfill \\ 0\hfill & (RR_c)\text{ .}\hfill \end{array}$$
(1)
Here $`R_{\mathrm{HI}}`$ is the scalelength of the gas disk, $`R_c`$ is a cut-off radius, and $`\beta `$ describes the magnitude of the central suppression in the HI distribution. As we show below, this simple parameterization yields extremely good fits to the observed HI surface brightness. The total HI mass that corresponds to this surface density is
$$M_{\mathrm{HI}}=2\pi \mathrm{\Sigma }_0R_{\mathrm{HI}}^2\gamma (\beta +2,R_c/R_{\mathrm{HI}})$$
(2)
with $`\gamma `$ the incomplete gamma function. The circular velocity of the gas is computed using equation \[2-146\] of Binney & Tremaine (1987), assuming that the disk is infinitesimally thin and that the total surface density of the gas is 1.3 times that of the HI to take account of the mass of helium.
For the stellar disk we assume a thick exponential
$$\rho ^{}(R,z)=\rho _0^{}\mathrm{e}^{R/R_d}\mathrm{sech}^2(z/z_0)$$
(3)
where $`R_d`$ is the scalelength of the disk. Throughout we set $`z_0=R_d/6`$. The exact value of this ratio does not influence the results to any significant degree. The circular velocity of the stellar disk is computed using equation \[A.17\] in Casertano (1983), and properly scaled with the stellar $`B`$-band mass-to-light ratio $`\mathrm{{\rm Y}}_B`$. Except for the two giant LSB galaxies, UGC 6614 and F568-6, none of the galaxies in our sample has a significant bulge component.
We assume that initially the dark and baryonic matter virialize to form a spherical halo with a density distribution given by
$$\rho (r)=\frac{\rho _0}{(r/r_s)^\alpha (1+r/r_s)^{3\alpha }},$$
(4)
with $`r_s`$ being the scale radius of the halo, such that $`\rho r^\alpha `$ for $`rr_s`$ and $`\rho r^3`$ for $`rr_s`$. For $`\alpha =1`$ equation (4) reduces to the NFW profile. We define the concentration parameter $`c=r_{200}/r_s`$, with $`r_{200}`$ the radius inside of which the mean density is 200 times the critical density for closure, i.e.,
$$\frac{r_{200}}{h^1\mathrm{kpc}}=\frac{V_{200}}{\mathrm{km}\mathrm{s}^1}.$$
(5)
Here $`V_{200}`$ is the circular velocity at $`r_{200}`$, and $`h=H_0/100\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$. The circular velocity as function of radius can be written as
$$V_{\mathrm{halo}}(r)=V_{200}\sqrt{\frac{\mu (xc)}{x\mu (c)}}$$
(6)
with $`x=r/r_{200}`$ and
$$\mu (x)=\underset{0}{\overset{x}{}}y^{2\alpha }(1+y)^{\alpha 3}dy$$
(7)
The initial collapse of the gas and stellar disks within the virialized halo leads to a contraction of the dark matter halo. We assume that the baryons collapse slowly, adiabatically compressing the halo. We treat this adiabatic contraction of the dark halo following the procedure in Blumenthal et al. (1986) and Flores et al. (1993). We assume that initially the gas follows the same density distribution as the dark matter (equation ), and assume that all baryons end up in either the gas or the stellar disk. We thus ignore any possible hot gas that might be present in the halo, or that might have been blown out of the halo owing to stellar feedback processes; given that $`V_{\mathrm{max}}80\mathrm{km}\mathrm{s}^1`$ for the galaxies in our sample, feedback is not expected to be very efficient (Dekel & Silk 1986; Mac Low & Ferrara 1999). Our baryon fraction, defined as $`f_{\mathrm{bar}}=(M_{\mathrm{gas}}+M_{\mathrm{stars}})/M_{200}`$, where $`M_{200}=r_{200}V_{200}^2/G`$ is the total mass of the galaxy (baryons plus dark matter), thus depends on the values of $`r_{200}`$ and $`V_{200}`$ that follow from fitting the rotation curves, and can be used as a sanity check of the best-fit model parameters.
### 3.2 Beam smearing
Because of the finite resolution of HI mapping, we need to convolve our model of the underlying surface brightness profile $`\mathrm{\Sigma }(r)`$ with the effective point spread function $`P`$ (i.e. the beam) of the interferometer, in order to derive the observed surface brightness at a position $`(x,y)`$ on the plane of the sky:
$$\stackrel{~}{\mathrm{\Sigma }}(x,y)=\underset{0}{\overset{\mathrm{}}{}}drr\underset{0}{\overset{2\pi }{}}d\theta \mathrm{\Sigma }(r^{})P(r,\theta \theta _0).$$
(8)
Here $`r^{}=\sqrt{x^2+y^2}`$ where $`x^{}=x+r\mathrm{cos}\theta `$ and $`y^{}=(y+r\mathrm{sin}\theta )/\mathrm{cos}i`$ are the Cartesian coordinates in the equatorial plane of the disk, $`i`$ is the disk’s inclination angle, and $`\theta _0`$ is the angle between the major axes of the beam and the galaxy. We describe the beam by a two-dimensional Gaussian:
$$P(r,\theta )=\frac{1}{2\pi q\sigma ^2}\mathrm{exp}\left(\frac{r^2}{2\sigma ^2}\left[\mathrm{cos}^2\theta +\frac{\mathrm{sin}^2\theta }{q^2}\right]\right),$$
(9)
where $`q`$ is the flattening of the beam, and $`\sigma `$ is the Gaussian dispersion of the beam along its major axis.
Beam smearing also affects the observed rotation curve, by allowing gas from a wide range of radii to contribute to the observed rotation velocity, $`\stackrel{~}{V}_{\mathrm{rot}}`$, at a position $`(x,y)`$ on the plane of the sky:
$$\stackrel{~}{V}_{\mathrm{rot}}(x,y)=\frac{1}{\stackrel{~}{\mathrm{\Sigma }}}\underset{0}{\overset{\mathrm{}}{}}drr\underset{0}{\overset{2\pi }{}}d\theta \mathrm{\Sigma }(r^{})V_{\mathrm{los}}(x^{},y^{})P(r,\theta \theta _0)$$
(10)
(cf. Begeman 1989) where $`V_{\mathrm{los}}`$ is the line of sight velocity. Throughout we assume that the gas moves on circular orbits in the plane of the disk.
### 3.3 Fitting the data
In order to determine the parameters which best reproduce the observed surface brightness distribution $`\mathrm{\Sigma }_{\mathrm{obs}}`$ and rotation curve $`V_{\mathrm{obs}}`$, we adjust our models to minimize
$$\chi _{\mathrm{HI}}^2=\underset{i=1}{\overset{N_{\mathrm{HI}}}{}}\left(\frac{\mathrm{\Sigma }_{\mathrm{obs}}(R_i)\stackrel{~}{\mathrm{\Sigma }}(R_i)}{\mathrm{\Delta }\mathrm{\Sigma }_{\mathrm{obs}}(R_i)}\right)^2,$$
(11)
and
$$\chi _{\mathrm{vel}}^2=\underset{i=1}{\overset{N_{\mathrm{vel}}}{}}\left(\frac{V_{\mathrm{obs}}(R_i)\stackrel{~}{V}(R_i)}{\mathrm{\Delta }V_{\mathrm{obs}}(R_i)}\right)^2.$$
(12)
Here $`\mathrm{\Delta }\mathrm{\Sigma }_{\mathrm{obs}}`$ and $`\mathrm{\Delta }V_{\mathrm{obs}}`$ are the errorbars of the HI surface brightness and rotation velocities, respectively, as quoted in the literature. $`N_{\mathrm{HI}}`$ and $`N_{\mathrm{vel}}`$ are the number of data points.
For a given choice of the Hubble constant the models described above have eight free parameters to fit the data: $`\mathrm{\Sigma }_0`$, $`R_{\mathrm{HI}}`$, $`\beta `$, $`R_c`$, $`\mathrm{{\rm Y}}`$, $`\alpha `$, $`c`$, and $`V_{200}`$ (or equivalently $`r_{200}`$). Note, however, that $`\mathrm{\Sigma }_0`$ is completely determined by normalizing the models to the total mass in HI, and can thus be ignored in the fitting routine. In principle one wants to fit the surface brightness of the gas and the rotation curve simultaneously. However, a $`\chi ^2`$-minimization analysis using all seven parameters shows a very broad and noisy minimum and is computationally expensive. We therefore split the fitting procedure into two steps. We start by fitting the HI surface brightness, after which the observed rotation velocities are fit, while holding $`R_{\mathrm{HI}}`$, $`\beta `$, and $`R_c`$ fixed at their best-fit values. In both cases we use the downhill-simplex method (e.g., Press et al. 1992) combined with a simple random-walk routine to minimize $`\chi ^2`$. Tests on model galaxies show that this routine is robust in recovering the global minimum. Throughout we use equation (5) with $`h_{70}=1`$ to set $`r_{200}`$, and we require $`c1`$ and $`0\alpha 2`$.
In order to obtain confidence levels around the best fitting kinematical model we proceed as follows. We set $`\alpha `$ fixed at its best fit value and compute $`\chi _{\mathrm{vel}}^2`$ on a $`50\times 50`$ grid of $`(c,\mathrm{{\rm Y}}_B)`$-values by means of a fast one-dimensional $`\chi ^2`$-minimization routine to find the best fitting value of $`V_{200}`$. We use the $`\mathrm{\Delta }\chi ^2`$-statistic to compute confidence levels on $`(c,\mathrm{{\rm Y}}_B)`$ for the given value of $`\alpha `$. The same procedure is repeated with either $`c`$, or $`\mathrm{{\rm Y}}_B`$ held fixed at their best fitting values.
These confidence levels should not be interpreted as strict levels of confidence on the actual physical parameters. We have not taken any uncertainties into account on the inclination angle, distance, or the rotation curves of the gas and stellar disks (i.e., we assume that the stellar mass-to-light ratio is constant throughout the galaxy). Nor in most cases are the individual data points independent of each other. Rather, the confidence levels indicate to what extent the model parameters can be changed under the given assumptions, to obtain fits to the rotation curve that are statistically equivalent.
## 4 Low resolution data
In order to assess the importance of beam-smearing on low-resolution data with $`_11`$ we first analyze the six galaxies from the sample of de Blok, McGaugh & van der Hulst (1996; hereafter BMH96) that have $`V_{\mathrm{max}}80\mathrm{km}\mathrm{s}^1`$. We do not analyze F571-8 because its high inclination angle ($`i90^\mathrm{o}`$) requires a somewhat different modeling technique than used here. The best fit parameters to the HI surface density profiles are listed in Table 2. The corresponding fits to the data are shown in Figure 1. In each case we obtain an excellent fit to the data, except for F563-1, where the wiggle in $`\mathrm{\Sigma }_{\mathrm{HI}}`$ at $`14h_{70}^1`$kpc is not reproduced by our parameterized model. With the exception of F563-1, which is virtually equally well fit by a simple exponential, we find clear evidence for the presence of large holes in the HI distributions.
The best fits to the observed rotation velocities are shown in Figure 2. The parameters of the best-fit models are listed in Table 3. Circular velocities of the stellar disk are computed from the $`B`$-band surface photometry (de Blok, van der Hulst & Bothun 1995). The best-fit models reproduce the observed rotation velocities extremely well. The only exception is F563-1, for which we fail to reproduce the steep inner rise of the rotation curve despite the steep central cusp of the best-fit model ($`\alpha =2`$).
The parameters of the best fit models span a wide range. The central cusp slope $`\alpha `$ spans the entire range of allowed values from $`\alpha =2`$ (F563-1 and F568-1) to $`\alpha =0`$ (F583-1). The mass-to-light ratio, $`\mathrm{{\rm Y}}_B`$, spans a range from zero (F563-1 and F583-1) up to $`6.2\mathrm{M}_{}/\mathrm{L}_{}`$ (F568-1). Finally, the baryon fraction, $`f_{\mathrm{bar}}`$, covers the range from $`0.01`$ (F568-3) to $`0.37`$ (F568-1). Note that for realistic cosmologies we expect $`f_{\mathrm{bar}}0.050.1`$, based on a baryon density of $`\mathrm{\Omega }_{\mathrm{bar}}=0.0125h^2`$ (Walker et al. 1991). Clearly, the best fit models are not always the most physically meaningful solutions. More interesting, therefore, is to explore the areas of parameter space that yield solutions that are consistent with the data.
Figures 3 and 4 show contour plots of confidence levels around the best fit model for three two-dimensional cuts through parameters space. In all cases we find that large areas of parameter space are consistent with the data. The two extreme cases are F568-1 and F574-1 where virtually the entire parameter space probed falls within the 68.3 percent confidence level. Clearly, the data of the BMH96 sample is of insufficient spatial resolution to meaningfully constrain the density distributions of the dark matter halos.
To emphasize this point, in Figure 5 we plot the results for three models of F574-1 that all fall within the 68.3 percent confidence level of the best fit model. These models are indicated by small labeled dots in Figure 4, and the corresponding parameters are listed in Table 3. Models $`a`$, $`b`$, and $`c`$ have wildly different mass profiles, yet they all fit the data remarkably well; the observed rotation velocities of F574-1 are of little use in constraining the mass model.
Another important result is that the maximum mass-to-light ratios which are consistent with the data are much larger than those derived by de Blok & McGaugh (1997) using a standard maximal-disk fit (i.e., by scaling the contribution of the disk to fit the inner-most data point of the rotation curve). This inner point, however, is the point most severely influenced by beam smearing, and is artificially reduced, leading to small mass-to-light ratios (cf., Blais-Ouellette, Carignan & Amram 1998). Our results indicate that, based on the observed rotation curves, one can currently not rule out that the central regions of most of the LSB galaxies are dominated by baryons rather than dark matter.
In our fitting procedure, we consider a range of $`c`$ and $`\mathrm{{\rm Y}}_B`$ that is much larger than the expected spread, leading sometimes to unphysical results; for instance, for small values of $`c`$ we often find unrealistically high values of $`V_{200}`$ (sometimes in excess of $`2000\mathrm{km}\mathrm{s}^1`$). Clearly, we do not expect LSB disks to be embedded in halos this massive. In addition, $`B`$-band mass-to-light ratios of $`10`$ are rather high given the typical colors of LSB galaxies, and might imply self-gravitating, and hence unstable, disks. Finally, we could be much more restrictive by requiring realistic baryon fractions. However, the main aim of this exercise is to illustrate that rotation curves with $`_11`$ themselves can not be used to obtain any meaningful constraints on the density profiles or masses of dark halos.
Recently, Swaters (1999) obtained H$`\alpha `$ rotation curves for five LSB galaxies in the BMH96 sample (four of which have been analyzed here). These high resolution data reveal much steeper inner rotation rotation curves than for the HI data of BMH96, clearly indicating that the HI rotation curves are severely affected by beam smearing. The H$`\alpha `$ rotation curves of Swaters (1999) are in good agreement with our unsmeared best-fit rotation curves.
Reliable and strict constraints on the dark matter halos of disk galaxies thus requires much higher spatial resolution than for the sample of BMH96. Except for NGC 247, all galaxies in Table 1 with $`V_{\mathrm{max}}80\mathrm{km}\mathrm{s}^1`$ have been observed with similar spatial resolution as the galaxies of the sample of BMH96 analyzed here. The two giant LSB galaxies, UGC 6614 and F568-6, have $`_1`$ somewhat larger than unity, but this owes mainly to the extremely large disk scalelength. These galaxies have significant bulge components which complicate constraining the density distribution of the halo component. We therefore refrain from analyzing any of the other low-resolution data, and shift our attention to the galaxy which has been observed with the highest spatial resolution: NGC 247.
## 5 High resolution data
### 5.1 The nearby LSB galaxy NGC 247
The LSB galaxy NGC 247 is a member of the nearby ($`2.5`$ Mpc) Sculptor group and has been observed in HI by Carignan & Puche (1998). Surface photometry is taken from Carignan (1985b). The proximity of NGC 247 ensures that its HI velocity field has been very well resolved. The FWHM of the beam extends only $`13`$ percent of the scalelength of the stellar disk, and the HI rotation curve is measured out to 25 beams. Therefore, this galaxy is the ideal candidate to obtain useful constraints on the density distribution of its dark matter halo. Unfortunately, to date it is also the *only* candidate with sufficient spatial resolution that also meets our selection criteria (see § 2).
Although for NGC 247 the effects of beam smearing can easily be ignored, we have used the same analysis for NGC 247 as for the low resolution data. As a first step we fit the observed HI surface brightness. No errorbars for $`\mathrm{\Sigma }_{\mathrm{HI}}`$ are given, so we assign each data point the same (arbitrary) error in our analysis. It turns out that the particular parameterization of equation (1) yields a very poor fit. After some experimenting, we found a good fit for a surface brightness profile of the form:
$$\mathrm{\Sigma }_{\mathrm{HI}}(R)=\mathrm{\Sigma }_0\mathrm{exp}\left(R_2\left[R_1^\beta +R_2^\beta \right]^{1/\beta }\right),$$
(13)
with $`R_1=R/R_{\mathrm{HI}}`$ and $`R_2=R_c/R_{\mathrm{HI}}`$. The best fit parameters are listed in Table 2, and the resulting fit is shown in the upper left panel of Figure 6.
The upper right panel shows the best fit model to the observed rotation curve. This model is completely dark matter dominated, and has a dark halo with $`\alpha =1.02`$, $`c=7.2`$, and $`V_{200}=93.1\mathrm{km}\mathrm{s}^1`$. The best-fit mass-to-light ratio of the stellar disk is $`\mathrm{{\rm Y}}_B=1.0\mathrm{M}_{}/\mathrm{L}_{}`$, resulting in a baryon fraction of $`f_{\mathrm{bar}}=0.011`$. Figure 7 shows that these parameters are well constrained. We can rule out models with $`\alpha <0.55`$ and $`\alpha >1.26`$ at the 99.73 percent confidence level; the improvement with respect to the low resolution data of BMH96 is apparent. Note that the parameters of the best fit model are remarkably consistent with a NFW profile with a concentration parameter that is typical for realistic cosmologies (Navarro, Frenk & White 1997).
### 5.2 Comparison to nearby dwarfs
In order to contrast the results for NGC 247 with dwarf galaxies, we have applied the same analysis to DDO 154 and NGC 3109. These galaxies, observed by Carignan & Beaulieu (1989) and Jobin & Carignan (1990) respectively, were shown by Moore (1994) to be inconsistent with a dark matter halo that follows a Hernquist (1990) profile. Both galaxies are nearby and have been observed with good spatial resolution (see Table 1). For NGC 3109 $`_1=5.2`$ and $`_2=27.8`$ such that it has one of the best resolved HI velocity fields. While DDO 154 has only $`_10.7`$, the radius where the rotation curve changes slope is significantly larger than the beam size.
Our best fit models to the data are shown in Figure 6 and listed in Tables 2 and 3. As for NGC 247, no errorbars for the HI surface density are available, and we assign each data point the same (arbitrary) error. The best fit kinematic models of both galaxies are remarkably similar. Both have $`\alpha =0`$ ($`\alpha 0.5`$ at the 99.73 percent confidence level), $`c12`$ and $`\mathrm{{\rm Y}}_B=0`$. The contour plots of the confidence levels are shown in Figure 7. Clearly, the observed rotation curves of these dwarfs constrain the models extremely well, although the unphysical mass-to-light ratio suggests that our underlying model for either the surface brightness or the detailed halo shape must not be quite correct. Although firm conclusions cannot be made on the basis of three galaxies only, these preliminary results suggest that constant density cores may be a feature of only the least massive galaxies. The low resolution data of BMH96 is at least consistent with this picture.
## 6 Discussion
The disk-halo decompositions of LSB galaxies with $`V_{\mathrm{max}}80\mathrm{km}\mathrm{s}^1`$ presented in §4 call into question a number of strong claims that have been made in the literature. Foremost is the conclusion that the HI rotation curves of LSB disk galaxies are inconsistent with dark matter halos with steep central cusps (McGaugh & de Blok 1998a; Navarro 1998; Salucci & Persic 1997), and by implication, that observations are inconsistent with the predictions of CDM cosmologies (Navarro, Frenk & White 1997; Fukushige & Makino 1997; Moore et al. 1998). In contrast to previous interpretations, we find that for all but the lowest mass galaxies ($`V_{\mathrm{max}}60\mathrm{km}\mathrm{s}^1`$) the observations are of sufficiently low resolution that they place little or no constraints on the inner shapes of dark matter density profiles. In fact, the single case which we have found in the literature which is capable of limiting the dark matter profile (NGC 247; §5) has an inner slope which is *exactly* in agreement with the CDM predictions of Navarro, Frenk & White (1997).
Recently, Kravtsov et al. (1998) claimed that the rotation curves of LSB galaxies are self-similar and have the same shape as for dwarf galaxies (see also Burkert 1995). This presents another problem for CDM, since the actual density profiles of dark halos (and hence the corresponding rotation curves) are expected to reveal a significant amount of scatter (Avila-Reese et al. 1999; Ying 1999; Bullock et al. 1999). However, our results suggest that the observed self-similarity is likely to be a consequence of the large amount of beam smearing in the existing LSB observations.
The apparent shallowness of LSB rotation curves had previously lead astronomers to place strong upper limits on $`\mathrm{\Omega }_0`$. For example, using halo-only NFW fits to a wide variety of disk galaxies, Pickering et al. (1997) and Navarro (1998) derived low upper limits for the concentration of the halo, and thus argued for a low-density Universe with $`\mathrm{\Omega }_00.2`$, with the strongest constraints coming from LSB galaxies from the samples of BMH96 and Pickering et al. (1997). As the fits to these samples did not include beam smearing, the results presented here suggest that the best-fit halo concentration is likely to have been severely underestimated, and thus much larger values of $`\mathrm{\Omega }_0`$ are permitted by the data. For comparison, our most restrictive case in our analysis, F568-3, has $`c<8.3`$ at the 68.3 percent confidence level; this limit would be even higher if we had not taken account of the contributions of the gas and stellar disks, and adiabatic contraction of the halo.
While our analysis suggests that the halos of massive galaxies ($`V_{\mathrm{max}}80\mathrm{km}\mathrm{s}^1`$) are fully consistent with centrally concentrated dark matter halos and thus with large values of $`\mathrm{\Omega }_0`$, we find that the halos of dwarf galaxies continue to tell a different story. Unlike their more massive LSB counterparts, there is compelling evidence that the halos of dwarf galaxies do indeed have shallow central cores. If future high-resolution studies reveal the existence of large cores in massive LSB galaxies as well, we will need to revise the standard model for the nature of dark matter (e.g., Spergel & Steinhardt 1999). If, however, these studies confirm the presence of steep central cusps in galaxies other than NGC 247, then the data on dwarf galaxies suggest that there is an intrinsic change in the structure of dark matter halos at $`V_{200}100\mathrm{km}\mathrm{s}^1`$. Although this violates the scale-free behavior predicted by almost any variant of collisionless dark matter models, it is not necessarily in contradiction with these models, if, for instance, it can be caused by baryonic physical processes. One of the most promising candidate processes is outflows (or “feedback”), a process which is known to exist in dwarf galaxies (e.g., Puche et al. 1992; Hunter, Hawley & Gallagher 1993; Martin 1999), and which is necessary to explain the slope and scatter of the Tully-Fisher relation (e.g., van den Bosch 1999). After outflows remove mass from the center of the galaxy on a short timescale, the remaining material revirializes to form a less centrally condensed system. $`N`$-body simulations of this process (Navarro et al. 1996) have shown that violent outflows of baryonic matter from the center of a dark halo do result in the creation of a constant density core (see also Gelato & Sommer-Larsen 1999).
We wish (1) to test if outflows can be sufficiently large to produce cores in dwarf galaxies, and (2) to measure the velocity scale at which this process becomes ineffective. In the Appendix, we use a simple prescription to estimate how the structure of the galaxy responds to a realistic outflow. We assume that supernovae expel some fraction of the baryonic mass from the galaxy on a time-scale that is short compared to the dynamical time (i.e., resulting in a non-adiabatic change of the potential), and we tune the feedback efficiency to match the Tully-Fisher relation. The blow-out of baryonic material leads to a change in the profile of dark matter. We assume that the mass of dark matter which can essentially be displaced by this process is proportional to the mass of baryonic material lost by the galaxy, with some proportionality constant $`\eta `$. Assuming that when the halo revirializes, the “rearranged” dark matter mass winds up outside of the scale radius $`r_s`$, one can compute the final cusp slope $`\alpha _f`$ of the revirialized halo.
The resulting cusp slopes are plotted in Figure 8 as function of $`V_{200}`$. The locations of NGC 247, DDO 154 and NGC 3109 are plotted for comparison. Results are shown for halos with $`c=5`$, 10, and 20, which roughly covers the range of expected values, and for initial halos with both $`\alpha _i=1`$ (NFW profile) and $`\alpha _i=1.5`$ (the value suggested by Moore et al. 1998). The models clearly show that the final cusp slope decreases with decreasing halo mass, such that dwarf galaxies could have significantly flatter cusps than more massive galaxies, in spite of having the same initial dark matter profile. If $`\eta 1`$, halos initially follow an NFW profile, and have $`c10`$ (as for NGC 3109, DDO 154, and NGC 247), this model predicts constant density cores in systems with $`V_{200}75\mathrm{km}\mathrm{s}^1`$, in reasonable agreement with observations. If, however, $`\alpha _i=1.5`$, there is too much mass in the center for violent outflows to destroy the central cusp; even systems with $`V_{200}=30\mathrm{km}\mathrm{s}^1`$ will not have constant density cores. In this case, the efficiency of relocating dark matter has to be larger. For $`\eta 2`$, the resulting cusp slopes are in good agreement with the high-resolution data analyzed here. These results suggest that even if the initial structure of dark matter halos is identical for all galaxies, “gastrophysics” can easily introduce a systematic change in the profile with mass.
While this adhoc procedure ignores the detailed physics of how the dark matter adjusts to an instantaneous mass loss, it is still useful for setting the overall scale over which outflows could be effective in creating dark matter cores. Detailed $`N`$-body simulations are required to test the importance of our assumptions (i.e., constant $`r_s`$ and $`c`$) and to obtain estimates of $`\eta `$. Our results also depend on the particular model for the feedback (i.e., $`ϵ_{\mathrm{SN}}^0`$ and $`\nu `$, see the Appendix). With these caveats in mind, it is nevertheless reassuring that this simple model, with feedback parameters that are motivated by empirical data on the Tully-Fisher relation, yields results that are at least of the correct order of magnitude to reconcile the constant density cores in dwarfs with CDM predictions.
## 7 Summary and Conclusions
We have re-examined the disk-halo decompositions of several LSB disk galaxies with $`V_{\mathrm{max}}80\mathrm{km}\mathrm{s}^1`$, improving upon previous studies by taking beam smearing and adiabatic contraction into account. Contrary to previous claims we find that the rotation curves of these systems are consistent with dark halos with steep density cusps. The actual slope of this density cusp is very poorly constrained: the data can not be used to argue against either constant density cores or against very steep cusps such as those suggested by Fukushige & Makino (1997) and Moore et al. (1998). Our results also call into question the observed self-similarity of LSB rotation curves (Kravtsov et al. 1998), and strongly reduce the constraints on $`\mathrm{\Omega }_0`$ (Pickering et al. 1997; Navarro 1998).
While we have shown that existing HI rotation curves for LSB galaxies are consistent with predictions of CDM and a universal baryon fraction, they cannot place any meaningful constraints. Observations at much higher resolution are required. Given the typical beam size of HI synthesis observations, this limits us to relatively nearby galaxies. Most nearby LSB galaxies, however, are dwarf galaxies with $`V_{\mathrm{max}}60\mathrm{km}\mathrm{s}^1`$. The only LSB galaxy that meets our selection criteria and is well resolved is NGC 247, for which the HI rotation curve extends for almost 25 beams, and constrains the model parameters very well ($`0.55\alpha 1.26`$ at the 99.73 percent confidence level with a best-fit cusp slope of $`\alpha =1.02`$). This contrasts strongly with the nearby dwarf galaxies DDO 154 and NGC 3109, where we find $`\alpha 0.5`$ at 99.73 percent confidence with a best-fit value of $`\alpha =0`$ (i.e., a constant density core). This suggests that constant density cores are only present in dwarf galaxies, whereas more massive galaxies have dark halos with a steep central cusp as expected for CDM. We have shown that violent outflows due to supernovae feedback can account for such a mass-dependence of the shape of dark matter halos.
Although the low-resolution data examined here is consistent with this picture, it is premature to draw conclusions about the density profiles at high mass based upon the single case of NGC 247. What are needed at this point are high resolution rotation curves for a sufficiently large sample of LSB disk galaxies with $`V_{\mathrm{max}}80\mathrm{km}\mathrm{s}^1`$. Unless these galaxies are relatively nearby ($`10`$ Mpc), there is little hope that HI observations will yield sufficient spatial resolution. We need to focus on observations in either H$`\alpha `$ (i.e., Blais-Ouellette et al. 1998; Swaters 1999) or CO (Sofue et al. 1999). These tracers of the velocity field can be observed at much higher spatial resolution, and, when combined with HI rotation curves to sample the outer parts of the velocity field, provide promising tools to examine whether more massive disk galaxies have cusped dark halos similar to NGC 247. Independent of the outcome, such studies will have strong implications for theories of galaxy formation and for the nature of the dark matter.
We are grateful to Craig Hogan for a critical reading of an early draft of the paper, and to George Lake and Julio Navarro for valuable discussions. FvdB was supported by NASA through Hubble Fellowship grant # HF-01102.11-97.A awarded by the Space Telescope Science Institute, which is operated by AURA for NASA under contract NAS 5-26555. BER was supported by the Washington Space Grant Consortium, under a grant from NASA, and by the Mary Gates Endowment for Students.
## Appendix A Supernovae feedback and the mass-dependence of halo cusp slopes
We examine if outflows can explain a change in the central cusp slope of dark matter halos from $`\alpha 0`$ for systems with $`V_{\mathrm{max}}70\mathrm{km}\mathrm{s}^1`$ (i.e., NGC 3109) to $`\alpha 1`$ for systems with $`V_{\mathrm{max}}100\mathrm{km}\mathrm{s}^1`$ (i.e., NGC 247).
As a mechanism for producing outflows of baryonic matter we consider feedback from supernovae (SN). The total energy produced by SN can be written as
$$E_{\mathrm{tot}}=\eta _{\mathrm{SN}}E_{\mathrm{SN}}M_{}$$
(A1)
with $`\eta _{\mathrm{SN}}`$ the number of SN produced per Solar mass of stars formed, $`E_{\mathrm{SN}}`$ the energy produced per SN, and $`M_{}`$ the total mass in stars. We assume that this energy expels a mass $`M_{\mathrm{fb}}`$ of baryons from the halo. Taking account of the system’s escape velocity and requiring energy balance, one obtains
$$M_{\mathrm{fb}}=\frac{ϵ_{\mathrm{SN}}\eta _{\mathrm{SN}}E_{\mathrm{SN}}M_{}}{V_{200}^2},$$
(A2)
(cf. Kauffmann et al. 1993; Natarajan 1999) where we introduce the parameter $`ϵ_{\mathrm{SN}}`$ which describes the efficiency with which the SN energy expels baryonic mass from the halo. Conservation of baryon mass<sup>1</sup><sup>1</sup>1Note that we make the oversimplifying assumption that all the baryonic matter is in stars or expelled; no cold gas is taken into account. requires that $`M_{}+M_{\mathrm{fb}}=f_{\mathrm{bar}}M_{200}`$. Setting $`\eta _{\mathrm{SN}}=4\times 10^3\mathrm{M}_{}^{}{}_{}{}^{1}`$, $`E_{\mathrm{SN}}=10^{51}`$ ergs, and writing
$$ϵ_{\mathrm{SN}}=ϵ_{\mathrm{SN}}^0\left(\frac{V_{200}}{250\mathrm{km}\mathrm{s}^1}\right)^\nu $$
(A3)
one obtains
$$M_{\mathrm{fb}}=\frac{f_{\mathrm{bar}}M_{200}}{1+\frac{1}{3.22ϵ_{\mathrm{SN}}^0}\left(\frac{V_{200}}{250\mathrm{km}\mathrm{s}^1}\right)^{2\nu }}.$$
(A4)
As a model for the feedback we consider $`ϵ_{\mathrm{SN}}^0=0.05`$ and $`\nu =0.3`$ and we set the baryon fraction to $`f_{\mathrm{bar}}=0.085`$. Semi-analytical models for the formation of disk galaxies, presented by van den Bosch (1999) and van den Bosch & Dalcanton (1999) have shown that for these values one obtains a near-infrared Tully-Fisher relation with the correct slope, scatter and normalization. In addition, adopting a feedback model with these parameters yields an absence of high surface brightness dwarf galaxies, as observed.
We now assume that the blow-out of mass $`M_{\mathrm{fb}}`$ results in the relocation of a dark matter mass of $`\mathrm{\Delta }M_{\mathrm{DM}}=\eta M_{\mathrm{fb}}`$ to larger radii, with $`\eta `$ of order unity. In order to quantify $`\mathrm{\Delta }M_{\mathrm{DM}}`$ in terms of the cusp slopes of the dark halos, we make the following simplifying assumption. We assume that both the initial and final halos have density profiles given by equation (4) with $`\alpha =\alpha _i`$ and $`\alpha =\alpha _f`$, respectively. In addition, we assume that both halos have identical dark matter masses, scale-radii, $`r_s`$, and concentration parameters, $`c`$. Defining $`\mathrm{\Delta }M_{\mathrm{DM}}`$ as the difference in halo mass within a radius $`r_s`$ we obtain
$`\mathrm{\Delta }M_{\mathrm{DM}}M_i(r_s)M_f(r_s)`$ (A5)
$`=(1f_{\mathrm{bar}})M_{200}\left[{\displaystyle \frac{\mu _i(1)}{\mu _i(c)}}{\displaystyle \frac{\mu _f(1)}{\mu _f(c)}}\right]`$
with $`\mu _i(x)`$ and $`\mu _f(x)`$ given by equation (7) with $`\alpha `$ set to the initial and final values, respectively.
Using equations (A2) and (A5) one can compute the cusp slope $`\alpha _f`$ as function of $`V_{200}`$, $`c`$, $`\alpha _i`$, and $`\eta `$ by solving for the root of $`\mathrm{\Delta }M_{\mathrm{DM}}\eta M_{\mathrm{fb}}=0`$. The results are shown in Figure 8 and discussed in § 6. |
no-problem/9911/quant-ph9911040.html | ar5iv | text | # Coloring the rational quantum sphere and the Kochen-Specker theorem
## 1 Colorings
In what follows we shall consider “rational rays.” A “rational ray” is the linear span of a non-zero vector of $`𝐐^n𝐑^n`$.
Let $`p`$ be a prime number. A coloring of the rational rays of $`𝐑^n`$, $`n1`$, using $`p^{n1}+p^{n2}+\mathrm{}+1`$ colors can be constructed in a straightforward manner. We refer to for the theoretical background of the following construction.
Each rational ray is the linear span of a vector $`(x_1,x_2,\mathrm{},x_n)𝐙^n`$, where $`x_1,x_2,\mathrm{},x_n`$ are coprime. Such a vector is unique up to a factor $`\pm 1`$.
Next, let $`𝐙_p`$ be the field of residue classes modulo $`p`$. The vector space $`𝐙_p^n`$ has $`p^n1`$ non-zero vectors; each ray through the origin of $`𝐙_p^n`$ has $`p1`$ non-zero vectors. So there are exactly $`(p^n1)/(p1)=p^{n1}+p^{n2}+\mathrm{}+1`$ distinct rays through the origin which can be colored with $`p^{n1}+p^{n2}+\mathrm{}+1`$ distinct colors.
Finally, assign to the ray $`Sp(x_1,x_2,\mathrm{},x_n)`$ (“$`Sp`$” denotes linear span) the color of the ray of $`𝐙_p^n`$ which is obtained by taking the modulus of the coprime integers $`x_1,x_2,\mathrm{},x_n`$ modulo $`p`$. Observe that $`x_1,x_2,\mathrm{},x_n`$ cannot vanish simultaneously modulo $`p`$ and that $`\pm (x_1,x_2,\mathrm{},x_n)`$ yield the same color. Obviously, all $`p^{n1}+p^{n2}+\mathrm{}+1`$ colors are actually used.
In what follows, we consider the case $`p=2`$, $`n=3`$. Here all rational rays $`Sp(x,y,z)`$ (with $`x,y,z𝐙`$ coprime) are colored according to the property which ones of the components $`x,y,z`$ are even (E) and odd (O). There are exactly $`7`$ of such triples OEE, EOE, EEO, OOE, EOO, OEO, OOO which are associated with one of seven different colors $`\mathrm{\#}1,\mathrm{\#}2,\mathrm{\#}3,\mathrm{\#}4,\mathrm{\#}5,\mathrm{\#}6,\mathrm{\#}7`$. Only the EEE triple is excluded. Those seven colors can be identified with the seven points of the projective plane over $`𝐙_2`$; cf. Fig. 1.
Next, we restrict our attention to those rays which meet the rational unit sphere $`S^2𝐐^3`$. The following statements on a triple $`(x,y,z)𝐙^3\{(0,0,0)\}`$ (not necessarily coprime) are equivalent:
(i) The ray $`Sp(x,y,z)`$ intersects the unit sphere at two rational points; i.e., it contains the rational points $`\pm (x,y,z)/\sqrt{x^2+y^2+z^2}S^2𝐐^3`$.
(ii) The Pythagorean property holds, i.e., $`x^2+y^2+z^2=n^2,n𝐍`$.
This equivalence can be demonstrated as follows. All points on the rational unit sphere can be written as $`𝐫=(\frac{a}{a^{}},\frac{b}{b^{}},\frac{c}{c^{}})`$ with $`a,b,c𝐙`$, $`a^{},b^{},c^{}𝐙\{0\}`$, and $`\left(\frac{a}{a^{}}\right)^2+\left(\frac{b}{b^{}}\right)^2+\left(\frac{c}{c^{}}\right)^2=1`$. Multiplication of $`𝐫`$ with $`a_{}^{}{}_{}{}^{2}b_{}^{}{}_{}{}^{2}c_{}^{}{}_{}{}^{2}`$ results in a vector of $`𝐙^3`$ satisfying (ii). Conversely, from $`x^2+y^2+z^2=n^2,n𝐍`$, we obtain the rational unit vector $`(\frac{x}{n},\frac{y}{n},\frac{z}{n})S^2𝐐^3`$.
Notice that this Pythagorean property is rather restrictive. Not all rational rays intersect the rational unit sphere. For a proof, consider $`Sp(1,1,0)`$ which intersects the unit sphere at $`\pm (1/\sqrt{2})(1,1,0)S^2𝐐^3`$. Although both the set of rational rays as well as $`S^2𝐐^3`$ are dense, there are “many” rational rays which do not have the Pythagorean property.
If $`x,y,z`$ are chosen coprime then a necessary condition for $`x^2+y^2+z^2`$ being a non-zero square is that precisely one of $`x`$, $`y`$, and $`z`$ is odd. This is a direct consequence of the observation that any square is congruent to 0 or 1, modulo 4, and from the fact that at least one of $`x`$, $`y`$, and $`z`$ is odd. Hence our coloring of the rational rays induces the following coloring of the rational unit sphere with those three colors that are represented by the standard basis of $`𝐙_2^3`$:
color #1 if $`x`$ is odd, $`y`$ and $`z`$ are even,
color #2 if $`y`$ is odd, $`z`$ and $`x`$ are even,
color #3 if $`z`$ is odd, $`x`$ and $`y`$ are even.
All three colors occur, since the vectors of the standard basis of $`𝐑^3`$ are colored differently.
Suppose that two points of $`S^2𝐐^3`$ are on rays $`Sp(x,y,z)`$ and $`Sp(x^{},y^{},z^{})`$, each with coprime entries. The inner product $`xx^{}+yy^{}+zz^{}`$ is even if and only if the inner product of the corresponding basis vectors of $`𝐙_2^3`$ is zero or, in other words, the points are colored differently. In particular, three points of $`S^2𝐐^3`$ with mutually orthogonal position vectors are colored differently.
From our considerations above, three colors are sufficient to obtain a coloring of the rational unit sphere $`S^2𝐐^3`$ such that points with orthogonal position vectors are colored differently, but clearly this cannot be accomplished with two colors. So the “chromatic number” for the rational unit sphere is three. This result is due to Godsil and Zaks ; they also showed that the chromatic number of the real unit sphere is four. However, they obtained their result in a slightly different way. Following all rational rays are associated with three colors by making the following identification:
$`\mathrm{\#}1,`$
$`\mathrm{\#}2=\mathrm{\#}4,`$
$`\mathrm{\#}3=\mathrm{\#}5=\mathrm{\#}6=\mathrm{\#}7.`$
This 3-coloring has the property that coplanar rays are always colored by using only two colors; cf. Fig. 1. According to our approach this intermediate 3-coloring is not necessary, since rays in colors #4, #5, #6, #7 do not meet the rational unit sphere.
### 1.1 Reduced two-coloring
As a corollary, the rational unit sphere can be colored by two colors such that, for any arbitrary orthogonal tripod spanned by rays through its origin, one vector is colored by color #1 and the other rays are colored by color #2. This can be easily verified by identifying colors #2 & #3 from the above scheme. (Two equivalent two-coloring schemes result from a reduced chromatic three-coloring scheme by requiring that color #1 is associated with $`x`$ or $`y`$ being odd, respectively.)
Kent has shown that there also exist dense sets in higher dimensions which permit a reduced two-coloring. Unpublished results by P. Ovchinnikov, O.G.Okunev and D. Mushtari state that the rational d-dimensional unit sphere is d-colorable if and only if it admits a reduced two-coloring if and only if $`d<6`$.
### 1.2 Denseness of single colors
It can also be shown that each color class in the above coloring schemes is dense in the sphere. To prove this, Godsil and Zaks consider $`\alpha `$ such that $`\mathrm{sin}\alpha =\frac{3}{5}`$ and thus $`\mathrm{cos}\alpha =\frac{4}{5}`$. $`\alpha `$ is not a rational multiple of $`\pi `$; hence $`\mathrm{sin}(n\alpha )`$ and $`\mathrm{cos}(n\alpha )`$ are non-zero for all integers $`n`$. Let $`F`$ be the rotation matrix about the $`z`$-axis through an angle $`\alpha `$; i.e.,
$$F=\left(\begin{array}{ccc}\mathrm{cos}\alpha & \mathrm{sin}\alpha & 0\\ \mathrm{sin}\alpha & \mathrm{cos}\alpha & 0\\ 0& 0& 1\end{array}\right).$$
Then the image $`I`$, under the powers of $`F`$, of the point $`(1,0,0)`$ is a dense subset of the equator.
Now suppose that the point $`u=(\frac{a}{c},\frac{b}{c},0)`$ is on the rational unit sphere and that $`a,c`$ are odd and thus $`b`$ is even. In the coloring scheme introduced above, $`u`$ has the same color as $`(1,0,0)`$ (identify $`a=c=1`$ and $`b=0`$); and so does $`Fu`$. This proves that $`I`$ (the image of all powers of $`F`$ of the points $`u`$) is dense. We shall come back to the physical consequences of this property later.
In the reduced two-color setting, if the two ”poles” $`\pm (0,0,1)`$ acquire color #1, then the entire equator acquires color #2. Thus, for example, for the two tripods spanned by $`\{(1,0,0),(0,1,0),(0,0,1)\}`$ and $`\{(3,4,0),(4,3,0),(0,0,1)\}`$, the first two legs have color #2, while $`(0,0,1)`$ has color #1.
### 1.3 The chromatic number of the real unit sphere in three dimensions is four
A proof that four colors suffice for the coloring of points of the unit sphere in three dimensions is constructive and rather elementary. Consider first the intersection points of the sphere with the the $`x`$, the $`y`$ and the $`zaxis`$, colored by green, blue and red, respectively. There are exactly three great circles which pass through two of these three pairs of points. The great circles can be colored with the two colors used on the four points they pass through. The three great circles divide the sphere into eight open octants of equal area. Four octants, say, in the half space $`z>0`$, are colored by the four colors red, white, green and blue. The remaining octants obtain their color from their antipodal octant.
Although Godsil and Zaks’ paper is not entirely specific, it is easy to write down an explicit coloring scheme according to the above prescription. Consider spherical coordinates: let $`\theta `$ be the angle between the $`z`$axis and the line connecting the origin and the point, and $`\phi `$ be the angle between the $`x`$axis and the projection of the line connecting the origin and the point onto the $`xy`$plane. In terms of these coordinates, an arbitrary point on the unit sphere is given by $`(\theta ,\phi ,r=1)(\theta ,\phi )`$.
* The colors of the cartesian coordinate axes $`(\pi /2,0)`$, $`(\pi /2,\pi /2)`$, $`(0,0)`$ are green, blue and red, respectively.
* The color of the octant $`\{(\theta ,\phi )0<\theta \pi /2,\mathrm{\hspace{0.33em}0}\phi <\pi /2\}`$ is green.
* The color of the octant $`\{(\theta ,\phi )0\theta <\pi /2,\pi /2\phi \pi \}`$ is red.
* The color of the octant $`\{(\theta ,\phi )0<\theta <\pi /2,\pi <\phi <3\pi /2\}`$ is white.
* The color of the octant $`\{(\theta ,\phi )0<\theta \pi /2,\pi /2\phi <0\}`$ is blue.
* The colors of the points in the half space $`z<0`$ are inherited from their antipodes. This completes the coloring of the sphere.
The fact that three colors are not sufficient is not so obvious. Here we shall not review Godsil and Zaks’ proof based on a paper by A. W. Hales and E. G. Straus , but refer to a result of S. Kochen and E. Specker , which is of great importance in the present debate on hidden parameters in quantum mechanics. They have proven that there does not exist a reduced two-coloring, also termed valuation, on the one dimensional subspaces of real Hilbert space in three dimensions.
Recall that a reduced two-coloring of the one dimensional linear subspaces with two colors could immediately be obtained from any possible appropriate coloring of the sphere with three colors by just identifying two of the three colors. Thus, the impossibility of a reduced two-coloring implies that three colors are not sufficient for an appropriate coloring of the threedimensional real unit sphere. (“Appropriate” here means: “points at spherical distance $`\pi /2`$ get different colors.”)
In the same article , Kochen and Specker gave an explicit example (their $`\mathrm{\Gamma }_3`$) of a finite point set of the sphere with weaker properties which suffice just as well for this purpose: the structure still allows for a reduced two-coloring, yet it cannot be colored by three colors. (The authors did not mention nor discuss this particular feature .)
The impossibility of a reduced two-coloring also rules out another attempt to “nullify” the Kochen-Specker theorem by identifying pairs of colors of an appropriate four-coloring of the real unit sphere. Any such identification would result in tripods colored by #1-#2-#2, as well as for instance #1-#1-#2, which is not allowed for reduced coloring schemes, which requires colorings of the type #1-#2-#2.
## 2 Physical aspects
### 2.1 Physical truth values
Based on Godsil and Zaks results, Meyer suggested that the physical impact of the Kochen-Specker theorem is “nullified,” since for all practical purposes it is impossible to operationalize the difference between any dense set of rays and the continuum of Hilbert space rays. (See also the subsequent papers by Kent and Clifton and Kent .) However, for the reasons given below, the physical applicability of these constructions remain questionable.
Let us re-state the physical interpretation of the coloring schemes discussed above. Any linear subspace $`Sp𝐫`$ of a vector $`𝐫`$ can be identified with the associated projection operator $`E_𝐫`$ and with the quantum mechanical proposition “the physical system is in a pure state $`E_𝐫`$. The coloring of the associated point on the unit sphere (if it exists) is equivalent with a valuation or two-valued probability measure
$$Pr:E_𝐫\{0,1\}$$
where $`0\mathrm{\#}2`$ and $`1\mathrm{\#}1`$. That is, the two colors #1, #2 can be identified with the classical truth values: “It is true that the physical system is in a pure state $`E_𝐫`$” and “It is false that the physical system is in a pure state $`E_𝐫`$,” respectively.
Since, as has been argued before, the rational unit sphere has chromatic number three, two colors suffice for a reduced coloring generated under the assumption that the colors of two rays in any orthogonal tripod are identical. This effectively generates consistent valuations associated with the dense subset of physical properties corresponding to the rational unit sphere .
Kent has shown that there also exist dense sets in higher dimensions which permit a reduced two-coloring. Unpublished results by P. Ovchinnikov, O.G.Okunev and D. Mushtari state that the rational unit sphere of the d-dimensional real Hilbert space is d-colorable if and only if it admits a reduced two-coloring if and only if $`d<6`$.
### 2.2 Sufficiency
The Kochen-Specker theorem deals with the nonembedability of certain partial algebras—in particular the ones arising in the context of quantum mechanics—into total Boolean algebras. $`01`$ colorings serve as an important method of realizing such embeddings: e.g., if there are a sufficient number of them to separate any two elements of the partial algebra, then embedability follows . Kochen and Specker’s original paper contains a much stronger result—the nonexistence of $`01`$ colorings—than would be needed for nonembedability. However, the mere existence of some homomorphisms is a necessary but no sufficient criterion for embedability. The Godsil and Zaks construction merely provides three homomorphisms which are not sufficient to guarantee embedability. This has already be pointed out by Clifton and Kent .
### 2.3 Continuity
The regress to unsharp measurements is rather questionable and would allow total arbitrariness in the choice of approximation. That is, due to the density of the coloring, depending on which approximation is chosen, one predicts different results. The probabilities resulting from these truth assignments are noncontinuous and arbitrary. This is even more problematic if one realizes that the each color of the $`01`$ coloring of the rational unit sphere is dense.
### 2.4 Closedness
The rational unit sphere is not closed under certain geometrical operations such as taking an orthogonal ray of the subspace spanned by two non-collinear rays (the cross product of the associated vectors). This can be easily demonstrated by considering the two vectors
$$(\frac{3}{5},\frac{4}{5},0),(0,\frac{4}{5},\frac{3}{5})S^2𝐐^3.$$
The cross product thereof is
$$(\frac{12}{25},\frac{9}{25},\frac{12}{25})S^2𝐐^3.$$
Indeed, if instead of $`S^2𝐐^3`$ one would start with three non-orthogonal, non-collinear rational rays and generate new ones by the cross product, one would end up with all rational rays .
In the Birkhoff-von Neumann approach to quantum logics , this non-closedness under elementary operations such as the nor-operation might be considered a serious deficiency which rules out the above model as an alternative candidate for Hilbert space quantum mechanics. (However, his argument does not apply to the Kochen-Specker partial algebra approach to quantum logic, since there operations among propositions are only allowed if the propositions are comeasurable.)
Informally speaking, the relative (with respect to other sets such as the rational rays) “thinness” guarantees colorability. In such a case, the formation of finite cycles such as the ones introduced by Kochen and Specker are impossible.
To put it differently: given any nonzero measurement uncertainty $`\epsilon `$ and any non-colorable Kochen-Specker graph $`\mathrm{\Gamma }(0)`$ , there exists another graph (in fact, a denumerable infinity thereof) $`\mathrm{\Gamma }(\delta )`$ which lies inside the range of measurement uncertainty $`\delta \epsilon `$ \[and thus cannot be discriminated from the non-colorable $`\mathrm{\Gamma }(0)`$\] which can be colored. Such a graph, however, might not be connected in the sense that the associated subspaces can be cyclically rotated into itself by local transformation along single axes. The set $`\mathrm{\Gamma }(\delta )`$ might thus correspond to a collection of tripods such that none of the axes coincides with any other, although all of those non-identical single axes are located within $`\delta `$ apart from each other. Indeed, this appears to be precisely how the Clifton-Kent construction works \[11, p. 2104\].
The reason why a Kochen-Specker type contradiction does not occur in such a scenario is the impossibility to “close” the argument; to complete the cycle: the necessary propositions are simply not available in the rational sphere model. For the same reason, an equilateral triangle does not exist in $`𝐐^2`$ . Yet, while these findings seem to contradict the conclusions of Clifton and Kent , we would like to emphasize that this does not relate to their formal arguments but rather is a matter of interpretation and a question of how much should be sacrificed for value definiteness.
Thus, although the colorings of rational spheres offer a rather unexpected possibility to define consistent classical models, a closer examination shows that any such colorings should be excluded for physical reasons.
## Acknowledgments
The authors would like to acknowledge stimulating discussions with Ernst Specker. |
no-problem/9911/astro-ph9911110.html | ar5iv | text | # Close Binaries with Two Compact Objects
## 1. Introduction
The late inspiral phase due to gravitational radiation of close binaries with two compact objects, neutron stars (NS) or black holes (BH), is expected to be a major source of gravitational waves for ground-based interferometers, such as LIGO and VIRGO). The binary radio pulsar PSR B1913+16 (Hulse & Taylor 1975) is the prototypical NS–NS system and has provided a remarkable empirical confirmation of general relativity with the measurement of orbital decay due to gravitational radiation (Taylor & Weisberg 1982). In addition to NS–NS binaries, the existence of close BH–NS and BH–BH systems has also been predicted theoretically, and they are considered as possible sources of gravitational radiation.
The expected detection rate of inspiral events depends crucially on the rate of such events in the Galaxy and by extrapolation out to the maximum distance an inspiral event could be detected. In what follows I review the current estimates of Galactic coalescence rates, obtained theoretically and empirically, and I discuss the associated uncertainties. For the case of NS–NS binaries, I also discuss a new way of obtaining an empirical upper bound to their coalescence rate. The scale for all these rates is set by requiring that the detection rate for the “enhanced” LIGO is $`23`$ events per year. Given the amplitude of the signals (e.g., Thorne 1996), the requirement for the NS–NS Galactic rate is $`10^5`$ yr<sup>-1</sup> (detected out to $`200`$ kpc) and for the BH–BH Galactic rate $`2\times 10^7`$ yr<sup>-1</sup> (detected out to $`700`$ kpc).
## 2. Theoretical Estimates
The formation rate of coalescing binary compact objects (i.e., systems in close enough orbits that coalesce within a Hubble time) can be calculated, given a sequence of evolutionary stages leading to binary compact object formation. Over the years a relatively standard picture (van den Heuvel 1976) has been formed describing the birth of such systems based on the consideration of NS–NS binaries (for variations of this picture see Brown 1995; Terman & Taam 1995): The initial binary progenitor consists of two binary members massive enough to eventually collapse into a NS or a BH. The evolutionary path involves multiple phases of stable or unstable mass transfer, common-envelope phases, and accretion onto compact objects, as well as two core collapse events. Theoretical modeling of this formation process has been undertaken by various authors using population synthesis techniques, Such studies provide us with ab initio predictions of coalescence rates. The evolution of an ensemble of primordial binaries with assumed initial properties is followed until a coalescing binary is formed. The changes in the binary properties at the end of each evolutionary stage are calculated based on our current understanding of the processes involved: wind mass loss from massive hydrogen- and helium-rich stars, mass and angular-momentum losses during mass transfer phases, dynamically unstable mass transfer and common-envelope evolution, effects of highly super-Eddington accretion onto NS, and supernova explosions with kicks imparted to newborn NS or even BH. Given that several of these phases are not well understood, the results of population synthesis are expected to depend on the assumptions made in the treatment of the various processes. Therefore, exhaustive parameter studies are required by the nature of the problem.
Recent studies of the formation rate of coalescing compact objects (Lipunov, Postnov, & Prokhorov 1997; Fryer, Burrows, & Benz 1998; Portegies-Zwart & Yungel’son 1998; Brown & Bethe 1998; Fryer, Woosley, & Hartmann 1999) have explored the input parameter space and the robustness of the results at different levels of (in)completeness. Kicks imparted to NS and BH at birth are the best studied of all model assumptions. In the case of NS–NS coalescence the range of predicted Galactic rates is $`<10^75\times 10^4`$ yr<sup>-1</sup>. This large range indicates the importance of supernovae (two in this case) in binaries. Variations in the assumed mass-ratio distribution for the primordial binaries can further change the predicted rate by about a factor of $`10`$, while assumptions of the common-envelope phase add another factor of about $`10100`$. Variation in other parameters typically affects the results by factors of two or less. Predicted rates for BH–NS and BH–BH binaries lie in the ranges $`<10^710^4`$ yr<sup>-1</sup> and $`<10^710^5`$ yr<sup>-1</sup>, respectively when the kick magnitude to both NS and BH is varied. Other uncertain factors such as the critical progenitor mass for NS and BH formation lead to variations of the rates by factors of $`1050`$.
Although results from population syntheses regarding binary properties, such as orbital sizes, eccentricities, center-of-mass velocities, can be quite robust, it is evident that predicted rates, i.e., the absolute normalization of the models, cover a wide range of values (typically 3-4 orders of magnitude). Given these results it seems fair to say that population synthesis calculations have a rather limited predictive power and provide fairly loose constraints on coalescence rates.
## 3. Empirical Estimates
The observed sample of coalescing NS–NS binaries (PSR B1913+16 and PSR B1534+12) provides us with another way of estimating their coalescence rate. An empirical estimate can be obtained using their observed pulsar and binary properties along with models of selection effects in radio pulsar surveys. For each observed object, a scale factor can be calculated based on the fraction of the Galactic volume within which pulsars with properties identical to those of the observed pulsar could be detected by any of the radio pulsar surveys, given their detection thresholds. This scale factor is a measure of how many more pulsars like the ones detected in the coalescing NS–NS systems exist in our galaxy. The coalescence rate can then be calculated based on the scale factors and estimates of detection lifetimes summed up for all the observed systems. This basic method was first used by Phinney (1991) and Narayan, Piran, & Shemi (1991) who estimated the Galactic rate to be $`10^6`$ yr<sup>-1</sup>.
Since then, estimates of the NS–NS coalescence rate have known a significant downward revision primarily because of (i) the increase of the Galactic volume covered by radio pulsar surveys with no additional coalescing NS–NS being discovered (Curran & Lorimer 1995), (ii) the increase of the distance estimate for PSR B1534+12 based on measurements of post-Newtonian parameters (Stairs et al. 1998) (iii) revisions of the lifetime estimates (van den Heuvel & Lorimer 1996; Arzoumanian, Cordes, & Wasserman 1999). In addition, a significant upward correction factor ($`710`$) has been used recently to account for the faint end of the radio pulsar luminosity function. The most recently published study (Arzoumanian et al. 1999) gives a lower limit of $`2\times 10^7`$ yr<sup>-1</sup> and a “best” estimate of $`610\times 10^7`$ yr<sup>-1</sup>, which agrees with other recent estimates of $`23\times 10^6`$ yr<sup>-1</sup> (Stairs et al. 1998; Evans et al. 1999). Additional uncertainties arise from estimates of pulsar ages and distances, the pulsar beaming fraction, the spatial distribution of DNS in the Galaxy.
Despite all these uncertainties the empirical estimates of the NS–NS coalescence rate appear to span a range of $`12`$ orders of magnitude, which is narrow compared to the range covered by the theoretical estimates.
### 3.1. Small Number Sample and Pulsar Luminosity Function
One important limitation of empirical estimates of the coalescence rates is that they are derived based on only two observed NS–NS systems, under the assumption that the observed sample is representative of the true population, particularly in terms of their radio luminosity. Assuming that DNS pulsars follow the radio luminosity function of young pulsars and that therefore their true population is dominated in number by low-luminosity pulsars, it can be shown that the current empirical estimates most probably underestimate the true coalescence rate. If a small-number sample is drawn from a parent population dominated by low-luminosity (hence hard to detect) objects, it is statistically more probable that the sample will actually be dominated by objects from the high-luminosity end of the population. Consequently, the empirical estimates based on such a sample will tend to overestimate the detection volume for each observed system, and therefore underestimate the scale factors and the resulting coalescence rate.
This effect can be clearly demonstrated with a Monte Carlo experiment (Kalogera et al. 1999) using simple models for the pulsar luminosity function and the survey selection effects. As a first step, the average observed number of pulsars is calculated given a known “true” total number of pulsars in the Galaxy (thick-solid line in Figure 1a). As a second step, a large number of sets consisting of “observed” (simulated) pulsars are realized using Monte Carlo methods. These pulsars are drawn from a Poisson distribution of a given mean number ($`<N_{\mathrm{obs}}>`$) and have luminosities assigned according to the assumed luminosity function. Based on each of these sets, one can estimate the total number of pulsars in the Galaxy using empirical scale factors, as is done for the real observed sample. The many (simulated) ‘observed’ samples can then be used to obtain the distribution of the estimated total Galactic numbers ($`N_{\mathrm{est}}`$) of pulsars. The median and 25% and 75% percentiles of this distribution are plotted as a function of the assumed number of systems in the (fake) ‘observed’ samples in Figure 1a (thin-solid and dashed lines, respectively).
It is evident that, in the case of small-number observed samples (less than $`10`$ objects), the estimated total number, and hence the estimated coalescence rate, can be underestimated by a significant factor. For a two-object sample, for example, the true rate maybe higher by more than a factor of ten. This correction factor associated with the faint-end of the luminosity function can be applied to the estimated NS–NS coalescence rate in place of the factor of $`710`$ used so far from a direct extrapolation of the luminosity function.
## 4. Limits on the NS–NS Coalescence Rate
Observations of NS-NS systems and isolated pulsars related to NS–NS formation allow us to obtain upper limits on their coalescence rate. Depending on how their value compares to the “enhanced” LIGO requirement such limits can provide us with valuable information about the prospects of gravitational-wave detection.
Bailes (1996) used the absence of any young pulsars detected in NS–NS systems and obtained a rough upper limit to the rate of $`10^5`$ yr<sup>-1</sup>, while recently Arzoumanian et al. (1999) reexamined this in more detail and claimed a more robust upper limit of $`10^4`$ yr<sup>-1</sup>.
An upper bound to the NS–NS coalescence rate can also be obtained by combining our theoretical understanding of orbital dynamics (for supernovae with NS kicks in binaries) with empirical estimates of the birth rates of other types of pulsars related to NS–NS formation (Kalogera & Lorimer 2000, hereafter KL00). Progenitors of NS–NS systems experience two supernova explosions. The second supernova explosion (forming the NS that is not observed as a pulsar) provides a unique tool for the study of NS–NS formation, since the post-supernova evolution of the system is simple, driven only by gravitational-wave radiation. There are three possible outcomes after the second supernova: (i) a coalescing NS–NS is formed (CB), (ii) a wide NS–NS (with a coalescence time longer than the Hubble time) is formed (WB), or (iii) the binary is disrupted (D) and a single pulsar similar to the ones seen in NS–NS systems is ejected. Based on supernova orbital dynamics we can accurately calculate the probability branching ratios for these three outcomes, $`P_{\mathrm{CB}}`$, $`P_{\mathrm{WB}}`$, and $`P_\mathrm{D}`$. For a given kick magnitude, we can calculate the maximum ratio $`(P_{\mathrm{CB}}/P_\mathrm{D})^{\mathrm{max}}`$ for the complete range of pre-supernova parameters defined by the necessary constraint $`P_{\mathrm{CB}}0`$ (Figure 1b). Given that the two types of systems have a common parent progenitor population, the ratio of probabilities is equal to the ratio of the birth rates $`(BR_{\mathrm{CB}}/BR_\mathrm{D})`$.
We can then use (i) the absolute maximum of the probability ratio ($`0.26`$ from Figure 1b) and (ii) an empirical estimate of the birth rate of single pulsars similar to those in NS–NS based on the current observed sample to obtain an upper limit to the coalescence rate. The selection of this sample involves some subtleties (KL00), and the analysis results in $`BR_{\mathrm{CB}}<1.5\times 10^5`$ yr<sup>-1</sup> (KL00). Note that this number could be increased because of the small-number sample and luminosity bias affecting this time the empirical estimate of $`BR_\mathrm{D}`$ by a factor of $`26`$ (Kalogera et al. 1999).
This is an example of how we can use observed systems other than NS–NS to improve our understanding of their coalescence rate. A similar calculation can be done using the wide NS–NS systems instead of the single pulsars (see KL00).
## 5. Conclusions
The current theoretical estimates of NS–NS coalescence rates appear to have a rather limited predictive power. They cover a range in excess of 3 orders of magnitude and most importantly this range includes the value of $`10^5`$ yr<sup>-1</sup> required for an “enhanced” LIGO detection rate of 2–3 events per year. This means that at the two edges of the range the conclusion swings from no detection to many per month, and therefore the detection prospects of NS–NS coalescence cannot be assessed firmly. On the other hand empirical estimates derived based on the observed sample appear to be more robust (estimates are all within a factor smaller than 100). Given these we would expect a detection of one event every few years up to ten events per year.
Estimates of the coalescence rate of BH–NS and BH–BH systems rely solely on our theoretical understanding of their formation. As in the case of NS–NS binaries, the model uncertainties are significant and the ranges extend to more than 2 orders of magnitude. However, the requirement on the Galactic rate is less stringent for 10 M BH–BH binaries, only $`2\times 10^7`$ yr<sup>-1</sup>. Therefore, even with the pessimistic estimates for BH–BH coalescence rates ($`10^7`$ yr<sup>-1</sup>), we would expect at least a few detections per year, which is quite encouraging. We note that a very recent examination of dynamical BH–BH formation in globular clusters (Portegies-Zwart & McMillan 1999) leads to detection rates as high as a few per day.
#### Acknowledgments.
I would like to thank the organizers of the meeting for their invitation and support of my participation. I also acknowledge support by the Smithsonian Institute in the form of a CfA Post-doctoral Fellowship.
## References
Arzoumanian, Z., Cordes, J.M., & Wasserman, I. 1999, ApJ, 520, 696
Bailes, M. 1996, in IAU Symp. 165, Compact Stars in Binaries, ed. J. van Paradijs, E. P. J. van den Heuvel & E. Kuulkers (Kluwer Academic Publishers, Dordrecht), 213
Brown, G.E. 1995, ApJ, 440, 270
Brown, G.E. & Bethe, H. 1998, ApJ, 506, 780
Curran, S.J. & Lorimer, D.R. 1995, MNRAS, 276, 347
Evans, T., et al. 1999, to appear in the proceedings of the XXXIVth Rencontres de Moriond on Gravitational Waves and Experimental Gravity, Les Arcs, France.
Fryer, C.L., Burrows, A., Benz, W. 1998, ApJ, 496, 333
Fryer, C.L., Woosley, S.E., Hartmann, D. 1999, ApJ, in press \[astro-ph/9904122\].
Hulse, R.A., & Taylor, J.H. 1975, ApJ, 195, L51
Kalogera, V. & Lorimer, D.R. 2000, ApJ, 530, in press \[astro-ph/9907426\].
Kalogera, V., Narayan, R., Spergel, D., & Taylor, J.H. 1999, ApJ, in preparation
Lipunov, V.M., Postnov, K.A., Prokhorov, M.E. 1997, MNRAS, 288, 245
Narayan, R., Piran, T., & Shemi, A. 1991, ApJ, 379, L17
Phinney, E.S. 1991, ApJ, 380, L17
Portegies-Zwart, S.F. & McMillan, S.L.W., ApJ, submitted \[astro-ph/9910061\].
Portegies-Zwart, S.F. & Yungel’son, L.R. 1998, A&A, 332, 173
Stairs, I.H., et al. 1998, ApJ, 505, 352
Terman, J.L., & Taam, R.E. 1995, MNRAS, 276, 1320
Thorne, K.S. 1996, in IAU Symp. 165, Compact Stars in Binaries, ed. J. van Paradijs, E. P. J. van den Heuvel & E. Kuulkers (Kluwer Academic Publishers, Dordrecht), 153
van den Heuvel, E.P.J. 1976, in IAU Symp. 73, Structure and Evolution of Close Binary Systems, ed. P. Eggleton, S. Mitton & J. Whelan (Kluwer Academic Publishers, Dordrecht).
van den Heuvel, E.P.J. & Lorimer, D.R. 1996, MNRAS, 283, 37 |
no-problem/9911/cond-mat9911090.html | ar5iv | text | # Untitled Document
REEXAMINATION OF SEVEN-DIMENSIONAL SITE PERCOLATION THRESHOLDS
Dietrich Stauffer<sup>1</sup> and Robert M. Ziff<sup>2</sup>
<sup>1</sup> Institute for Theoretical Physics, Cologne University, D-50923 Köln, Euroland
<sup>2</sup> Department of Chemical Engineering, University of Michigan, Ann Arbor, MI 48109-2136, USA
e-mail: stauffer@thp.uni-koeln.de, rziff@umich.edu
Monte Carlo simulations alone could not clarify the corrections to scaling for the size-dependent $`p_c(L)`$ above the upper critical dimension. Including the previous series estimate for the bulk threshold $`p_c(\mathrm{})`$ gives preference for the complicated corrections predicted by renormalization group and against the simple $`1/L`$ extrapolation. Additional Monte-Carlo simulations using the Leath method corroborate the series result for $`p_c`$.
Keywords: Monte Carlo, 1/d expansion, finite-size scaling
How to extrapolate from finite samples of linear dimension $`L`$ to the thermodynamic limit $`L\mathrm{}`$ is an important problem in physics simulations. Above the upper critical dimension (4 in usual Ising models, 6 in random percolation), the critical exponents are known but nevertheless controversies remain<sup>1</sup> for the five-dimensional Ising model. Earlier, for site percolation in seven dimensions, the numerical variation of the apparent threshold $`p_c(L)=p_c`$ on a hypercubic lattice with $`L^7`$ sites could be fitted<sup>2</sup> on $`p_c(\mathrm{})p_c(L)1/L`$, but also<sup>3</sup> on the theoretical prediction $`p_c(\mathrm{})p_c(L)1/L^2\mathrm{const}/L^{7/3}`$. Also for three-dimensional self-avoiding walks one may assume one empirical correction term, or many terms as predicted by renormalization methods with more free parameters<sup>4</sup>. Therefore the present note reexamines seven-dimensional site percolation in the search of a clarifying example, using a much better Cray-T3E computer and a slightly improved program<sup>5</sup> compared with ref.2.
We checked if the top of the hypercube is connected with the bottom hyperplane, using helical boundary conditions in five and free boundaries in the two remaining directions<sup>5</sup> and a Hoshen-Kopelman algorithm without recycling<sup>6</sup>. Random number generation by integer multiplication with 16807 gave problems for $`L=4,\mathrm{\hspace{0.17em}8},\mathrm{\hspace{0.17em}16}`$ even for 64-bit arithmetic, but otherwise agreed with results from the Kirkpatrick-Stoll R250 generator; only the latter method was used for the data shown here. About 640 samples were averaged over for each $`L`$, giving an accuracy of about $`10^4`$ for $`p_c`$ in large lattices. ($`[<p_c^2><p_c>^2]^{1/2}`$ was not investigated in detail since it seems to vary as $`1/L^2`$.) Fig.1 shows our results for $`L`$ up to 20, together with a linear fit
$$p_c(L)=0.09090.12/L$$
$`(1)`$
and the same fit as published before<sup>3</sup>:
$$p_c(L)=0.08873.3/L^2+5/L^{7/3}.$$
$`(2)`$
We see that for large $`L`$ both fits agree nicely with our new Monte Carlo results: On the basis of these simulations alone no preference for one fit over the other is visible. However, when we extract from series expansions<sup>7</sup> (in terms of reciprocal dimensionality) a bulk $`p_c`$ of about 0.089, then the more complicated fit (2) is clearly preferred over the linear fit (1). This is a nice example how the combination of series with Monte Carlo techniques can give some answers which are not clear from each technique separately<sup>8</sup>. However, a more direct and new series expansion determination of $`p_c`$ would be desirable.
If we would have simulated $`L=40`$ instead of $`L=20`$ we might have chosen between the two fits on the basis of simulations alone. Such a size would have required about two orders of magnitude more in computer time and memory than the about 4000 processor hours (512 Mbyte each) used here.
(About boundary conditions: If we check whether top and bottom are connected we cannot use vertical periodic boundary conditions. As long as not all directions are treated with periodic boundary conditions, and at least one direction uses free boundaries, the finite-size effects are expected to be dominated by the free boundaries. The bulk $`p_c`$ value is independent of the boundary conditions, and so may be the powers of $`L`$ in finite-size corrections, but finite-size amplitudes as well as the fraction of samples spanning at $`p_c`$ depend on such details.<sup>9</sup> The Hoshen-Kopelman algorithm stores only one line of a square lattice at any one moment, and then it is simplest to use free boundaries also in this direction. In higher dimensions it is practical to store the $`L^{d1}`$ sites of the one hyperplane kept in memory by a one-dimensional index $`i=1,2,\mathrm{},L^{d1}`$. In this way <sup>5</sup> we come to our mixture of helical and free boundary conditions.)
Thus, to show that Monte Carlo can compete with series expansions, we made a different determination of the bulk $`p_c`$ using the Leath cluster growth algorithm<sup>10</sup>. Here, when a cluster stops growing before touching the boundaries, its properties are completely free of finite-size effects, and such data cannot be directly compared with the above Hoshen-Kopelman data for $`p_c(L)`$. However, the bulk $`p_c`$ must be the same. We determined $`p_c`$ by the methods given recently for three-dimensional lattices<sup>11</sup>. We considered a virtual lattice of size $`32^7`$ and a maximum size cut-off of 16384 sites, and simulated $`10^6`$$`10^7`$ clusters each at various values of $`p`$. The clusters are grown independently for each value of $`p`$, since in this method the random numbers are not assigned to all lattice sites but only to those the cluster visits.
It turned out that some of the largest clusters clusters (of maximum size of 16384 sites) wrapped a bit around the periodic boundary of the system, which could conceivably lead to the cluster touching itself and cause a bias. However, because of the very large number of sites in the lattice, the probability of this occurring is very low. (We could not easily check for wraparound error in our program.) In any case, the data for smaller clusters (where no wraparound was even possible) and for $`s=16485`$ was completely consistent.
Near the critical point, $`P_s`$, defined as the probability that a cluster grows to a size greater than or equal to $`s`$, behaves as
$$P_ss^{2\tau }f((pp_c)s^\sigma )s^{2\tau }(A+B(pp_c)s^\sigma +\mathrm{})$$
$`(3)`$
where $`\tau =5/2`$ and $`\sigma =1/2`$ for $`d6`$. Thus in Fig. 2 we plot $`s^{1/2}P_s`$ vs. $`s^{1/2}`$ for $`p=0.0885,`$ 0.0888, 0.0889 and 0.0890, and find good agreement with the expected behavior with $`A1.44`$, $`B20`$, and
$$p_c=0.08893\pm 0.00002$$
$`(4)`$
which is in excellent agreement with the series extrapolation<sup>7</sup>. With this value, the simple fit of eq (1) can be excluded since it requires $`p_c=0.091`$.
Acknowledgements
We thank A. Aharony, D.L. Hunter and N. Jan for discussions leading to this work, which was supported by GIF and the German Supercomputer Center, Jülich. RZ acknowledges support from the U. S. National Science Foundation under grant DMR-9502700.
References:
1. H.W.J.Blöte and E. Luijten, Phys. Rev. Lett. 76, 3662 (1996) and Europhys. Lett. 38, 565 (1997); X.S. Chen and V. Dohm, Int. J. Mod. Phys. C 9, 1007 and 1073 (1998); E. Luijten, K. Binder and H. W. J. Blöte, Eur. Phys. J. B 9, 289 (1999); M. Cheon, I. Chang and D. Stauffer, Int. J. Mod. Phys. C 10, 131 (1999).
2. D. Stauffer, Physica A 210, 317 (1994).
3. A. Aharony and D. Stauffer, Physica A 215, 342 (1995).
4. S. Joseph, D. L. Hunter, D. MacDonald, L.L. Moseley, N. Jan and T. Guttmann, preprint
5. D. Stauffer and N. Jan, in: Annual Reviews of Computational Physics, vol. VIII (Zanjan proceedings), World Scientific, Singapore 2000.
6. D. Stauffer and A. Aharony, Introduction to Percolation Theory, Taylor and Francis, London 1994; A. Bunde and S. Havlin, Fractals and Disordered Systems, Springer, Berlin-Heidelberg 1996; M. Sahimi, Applications of Percolation Theory, Taylor and Francis, London, 1994. For recent percolation thresholds see S.C. van der Marck, Int. J. Mod. Phys. C 9, 529 (1998).
7. D.S. Gaunt, M.F. Sykes and H.J. Ruskin, J. Phys. A 9, 1899 (1976).
8. J. Adler, page 241 in: Annual Reviews of Computational Physics, vol. IV, World Scientific, Singapore 1996.
9. M. Ford, D.L. Hunter, and N. Jan, Int. J. Mod. Phys. C 10, 183 (1999).
10. P.L. Leath, Phys. Rev. B 14, 5046 (1976).
11. C.D. Lorenz and R.M. Ziff, Phys. Rev. E 57, 230 (1998).
Fig.1: Effective thresholds $`p_c(L)`$ versus $`1/L`$, together with eqs.(1,2). Part a gives the overall trend, part b expands the region of large lattices. The horizontal line is the series extrapolation for infinite $`L`$.
Fig.2: Plots of $`s^{1/2}P_s`$ vs $`s^{1/2}`$ from the Leath simulations, for $`p=0.0890,`$ 0.0889, 0.0888, and 0.0885 from top to bottom. The slopes of the three upper curves are 0.0014, -0.0006 and -0.0026 respectively. |
no-problem/9911/math9911114.html | ar5iv | text | # Acknowledgment
## Acknowledgment
The research of this paper was made possible in part by Award UP1-309 of Civilian Research and Development Foundation and by Award 1.4/206 of Ukrainian DFFD. |
no-problem/9911/hep-ph9911238.html | ar5iv | text | # REFERENCES
DSF-T-99/36
Neutrino masses and mixings in $`SO(10)`$
M. Abud, F. Buccella<sup>∗∘</sup>, D. Falcone, G. Ricciardi and F. Tramontano
Dipartimento di Scienze Fisiche, Università di Napoli,
Mostra d’Oltremare, Pad. 19, I-80125, Napoli, Italy,
and INFN, Sezione di Napoli, Italy;
Istituto di Fisica Teorica, Università di Napoli, Italy
## Abstract
Assuming a Zee-like matrix for the right-handed neutrino Majorana masses in the see-saw mechanism, one gets maximal mixing for vacuum solar oscillations, a very small value for $`U_{e3}`$ and an approximate degeneracy for the two lower neutrino masses. The scale of right-handed neutrino Majorana masses is in good agreement with the value expected in a $`SO(10)`$ model with Pati-Salam $`SU(4)\times SU(2)\times SU(2)`$ intermediate symmetry.
The evidence for neutrino oscillation, a phenomenon discussed many years ago , in solar and atmospheric neutrinos, with square mass differences smaller than $`1(eV)^2`$, has stimulated the search for theoretical framework . Unified theories with $`SO(10)`$ as a gauge group are a natural choice, since there one expects neutrinos to be much lighter than the other fermions as a consequence of the see-saw mechanism .
$`SO(10)`$ is a suitable framework to discuss fermion masses, because all the left(right)-handed fermions of each family belong to a single irreducible representation, namely the spinorial $`\mathrm{𝟏𝟔}(\overline{\mathrm{𝟏𝟔}})`$. However, the spectrum and the mixing (the $`CKM`$ matrix) of the fermions do not show a clear $`SO(10)`$ pattern. In fact, by assuming that the electroweak Higgs belongs to a single real 10 representation, one gets the same spectrum for the charge $`\frac{2}{3}`$ and $`\frac{1}{3}`$ quarks (and at the $`SO(10)`$ unification scale also for the Dirac neutrino and charged lepton masses) and a trivial $`CKM`$ matrix. This last prediction can be thought as a zero order approximation in an expansion in $`\lambda =\mathrm{sin}(\theta _c)`$, but in order to account for the large difference between $`m_t`$ and $`m_b`$, at least one needs to assume that the two directions of the 10 with vanishing colour and electric charge have VEV’s not belonging to the same real 10 representations. In absence of components of the electroweak VEV along representations higher than 10 one keeps the equalities at the $`SO(10)`$ scale of the quark with charge $`\frac{1}{3}`$ and of the charged lepton mass matrix, as well for the quark with charge $`\frac{2}{3}`$ and the Dirac neutrino mass matrices. In that case one would have the same mixings for the quark and the Dirac lepton mass matrix and, especially for the two heaviest families, we can neglect the mixing of Dirac leptons ($`V_{cb}0.04`$).
The effective Majorana mass matrix for the light left-handed neutrinos is given by:
$$M_\nu ^{(L)}=M_D^TM_\nu ^{(R)1}M_D.$$
(1)
$`M_\nu ^{(R)}`$ has to be symmetric and if we assume, for simplicity, CP symmetry, it is a real matrix
$`M_\nu ^{(R)}=\left(\begin{array}{ccc}M_1& \rho & \nu \\ \rho & M_2& \mu \\ \nu & \mu & M_3,\end{array}\right).`$ (5)
The Dirac mass matrix $`M_D`$ is assumed real and diagonal, according to our approximation of neglecting the mixings:
$`M_D=\left(\begin{array}{ccc}M_{\nu _e}^D& 0& 0\\ 0& M_{\nu _\mu }^D& 0\\ 0& 0& M_{\nu _\tau }^D\end{array}\right).`$ (9)
From the previous equations it is easy to get
$`M_\nu ^{(L)}={\displaystyle \frac{1}{D}}\left(\begin{array}{ccc}\left(M_2M_3\mu ^2\right)\left(M_{\nu _e}^D\right)^2& \left(\mu \nu M_3\rho \right)M_{\nu _e}^DM_{\nu _\mu }^D& \left(\rho \mu M_2\nu \right)M_{\nu _e}^DM_{\nu _\tau }^D\\ \left(\mu \nu M_3\rho \right)M_{\nu _e}^DM_{\nu _\mu }^D& \left(M_1M_3\nu ^2\right)\left(M_{\nu _\mu }^D\right)^2& \left(\rho \nu M_1\mu \right)M_{\nu _\mu }^DM_{\nu _\tau }^D\\ \left(\rho \mu M_2\nu \right)M_{\nu _e}^DM_{\nu _\tau }^D& \left(\rho \nu M_1\mu \right)M_{\nu _\mu }^DM_{\nu _\tau }^D& \left(M_1M_2\rho ^2\right)\left(M_{\nu _\tau }^D\right)^2\end{array}\right)`$ (13)
where
$$D\mathrm{det}M_\nu ^{(R)}=M_1M_2M_3+2\rho \nu \mu M_1\mu ^2M_2\nu ^2M_3\rho ^2.$$
(14)
Let us make the approximation of neglecting $`M_{\nu _e}^D`$ with respect to the Dirac masses of the other neutrinos, which may be reasonable if the electroweak doublet has components only along 10 representations, which would imply
$$M_D=\frac{m_\tau }{m_b}M_{q=\frac{2}{3}},$$
(15)
but may hold more generally, since all the charged fermions of the first family have masses smaller than those of the other two families. In that limit the matrix elements of the first row and of the first column of eq.(4) vanish and the remaining $`2\times 2`$ matrix has the property of having a vanishing eigenvalue if one takes only the contributions proportional to $`\nu ^2`$, $`\nu \rho `$ and $`\rho ^2`$. Since one of the solutions for the neutrino spectrum is the hierarchical one with the highest mass $`\frac{1}{16}eV`$, we conclude that it is an intriguing possibility to have only non-diagonal matrix elements in $`M_\nu ^{(R)}`$. A matrix with this property has been proposed many years ago by Zee for left-handed neutrinos and its phenomenological consequences have been discussed by Frampton and Glashow in the case of light Majorana neutrinos. We also assume a non-vanishing $`\mu `$ in order to have $`D0`$. We define
$$A_e=\mu M_{\nu _e}^D=M_0^2\mathrm{sin}\beta $$
(16)
$$A_\mu =\nu M_{\nu _\mu }^D=M_0^2\mathrm{cos}\beta \mathrm{cos}\alpha $$
(17)
$$A_\tau =\rho M_{\nu _\tau }^D=M_0^2\mathrm{cos}\beta \mathrm{sin}\alpha $$
(18)
and get
$$M_\nu ^{(L)}=\frac{1}{D}\left(\begin{array}{ccc}A_e^2& A_eA_\mu & A_eA_\tau \\ A_eA_\mu & A_\mu ^2& A_\mu A_\tau \\ A_eA_\tau & A_\mu A_\tau & A_\tau ^2\end{array}\right)$$
$$=\frac{M_0^4}{D}\left(\begin{array}{ccc}\mathrm{sin}^2\beta & \mathrm{sin}\beta \mathrm{cos}\beta \mathrm{cos}\alpha & \mathrm{sin}\beta \mathrm{cos}\beta \mathrm{sin}\alpha \\ \mathrm{sin}\beta \mathrm{cos}\beta \mathrm{cos}\alpha & \mathrm{cos}^2\beta \mathrm{cos}^2\alpha & \mathrm{cos}^2\beta \mathrm{cos}\alpha \mathrm{sin}\alpha \\ \mathrm{sin}\beta \mathrm{cos}\beta \mathrm{sin}\alpha & \mathrm{cos}^2\beta \mathrm{cos}\alpha \mathrm{sin}\alpha & \mathrm{cos}^2\beta \mathrm{sin}^2\alpha \end{array}\right),$$
(19)
whose eigenvalues obey the secular equation
$$m_i^3m_i^2\frac{A_e^2+A_\mu ^2+A_\tau ^2}{D}+\frac{4A_e^2A_\mu ^2A_\tau ^2}{D^3}=0(i=1,2,3).$$
(20)
The absence of the linear term in eq.(11) implies for the three solutions the relation
$$m_1m_2+m_1m_3+m_2m_3=0.$$
(21)
If $`m_3`$ is larger than the other two solutions, one has that $`m_1`$ and $`m_2`$ are almost opposite, more precisely
$$m_1+m_2=\frac{m_1m_2}{m_3}.$$
(22)
The ratios of the solutions of eq.(11) depend on the quantity
$$k^2=\frac{4A_e^2A_\mu ^2A_\tau ^2}{\left(A_e^2+A_\mu ^2+A_\tau ^2\right)^3}=\frac{1}{4}\mathrm{sin}^22\alpha \mathrm{sin}^22\beta \mathrm{cos}^2\beta $$
(23)
which should be small in order to have one solution larger than the other two. In this case one has approximately, to lowest order in $`k`$
$$m_3\frac{A_e^2+A_\mu ^2+A_\tau ^2}{D}=\frac{M_0^4}{D}$$
(24)
$$|m_1||m_2|\frac{2A_eA_\mu A_\tau }{D\sqrt{A_e^2+A_\mu ^2+A_\tau ^2}}=k\frac{M_0^4}{D}$$
(25)
and
$$m_1+m_2\frac{4A_e^2A_\mu ^2A_\tau ^2}{D\left(A_e^2+A_\mu ^2+A_\tau ^2\right)^2}=k^2\frac{M_0^4}{D}$$
(26)
which imply
$$m_{1,2}k\frac{M_0^4}{D}\left(1+\frac{k}{2}\right)$$
(27)
and
$$m_2^2m_1^2=2k^3\frac{M_0^8}{D^2}.$$
(28)
Under these assumptions, $`m_3^2`$ and $`m_2^2m_1^2`$ have to be identified with the values of $`\mathrm{\Delta }m_{atm}^2`$ and $`\mathrm{\Delta }m_{sun}^2`$, the quantities that are relevant for the oscillations of the atmospheric and solar neutrinos. The parameter $`k`$ is so recovered to be
$$k^3=\frac{1}{2}\frac{\mathrm{\Delta }m_{sun}^2}{\mathrm{\Delta }m_{atm}^2}.$$
(29)
The absence of the distortion (which is predicted by the MSW solution) of the spectrum, according to the new data of SuperKamiokande with a lower threshold (5 MeV) for the electron detection, has brought Bilenky, Giunti and Grimus to consider the vacuum solution as the most probable with the following parameters
$$\mathrm{sin}^22\theta _{atm}=1,\mathrm{\Delta }m_{atm}^2=3.5\times 10^3\mathrm{eV}.$$
(30)
$$\mathrm{sin}^22\theta _{sun}=0.8,\mathrm{\Delta }m_{sun}^2=4.3\times 10^{10}\mathrm{eV}.$$
(31)
From eq.(20),(21) and (22) we get
$$k=4\times 10^3.$$
(32)
In order to have a small value for $`k`$, at least one of the $`A`$’s defined in eqs.(7),(8) and (9) should be smaller than the others. The most natural choice is $`A_e`$, since it is proportional to the Dirac mass of $`\nu _e`$, that corresponds to a small value for the $`\beta `$ angle. In this case $`k\beta \mathrm{sin}2\alpha `$, and we are naturally brought to a small value for $`U_{e3}`$ and to a large value for the angle intervening in the solar neutrino oscillations.
In fact, the exact eigenvalues of eq.(11) are
$$m_3=\frac{M_0^4}{D}\left(\frac{1}{3}+\frac{2}{3}\mathrm{cos}\psi \right)$$
(33)
$$m_{2,1}=\frac{M_0^4}{D}\left[\frac{1}{3}(1\mathrm{cos}\psi )\pm \frac{\sqrt{3}}{3}\mathrm{sin}\psi \right]$$
(34)
with $`\psi `$ given by
$$\mathrm{cos}3\psi =1\frac{27}{2}k^2.$$
(35)
The corresponding eigenvectors, in terms of $`\lambda _i=m_i\frac{D}{M_0^4}`$ are given by
$$a_i\left(\begin{array}{c}2\lambda _i(\lambda _i\mathrm{cos}^2\beta )\\ \mathrm{sin}2\beta \mathrm{cos}\alpha (\lambda _i+2\mathrm{cos}^2\beta \mathrm{sin}^2\alpha )\\ \mathrm{sin}2\beta \mathrm{sin}\alpha (\lambda _i2\mathrm{cos}^2\beta \mathrm{cos}^2\alpha )\end{array}\right)$$
(36)
where the $`a_i`$ are the appropriate normalization factors, to have unit norm vectors. Developing in the small parameter $`\beta `$, to lowest order, we get:
$$|U_{e3}|\beta \mathrm{cos}2\alpha k\mathrm{cot}2\alpha $$
(37)
$$\mathrm{sin}^22\theta _{sun}=\frac{4U_{e1}^2U_{e2}^2}{(|U_{e1}|^2+|U_{e2}|^2)^2}\left(1\frac{k^2}{4}\right)+O(k^3).$$
(38)
For small $`k`$, the mixing angle for atmospheric neutrinos may be identified with $`\alpha `$ and the high value for $`\mathrm{sin}^22\theta _{atm}`$ implies a very small value for $`|U_{e3}|`$, well below the limit $`|U_{e3}|^2<0.05`$ , from the CHOOZ experiment . Eq. (38) implies that $`\mathrm{sin}^22\theta _{sun}`$ is practically equal to the maximal value.
For $`\alpha =\pi /4`$ one has, exactly to all orders, $`U_{e3}=0`$ and maximal mixing for atmospheric neutrinos. In that case the mixing matrix, in the limit $`k=0`$, is the one with bimaximal mixing proposed in ref.
$`\left(\begin{array}{ccc}\frac{1}{\sqrt{2}}& \frac{1}{\sqrt{2}}& 0\\ \frac{1}{2}& \frac{1}{2}& \frac{1}{\sqrt{2}}\\ \frac{1}{2}& \frac{1}{2}& \frac{1}{\sqrt{2}}\end{array}\right).`$ (42)
A scenario similar to the one proposed here can be found in ref. with an appropriate choice of the Dirac neutrino mass and an antidiagonal form for the Majorana mass matrix for right-handed neutrinos (in our notation with only $`\nu `$ and $`M_2`$ different from zero). Also in that paper the ratio of the scales for solar and atmospheric neutrino oscillations are given in terms of an expansion parameter in such a way that, when it vanishes gives also rise to the bimaximal mixing (30). Also there the two lower mass neutrino eigenstates are almost degenerate. The resulting matrices for the two cases are different as well the secular equation for the mass eigenstates, which brings to a value for the mass of the almost degenerate lighter neutrinos $`\sqrt[3]{2}`$ larger than the value given in ref., a difference which unfortunately seems very far from experimental detection. Another difference concerns the matrix element $`M_{ee}`$, which appears in the double $`\beta `$-decay, which is predicted to vanish in , but as we shall see later, the non-vanishing value predicted here is completely negligible for the vacuum solution.
It is the right moment to remind that we have neglected the rotation between the charged lepton and the Dirac neutrino mass matrices. This approximation is reasonable if we assume that electroweak Higgs has components only along 10 representations of $`SO(10)`$ since in that case this solution is given by the CKM matrix, which has small non-diagonal matrix elements. We are aware that the inequality $`m_\mu m_b>m_\tau m_s`$ requires some component of the electroweak Higgs along some higher representation .
In $`SO(10)`$ the scale for the right-handed neutrino masses is related to the scale of $`BL`$ symmetry breaking and therefore it is interesting to see which values of $`\mu `$, $`\nu `$, $`\rho `$ are needed to get the values of $`\mathrm{\Delta }m_{atm}^2`$ and $`\mathrm{\Delta }m_{sun}^2`$. To the purpose of getting the right order of magnitude, one can assume eq.(6) and $`\alpha =\pi /4`$, which implies the value of $`\mu `$
$$\mu =\left(\frac{m_\tau }{m_b}\right)^2\frac{m_cm_t}{m_3}7.5\times 10^{11}GeV.$$
(43)
We have taken $`m_3=\sqrt{\mathrm{\Delta }m_{atm}^2}=6\times 10^2`$ eV and, as reference values, the masses at the scale $`M_Z`$: $`m_u=2.3\times 10^3`$ GeV, $`m_c=0.67`$ GeV, $`m_b=3.0`$ GeV, $`m_t=181`$ GeV and $`m_\tau =1.75`$ GeV . Anyway, since our quantities depend only on the ratios of the quark masses, a strong dependence on the scale is not expected.
We also have
$$\nu =\mu \frac{m_u}{m_c}\frac{A_\mu }{A_e}=4.6\times 10^{11}\mathrm{GeV},$$
(44)
$$\rho =\nu \frac{m_c}{m_t}=1.7\times 10^9\mathrm{GeV}.$$
(45)
The rather moderate values for the matrix elements of $`M_\nu ^{(R)}`$ are a consequence of the approximate degeneracy of the two lower neutrino mass eigenstate, which makes
$$m_{1,2}\sqrt{\frac{\mathrm{\Delta }m_{sun}^2}{2}}\left(\frac{2\mathrm{\Delta }m_{atm}^2}{\mathrm{\Delta }m_{sun}^2}\right)^{1/6}\sqrt{\mathrm{\Delta }m_{sun}^2}$$
(46)
while with the hierarchical relationship $`m_2>m_1`$ one should have
$$m_1m_2\sqrt{\mathrm{\Delta }m_{sun}^2},$$
(47)
and the opposite inequality with respect to eq.(34). In fact from the eqs.(1),(2),(5) and (6), and $`m_3\sqrt{\mathrm{\Delta }m_{atm}^2}`$, one would get
$$D=\frac{R^3m_u^2m_c^2m_t^2}{m_1m_2m_3}2\times 10^{35}\frac{\mathrm{\Delta }m_{sun}^2}{m_1m_2}GeV^3$$
(48)
($`R=(m_\tau /m_b)^2`$) and the inequality (34) brings to a lower value for D. Also the approximate degeneracy between $`m_2`$ and $`m_1`$ allows to get a not too broad neutrino mass spectrum.
For the first diagonal matrix elements of the Majorana mass one finds as order of magnitude
$$M_{ee}=\frac{km_u^2\mu }{2\rho \nu }1.2\times 10^6eV;$$
(49)
much smaller than the present experimental limit ($`0.2eV`$).
It is worth to recall that a value $`2.8\times 10^{11}GeV`$ has been found for the scale of spontaneous breaking of $`BL`$ in the $`SO(10)`$ model with $`SU(4)\times SU(2)\times SU(2)`$ intermediate symmetry , which is broken by a VEV of the 126 representation, endowed with the right quantum numbers to give Majorana masses to the right-handed neutrino. Above that scale, $`SU(4)_{PS}`$ for quarks and leptons implies that their mass ratios do not change. The narrow range for the evolution of $`m_b/m_\tau `$ is not enough to get $`m_b=m_\tau `$ at that scale as required by the hypothesis that the electroweak Higgs VEV’s are only along 10 representations, but at least a smaller contribution from other representation is needed with respect to the case where the range of RGE for the mass ratio extends to the unification scale, at which the lepto-quarks responsible for proton decay take their mass.
As a conclusion we think that the choice of the Zee matrix for the right-handed neutrino Majorana masses seems very appealing for the vacuum solution of the solar neutrino problem and well consistent with the scale found in the $`SO(10)`$ model with Pati-Salam intermediate symmetry.
One of us (F. B.) gratefully acknowledges stimulating discussions with Profs. Z. Berezhiani and A. Masiero. |
no-problem/9911/cond-mat9911184.html | ar5iv | text | # On the Helix-Coil transition in grafted chains
## Abstract
The helix-coil transition is modified by grafting to a surface. This modification is studied for short peptides capable of forming $`\alpha `$-helices. Three factors are involved: (i) the grafting can induced change of the boundary free energy of the helical domain (ii) the van der Waals attraction between the helices and (iii) the crowding induced stretching of the coils. As a result the helix-coil transition acquires “all or nothing” characteristics. In addition the transition temperature is elevated and the transition itself sharpens as the grafting density increases.
During the past two decades the physics of polymer brushes formed by terminally anchored chains were studied extensively. Most of the research effort dealt with brushes of flexible, synthetic polymers, devoid of internal degrees of freedom. In contrast, this letter concerns brushes formed by biopolymers capable of undergoing a cooperative helix-coil transition. It is motivated by two experimental observations. First, the promotion of the adhesion and spreading of cells by brushes of collagen model peptides. Second, membrane fusion induced by model fusogenic peptides grafted to vesicles. In both cases, the function of the short peptide chains was correlated with a helical state. Furthermore, the helical state was favored by the grafting. With this in mind we present a highly simplified theoretical model for the helix-coil transition in brushes of short, laterally immobile peptides. In particular, we discuss the transition temperature $`T_t`$ and the width of the transition $`\mathrm{\Delta }T`$ as a function of grafting density, $`\mathrm{\Sigma }^1`$. We focus on the simplest situation, of short neutral homopeptides forming a single stranded $`\alpha `$-helix. As we shall see, the grafting of the chains can lead to qualitative modifications of the helix-coil transition: (i) in marked distinction to the case of a free peptide, the helix-coil transition in an isolated grafted peptide can acquire “all or nothing” characteristics; (ii) $`T_t`$ is elevated with the grafting density while $`\mathrm{\Delta }T`$ decreases; (iii) eventually, for high grafting density the transition may take place as a first-order phase transition. These distinctive features arise because of three factors: The lower configurational entropy of the monomer at the grafting site favors helix formation even in isolated chains. In a brush, the helical state is also promoted by the crowding induced stretching and by the attractive van der Waals interaction between the helices.
In a $`\alpha `$-helix the $`i`$th monomer forms $`H`$-bonds with the $`(i3)`$th and the $`(i+3)`$th monomers. Overall, a helical domain consisting of $`n`$ monomers contains $`n2`$ $`H`$-bonds. It is convenient to consider the chain as a sequence of bonds. Choosing the coil state as a reference state, each helical bond is associated with an excess free energy $`\mathrm{\Delta }f`$ reflecting the formation of $`H`$-bonds, the accompanying change in solvation and the loss of configurational entropy. $`\mathrm{\Delta }f`$ is a function of the temperature, $`T`$. $`\mathrm{\Delta }f`$ vanishes at the transition temperature $`T_{}`$, while $`\mathrm{\Delta }f<0`$ when $`T<`$ $`T_{}`$ and $`\mathrm{\Delta }f>0`$ for $`T>T_{}`$. The terminal bonds, at the boundary of the helical domains, are associated with an additional free energy penalty $`\mathrm{\Delta }f_t`$. This arises because the terminal bonds lose their configurational entropy but do not contribute $`H`$-bonds. $`\mathrm{\Delta }f_t`$ plays the role of an interfacial free energy associated with the helix-coil boundary. It is customary to formulate the theory of the helix-coil transition in terms of the Bragg-Zimm parameters $`s=\mathrm{exp}(\mathrm{\Delta }f/kT)`$ and $`\sigma =\mathrm{exp}(2\mathrm{\Delta }f_t/kT)`$. $`\sigma `$ is independent of $`T`$ and is typically of order of $`10^310^4`$, depending on the identity of the amino acid residues forming the peptide. On the other hand, $`s`$ is a function of $`T`$ and in the vicinity of $`T_{}`$ it varies as $`s1(T_{}T)/T_{}`$. In an infinite chain, the plot of the fraction of helical bonds, $`\theta `$ vs. $`s`$ is sigmoid and the width of the transition is $`\mathrm{\Delta }TT_{}\sigma ^{1/2}`$. Since the chain is one dimensional object, a first order phase transition is impossible. As a result the chain consists of an alternating sequence of helical and coil regions. The minimal size of a domain comprising a helical and a coil regions, as obtained at $`T_{}`$, is $`\sigma ^{1/2}`$. When the polymerization degree, $`N`$ is much larger than $`\sigma ^{1/2}`$ the chain may be considered as infinite. It consists of a large number of domains and the width of the transition, $`\sigma ^{1/2}`$, is independent of $`N`$. On the other hand, when $`N\sigma ^{1/2}`$ the chain incorporates typically only one helical region and the width of the transition is of order $`1/N`$. The statistical physics of such short oligopeptides are well described by the “one sequence approximation” where the chain is assumed to contain a single helical region . The customary formulation of this approximation utilizes the appropriate partition function. For the purposes of our discussion it is convenient to utilize the corresponding free energy. The chemical potential of a chain supporting a helical domain consisting of $`n`$ bonds is $`\mu _0(n)=n\mathrm{\Delta }f+2\mathrm{\Delta }f_tkT\mathrm{ln}(Nn1)`$ where the last term is the entropy associated with the placement of the helical segment along the chain. Altogether there are $`N2`$ bonds and the first monomer of the helical segment of length $`n`$ can be placed in any of the $`(N2)n+1`$ sites. For $`n=0`$ we have $`\mu _0(0)=0`$. The fraction of chains with a given $`n`$, $`p_n`$, is determined by minimizing $`\mathrm{\Omega }=\mathrm{\Omega }_0+M_np_n[\mu _0(n)+kT\mathrm{ln}p_n]`$ subject to the constraint $`_np_n=1`$. This leads to $`\mu =\mu _0(n)+kT\mathrm{ln}p_n=const^{}`$ since all species coexist in equilibrium. For our discussion it is sufficient to consider the dominant term in $`\mathrm{\Omega }`$, as specified by $`\mu _0(n)/n=0`$. When $`\mathrm{\Delta }f0`$ the minimal $`\mu _0(n)`$ is $`\mu _0(0)=0`$ and the majority of chains contain no helical region. On the other hand, for $`\mathrm{\Delta }f<0`$ the probability distribution peaks at $`n_{}=N1kT/\mathrm{\Delta }f`$. Accordingly, $`n_{}`$ attains its maximum value, $`n_{}=N2`$, when $`\mathrm{\Delta }f=kT`$. The discussion as presented above applies to free chains. Two features are of special importance for future reference: (i) $`\mathrm{\Delta }f_t`$ is independent of the position of the helical region and, as a result (ii) all placements of the helical segments are equally probable. Because of the associated entropy $`n_{}`$ is lower than $`N2`$ for $`0>\mathrm{\Delta }f>kT`$.
The situation described above is modified significantly when the chain is grafted, terminally anchored, to a surface. We first discuss the case of a single chain grafted to a flat surface. The grafting gives rise to two effects. First, the overall number of configurations of the chain is reduced because of the presence of an impenetrable wall. As a result, the overall configurational entropy is reduced by $`\mathrm{\Delta }S\mathrm{ln}N`$. This is expected to lead to a small changes in $`s`$ and $`\sigma `$ which we will ignore since they do not give rise to qualitative effects. Second, the two boundaries of the helical region are no longer equivalent because of the presence of the surface. The surface effect is expected to decay with the length of the coil region separating the helical region from the wall. The maximal effect is attained when the helical region is initiated at the wall. The penalty of the “free” boundary remains $`\mathrm{\Delta }f_t.`$ However, the terminal penalty at the wall, $`\mathrm{\Delta }f_w,`$ can be different. The difference can arise from two sources. First a modification of the torsional potential for the terminal monomer and the consequent reduction of its configurational entropy. Second, change in the electrostatic interactions experienced by the terminal monomer at the grafting site. As a result, the Bragg-Zimm parameters are modified and for a helical sequence at the wall $`\sigma _g=\mathrm{exp}[(\mathrm{\Delta }f_w+\mathrm{\Delta }f_t)/kT].`$ In the following we consider the physically plausible case of $`\mathrm{\Delta }f_w<\mathrm{\Delta }f_t`$. For simplicity we assume that $`\mathrm{\Delta }f_w`$ is attained only when the helical sequence begins at the grafting site. In this situation a helical sequence of length $`n`$ may assume two states: (i) When the terminal monomer is not at the wall, the helical sequence can be freely placed at the available remaining sites. The “free” sequence is associated with $`\mu _{f0}=n\mathrm{\Delta }f+2\mathrm{\Delta }f_tkT\mathrm{ln}(Nn2)`$. (ii) A “bound” state when the terminal monomer is at the grafting site and $`\mu _{b0}=n\mathrm{\Delta }f+\mathrm{\Delta }f_t+\mathrm{\Delta }f_w`$. The two states are equally stable when $`\mu _{f0}=\mu _{b0}`$. This condition is satisfied for $`n_{}`$ given by $`n_{}=N2\mathrm{exp}[(\mathrm{\Delta }f_t\mathrm{\Delta }f_w)/kT]`$. For $`n>n_{}`$ the bound state is favored. Accordingly, the bound state is always dominant when $`n_{}=1`$ or $`\mathrm{ln}(N3)(\mathrm{\Delta }f_t\mathrm{\Delta }f_w)/kT`$. Furthermore, among the bound states, the fully helical state is always the most stable i.e., $`\mu _{b0}(N)<\mu _{b0}(n)`$ whenever $`\mathrm{\Delta }f<0`$. For this choice of $`N`$ the coil state remains the most probable state, $`n_{}=0`$ when $`\mathrm{\Delta }f>0`$ while for any $`\mathrm{\Delta }f<0`$ the most probable state is fully helical, $`n_{}=N2`$. Thus, the grafting of short chains modifies the nature of the helix-coil transition that acquires “all or nothing” characteristics. In the remainder of this letter we focus on brushes of such chains.
Our earlier considerations concerned a single grafted chain. With them in mind we model a grafted layer consisting of many chains as a mixture of chains in a coil state and chains in a fully helical state. In this case, additional contributions come into play. One is the van der Waals attraction between the rods. Another important contribution is the crowding induced extension of the coils in the brush regime. To quantify this picture it is necessary to introduce further assumptions concerning the orientation imposed on the helices by the grafting. For simplicity we will consider the case of grafting sites enforcing perpendicular orientation of the helices with respect to the surface. We further limit the discussion to the case of rigid junctions that is, the helices cannot bend at the grafting sites. Thus far, our considerations were directed at chains with $`N\sigma ^{1/2}10^2`$. At this point it is necessary to limit our discussion further to chains with $`NN_{}(a/a_h)^2`$ where $`a`$ is the effective monomer size in the coil state while $`a_h=1.5`$Å is the projection of a residue on the axis of the helix. While $`a`$ depends on the identity of the residue, $`a/a_h>1`$ and typically $`N_{}50`$. This choice is necessary because the length of a fully helical chain, $`L`$, is smaller than the radius of the corresponding random coil, $`R_0N^{1/2}a`$ when $`NN_{}`$ while $`L>R_0`$ if $`N>N_{}`$. Thus, in one case the helices are submerged in the brush formed by the coils while in the other the helices can be partially exposed (Figure 1). This choice of the $`N`$ range is dictated by the experimental systems motivating this work . With the system fully specified we are in a position to write down the free energy per chain, $`F`$. For simplicity we use a modification of the Alexander model , assuming that the coils are uniformly stretched and that concentration profile of the monomers of the coils is step-like. We consider a planar grafted layer such that the area per chain is $`\mathrm{\Sigma }`$ and the fraction of chains in a helical state is $`\theta `$. $`F`$ consists of four terms, $`F=F_0+\theta F_{vdW}+(1\theta )F_{cc}+\theta F_{hc}`$. $`F_0`$ reflects the contribution of the non interacting chains $`F_0\theta (N\mathrm{\Delta }f+\mathrm{\Delta }f_t+\mathrm{\Delta }f_w)+kT[\theta \mathrm{ln}\theta +(1\theta )\mathrm{ln}(1\theta )]`$ where the last term is the mixing entropy associated with the possible placements of chains in the two possible states. The van der Waals attraction between two oriented helices is modeled as the interaction between two parallel rods of diameter $`d`$. When the distance between two neighboring rods, $`D`$, is small in comparison to $`L`$, their length, $`F_{vdW}L/D^5`$ while in the opposite limit $`F_{vdW}L^2/D^6`$. In the first case, an element of the rod experiences, in effect, an interaction with an infinite neighboring rod. In the second limit, $`F_{vdW}`$ reflects the sum of all pairwise interactions between the elements of the two rods. The van der Waals energy of a rod within the grafted layer, where the minimal distance between two neighboring helices is $`(4\mathrm{\Sigma }/\pi \theta )^{1/2},`$ is
$$F_{vdW}=\frac{ALd^4}{18}\left(\frac{\pi \theta }{4\mathrm{\Sigma }}\right)^{5/2}\mathrm{arctan}\sqrt{\frac{\pi \theta L^2}{4\mathrm{\Sigma }}}$$
(1)
where $`A`$ is the Hamaker constant. In a “normal” brush, the overlap between the coils gives rise to chain extension along the normal to the surface. The stretching reduces the number of monomer-monomer contacts at the price of an extension penalty. In our situation the picture is somewhat different. First, it is necessary to allow for both coil-coil and coil-helix interactions. Second, only the coils are extendible. Finally, for our choice of $`N`$, the helices are fully immersed in a brush of coils. $`F_{cc}`$ allows for the elastic free energy of the coils as well as for monomer-monomer interactions involving the coils. This contribution has the form of the free energy of a brush formed by coils that is, $`F_{cc}/kTH^2/Na^2+v_{cc}N^2a^3(1\theta )/H\mathrm{\Sigma }`$ where $`H`$ is the thickness of the brush and $`v_{cc}`$ is the virial coefficient for monomer-monomer interactions involving coils. The first term reflects the Gaussian stretching penalty while the second allows for the repulsive monomer-monomer contacts. $`F_{cc}`$ must be supplemented by $`F_{hc}`$ that reflects the interactions between the coils and the helices, $`F_{hc}/kTv_{hc}N^2a^3(1\theta )/H\mathrm{\Sigma }`$. Here $`v_{hc}`$ is the virial coefficient for the binary interactions between helical monomers and coil monomers. For simplicity we consider the case of $`v_{hc}=v_{cc}=v`$. In this case the equilibrium thickness of the brush $`H_{eq}/aN(va^2/\mathrm{\Sigma })^{1/3}`$, as specified by $`F/H=0`$, is independent of $`\theta `$ so long as $`\theta `$ is small enough to ensure overlap between the coils. For such a choice of $`\theta `$ the equilibrium form of the free energy per chain, as obtained by substituting $`H_{eq}`$ into $`F`$, is
$`{\displaystyle \frac{F}{NkT}}`$ $``$ $`{\displaystyle \frac{\theta \mathrm{\Delta }f}{kT}}+{\displaystyle \frac{\theta (\mathrm{\Delta }f_t+\mathrm{\Delta }f_w)}{NkT}}\lambda \theta ^{7/2}\left({\displaystyle \frac{a^2}{\mathrm{\Sigma }}}\right)^{5/2}+(1\theta )\left({\displaystyle \frac{va^2}{\mathrm{\Sigma }}}\right)^{2/3}`$ (2)
$`+{\displaystyle \frac{1}{N}}\left[\theta \mathrm{ln}\theta +(1\theta )\mathrm{ln}(1\theta )\right]`$
where $`\lambda \frac{ALd^4}{NkTa^5}`$. $`F_{vdW}`$ in (2) is approximated by the $`\pi \theta L^2/\mathrm{\Sigma }1`$ limit of (1). As a result, $`F_{vdW}`$ is overestimated for $`\theta 1`$. This does not affect our analysis since the contribution of the van der Waals attraction in this regime is negligible. The corresponding spinodal condition, $`^2F/\theta ^2=0`$, leads to $`f(\theta )=\theta ^{5/2}(1\theta )(4/35)(\mathrm{\Sigma }/a^2)^{5/2}(\lambda N)^1`$ revealing a critical point at $`\theta _c=5/7`$ where $`f(\theta )`$ exhibits a maximum. The critical grafting density is thus specified by
$$\mathrm{\Sigma }_c/a^2\lambda ^{2/5}N^{2/5}$$
(3)
and for $`\mathrm{\Sigma }<\mathrm{\Sigma }_c`$ the helix-coil transition within the layer proceeds as a first-order phase transition. For lower grafting densities, $`\mathrm{\Sigma }>\mathrm{\Sigma }_c,`$ the helix-coil transition is cooperative but no phase transition is involved. In this last regime it is of interest to characterize the dependence of the transition temperature $`T_t`$ and the width of the transition $`\mathrm{\Delta }T`$ on $`\mathrm{\Sigma }`$ and $`N`$. To this end it is helpful to recast the equilibrium condition, $`F/\theta =0`$, in the form
$$\frac{\theta }{1\theta }=K(\theta )s^N\sigma _g\mathrm{exp}\left[\frac{7}{2}\lambda N\left(\frac{\theta a^2}{\mathrm{\Sigma }}\right)^{5/2}+N\left(\frac{va^2}{\mathrm{\Sigma }}\right)^{2/3}\right]$$
(4)
where $`K(\theta )`$ is the equilibrium constant governing the ratio of helices and coils. A rough idea concerning the transition is obtained by identifying it with the condition $`\theta =1/2`$ i.e., $`K(\theta )=1`$. It is helpful to consider first the case of non interacting mushrooms, $`\mathrm{\Sigma }\mathrm{}`$, when $`K(\theta )=s^N\sigma _g`$. Since $`s`$ in the vicinity of $`T_{}`$ is given by $`s\mathrm{exp}\left[r\left(\frac{T}{T_{}}1\right)\right]`$ where $`r`$ is a phenomenological constant, this condition leads to
$$T_t(\mathrm{})T_{}+\frac{T_{}}{Nr}\mathrm{ln}\sigma _gT_{}.$$
(5)
Similarly we define the width of the transition as $`\mathrm{\Delta }T=T_{}T_+`$ where $`T_+`$ is the temperature for which $`\theta _+=9/10`$ while $`T_{}`$ is the temperature corresponding to $`\theta _{}=1/10`$. The ratio $`K(\theta _+)/K(\theta _{})=\left(s_+/s_{}\right)^N=81`$ may be rewritten as $`\mathrm{ln}81Nr\mathrm{\Delta }T(\mathrm{})/T_{}`$ or
$$\mathrm{\Delta }T(\mathrm{})\frac{T_{}}{Nr}.$$
(6)
By using the same argument for a brush with $`\mathrm{\Sigma }R_0^2`$ we obtain
$$T_t=T_t(\mathrm{})+\frac{T_{}}{r}\left[\frac{7}{2}\lambda \left(\frac{a^2}{2\mathrm{\Sigma }}\right)^{5/2}+\left(\frac{va^2}{\mathrm{\Sigma }}\right)^{2/3}\right]$$
(7)
and
$$\mathrm{\Delta }T=\mathrm{\Delta }T(\mathrm{})\frac{7}{2}\frac{T_{}}{r}\lambda \left(\frac{9a^2}{10\mathrm{\Sigma }}\right)^{5/2}.$$
(8)
As $`\mathrm{\Sigma }`$ decreases, $`T_t`$ increases while $`\mathrm{\Delta }T`$ decreases. In other words, the stability of the helical state grows and the transition becomes sharper when the grafting density increases. Eventually, at the critical point, $`\mathrm{\Delta }T=0`$ thus signaling the onset of a phase transition. Note that within our model the decrease of $`\mathrm{\Delta }T`$ is due only to $`F_{vdW}`$ while the increase of $`T_t`$ results from contributions from $`F_{vdW}`$ as well as the brush penalty. This is because our choice of $`v_{hc}=v_{cc}`$ leads to a $`\theta `$ independent $`H_{eq}`$. In the general case $`\mathrm{\Delta }T`$ should reflect both contributions.
The analysis presented above focused on the case of short, immobile, grafted chains that form a single-stranded helix. In this system $`T_t`$ increases while $`\mathrm{\Delta }T`$ decreases as the grafting density grows. The effect of grafting on $`\sigma `$, the van der Waals attraction and the crowding induced stretching should however play a role irrespective of the precise specifications of the grafted layer. Additional factors may contribute when chain mobility, multiple stranded helices and longer chains are considered. Lateral chain mobility may give rise to in-plane phase separation driven by the van der Waals attraction. The discussion of multiple stranded helices should allow for the effect of loops on $`\sigma `$. It may also be necessary to allow for loss of translational entropy due to the formation of a multiple helix. The “all or nothing” model described above is clearly limited to brushes of short chains. A discussion of brushes formed by long helix-forming chains should allow for helix-coil coexistence on each of the polymers. While the discussion focused on brushes of peptides, it should be noted that similar situation is encountered in DNA chips undergoing hybridization. Finally, it is of interest to note that our considerations are somewhat similar of the analysis of the coupling between the helix-coil transition and the onset of liquid crystalline order in peptides solutions and in a collapsing chain. From this perspective, the distinguishing features of the brush case are due to the grafting modification of $`\sigma `$ and to the crowding induced chain stretching.
\***
The authors benefited from instructive discussions with M. Tirrell and D. Leckband. |
no-problem/9911/astro-ph9911460.html | ar5iv | text | # Hierarchical clustering and the baryon distribution in galaxy clusters
## 1 INTRODUCTION
Since the early hydrodynamic simulations of galaxy clusters (Evrard, 1990), the global distribution of the baryons has been of interest. In particular, the local enhancement or deficit of the baryons compared with the dark matter has profound implications for inferences of the universal baryon fraction as parametrized by $`\mathrm{\Omega }_b`$. Recall that $`\mathrm{\Omega }`$ is the ratio of the mass-energy density of the universe to the critical density required to close the universe. The ratio, $`\mathrm{\Omega }_b`$, is the contribution of the baryonic component to $`\mathrm{\Omega }`$, while $`\mathrm{\Omega }_m`$ is the contribution of all mass. The fraction, $`\mathrm{\Omega }_b`$, is constrained by primordial nucleosynthesis calculations which are sensitive to the cosmological model. Currently, estimates vary between a reported low of $`\mathrm{\Omega }_b=0.013\pm 0.003h^2`$ (White and Fabian, 1995) to $`\mathrm{\Omega }_b=(0.020\pm 0.002)h^2`$ (Bludman, 1998). Since galaxy clusters are the largest objects in the universe for which one can observe the baryon content as well as derive the total mass, they provide the most unbiased samples from which to calculate the baryon fraction of the universe. The values for $`\mathrm{\Omega }_b`$ and the baryon fraction, $`f_b{}_{}{}^{\mathrm{\Omega }_b}/_{\mathrm{\Omega }_m}^{}`$, derived from observations of clusters, may be combined to derive $`\mathrm{\Omega }_m`$, the total contribution of matter to $`\mathrm{\Omega }`$. Observations of this sort find baryon fractions of 10 to 22% (White and Fabian, 1995; White et al., 1997). This implies $`\mathrm{\Omega }_m<0.9`$ for $`f_b=0.1`$ and $`h=0.5`$ while suggesting it is probably closer to 0.3 if we take $`f_b=0.15`$ and $`h=0.65`$. However, the derivation of $`f_b`$ from the observations of clusters assumes the dark matter and hot gas are distributed in constant proportions within the cluster. Consequently, an understanding of the concentration factor of the gas in clusters is required since it represents the degree of biasing. As well, this biasing can have consequences in regards to constraining the deceleration parameter, $`q_0`$ (Rines et al., 1998).
Numerical simulations generally agree that the gas is anti-biased in clusters of galaxies (Evrard, 1990; Thomas and Couchman, 1992; Cen and Ostriker, 1993; Kang et al., 1994; Metzler and Evrard, 1994; Pearce et al., 1994; Navarro et al., 1995; Anninos and Norman, 1996; Lubin et al., 1996; Pildis et al., 1996). Results have been produced to the contrary (Owen and Villumsen, 1997) using two-dimensional codes. Pearce et al. (1994) explains the anti-bias as a result of the merging process in which gas is shocked, permanently removing energy from the dark matter component and passing it to the gas.
However, there is an uncertainty in the actual amount of bias. The bias between the universal baryon fraction and that found in clusters cannot be measured directly but must be inferred from numerical simulations. The numerical simulations must assume a model for the universe. Since the actual cosmogony of the universe is unknown in the details, variation among baryon fractions found in different model universes adds uncertainty. A better understanding of how this bias depends on cosmological models would assist in understanding the actual baryon fraction of the universe.
In hierarchical clustering, the largest structures forming at a given time do so via the amalgamation of many smaller structures which have formed at an earlier time. This is owing to the form of the initial density perturbation spectrum in which small-scale perturbations have higher initial amplitudes than large-scale. In contrast, non-hierarchical clustering involves structure formation from the collapse of large structures with smooth density distributions. The results of numerical simulations compared with observations support the theory that we live in a universe in which structure is formed hierarchically. The degree to which the hierarchical nature affects the baryon distribution in galaxy clusters is not entirely clear. In this paper, a comparison of the baryon fraction bias will be made using numerical simulations of galaxy clusters formed hierarchically and non-hierarchically. By using these extreme cases, the significance of hierarchical clustering itself will be determined. The analysis will concentrate not on the properties of individual clusters, but on the global mean properties of scaled quantities.
The layout of the paper is as follows. In Sec. 2, the simulations are described. A brief description of the analysis methods are given in Sec. 3. The effect of hierarchical clustering on the baryon fraction is examined in Sec. 4. The results are discussed in Sec. 5.
## 2 SIMULATIONS
The simulations used for this study are the same as described in Tittley and Couchman (1999). They comprise a set of five high resolution simulations in which hierarchical clustering was reduced or eliminated in four of the simulations via smoothing of the density perturbations in the initial conditions file.
The initial, unsmoothed density distribution was scale-invariant, with a spectral index of $`n=1`$, such that the density fluctuation power-spectrum had the form of the power law, $`P(k)k^n`$. It was normalised to set the variance on the scale of $`8h^1Mpc`$ to $`\sigma _8=0.935`$.
The smoothing was done using a top-hat method for two of the simulations and a low-bandpass filter for the other two. For the top-hat method, smoothing over radii of $`3h^1Mpc`$ and $`7h^1Mpc`$ was done. The frequency cutoffs were $`2\pi (7h^1Mpc)`$ and $`2\pi (14h^1Mpc)`$. Subsequent references to these simulations will omit the $`h^1`$ and $`2\pi `$ factors. The Hubble constant parameter, $`h=0.65`$, was adopted for these simulations. The details of the simulations are given in Table 1.
The simulations contained $`2\times 64^3`$ particles, half of which were collisionless particles representative of the dark matter component while the other half were collisional which represented a gas phase. The particles had masses of $`9.27\times 10^{10}\mathrm{M}_{}`$ for the dark matter and $`1.03\times 10^{10}\mathrm{M}_{}`$ for the gas. The total mass density is sufficient to produce a flat cosmology, without a cosmological constant.
The matter was evolved in a box with periodic boundary conditions on a scale of $`40h^1Mpc`$ in co-moving coordinates. The N-body $`\mathrm{AP}^3\mathrm{M}\mathrm{SPH}`$ code, hydra (Couchman et al., 1995) was used for all simulations. Two-body interactions were reduced by the use of gravitational softening. This was provided by the softening parameter, $`ϵ=20h^1kpc`$. Cooling was neglected. The cooling time for the bulk of the cluster gas is estimated to be well over the age of the clusters.
## 3 ANALYSIS
The same sample of clusters as selected in Tittley and Couchman (1999) was used here. The details of the sets of clusters for each simulation are given in Table 2. Clusters were selected by a tomographic deprojection method.
For each cluster, masses and radii were calculated at overdensities of 200 and 500. The radius for an overdensity of 200, $`R_{200}`$, is the radius of a cluster interior to which the mean density is 200 times the background density, which for these simulations is the critical density, $`\rho _c`$. A similar definition holds for $`R_{500}`$. The mass interior to these radii are denoted, $`M_{200}`$ and $`M_{500}`$, respectively. The overdensity of 200 was chosen since it is on the conservative side of 178, the overdensity at which a spherically symmetric sphere of matter virialises, according to analytic models of top-hat collapses. The masses of both the dark matter, $`M_{dark}`$, and gas, $`M_{gas}`$, were found at this virial radius, along with the total mass, $`M_{200}`$.
A lower limit to the size of the clusters was set by the requirement that each cluster, within the overdensity radius of $`R_{500}`$, have at least 300 gas particles and 300 dark matter particles. This ensures the densities are calculated correctly (Tittley and Couchman, 1999).
Profiles of the dark matter and gas densities were calculated for each of the clusters. This was done by summing the mass contributions of the particles falling in exponentially separated radial bins centred on the clusters. The profiles were scaled by the approximate virial radius, $`R_{200}`$.
Mean density profiles for each simulations were then calculated from the profiles of the individual clusters in each simulation. For the calculation of the mean density profile, the individual cluster profiles were weighted by their respective mass, $`M_{200}`$. This weighting removes the bias towards the more plentiful low-mass clusters. However, for the density profiles, this weighting was found to have only a small effect, demonstrating the lack of dependency on mass for the profiles.
## 4 BARYON FRACTION
### 4.1 The concentration parameter, $`\mathrm{{\rm Y}}`$
If we define $`\mathrm{{\rm Y}}_{200}`$ as the gas fraction in a spherical region of overdensity 200 normalised to the cosmic value, then we have (White et al., 1993)
$$\mathrm{{\rm Y}}_{200}\frac{M_{gas}}{M_{gas}+M_{dark}}\frac{\mathrm{\Omega }_m}{\mathrm{\Omega }_b}.$$
(1)
Similarly, we may define $`\mathrm{{\rm Y}}(r)`$ as the normalised gas fraction in a shell of radius $`r`$. Here, $`r`$ is replaced by the dimensionless parameter $`r/R_{\overline{\delta }}`$ since radial profiles of both the dark matter and gas densities are characterised by the virial radius.
### 4.2 Variation of $`\mathrm{{\rm Y}}`$ on cluster mass
The values of $`\mathrm{{\rm Y}}_{200}`$, when plotted versus $`M_{200}`$ (1), support the values found in Evrard (1997) and Eke et al. (1998) of $`\mathrm{{\rm Y}}_{200}=0.85`$ to $`0.90`$ if the sample is restricted to the largest clusters.
For the lower mass clusters, the clusters in the unsmoothed model have values ranging down to $`0.75`$. There is no trend with mass for this model other than an increase in dispersion. The mean for the hierarchically formed clusters is $`0.84`$ with the standard deviation varying from $`\pm 0.04`$ for clusters less massive than $`10^{14}\mathrm{M}_{}`$ to $`\pm 0.01`$ for clusters more massive than $`10^{15}\mathrm{M}_{}`$.
For the smoothed models, the trend is for $`\mathrm{{\rm Y}}_{200}`$ to decrease with mass with little change in the dispersion. For clusters more massive than $`10^{15}\mathrm{M}_{}`$, $`\mathrm{{\rm Y}}=0.87`$ to $`0.97`$. Below $`10^{14}\mathrm{M}_{}`$, this drops down to the range $`0.82`$ to $`0.92`$.
### 4.3 Baryon concentration profiles
Since it is found that both the dark matter and gas density profiles scale with the virial radius, it stands to reason that the radial profile of the baryon concentration should also scale with radius. Numerical simulations of galaxy clusters using a 1-D Lagrangian code indicate that this is not true. Knight and Ponman (1997) reports that $`f_b`$ varies by a factor of about two at a given scaled radius between clusters that differ in mass by two orders of magnitude, with the less massive cluster having the smaller baryon fraction. This trend is consistent with the variation in $`\mathrm{{\rm Y}}_{200}`$ with the mass of the cluster seen for the clusters formed from smoothed initial conditions. However, since this factor is comparable to the cluster to cluster variation, it is justified once again to look at the mean radial profile of all the clusters in a model.
The baryon fraction profiles vary considerably for those clusters formed in the simulation with hierarchical clustering from those formed otherwise (2, top) particularly around the virial radius. The baryon fraction is enhanced in a much deeper region ($`0.7<r/R_{200}<10`$) in the hierarchical case than in the case with smoothed initial conditions ($`1<r/R_{200}<4`$) and the enhancement is almost 4 times greater. In both cases, the peak in the baryon enrichment occurs just beyond the virial radius. Towards the more central regions, the enrichment in the hierarchical clusters drops off slightly more steeply.
The discrepancies are not so large for the cumulative profile (2, bottom). The clusters in the hierarchical model are generally more depleted in baryons than those in the smoothed models, in agreement with the results in Sec. 4.2. The mean profile for the unsmoothed model compares well those of the simulated clusters discussed in Eke et al. (1998). Those are simulated for a low-density universe.
It must be noted that there are large deviations in these profiles among the many clusters in the hierarchical scenario. These mean profiles represent trends. For the clusters formed non-hierarchically, however, the baryon profiles are fairly consistent among clusters.
Interior to $`1.5R_{200}`$, the mean $`\mathrm{{\rm Y}}`$ profile for the unsmoothed model may be fit by $`\mathrm{{\rm Y}}(r)=0.67\pm 0.03+(0.52\pm 0.05){}_{}{}^{r}/_{R_{200}}^{}`$, non-cumulative. For the mean of the profiles for the non-hierarchically formed clusters, the fit takes a shallower form, $`\mathrm{{\rm Y}}(r)=0.76\pm 0.02+(0.30\pm 0.07){}_{}{}^{r}/_{R_{200}}^{}`$.
## 5 DISCUSSION
Using a series of simulations in which clusters were formed either hierarchically and non-hierarchically, the relevance of hierarchical structure formation to the baryon distribution in clusters was explored. Non-hierarchical clustering was achieved by smoothing the initial density distribution.
It is found that the baryons are anti-biased in clusters of galaxies and this biasing is dependent on the presence of smoothing in the initial conditions. The bias parameter $`\mathrm{{\rm Y}}0.85`$ for the hierarchically formed clusters. For the clusters formed from smoothed initial conditions, $`\mathrm{{\rm Y}}0.92`$, indicating less of an anti-bias. For the non-hierarchically formed clusters, the normalised baryon fraction is dependent on the mass of the cluster, as well, which is not observed for the hierarchically formed clusters.
In all cases, the baryon fraction was found to increase from the centre of the clusters, outward. Interior to $`0.1R_{200}`$, the normalised factor spans $`\mathrm{{\rm Y}}=0.5`$ to $`0.7`$. We determined that the baryon fraction profile, as parametrized by the normalised fraction, is steeper for hierarchically formed clusters. For these, it was found that $`f_b(0.52\pm 0.05)r`$ interior to $`1.5R_{200}`$. For non-hierarchically formed clusters, the profiles were shallower, regardless of the degree of initial smoothing, with a mean dependency of $`f_b(0.30\pm 0.07)r`$. The baryon fractions peak between 1 to 2 virial radii. For the clusters formed hierarchically, the baryons ‘pile up’ in a deep region spanning $`0.7<r/R_{200}<10`$. This implies that measurements of the baryon fraction of clusters are sensitive to the radius outward to which the baryon content is integrated.
Baryon fraction profiles determined from observations of clusters via the deprojection of the gas density (White et al., 1997) support the result that the baryon fractions increase with increasing radius (Markevitch and Vikhlinin, 1997). However, the profiles are often shallower than those found here. White and Fabian (1995) finds from a sample of 19 clusters that the cumulative measurement of $`f_b`$ increases with radius following roughly the $`f_b=0.06R`$ with $`R`$ in $`Mpc`$. Similarly, White et al. (1997) finds a fit of $`f_b=(0.12\pm 0.04)R^{0.7\pm 0.3}`$ for a sample of 28 clusters without cooling flows. With cooling flows, a higher central value for $`f_b`$ and a shallower profile are reported, as would be expected. The fractional slopes of both these fits are steeper than that of the mean $`\mathrm{{\rm Y}}`$ profile of all the models. This is likely due to the lack of cooling in the simulations, which would deposit a concentrated amount of gas near the cores of the clusters. This gas would represent a small fraction of the total mass of the halo gas, but would elevate the baryon fraction in the interior. This should not greatly affect the results at radii near the virial radius, since the cooled gas would constitute a small fraction of the total gas mass at that point.
These results indicate that the observed baryon fractions in galaxy clusters lead to under-estimates of the universal baryon fraction by a factor of $`5\%`$ to $`20\%`$ and that this factor is dependent on the cosmology. The anti-biasing of the gas is greatest in the hierarchical run, consistent with the picture of Pearce et al. (1994) in which the bias is due to transfer of energy from the dark matter to the gas during merging. Since cooling flows are not created in these simulations, the true bias factor is not well established. However, cooling flows will only alter the baryon fractions in the inner radii since the total mass cooled is only on the order of $`10^{12}\mathrm{M}_{}`$ or less. Ultimately, this result exacerbates the baryon overdensity problem in galaxy clusters. The variation in the enrichment of baryons with distance from the cluster centres also makes measurements of $`f_b`$ problematic. It is crucial that the cluster be observed to radii of up to two virial radii to reduce the error to less than $`10\%`$.
Observations of the baryon fraction in clusters will be hampered by this biased distribution of gas in clusters. Measurements must be made to more than twice the virial radius, sampling regions for which accurate densities of both mass and gas are difficult. This uneven distribution is not as significant if there is any degree of smoothing, implying that any attempt to formulate a correction factor will be dependent on the cosmology at the $`10\%`$ level. Fortunately, this is at the present level of measurement error. |
no-problem/9911/hep-th9911157.html | ar5iv | text | # References
Holography and the generalized second law of thermodynamics in (2+1)-dimensional cosmology
Bin Wang<sup>a,b,</sup><sup>1</sup><sup>1</sup>1e-mail:binwang@fma.if.usp.br, Elcio Abdalla<sup>a,</sup><sup>2</sup><sup>2</sup>2e-mail:eabdalla@fma.if.usp.br
<sup>a</sup> Instituto De Fisica, Universidade De Sao Paulo, C.P.66.318, CEP 05315-970, Sao Paulo, Brazil
<sup>b</sup> Department of Physics, Shanghai Teachers’ University, P. R. China
## Abstract
The Fischler-Susskind entropy bound has been studied in (2+1)-dimensional universes with negative cosmological constant. As in all contracting universes, that bound is not satisfied. Furthermore, we found that the Fischler-Susskind bound is not compatible with a generalized second law of thermodynamics in (2+1)-dimensional cosmology, neither the classical nor the quantum version. On the other hand, the Hubble entropy bound has been constructed in (2+1)-dimensional cosmology and it is shown compatible with the generalized second law of thermodynamics.
PACS number(s): 04.70.Dy, 98.80.Cq
Motivated by the well-known result in black hole theory that the total entropy of matter inside a black hole cannot exceed the Bekenstein-Hawking entropy, a conceptual change in our thinking about gravity has recently been put forward by the so called “holographic principle” . According to this principle, all the degrees of freedom inside a volume is expressed on its boundary, implying that the entropy of a system cannot be larger than its boundary area. A specific generalization of the holographic principle to cosmology was realized by Fischler and Susskind (FS) . A remarkable point of their proposal is that the holographic principle is valid for flat or open universes with the equation of state satisfying the condition $`0P\rho `$. However, for closed universes the principle is violated. The problem becomes even more serious if one investigates the universe with a negative cosmological constant . In that case the holographic principle fails, independently of whether the universe is closed, open or flat. Various different modifications of the FS version of the holographic principle have been raised recently, such as replacing the holographic principle by the generalized second law of thermodynamics , using the cosmological apparent horizon instead of the particle horizon in the formulation of holographic principle , changing the definition of “degrees of freedom” etc. A very recent result claimed that the holographic principle in a closed universe can be obeyed if the universe contains strange negative pressure matter . The study of the cosmic holography has also been extended to Pre-big-bang string cosmological models . All these studies have concentrated on (3+1)-dimensional (4D) cosmology.
In our previous work, we have considered the investigations on cosmic holography in (2+1)-dimensional (3D) cosmological models . Analogously to the 4D counterpart, the holographic principle is satisfied in all 3D flat and open universes, but breaks down for 3D closed universes. Attempts to uphold the holographic principle by introducing negative pressure matter as well as matter with very unconventional high pressure failed, because they cannot accomodate any classical description after the big bang. It is of interest to generalize our discussions to 3D universes with a negative cosmological constant. There has been many successful applications of the holographic principle for 3D pure Anti-de Sitter (AdS) space from string theory \[11-14\]. Thus we have the motivation to investigate whether the holographic principle holds in 3D AdS cosmology.
Recently a generalized second law (GSL) of thermodynamics in 4D cosmologies has been put forward , and its relation to the Hubble entropy bound (HE) suggested by Veneziano has also been discussed. Further study of the GSL in 4D string cosmology has been addressed as well . The second purpose of the present paper is to consider their discussions in 3D cosmological models. By establishing the GSL in 3D universes, and studying its relation to FS entropy bound, we find that the FS bound is not compatible with neither the classical nor the quantum mechanical version of GSL. Since the second law of thermodynamics is more fundamental, the incompatible result between FS bound and GSL gives us additional reason to look for a reformulation of the cosmological holographic principle.
The conflict result between the FS bound and the GSL can be attributed to the fact that the FS bound is too strong and a weaker entropy bound in cosmology is called for. In 4D universes it was claimed that the HE bound, which is looser than the FS bound, is sufficient to avoid any problem with entropy produced at reheating after inflation . A generalization of these studies to 3D universes is appealing and will be carried out in this paper. We are going to define the HE bound and discuss its relation with the FS bound and the Bekenstein entropy (BE) bound in 3D cosmological models. The relation between HE bound and GSL will also be addressed and the compatible result will be reached. These results support the argument that the HE bound may be a candidate to replace the FS bound and describe the cosmic holography.
Cosmological solutions in (2+1)-dimensional Einstein gravity have been proposed in . In terms of the (2+1)-dimensional Robertson-Walker line element
$$\mathrm{d}s^2=\mathrm{d}t^2a^2(t)(\frac{\mathrm{d}r^2}{1kr^2}+r^2\mathrm{d}\theta ^2),$$
(1)
the Einstein field equations become
$`({\displaystyle \frac{\dot{a}}{a}})^2+{\displaystyle \frac{k}{a^2}}`$ $`=`$ $`2\pi G\rho ,`$ (2)
$`{\displaystyle \frac{\ddot{a}}{a}}`$ $`=`$ $`2\pi GP,`$ (3)
$`{\displaystyle \frac{d}{dt}}(\rho a^2)+P{\displaystyle \frac{d}{dt}}a^2`$ $`=`$ $`0.`$ (4)
When the material content is a perfect fluid with equation of state
$$P=(\gamma 1)\rho ,$$
(5)
where $`\gamma `$ is a constant, we derive the relation
$$\rho a^{2\gamma }=const=\rho _0a_0^{2\gamma }.$$
(6)
The scale factor is determined by a Friedmann-like equation
$$(\frac{\dot{a}}{a})^2=\frac{2GM_0}{a^{2\gamma }}\frac{k}{a^2},$$
(7)
where $`M_0=\pi \rho _0a_0^{2\gamma }`$. For $`\gamma =1`$, the universe is dust-filled and always expands regardless of the value of $`k`$. However, for $`1<\gamma 2`$, the solutions of (7) are closed, open or flat cosmological models according to whether $`k`$ is $`1,1,`$ or $`0`$, respectively. The case $`\gamma =3/2`$ corresponds to the radiation-dominated universe.
It is of interest to investigate the holographic principle in a universe with negative vacuum energy. A universe with negative cosmological constant is contracting, independently of the value of $`k`$. For simplicity, we just consider the flat universe ($`k=0`$) with general equation of state ($`1<\gamma 2`$). The vacuum energy density is negative, $`\lambda <0`$, so that in the expanding universe $`\rho ^{}=\rho \lambda ={\displaystyle \frac{\rho _0a_0^{2\gamma }}{a^{2\gamma }}}\lambda `$, and the Friedmann equation can be expressed as
$$\dot{a}^2=\frac{2GM_0}{a^{2\gamma 2}}\lambda a^2.$$
(8)
The scale factor can be calculated and has the form
$$a(t)=(2GM_0/\lambda )^{1/(2\gamma )}\{\mathrm{sin}[\gamma \sqrt{\lambda }t]\}^{1/\gamma }.$$
(9)
As in the case discussed in , we find that $`\dot{a}`$ vanishes at $`a=(2GM_0/\lambda )^{1/(2\gamma )}`$, and after that point $`\dot{a}`$ becomes negative and the universe collapses. This happens within a finite time after the beginning of the expansion. One can find the value of $`L_H`$ at the turning point:
$$L_H(turning)=\frac{1}{2\gamma \lambda ^{1/2}}B((\gamma 1)/2\gamma ,1/2)$$
(10)
where $`B(p,q)`$ is the Euler Beta function. It is worth noting that the particle horizon here has the same dependence on $`\lambda `$ as that in 4D case . Putting these formulas together, we get at the turning point
$$\frac{S}{A}\lambda ^{1/\gamma 1/2}.$$
(11)
Considering $`\gamma 2`$, for small value of $`\lambda `$ (usually believed to be smaller than $`10^{122}`$), the entropy over area bound is satisfied at the turning point. Now we can consider what happens near the final stage of collapse, where the universe shrinks to the Planck scale. By symmetry, $`L_H{\displaystyle \frac{2a_0}{a(turning)}}L_H(turning)\lambda ^{1/2\gamma 1/2}`$ at this time. The scale factor at the Planck time $`t=1`$ is $`a(t=1)=(2GM_0/\lambda )^{1/2\gamma }\{\mathrm{sin}[\gamma \sqrt{\lambda }]\}^{1/\gamma }`$. The ratio is $`S/A\lambda ^{1/2\gamma 1/2}`$. Hence for small $`\lambda `$, the ratio is much bigger than unity for the universe with general equation of state $`1<\gamma 2`$. We find that prior to the point of maximal expansion, the holographic constraints hold. However once after the point where the ratio exceeds unity the holographic bound cannot be further maintained. This result is in agreement with that in (3+1)-dimensions.
For comparison, we can easily see that in the general d-dimensional case, eq.(4) gets modified to $`{\displaystyle \frac{d}{dt}}(\rho a^{d1})+P{\displaystyle \frac{d}{dt}}(a^{d1})=0`$, which in view of the equation of state (5) has a solution $`\rho a^{\gamma (d1)}`$. Going through calculations, we learn that Eq(11) has the same form, leading above to similar conclusions. From this point however, we rather stay in 2+1-dimensions where we are able to draw further conclusions.
As pointed out in , the problem can not be expected to be cured by replacing particle horizon by the apparent horizon proposed in , because at the turning point $`\dot{a}=0`$, and Eq.(16) from ref. diverges, that is,
$$\frac{4\sigma }{3a^2(t)\dot{a}(t)}\mathrm{}.$$
(12)
Therefore the new holographic principle is violated even earlier.
We cannot naively expect the bound to be saved by considering the universe with the unusual negative pressure matter analoguous to the 4D closed universe case , because as we see from (8), after the big bang the universe cannot have sufficient expansion for $`\gamma <1`$, therefore the classical description is not valid.
It is appealing to establish the GSL in 3D universes and study the relation between the GSL and the FS bound. Using the idea proposed in , the definition of the total entropy of a domain in 3D universes containing more than one cosmological horizons is given by $`S=n_HS^H`$, where $`n_H`$ is a number of cosmological horizons within a given comoving “volume” divided by a “volume” of a single horizon $`n_H=a(t)^2/|H(t)|^2`$; $`S^H`$ is the entropy within a given horizon. The classical GSL requires that the cosmological evolution must obey $`dS0`$, which corresponds to
$$n_H_tS^H+_tn_HS^H0$$
(13)
Adopting the idea used in the GSL for black hole, we consider that there could be many sources and types of entropy, and the total entropy is the sum of their contributions. Supposing a single type of entropy is dominant, $`S^H=|H|^\alpha `$, where $`\alpha `$ indicates the type of the entropy source, therefore $`S=(a|H|)^2|H|^\alpha `$, and Eq(13) can be rewritten as
$$2H+(2+\alpha )\frac{\dot{H}}{H}0.$$
(14)
Let us reexpress Eqs.(2-4) in the forms,
$`H^2`$ $`=`$ $`2\pi G\rho {\displaystyle \frac{k}{a^2}}`$
$`\dot{H}`$ $`=`$ $`2\pi G(P+\rho )+{\displaystyle \frac{k}{a^2}}`$ (15)
$`\dot{\rho }`$ $`+`$ $`2H(P+\rho )=0`$
and substitute them into (14). We find that the relations for equations of state determined by the GSL are
$`{\displaystyle \frac{P}{\rho }}`$ $``$ $`{\displaystyle \frac{2}{2+\alpha }}1+{\displaystyle \frac{\alpha k}{2\pi G(2+\alpha )a^2\rho }},forH>0,`$ (16)
$`{\displaystyle \frac{P}{\rho }}`$ $``$ $`{\displaystyle \frac{2}{2+\alpha }}1+{\displaystyle \frac{\alpha k}{2\pi G(2+\alpha )a^2\rho }},forH<0.`$ (17)
In the last terms of Eqs(16,17), $`a^2\rho `$ corresponds to the energy of the whole universe, so $`E=a^2\rho k`$, thus the last terms in Eqs. (16,17) can be neglected.
Employing the first law of thermodynamics, $`TdS=dE+PdV=(\rho +P)dV+Vd\rho `$, the temperature can be obtained by $`T^1=({\displaystyle \frac{S}{E}})_V={\displaystyle \frac{s}{\rho }}`$, where $`E=\rho V,S=sV`$, therefore
$$T=\frac{1}{\pi G(2+\alpha )|H|^\alpha }$$
(18)
To ensure that singularities are avoided for the expressions of the total entropy $`S`$, $`_tS`$ and the temperature $`T`$ when a flat space limit of vanishing $`H`$ is taken into account, the reasonable physical range of $`\alpha `$ should lie within the region $`1\alpha 0`$.
Let us now consider that the dominant contribution to the entropy of the universe is given by the geometric entropy $`S_g`$ whose source is the existence of a cosmological horizon . This is a speculative notion introduced in . Let us herewith suppose that a component of entropy arises from geometry. We are thus in a position to discuss the relation between the GSL and the FS entropy bound. For a system with a cosmological horizon, $`S_g^H`$ is given by
$$S_g^H=|H|^1G_N^1,$$
(19)
which corresponds to $`\alpha =1`$. Substituting this value of $`\alpha `$ into Eqs.(16,17), equations of state corresponding to adiabatic evolution with dominant $`S_g`$ are obtained. For the expanding universe $`H>0`$, and GSL requires $`P\rho `$, which is in agreement with the results obtained in for the FS bound. However for negative $`H`$, which corresponding to the contracting universe, GSL requires $`P/\rho (1,\mathrm{})`$. This range corresponds to $`\gamma >2`$ in Eq.(5), which is ruled out in any 3D contracting universes in FS bound discussions. Therefore, for the 3D contracting universes the FS bound is not compatible with the GSL.
Whether adding a missing quantum entropy term and developing the quantum mechanical version of the GSL can help us arriving at the compatibility between the FS bound and the GSL for contracting 3D universes is still not clear. Using the definition for the quantum entropy in 4D cases , $`dS_{Quan.}=\mu dn_H`$, where $`\mu `$ is a “chemical potential”, always taken to be positive, $`n_H=(aH)^2`$, we obtain
$`dS`$ $`=`$ $`dS_{Class.}+dS_{Quan.}`$ (20)
$`=`$ $`dn_HS^H+n_HdS^H\mu dn_H`$
where $`S^H`$ is the classical entropy within a cosmological horizon and $`S^H=|H|^\alpha `$ if the classical entropy is dominated by a single source. The quantum modified GSL can be expressed as
$$(2H+2\frac{\dot{H}}{H})n_H(S^H\mu )+\alpha \frac{\dot{H}}{H}n_HS^H0.$$
(21)
Considering that geometric entropy still dominates the classical entropy, $`\alpha =1`$, we learn from (21) that
$$\frac{P}{\rho }\frac{S^H}{S^H2\mu }$$
(22)
for the contracting universe, $`H<0`$.
If $`\mu S^H`$, it returns to the classical case and leads to $`{\displaystyle \frac{P}{\rho }}(1,\mathrm{})`$; if quantum effects are comparable to classical effects, $`\mu S^H`$, $`{\displaystyle \frac{P}{\rho }}(1,\mathrm{})`$; and if quantum effect dominates, say $`\mu S^H`$, $`{\displaystyle \frac{P}{\rho }}(0,\mathrm{})`$. These results are never compatible with the requirement of the FS bound, because $`{\displaystyle \frac{P}{\rho }}>1`$ or $`1<{\displaystyle \frac{P}{\rho }}<0`$ corresponds to $`\gamma >2`$ or $`0<\gamma <1`$, respectively, which are all unacceptable by the FS bound in all contracting universes. Therefore the quantum consideration still cannot lead to the compatibility between the FS bound and the GSL.
Since we believed that the second law of thermodynamics is a fundamental principle in physics, such an incompatible result gives us additional motivation to seek for the reformulation of the cosmic holographic principle. One may attribute the incompatibility to the argument that FS bound is too strong, a looser cosmic entropy bound is called for to solve this conflict with GSL. A possible way was suggested by Veneziano for 4D cosmology by replacing the FS bound by the HE bound. Based upon the argument that a black hole larger than $`H^1`$ cannot form, generalizing to 3D cases we find the largest entropy in a region corresponding to have just one black hole per Hubble “volume” $`H^2`$ is that $`sM_p^2|H|`$, where $`M_p=G_N^{1/2}`$, and the HE bound in 3D universes can be defined as
$$S^HM_p^2|H|^1S_{HE}$$
(23)
Recalling the definitions for the BE bound and the FS bound, in 3D they can be expressed as
$`S_{BE}`$ $``$ $`EL/\mathrm{}Md_p/\mathrm{}\rho d_p^2d_p/\mathrm{}H^2d_p^2d_p/l_p^2`$ (24)
$`S_{FS}`$ $``$ $`d_p/\mathrm{}d_p/l_p^2`$ (25)
where $`d_p`$ is the particle horizon, $`l_p\sqrt{\mathrm{}}M_p`$. We substitute the particle horizon $`d_p`$ by $`H^1`$, and consider that when applied to non-inflationary cosmology, they are about the same . Thus in 3D cases we can reproduce the relation among different entropy bounds first obtained in 4D cases ,
$$S_{HE}=S_{FS}^{1/2}S_{BE}^{1/2}$$
(26)
It is easy to see that the HE bound is much looser than the FS bound.
The most attractive point now is to study whether the HE bound is compatible with GSL in 3D cosmological models. Assuming a single dominant entropy form, $`S^H=({\displaystyle \frac{|H|}{M_p}})^\alpha `$, bound (23) can be written as
$$(\frac{|H|}{M_p})^\alpha M_p^2|H|^1$$
(27)
Considering the reasonable physical region of $`\alpha `$, we need
$$|H|M_pM_p^{1/\alpha +1}.$$
(28)
From the HE bound, we learn that $`|H|`$ has a maximum value, therefore in order not to violate this bound we need for expanding universe $`H>0`$, the evolution of the 3D universe undergoes decelerated expansion, say $`H>0,\dot{H}<0`$. GSL allows such an evolution. Since in the physical region $`1\alpha 0`$, Eq(14) reads
$$2H+(2+\alpha )\frac{\dot{H}}{H}>2H+\frac{\dot{H}}{H}0$$
(29)
For $`\dot{H}<0`$, it leads directly to a requirement for the equation of state, that is, $`\rho P`$. This is quite natural and thus the entropy bound is valid without violation of GSL.
For a contracting universe with $`H<0`$, to satisfy the entropy bound, we need the universe to experience decelerated contraction, namely $`\dot{H}>0`$. This requirement is obviously not compatible with GSL, because Eqs(14)and (29) tell us that
$$2|H|^2\dot{H}$$
(30)
which is false for $`\dot{H}>0`$.
Luckily this conflict between HE bound and GSL can be resolved by considering the quantum modified GSL (21). For contracting 3D universes, $`H<0`$, and Eq(21) can be rewritten as
$$2|H|^2n_H(S^H\mu )\dot{H}[2n_HS^H2n_H\mu +\alpha n_HS^H].$$
(31)
Neglecting the quantum effect, $`\mu S^H`$, Eq(31) boils down to (30) for $`\alpha =1`$. However if we consider that the quantum effect is strong, namely $`S^H\mu <0`$, we can rewrite Eq(31) as
$$2|H|^2\dot{H}2+\frac{\alpha S^H}{S^H\mu }.$$
(32)
For decelerated contraction $`\dot{H}>0`$, Eq(32) certainly holds. Thus considering the quantum effect, the conflict between HE bound and GSL can be overcome.
In summary we have found that analogously to 3D closed universes, FS bound in 3D cosmological models with negative cosmological constant breaks down regardless of the value of $`k`$. Unlike the 4D closed universe , here the negative pressure matter cannot be used to save the FS bound. Establishing the GSL in 3D cosmologies, we have shown that the state equations required are not consistent with those needed by FS bound, which shows that FS bound and GSL are not compatible. Furthermore we have shown that the conflict cannot be resolved by taking account of the quantum modified version of GSL, which has not been addressed in 4D cases. Considering that the GSL is a fundamental principle in physics, we have further motivations to rethink the expression of cosmic holography. Extending different cosmic entropy bounds to 3D cosmology, we have reproduced the relation between HE bound, BE bound and FS bound first obtained in 4D cases . Compared to FS bound, HE bound is looser, and compatible with GSL. This result agrees with that claimed in 4D cosmology and supports the argument that HE bound is a candidate for describing cosmic holography. It is of interest to consider generalization of our discussions on the relation between GSL and cosmic entropy bound to higher dimensional spacetimes.
ACKNOWLEDGEMENT: This work was partically supported by Fundac$`\stackrel{~}{a}`$o de Amparo $`\stackrel{`}{a}`$ Pesquisa do Estado de S$`\stackrel{~}{a}`$o Paulo (FAPESP) and Conselho Nacional de Desenvolvimento Cient$`\stackrel{´}{i}`$fico e Tecnol$`\stackrel{´}{o}`$gico (CNPQ). B. Wang would also like to acknowledge the support given by Shanghai Science and Technology Commission. |
no-problem/9911/math9911021.html | ar5iv | text | # Slices in the unit ball of a uniform algebra
## Abstract.
We show that every nonvoid relatively weakly open subset, in particular every slice, of the unit ball of an infinite-dimensional uniform algebra has diameter $`2`$.
###### Key words and phrases:
Uniform algebra, slice, denting point
It is an important task in Banach space theory to determine the extreme point structure of the unit ball for various examples of Banach spaces. The most common way to describe “corners” of convex sets is by looking for extreme points, exposed points, denting points and strongly exposed points. Every strongly exposed point is both denting and exposed and every denting or exposed point is extreme.
In this note $`𝒜`$ denotes an infinite-dimensional uniform algebra, i.e., an infinite-dimensional closed subalgebra of some $`C(K)`$-space which separates the points of $`K`$ and contains the constant functions. In Beneker and Wiegerinck demonstrated the non-existence of strongly exposed points in $`B_𝒜`$, the closed unit ball of $`𝒜`$. Here we shall prove a stronger result by more elementary means. A corollary of our result is that the set of denting points is, in fact, also empty.
Recall that we may assume that $`K`$ is the Silov boundary of $`𝒜`$. It is a fundamental result in the theory of uniform algebras that then the set of strong boundary points is dense in $`K`$; cf. \[4, p. 48 and p. 78\]. (A point $`xK`$ is a strong boundary point if for every neighbourhood $`V`$ of $`x`$ and every $`\delta >0`$ there is some $`f𝒜`$ such that $`f(x)=f=1`$ and $`|f|\delta `$ off $`V`$.)
We now turn to our first result, which gives a quantitative statement of the non-dentability of $`B_𝒜`$.
###### Theorem 1.
Every slice of the unit ball of an infinite-dimensional uniform algebra $`𝒜`$ has diameter $`2`$.
###### Proof.
Take an arbitrary slice $`S=\{aB_𝒜:re\mathrm{}(a)1\epsilon \}`$, where $`\mathrm{}=1`$. We will produce two functions in $`S`$ having distance nearly $`2`$.
Let $`0<\delta \epsilon /11`$. We first pick some $`fB_𝒜`$ such that
$$re\mathrm{}(f)1\delta .$$
The functional $`\mathrm{}`$ can be represented by a regular Borel measure $`\mu `$ on $`K`$ with $`\mu =1`$, i.e., $`\mathrm{}(a)=_Ka𝑑\mu `$ for all $`a𝒜`$. Let $`\mathrm{}V_0K`$ be an open set with $`|\mu |(V_0)\delta `$; such a set exists since $`K`$ is infinite. Fix a strong boundary point $`x_0V_0`$. Using the definition of a strong boundary point, inductively construct functions $`g_1,g_2,\mathrm{}𝒜`$ and nonvoid open subsets $`V_0V_1V_2\mathrm{}`$ such that
$$g_n(x_0)=g_n=1,|g_n|\delta \text{ on }KV_{n1}$$
and
$$V_n=\{xV_{n1}:|g_n(x)1|<\delta \}.$$
Let $`N>1/\delta `$ and define
$$g=\frac{1}{N}\underset{k=1}{\overset{N}{}}g_k,h=f(1g)𝒜.$$
By construction, $`|h|\delta `$ on $`V_N`$ and $`|h|1+\delta `$ on $`KV_0`$. We claim that $`h1+3\delta `$. In fact, if $`xV_{r1}V_r`$, then $`|1g_k(x)|\delta `$ if $`1k<r`$, $`|g_r(x)|1`$ and $`|g_k(x)|\delta `$ if $`r<kN`$, and therefore
$$|h(x)|\frac{1}{N}\underset{k=1}{\overset{N}{}}|1g_k(x)|\frac{(N1)(1+\delta )+2}{N}1+3\delta .$$
We now estimate $`|\mathrm{}(f)\mathrm{}(h)|`$:
$`|\mathrm{}(f)\mathrm{}(h)|`$ $`{\displaystyle _{KV_0}}|fh|d|\mu |+{\displaystyle _{V_0}}|fh|d|\mu |`$
$`{\displaystyle _{KV_0}}|g|d|\mu |+{\displaystyle _{V_0}}(|f|+|h|)d|\mu |`$
$`\delta +(2+3\delta )|\mu |(V_0)4\delta .`$
Next, we produce a function $`\phi 𝒜`$ such that
$$\phi (x_0)=\phi =1,|\phi |\delta \text{ on }KV_N.$$
We then have $`h\pm \phi 1+4\delta `$, and the functions $`\psi _\pm =(h\pm \phi )/(1+4\delta )`$ are in the unit ball of $`𝒜`$. We have $`|\mathrm{}(\phi )|2\delta `$ and thus
$$|\mathrm{}(\psi _\pm )\mathrm{}(h)||\mathrm{}(h)|\frac{4\delta }{1+4\delta }+\frac{2\delta }{1+4\delta }6\delta .$$
Consequently,
$$re\mathrm{}(\psi _\pm )re\mathrm{}(f)10\delta 111\delta 1\epsilon $$
so that $`\psi _\pm S`$; but $`\psi _+\psi _{}=2/(1+4\delta )2`$ as $`\delta 0`$. Hence $`diamS=2`$. ∎
The point of working with $`g`$ rather than $`g_1`$ in the proof is to control $`1g`$. Another way to achieve this is to construct a suitable conformal map $`\varphi `$ from the unit disk to a neighbourhood of $`[0,1]`$ in $``$ and to consider $`\varphi g_1`$.
We now extend Theorem 1 to relatively weakly open subsets.
###### Theorem 2.
Every nonvoid relatively weakly open subset $`W`$ of the unit ball of an infinite-dimensional uniform algebra $`𝒜`$ has diameter $`2`$.
###### Proof.
Every nonvoid relatively weakly open subset of the unit ball of a Banach space contains a convex combination of slices, see \[2, Lemma II.1\] or . Thus, if $`WB_𝒜`$ is given as above, there are slices $`S^{(1)},\mathrm{},S^{(n)}`$ and $`0\lambda _j1`$, $`_{j=1}^n\lambda _j=1`$, such that $`_{j=1}^n\lambda _jS^{(j)}W`$.
Let $`S^{(j)}=\{aB_𝒜:re\mathrm{}_j(a)1\epsilon _j\}`$ with $`\mathrm{}_j=1`$ and representing measures $`\mu _j`$. We now perform the construction of the proof of Theorem 1 with $`\epsilon =\mathrm{min}\epsilon _j`$, $`0<\delta \epsilon /11`$ as before and a nonvoid open set $`V_0K`$ such that $`|\mu _j|(V_0)\delta `$ for all $`j`$. We obtain functions $`h^{(j)}`$ and $`\phi `$ (independently of $`j`$) such that $`(h^{(j)}\pm \phi )/(1+4\delta )S^{(j)}`$ and $`\phi =1`$. Therefore $`_{j=1}^n\lambda _jh^{(j)}\pm \phi (1+4\delta )W`$, and $`diamW=2`$. ∎
In case $`K`$ does not have isolated points, Theorem 1 is a formal consequence of the Daugavet property of $`𝒜`$, proved in or , and \[3, Lemma 2.1\]. Likewise Theorem 2 follows from .
###### Corollary 3.
The unit ball of an infinite-dimensional uniform algebra does not contain any denting points or merely points of continuity for the identity mapping with respect to the weak and the norm topology.
We would also like to remark that T.S.S.R.K. Rao has shown that $`B_{WC(K,X)}`$, the unit ball in the space of continuous functions from a compact Hausdorff space into a Banach space equipped with its weak topology, has no denting points. He has also given a proof of Corollary 3 based on techniques from that paper. |
no-problem/9911/astro-ph9911367.html | ar5iv | text | # Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars
## 1 Introduction
Color-temperature (CT) relations and bolometric corrections (BCs) are often used to infer the physical characteristics of stars from their photometric properties and, even more commonly, to translate isochrones from the theoretical (effective temperature, luminosity) plane into the observational (color, magnitude) plane. The latter allows the isochrones to be compared to observational data to estimate the ages, reddenings and chemical compositions of star clusters and to test the theoretical treatment of such stellar evolutionary phenomena as convection and overshooting. Isochrones are also used in evolutionary synthesis to model the integrated light of simple stellar populations, coeval groups of stars having the same (initial) chemical composition, and accurate CT relations and BCs are needed to reliably predict the observable properties of these stellar systems.
Theoretical color-temperature relations are generally produced by convolving models of photometric filter-transmission-profiles with synthetic spectra of stars having a range of effective temperatures, surface gravities, and/or chemical compositions. Stellar atmosphere models are an integral part of this process because the synthetic spectra are either produced as part of the model atmosphere calculations themselves or come from later computations in which the model atmospheres are used. Indeed, recent tabulations of theoretical color-temperature relations, such as those of Buser & Kurucz (1992) and Bessell et al. (1998), differ mainly in the details of the model atmosphere calculations (input physics, opacities, equation of state, etc.), although there are differences in the adopted filter profiles and other aspects of the synthetic color measurements as well.
VandenBerg & Bell (1985; hereafter VB (85)) and Bell & Gustafsson (1989; hereafter BG (89)) have published color-temperature relations for cool dwarfs and cool giants, respectively, which were derived from a combination of MARCS model atmospheres (Gustafsson et al. (1975), Bell et al. (1976)) and synthetic spectra computed with the SSG spectral synthesis code (Bell & Gustafsson 1978; hereafter BG (78); Gustafsson & Bell (1979); BG (89)). However, since the work of VB (85) and BG (89), substantial improvements have been made to the MARCS and SSG computer codes, the opacity data employed by each (especially at low temperatures) and the spectral line lists used in SSG. In addition, some of the photometric filter-transmission-profiles used in the synthetic color measurements have been replaced by more recent determinations, and the calibration of the synthetic colors has been improved. Thus, as part of our evolutionary synthesis program, which will be fully described in a forthcoming paper (Houdashelt et al. 2001; hereafter Paper III ), we have calculated an improved, comprehensive set of MARCS/SSG color-temperature relations and bolometric corrections for stars cooler than 6500 K.
We present here the colors and BCs for a grid of F, G and K stars having 4000 K $``$ T<sub>eff</sub>$``$ 6500 K, 0.0 $``$ log g $``$ 4.5 and –3.0 $``$ \[Fe/H\] $``$ 0.0. We discuss the improvements made to the stellar modelling and compare our synthetic spectra of the Sun and Arcturus to spectral atlases of these stars at selected infrared wavelengths. We also describe the measurement and calibration of the synthetic colors and show the good agreement between the resulting CT relations and their empirical counterparts derived from observations of field stars. A cooler grid, representing the M giants, is presented in a companion paper (Houdashelt et al. 2000; hereafter Paper II ).
The content of this paper is structured as follows. In Section 2, the computation of the stellar atmosphere models and synthetic spectra is described, emphasizing the updated opacity data and other recent improvements in these calculations. Section 3 discusses the calibration of the synthetic colors. We first reaffirm the effective temperature scale adopted by BG (89) and then use it to derive the color calibrations required to put the synthetic colors onto the observational systems. The significance of these calibrations is illustrated through comparisons of a 4 Gyr, solar-metallicity isochrone and photometry of M67. In Section 4, we present the improved grid of color-temperature relations and compare them to observed trends and to previous MARCS/SSG results. A summary is given in Section 5.
## 2 The Models
To model a star of a given effective temperature, surface gravity and chemical composition, a MARCS stellar atmosphere is calculated and is then used in SSG to produce a synthetic spectrum. In this section, we describe these calculations more fully and how they improve upon previous MARCS/SSG work. Because we display and discuss some newly-calculated isochrones later in this paper, we start by briefly describing the derivation of these isochrones and the stellar evolutionary tracks used in their construction.
### 2.1 The Stellar Evolutionary Tracks and Isochrones
The stellar evolutionary tracks were calculated with the computer code and input physics described by Sweigart (1997) and references therein. The tracks were calibrated by matching a 4.6 Gyr, 1 M model to the known properties of the Sun. To simultaneously reproduce the solar luminosity, radius and Z/X ratio of Grevesse & Noels (1993) required Z = 0.01716, Y = 0.2798 and $`\alpha `$ = 1.8452, where X, Y and Z are the mass fractions of hydrogen, helium and metals, respectively, and $`\alpha `$ is the convective mixing-length-to-pressure-scale-height ratio. The surface pressure boundary condition was specified by assuming a scaled, solar T($`\tau `$) relation. Diffusion and convective overshooting were not included in these calculations.
The solar-metallicity isochrones shown later in this paper are taken from the set of isochrones we calculated for use in our evolutionary synthesis program. They were produced by interpolating among stellar evolutionary tracks with masses ranging from 0.2 M through 1.5 M using the method of equivalent-evolutionary points (see e.g., Bergbusch & VandenBerg (1992)). Further details of the derivation of these isochrones and evolutionary tracks will be given in a future paper presenting our evolutionary synthesis models (Paper III ).
### 2.2 The MARCS Stellar Atmosphere Models
MARCS computes a flux-constant, homogeneous, plane-parallel atmosphere assuming hydrostatic equilibrium and LTE. The continuous opacity sources used in the Maryland version of this program include H<sup>-</sup>, H I, H$`{}_{}{}^{}{}_{2}{}^{}`$, H$`{}_{}{}^{+}{}_{2}{}^{}`$, He<sup>-</sup>, C I, Mg I, Al I, Si I, Fe I, electron scattering, and Rayleigh scattering by H I and H<sub>2</sub>. In addition, for $`\lambda <`$ 7200 Å, the opacity from atomic lines, as well as that due to molecular lines of MgH, CH, OH, NH, and the violet system of CN, is included in the form of an opacity distribution function (ODF). At longer wavelengths, an ODF representing the molecular lines of CO and the red system of CN supplements the continuous opacities. The main improvements made to the MARCS opacity data since the work of BG (89) are the use of the H<sup>-</sup> free-free opacity of Bell & Berrington (1987), replacing that of Doughty & Fraser (1966); the addition of continuous opacities from the Opacity Project for Mg I, Al I and Si I and from Dragon & Mutschlecner (1980) for Fe I; and the use of detailed ODFs over the entire wavelength range from 900–7200 Å (earlier MARCS models used only schematic ODFs between 900 and 3000 Å).
For all of the MARCS models, a value of 1.6 was used for the mixing-length-to-pressure-scale-height ratio, and the y parameter, which describes the transparency of convective bubbles (Henyey et al. (1965)), was taken to be 0.076. In general, we have calculated an ODF of the appropriate metallicity to use in computing the model atmospheres, and we have adopted solar abundance ratios (but see Sections 3.2.2 and 2.4 for exceptions to these two guidelines, respectively).
### 2.3 The SSG Synthetic Spectra
Unless otherwise specified, the synthetic spectra discussed in this paper have been calculated at 0.1 Å resolution and in two pieces, optical and infrared (IR). The optical portion of the spectrum covers wavelengths from 3000–12000 Å, and the IR section extends from 1.0–5.1 $`\mu `$m (the overlap is required for calculating J-band magnitudes). In addition, the microturbulent velocity, $`\xi `$, used to calculate each synthetic spectrum has been derived from the star’s surface gravity using the field-star relation, $`\xi `$ = 2.22 – 0.322 log g (Gratton et al. 1996), and the chemical composition used in the spectral synthesis was always the same as that adopted for the corresponding MARCS model atmosphere.
The spectral computations used a version of SSG which has been continually updated since the work of BG (89). Here, the Bell & Berrington (1987) H<sup>-</sup> free-free opacity data replaced that of Bell et al. (1975), although the differences are small. The continuous opacities of Mg I, Al I, Si I and Fe I described for the MARCS models have also been incorporated into SSG, as have continuous opacities for OH and CH (Kurucz et al. (1987)).
We used an updated version of the atomic and molecular spectral line list denoted as the Bell “N” list by Bell et al. (1994; hereafter BPT (94)). This line list has been improved through further detailed comparisons of synthetic and empirical, high-resolution spectra. For $`\lambda <`$ 1.0 $`\mu `$m, it is supplemented by spectral lines of Ca, Sc, Ti, V, Cr, Mn, Fe, Co and Ni which have been culled from the compilation of Kurucz (1991) in the manner described in BPT (94). Molecular data for the vibration-rotation bands of CO were taken from Goorvitch (1994). We omit H<sub>2</sub>O lines from our calculations, but molecular lines from the $`\alpha `$, $`\beta `$, $`\gamma `$, $`\gamma ^{}`$, $`\delta `$, $`ϵ`$ and $`\varphi `$ bands of TiO have been included in all of the synthetic spectra; the latter have been given special consideration so that the observed relationship between TiO band depth and spectral type in M giants is reproduced in the synthetic spectra. A complete explanation of the sources of the TiO line data and the treatment of the TiO bands in general is given in our companion paper presenting synthetic spectra of M giants (Paper II ).
Spectral line list improvements have also been determined by comparing a synthetic spectrum of Arcturus ($`\alpha `$ Boo) to the Arcturus atlas (Hinkle et al. (1995)) and by comparing a synthetic solar spectrum to the solar atlases of Delbouille et al. (1973), its digital successors from the National Solar Observatory, and the solar atlas obtained by the ATMOS experiment aboard the space shuttle (Farmer & Norton (1989)). Identification of the unblended solar spectral lines, especially in the J and H bandpasses, was aided by the compilations of Solanki et al. (1990) and Ramsauer et al. (1995); the relevant atomic data for these lines was taken from Biémont et al. (1985a,) 1985b, 1986) . Geller (1992) has also identified many of the lines in the ATMOS spectrum, but few laboratory oscillator strengths are available for these lines. In addition, Johansson & Learner (1990) reported about 360 new Fe I lines in the infrared, identifying them as transitions between the 3d<sup>6</sup>4s(<sup>6</sup>D)4d and 3d<sup>6</sup>4s(<sup>6</sup>D)4f states. More than 200 of these lines coincide in wavelength with lines in the solar spectrum, but only 16 of them are included in the laboratory gf measurements of Fe I lines by O’Brian et al. (1991). While some of the line identifications may be in error, Johansson & Learner have checked their identifications by comparing the line intensities from the laboratory source with those in the solar spectrum. They found that only four of their lines were stronger in the solar spectrum than inferred from the laboratory data, indicating that coincidence in wavelength implies a high probability of correct identification.
Additional sources of atomic data were Nave et al. (1994) for Fe I, Litzen et al. (1993) for Ni I, Davis et al. (1978) for V I, Taklif (1990) for Mn I, Forsberg (1991) for Ti I, and Kurucz (1991). Opacity Project gf-value calculations were used for lines of Na I, Mg I, Al I, Si I, S I and Ca I. However, in view of the overall dearth of atomic data, “astrophysical” gf values have been found for many lines by fitting the synthetic and observed solar spectra.
Probably the greatest uncertainty remaining in the synthetic spectra is the “missing ultraviolet (UV) opacity problem,” which has been known to exist for some time (see e.g., Gustafsson & Bell (1979)). Holweger (1970) speculated that this missing opacity could be caused by a forest of weak Fe lines in the UV, and in fact, Buser & Kurucz (1992) have claimed to have “solved” the problem through the use of a new, larger spectral line list. However, this claim rests solely on the fact that the models calculated using this new list produce better agreement between the synthetic and empirical color-color relations of field stars. While this improvement is evident and the UV opacity has definitely been enhanced by the new line list, it is also clear that the missing UV opacity has not been “found.” BPT (94) have shown that many of the spectral lines which appear in the new list are either undetected or are much weaker in the observed solar spectrum than they are in the solar spectrum synthesized from this line list.
Balachandran & Bell (1998) have recently examined the solar abundance of Be, which has been claimed to be depleted with respect to the meteoritic abundance. They argue that the OH lines of the A–X system, which appear in the same spectral region as the Be II lines near 3130 Å, should yield oxygen abundances for the Sun which match those derived from the vibration-rotation lines of OH in the near-infrared. To produce such agreement requires an increase in the continuous opacity in the UV corresponding, for example, to a thirty-fold increase in the bound-free opacity of Fe I. Such an opacity enhancement not only brings the solar Be abundance in line with the meteoritic value, but it also improves the agreement of the model fluxes with the limb-darkening behavior of the Be II lines and the solar fluxes measured by the Solstice experiment (Woods et al. (1996)). However, this is such a large opacity discrepancy that we have not included it in the models presented here. Thus, we expect our synthetic U–V (and possibly B–V) colors to show the effects of insufficient UV opacity. For this reason, we recommend that the U–V colors presented in this paper be used with caution.
After this paper was written, it was found that recent Fe I photoionization cross-sections calculated by Bautista (1997) are much larger than those of Dragon & Mutschlecner (1980), which we have used in our models. In the region of the spectrum encompassing the aforementioned Be II lines ($``$3130 Å), these new cross-sections cause a reduction of 15% in the solar continuous flux; this represents about half of the missing UV opacity at these wavelengths. Further work is being carried out using Bautista’s opacity data, with the expectation of improving both our model atmospheres and synthetic spectra.
### 2.4 Mixing
The observed abundance patterns in low-mass red giants having approximately solar metallicity indicate that these stars can mix CNO-processed material from the deep interior outward into the stellar atmosphere. This mixing is in addition to that which accompanies the first dredge-up at the beginning of the red-giant-branch (RGB) phase of evolution. We have included the effects of such mixing in both the stellar atmosphere models and the synthetic spectra of the brighter red giants by assuming \[C/Fe\] = –0.2, \[N/Fe\] = +0.4 and <sup>12</sup>C/<sup>13</sup>C = 14 (the “unmixed” value is 89) for these stars. These quantities are deduced from abundance determinations in G8–K3 field giants (Kjærgaard et al. (1982)) and in field M giants (Smith & Lambert (1990)), as well as from carbon isotopic abundance ratios in open cluster stars (e.g., Gilroy & Brown (1991) for M67 members).
In our evolutionary synthesis program, we incorporate mixing for all stars more evolved than the “bump” in the RGB luminosity function which occurs when the hydrogen-burning shell, moving outward in mass, encounters the chemical composition discontinuity produced by the deep inward penetration of the convective envelope during the preceding first-dredge-up. Sweigart & Mengel (1979) have argued that, prior to this point, mixing would be inhibited by the mean molecular weight barrier caused by this composition discontinuity. The M67 observations mentioned above and the work of Charbonnel (1994,) 1995) and Charbonnel et al. (1998) also indicate that this extra mixing first appears near the RGB bump. For simplicity, we assume that the change in composition due to mixing occurs instantaneously after a star has evolved to this point.
When modelling a group of stars of known age and metallicity, it is straightforward to determine where the RGB bump occurs along the appropriate isochrone and then to incorporate mixing in the models of the stars more evolved than the bump. However, for a field star of unknown age, this distinction is not nearly as clear because the position of the RGB bump in the HR diagram is a function of both age and metallicity. To determine which of our field-star and grid models should include mixing effects, we have used our solar-metallicity isochrones as a guide.
The RGB bump occurs near log g = 2.38 on our 4 Gyr, solar-metallicity isochrone and at about log g = 2.46 on the corresponding 16 Gyr isochrone. Based upon these gravities, we only include mixing effects in our field-star and grid models having log g $``$ 2.4. Using this gravity threshold should be reasonable when modelling the field stars used to calibrate the synthetic colors (see Section 3.2) because the majority of these stars have approximately solar metallicities. However, this limit may not be appropriate for all of our color-temperature grid models (see Section 4). Our isochrones show that the surface gravity at which the RGB bump occurs decreases with decreasing metallicity at a given age. Linearly extrapolating from these isochrones, which to-date only encompass metallicities greater than about –0.5 dex in \[Fe/H\], it appears that perhaps the models having log g = 2.0 and \[Fe/H\] $``$ –2.0 should have included the effects of mixing as well.
Overall, including mixing in our models only marginally affects the resulting broad-band photometry (see Section 3.3) but has a more noticeable influence on the synthetic spectra and some narrow-band colors.
### 2.5 Spectra of the Sun and Arcturus
To judge the quality of the synthetic spectra on which our colors are based, we present some comparisons of observed and synthetic spectra of the Sun and Arcturus in the near-infrared. At optical wavelengths, the agreement between our synthetic solar spectrum and the observed spectrum of the Sun is similar to that presented by BPT (94), Briley et al. (1994) and Bell & Tripicco (1995). Refinements of the near-infrared line lists are much more recent, and some examples of near-IR fits can be found in Bell (1997) and in Figures 1 and 2, which compare our synthetic spectra of the Sun (5780 K, log g = 4.44) and Arcturus (4350 K, log g = 1.50, \[Fe/H\] = –0.51, \[C/H\] = –0.67, \[N/H\] = –0.44, \[O/H\] = –0.25, <sup>12</sup>C/<sup>13</sup>C = 7) to data taken from spectral atlases of these stars. In each figure, the synthetic spectra are shown as dotted lines, and the observational data are represented by solid lines; the former have been calculated at 0.01 Å resolution and convolved to the resolution of the empirical spectra.
The upper panel of Figure 1 shows our synthetic solar spectrum and the ATMOS spectrum of the Sun (Farmer & Norton (1989)) near the bandhead of the <sup>12</sup>CO(4,2) band. The lower panel of this figure is a similar plot for Arcturus, but the observed spectrum is taken from NOAO (Hinkle et al. (1995)). In Figure 2, NOAO spectra of the Sun (Livingston & Wallace (1991)) and Arcturus (Hinkle et al. (1995)) are compared to the corresponding synthetic spectra in a region of the H band. The NOAO data shown in these two figures are ground-based and have been corrected for absorption by the Earth’s atmosphere. To allow the reader to judge its importance, the telluric absorption at these wavelengths is also displayed, in emission and to half-scale, across the bottoms of the appropriate panels; these telluric spectra have been taken from the same sources as the observational data. The slight wavelength difference between the empirical and synthetic spectra in the lower panel of Figure 1 occurs because the observed spectrum is calibrated using vacuum wavelengths, while the spectral line lists used to calculate the synthetic spectra assume wavelengths in air; this offset has been removed in the lower panel of Figure 2.
Note that the oxygen abundance derived for the Sun from the near-infrared vibration-rotation lines of OH is dependent upon the model atmosphere adopted. The Holweger & Müller (1974) model gives a result 0.16 dex higher than the OSMARCS model of Edvardsson et al. (1993). We use a logarithmic, solar oxygen abundance of 8.87 dex (on a scale where H = 12.0 dex), only 0.04 dex smaller than the value inferred from the Holweger & Müller model. Presumably, this is part of the reason that our CO lines are slightly too strong in our synthetic solar spectrum while being about right in our synthetic spectrum of Arcturus (see Figure 1). Overall, though, the agreement between the empirical and synthetic spectra illustrated in Figures 1 and 2 is typical of that achieved throughout the J, H and K bandpasses.
## 3 The Calculation and Calibration of the Synthetic Colors
Colors are measured from the synthetic spectra by convolving the spectra with filter-transmission-profiles. To put the synthetic colors onto the respective photometric systems, we then apply zero-point corrections based upon models of the A-type standard star, $`\alpha `$ Lyrae (Vega), since the colors of Vega are well-defined in most of these systems. In this paper, we measure synthetic colors on the following systems using the filter-transmission-profiles taken from the cited references – Johnson UBV and Cousins VRI: Bessell (1990); Johnson-Glass JHK: Bessell & Brett (1988); CIT/CTIO JHK: Persson (1980); and CIT/CTIO CO: Frogel et al. (1978). Gaussian profiles were assumed for the CO filters. In Figure 3, these filter-transmission-profiles overlay our synthetic spectrum of Arcturus.
The synthetic UBVRI colors are initially put onto the observed Johnson-Cousins system using the Hayes (1985) absolute-flux-calibration data for Vega. Since Hayes’ observations begin at 3300 Å, they must be supplemented by the fluxes of a Vega model (9650 K, log g = 3.90, \[Fe/H\] = 0.0; Dreiling & Bell (1980)) shortward of this. The colors calculated from the Hayes data are forced to match the observed colors of Vega by the choice of appropriate zero points; we assume U–B = B–V = 0.0 for Vega and take V–R = –0.009 and V–I = –0.005, as observed by Bessell (1983). The resulting zero points are then applied to all of the synthetic UBVRI colors. While this results in UBVRI colors for our Vega model which are slightly different than those observed, we prefer to tie the system to the Hayes calibration, which is fixed, rather than to the Vega model itself because the latter will change as our modelling improves. The use of our model Vega fluxes shortward of 3300 Å introduces only a small uncertainty in the U–B zero points.
Since there is no absolute flux calibration of Vega as a function of wavelength in the infrared, the VJHK and CO colors are put onto the synthetic system by applying zero point corrections which force the near-infrared colors of our synthetic spectrum of Vega to all become 0.0. This is consistent with the way in which the CIT/CTIO and Johnson-Glass systems are defined by Elias et al. (1982) and Bessell & Brett (1988), respectively.
### 3.1 Testing the Synthetic Colors
As a test of our new isochrones and synthetic color calculations, we decided to compare our 4 Gyr, solar-metallicity isochrone to color-magnitude diagrams (CMDs) of the Galactic open cluster M67. We chose this cluster because it has been extensively observed and therefore has a very well-determined age and metallicity. Recent metallicity estimates for M67 (Janes & Smith (1984), Nissen et al. (1987), Hobbs & Thorburn (1991), Montgomery et al. (1993), Fan et al. (1996)) are all solar or slightly subsolar, and the best age estimates of the cluster each lie in the range of 3–5 Gyr, clustering near 4 Gyr (Nissen et al. (1987), Hobbs & Thorburn (1991), Demarque et al. (1992), Montgomery et al. (1993), Meynet et al. (1993), Dinescu et al. (1995), Fan et al. (1996)). Thus, we expect this particular isochrone to be a good match to all of the CMDs of M67.
We translated the theoretical isochrone to the color-magnitude plane by calculating synthetic spectra for effective temperature/surface gravity combinations lying along it and then measuring synthetic colors from these spectra. Absolute V-band and K-band magnitudes were derived assuming M<sub>V,⊙</sub> = +4.84 and BC<sub>V,⊙</sub> = –0.12. After doing this, we found that our isochrone and the M67 photometry differed systematically in some colors (see Section 3.2.3), which led us to more closely examine the synthetic color calculations. By modelling field stars with relatively well-determined physical properties, we learned that, after applying the Vega zero-point corrections, some of the synthetic colors were still not on the photometric systems of the observers. Using these field star models, we have derived the linear color calibrations which are needed to put the synthetic colors onto the observational systems. In the following sections, we describe the derivation of these color calibrations and show how the application of these relations removes most of the disagreement between the 4 Gyr isochrone and the M67 colors, especially near the main-sequence turnoff.
Note, however, that the location of the isochrone in the HR diagram does depend upon the details of the stellar interior calculations, such as the surface boundary conditions and the treatment of convection, particularly at cooler temperatures. Near the main-sequence turnoff, its detailed morphology could also be sensitive to convective overshooting (see Demarque et al. (1992) and Nordström et al. (1997)), which we have omitted. The primary purpose in making comparisons between this isochrone and the M67 photometry is to test the validity of using this and older isochrones to model the integrated light of galaxies.
### 3.2 The Color Calibrations
It should not be too surprising, perhaps, that the colors measured directly from the synthetic spectra are not always on the systems defined by the observational data. Bessell et al. (1998) present an excellent discussion of this topic in Appendix E of their paper. They conclude that, “As the standard systems have been established from natural system colors using linear and non-linear corrections of at least a few percent, we should not be reluctant to consider similar corrections to synthetic photometry to achieve good agreement with the standard system across the whole temperature range of the models.” Thus, we will not explore why the synthetic colors need to be calibrated to put them onto the observational systems but instead will focus upon the best way to derive the proper calibrations.
Paltoglou & Bell (1991), BPT (94) and Tripicco & Bell (1991) have also looked at the calibration of synthetic colors. They determined the relations necessary to make the colors measured from the Gunn & Stryker (1983; hereafter GS (83)) spectral scans match the observed photometry of the GS (83) stars. This approach has the advantage that it is model-independent, i.e., it is not tied in any way to the actual spectral synthesis, and should therefore primarily be sensitive to errors in the filter-transmission-profiles. Of course, it is directly dependent upon the accuracy of the flux calibration of the GS (83) scans.
The GS (83) scans extend from 3130–10680 Å and are comprised of blue and red sections; these pieces have 20 Å and 40 Å resolution, respectively, and are joined at about 5750 Å. We have made comparisons of the GS (83) scans of stars of similar spectral type and found the relative flux levels of the blue portions to be very consistent from star to star; however, there appear to be systematic differences, often quite large, between the corresponding red sections. This “wiggle” in the red part of the GS (83) scans usually originates near 5750 Å, the point at which the blue and red parts are merged. Because of these discrepancies and other suspected problems with the flux calibration of the GS (83) scans (see e.g., Rufener & Nicolet (1988), Taylor & Joner (1990), Worthey (1994)), we have chosen instead to calibrate the synthetic colors by calculating synthetic spectra of a group of stars which have well-determined physical properties and comparing the resulting synthetic colors to photometry of these stars. The drawback of this method is that the calibrations are now model-dependent. However, since we have decoupled the color calibrations and the GS (83) spectra, this allows us to calibrate both the optical and infrared colors in a consistent manner. In addition, we use more recent determinations of the photometric filter-transmission-profiles than the previous calibrations cited above.
#### 3.2.1 Effective Temperature Measurements and the Field Star Sample
To calibrate the synthetic colors, we have chosen a set of field stars which have well-determined effective temperatures, since T<sub>eff</sub> is the most critical determinant of stellar color. Most stars in this set have surface gravity and metallicity measurements as well, although we have not always checked that the log g and \[Fe/H\] values were found using effective temperatures similar to those which we adopt.
The most straightforward way to derive the effective temperature of a star is through measurement of its angular diameter and apparent bolometric flux. The effective temperature can then be estimated from the angular diameter through the relation,
$$\mathrm{T}_{\mathrm{eff}}\left(\frac{\mathrm{f}_{\mathrm{bol}}}{\varphi ^2}\right)^{0.25},$$
(1)
where $`\varphi `$ is the limb-darkened angular diameter of the star, hereafter denoted $`\varphi _{\mathrm{LD}}`$, and f<sub>bol</sub> is its apparent bolometric flux. Obviously, this method can only be used for nearby stars.
The infrared flux method (IRFM) of Blackwell & Shallis (1977) is regarded as one of the more reliable ways to estimate T<sub>eff</sub> for stars which are not near enough to have their angular diameters measured. It relies on the temperature sensitivity of the ratio of the apparent bolometric flux of a star to the apparent flux in an infrared bandpass (usually the K band). Model atmosphere calculations are employed to predict the behavior of this flux ratio with changing T<sub>eff</sub>, and the resulting calibration is then used to infer effective temperatures from observed fluxes. IRFM temperatures have been published by Blackwell et al. (1990), Blackwell & Lynas-Gray (1994), Saxner & Hammarb$`\ddot{\mathrm{a}}`$ck (1985; hereafter SH (85)), Alonso et al. (1996), and BG (89), among others, although BG (89) used their color calculations to adopt final temperatures which are systematically 80 K cooler than their IRFM estimates. BG (89) also calculated angular diameters for the stars in their sample from the apparent bolometric fluxes that they used in the IRFM and their “adopted” effective temperatures. We can check the accuracy of these BG (89) T<sub>eff</sub> values by comparing their angular diameter predictions to recent measurements.
Pauls et al. (1997; hereafter PMAHH ) have used the U.S. Navy Prototype Optical Interferometer to make observations in 20 spectral channels between 5200 Å and 8500 Å and have measured $`\varphi _{\mathrm{LD}}`$ for two of the stars with IRFM temperatures quoted by BG (89). In addition, Mozurkewich et al. (1991) and Mozurkewich (1997) have presented uniform-disk angular diameters ($`\varphi _{\mathrm{UD}}`$), measured with the Mark III Interferometer at 8000 Å, for 20 of BG (89)’s stars, including the two observed by PMAHH ; these data, which we will hereafter refer to collectively as M (97), must be converted to limb-darkened diameters before comparing them to the other estimates.
To determine the uniform-disk-to-limb-darkened conversion factor, we performed a linear, least-squares fit between the ratio $`\varphi _{\mathrm{LD}}`$/$`\varphi _{\mathrm{UD}}`$ and spectral type for the giant stars listed in Table 3 of Mozurkewich et al. (1991), using their angular diameters measured at 800 nm. In deriving this correction, $`\delta `$ And was omitted; $`\varphi _{\mathrm{LD}}`$ for $`\delta `$ And (4.12 mas) does not follow the trend of the other giants, possibly due to a typographical error (4.21 mas would fit the trend). We applied the resulting limb-darkening correction:
$$\varphi _{\mathrm{LD}}/\varphi _{\mathrm{UD}}=1.078+0.002139\times \mathrm{SP},$$
(2)
where SP is the spectral type of the star in terms of its M subclass (i.e., SP = 0 for an M0 star, SP = –1 for a K5 star, etc.), to the M (97) angular diameters to derive $`\varphi _{\mathrm{LD}}`$ for these stars, keeping in mind that the predicted correction factor may not be appropriate for all luminosity classes; for example, the above equation gives $`\varphi _{\mathrm{LD}}`$/$`\varphi _{\mathrm{UD}}`$ = 1.033 for $`\alpha `$ CMi, an F5 IV–V star, while Mozurkewich et al. (1991) used 1.047.
The limb-darkened angular diameters of M (97) and PMAHH are compared to the BG (89) estimates in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars; the M (97) and BG (89) comparison is also illustrated in Figure 4. Note that the PMAHH value for $`\alpha `$ UMa is in much better agreement with the BG (89) estimate than is the M (97) diameter. The average absolute and percentage differences between the BG (89) and M (97) angular diameters, in the sense BG (89) – M (97), are only –0.006 mas and 0.87%, respectively. Translated into effective temperatures, these differences become –23 K and 0.44%. For the PMAHH data, the analogous numbers are +0.03 mas (0.51%) and +13 K (0.30%), albeit for only two stars. This exceptional agreement between the IRFM-derived and observed angular diameters is a strong confirmation of the “adopted” effective temperatures of BG (89) and gives us faith that any errors in the synthetic colors which we measure for the stars taken from this source are not dominated by uncertainties in their effective temperatures.
Since we are confident that the BG (89) “adopted” T<sub>eff</sub> scale for giant stars is accurate, we have used a selection of their stars as the basis of the synthetic color calibrations. Specifically, we chose all of the stars for which they derived IRFM temperatures. However, BG (89) included only G and K stars in their work. For this reason, we have supplemented the BG (89) sample with the F and G dwarfs studied by SH (85); before using the latter data, of course, we must determine whether the IRFM temperatures of SH (85) are consistent with those of BG (89).
There is only one star, HR 4785, for which both BG (89) and SH (85) estimated T<sub>eff</sub>. The IRFM temperature of SH (85) is 5842 K, while BG (89) adopt 5861 K. In addition, M (97) has measured an angular diameter for one of the SH (85) stars, $`\alpha `$ CMi (HR 2943). Applying the limb-darkening correction used by Mozurkewich et al. (1991) for this star and using the bolometric flux estimated by SH (85) gives T<sub>eff</sub> = 6569 K; SH (85) derive 6601 K. Based upon these two stars, it appears that the SH (85) effective temperature scale agrees with the BG (89) scale to within $``$20–30 K, which is well within the expected uncertainties in the IRFM effective temperatures. Thus, we conclude that the two sets of measurements are in agreement and adopt the IRFM temperatures of SH (85) as given.
Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars lists the stars for which we have calculated synthetic spectra for use in calibrating the synthetic color measurements and gives the effective temperatures, surface gravities and metallicities used to model each. From BG (89), we used the “adopted” temperatures and metallicities given in their Table 3. In most cases, we used the surface gravities and metallicities from this table as well. However, for seven of the stars, we took the more recent log g determinations of Bonnell & Bell (1993a), 1993b) . For the SH (85) stars, we assumed the average effective temperatures presented in their Table 7, omitting stars HR 1008 and HR 2085, which may be peculiar (see SH (85)); the surface gravities and metallicities of these stars were taken from Table 5 of SH (85). We adopted the BG (89) parameters for HR 4785. For those stars with unknown surface gravities, log g was estimated from our new isochrones, taking into account the metallicity and luminosity class of the star, or from other stars of similar effective temperature and luminosity class. Because most of the stars with well-determined chemical compositions have \[Fe/H\] $``$ 0.0, solar abundances were assumed for those stars for which \[Fe/H\] had not been measured.
In Figure 5, we show where the stars used to calibrate the synthetic colors lie in the log T<sub>eff</sub>, log g plane. Our 3 Gyr and 16 Gyr, solar-metallicity isochrones are also shown in this figure for reference. Figure 5a shows the entire sample of stars from Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars; the other panels present interesting subsets of the larger group. In panels b) and c), the metal-poor (\[Fe/H\] $`<`$ –0.3) and metal-rich (\[Fe/H\] $`>`$ +0.2) stars are shown, respectively. In panels d), e) and f), the solar metallicity (–0.3 $``$ \[Fe/H\] $``$ +0.2) stars are broken up into the dwarfs and subgiants (luminosity classes III–IV, IV, IV–V and V), the normal giants (classes II–III and III), and the brighter giants (classes I and II), respectively.
Even though many of the field stars in Figure 5a do not fit the solar-metallicity isochrones as well as one might expect, especially along the red-giant branch, examining the subsets of these stars shows that the “outliers” are mostly brighter giants and/or metal-poor stars; at a given surface gravity, these stars generally lie at higher temperatures than the solar-metallicity isochrones, as would be expected from stellar-evolution theory and observational data. The stars located near log T<sub>eff</sub> = 3.7, log g = 2.7 in panel e) of Figure 5 (the solar-metallicity giants) are probably clump stars. In fact, there are only three stars in Figure 5 which lie very far from the positions expected from their physical properties or spectral classification, and even their locations are not unreasonable. The three stars in question are represented by filled symbols and are the metal-poor star, $`\gamma `$ Tuc; the metal-rich giant, 72 Cyg; and the solar-metallicity giant, 31 Com. $`\gamma `$ Tuc is unusual in that SIMBAD assigns it the spectral type F1 III, yet it lies very near the turnoff of the 3 Gyr, Z isochrone. The Michigan Spectral Survey (Houk & Cowley (1975)) classifies this star as an F3 IV–V star, and SH (85) listed it as an F-type dwarf, which appears to be appropriate for its temperature and surface gravity, indicating that the SIMBAD classification is probably incorrect. 72 Cyg is suspect only because it is hotter than the solar-metallicity isochrones while having a supersolar metallicity. Since it could simply be younger (i.e., more massive) than the other stars in the sample, we do not feel justified in altering the parameters adopted for this star. The third star, 31 Com, lies further from the isochrones than any other star, but it is classified as a G0 III peculiar star, so perhaps its log g and T<sub>eff</sub> are not unreasonable either.
Overall, Figure 5 indicates that the temperatures and gravities assigned to the stars in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars are good approximations to those one would infer from their spectral types and metallicities.
#### 3.2.2 Derivation of the Color Calibrations
We have calculated a MARCS stellar atmosphere model and an SSG synthetic spectrum for each of the stars in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars. However, rather than compute a separate ODF for each star, we chose to employ only three ODFs in these model atmosphere calculations – the same ODFs used in our evolutionary synthesis work (Paper III ). For stars with \[Fe/H\] $`<`$ –0.3, a Z = 0.006 (\[Fe/H\] = –0.46) ODF was used; stars having –0.3 $``$ \[Fe/H\] $``$ +0.2 were assigned a solar-metallicity (Z = 0.01716) ODF; and a Z = 0.03 (\[Fe/H\] = +0.24) ODF was used for stars with \[Fe/H\] $`>`$ +0.2. The synthetic spectra were then calculated using the metallicities of the corresponding ODFs.
To calibrate the synthetic colors, photometry of the field stars in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars has been compiled from the literature. Since all of these stars are relatively nearby, no reddening corrections have been applied to this photometry. The UBV photometry comes from Mermilliod (1991) and Johnson et al. (1966). The VRI data is from Cousins (1980) and Johnson et al. (1966); the latter were transformed to the Cousins system using the color transformations of Bessell (1983). The VJHK photometry has been derived from colors observed on the Johnson system by Johnson et al. (1966), Johnson et al. (1968), Lee (1970) and Engels et al. (1981) and on the SAAO system by Glass (1974); the color transformations given by Bessell & Brett (1988) were used to calculate the observed Johnson-Glass and CIT/CTIO colors. Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars cites the specific sources of the photometry used for each field star. Unfortunately, we were unable to locate CO observations for a sufficient number of these stars to be able to calibrate this index.
Although a few colors show a hint that a higher-order polynomial may be superior, we have fit simple linear, least-squares relations to the field-star data, using the photometric color as the dependent variable and the synthetic color as the independent variable in each case and omitting the three coolest dwarfs. Figure 6 shows some of these color-calibration fits (solid lines), and the zero points and slopes of all of the resulting relations are given in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars; the $`\sigma `$ values listed there are the 1$`\sigma `$ uncertainties in the coefficients of the least-squares fits. The number of stars used to derive each calibration equation, n, and the colors spanned by the photometric data are also given in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars.
For the optical colors, the coolest dwarfs appear to follow a different color calibration than the other field stars, so we have also calculated a separate set of color calibrations for cool dwarfs. This was done by fitting a linear relation to the colors of the three reddest field dwarfs but forcing this fit to intersect the “main” calibration relation of Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars at the color corresponding to a 5000 K dwarf. These cool-dwarf color calibrations are shown as dashed lines in Figure 6. However, we wish to emphasize that the cool-dwarf color calibrations are very uncertain, being derived from photometry of only three stars (HR 8085, HR 8086 and HR 8832); the choice of 5000 K as the temperature at which the cool dwarf colors diverge is also highly subjective. Thus, since it is not clear to us whether the use of different optical color calibrations for cool dwarfs is warranted, we will proceed by discussing and illustrating the calibrated colors of the cool dwarf models which result when both the color calibrations of Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars and the separate cool-dwarf calibrations are adopted. In Section 4.1.1, we further discuss the coolest dwarfs in our sample and possible explanations for their erroneous colors.
Due to the small color range of the H–K photometry and the uncertainties in the Johnson/SAAO H–K photometry and color transformations, the calibrations derived for the H–K colors are highly uncertain; application of the corresponding calibrations may not improve the agreement with the observational data significantly. It is also apparent that our U–V calculations are not very good, but we expect them to improve substantially after the aforementioned Fe I data of Bautista (1997) have been incorporated into both MARCS and SSG. Additionally, we want to emphasize that the mathematical form of the color calibrations is presented here only to illustrate their significance. We caution others against the use of these specific relations to calibrate synthetic colors in general because the coefficients are dependent, at least in part, on our models.
The poor agreement between the synthetic and observed CIT/CTIO J–K colors, contrasted with the relatively good fit for Johnson-Glass J–K, is probably due to better knowledge of the J-band filter-transmission-profile of the latter system. This is suggested by the ratio of the scale factors (slopes) in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars (JG/CIT = 0.976/0.895 = 1.090), which is very close to the slope of the color transformation (1.086) between the two systems found by Bessell & Brett (1988). In other words, our synthetic colors for the two systems are very similar, while the Bessell & Brett observational transformation indicates that they should differ.
Even though most of our color calibrations call for only small corrections to the synthetic colors, we have chosen to apply each of them because every relation has either a slope or a zero point which is significant at greater than the 1$`\sigma `$ level. We believe that using these calibration equations will help us to reduce the uncertainties in the integrated galaxy colors predicted by our evolutionary synthesis models (Paper III ). Indeed, the importance of calibrating the synthetic colors when modelling the integrated light of galaxies and other stellar aggregates becomes apparent when the uncalibrated and calibrated colors of our 4 Gyr, solar-metallicity isochrone are compared to color-magnitude diagrams of M67.
#### 3.2.3 The M67 Color-Magnitude Diagrams
The (uncalibrated) colors of our 4 Gyr, solar-metallicity isochrone were determined by calculating synthetic spectra for effective temperature/surface gravity combinations lying along it, measuring synthetic colors from these spectra and then applying the Vega-based zero-point corrections to these colors. In Figure 7, we show the effects of applying the color calibrations to the synthetic colors of the 4 Gyr, Z isochrone. In each panel of this figure, the uncalibrated isochrone is shown as a dotted line, while the calibrated isochrones are represented by a solid line (Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars color calibrations throughout) and a dashed line (Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars relations + cool-dwarf calibrations for optical colors of main-sequence stars cooler than 5000 K), respectively. The M67 UBVRI photometry of Montgomery et al. (1993) is shown as small crosses in Figure 7, and the giant-star data tabulated by Houdashelt et al. (1992) are represented by open circles.
We have corrected the M67 photometry for reddening using E(B–V) = 0.06 and the reddening ratios of a K star as prescribed by Cohen et al. (1981); we also assumed E(U–V)/E(B–V) = 1.71. This reddening value is on the upper end of the range of recent estimates for M67 (Janes & Smith (1984), Nissen et al. (1987), Hobbs & Thorburn (1991), Montgomery et al. (1993), Meynet et al. (1993), Fan et al. (1996)), but we found that a reddening this large produced the best overall agreement between the 4 Gyr, solar-metallicity isochrone and the M67 photometry in the region of the main-sequence turnoff for the complete set of CMDs which we examined. This reddening leaves the isochrone turnoff a bit bluer than the midpoint of the turnoff-star color distribution in V–I and a bit redder than the analogous M67 photometry in U–V, but it gives a reasonable fit for the other colors. Because the turnoff stars in the M<sub>V</sub>, U–V CMD suggest a smaller reddening is more appropriate, we also examined the possibility that the Montgomery et al. (1993) V–I colors are systematically in error; however, they are consistent with the photometry of M67 reported by Chevalier & Ilovaisky (1991) and by Joner & Taylor (1990). Most importantly, we simply trust our synthetic V–I colors (and the corresponding reddening ratios) more than those in U–V, given the aforementioned uncertainties in the bound-free Fe I opacity, so we accepted the poorer turnoff fit in the U–V CMD. We derived the absolute magnitudes of the M67 stars, after correcting for extinction, assuming (m–M)<sub>0</sub> = 9.60 (Nissen et al. (1987), Montgomery et al. (1993), Meynet et al. (1993), Dinescu et al. (1995), Fan et al. (1996)).
The agreement between the isochrone and the M67 observations is improved in all of the CMDs (except perhaps in the H–K diagram, which is not shown) after the isochrone colors have been calibrated to the photometric systems using the color calibrations derived in Section 3.2.2. In fact, if we concentrate upon the main-sequence-turnoff region, there is no further indication of problems with the synthetic colors, with the possible exception of U–V, which again is sensitive to opacity uncertainties. The level of agreement between the photometry and the isochrone along the red-giant branch strengthens this conclusion. Of course, more R-band and deeper near-infrared photometry of M67 members would be helpful in verifying this result for the respective colors.
We also note that the detailed fit of the isochrone to the M67 turnoff could possibly be improved by including convective overshooting in our stellar interior models. While there is no general agreement regarding the importance of convective overshooting in M67 (see Demarque et al. (1992), Meynet et al. (1993), and Dinescu et al. (1995) for competing views), solar-metallicity turnoff stars only slightly more massive than those in M67 do show evidence for significant convective overshooting (Nordström et al. (1997), Rosvick & VandenBerg (1998)). To help us examine this further, VandenBerg (1999) has provided us with a 4 Gyr, solar-metallicity isochrone which includes the (small) amount of convective overshooting which he considers to be appropriate for M67. This isochrone is in excellent agreement with ours along the lower main-sequence but contains a “hook” feature at the main-sequence turnoff, becoming cooler than our 4 Gyr, Z isochrone at M<sub>V</sub>=3.5 and then hooking back to be hotter than ours at M<sub>V</sub>=3.0; this “hook” does appear to fit the M67 photometry slightly better. Still, we do not expect convective overshooting to significantly affect our evolutionary synthesis models of elliptical galaxies because it will have an even smaller impact upon the isochrones older than 4 Gyr.
In those colors for which the deepest photometry is available, the calibrated isochrone is still bluer than the M67 stars on the lower main-sequence when only the color calibrations of Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars are applied. Using separate calibrations for the cool dwarfs makes the faint part of the calibrated isochrone agree much better with the U–V and B–V colors of the faintest M67 stars seen in Figure 7 but appears to overcorrect their V–I colors. This improved fit between the calibrated isochrone and the M67 photometry supports the use of different color calibrations for the optical colors of the cool dwarfs, although there is a hint in Figure 7 that perhaps the cool-dwarf calibrations should apply at an effective temperature slightly hotter than 5000 K. Nevertheless, we do not find lower main-sequence discrepancies as worrisome as color differences near the turnoff would be, since the lowest-mass stellar interior models are sensitive to the assumed low-temperature opacities, equation of state and surface pressure boundary treatment. In addition, these faint, lower-main-sequence stars make only a small contribution to the integrated light of our evolutionary synthesis models when a Salpeter IMF is assumed. Consequently, small errors in their colors will not generally have a detectable effect on the integrated light of the stellar population represented by the entire isochrone.
### 3.3 Uncertainties in the Synthetic Colors
While the color calibrations are technically appropriate only for stars with near-solar metallicities, there is no evidence from the photometry of the field star sample that the synthetic color calibrations depend upon chemical composition. In fact, it is likely that we are at least partially accounting for metallicity effects by applying color corrections as a function of color rather than effective temperature. If the differences between the uncalibrated, synthetic colors and the photometry are due to errors in the synthetic spectra caused by missing opacity, for example, then different color corrections would be expected for two stars having the same temperature but different metallicities; the more metal-poor star should require a smaller correction. This is in qualitative agreement with the calibrations derived here, since this star would be bluer than its more metal-rich counterpart.
In this section, we present the sensitivities of the colors to uncertainties in some of the model parameters: effective temperature, surface gravity, metallicity, microturbulent velocity ($`\xi `$) and mixing. We estimate the uncertainties in these stellar properties to be $`\pm `$80 K in T<sub>eff</sub>, $`\pm `$0.3 dex in log g, $`\pm `$0.25 dex in \[Fe/H\], and $`\pm `$0.25 km s<sup>-1</sup> in $`\xi `$. To examine the implications of these uncertainties, we have varied the model parameters of the coolest dwarf (61 Cyg B), the coolest giant ($`\alpha `$ Tau) and the hottest solar-metallicity dwarf (HR 4102) listed in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars by the estimated uncertainties, constructing new model atmospheres and synthetic spectra of these stars. We also constructed an $`\alpha `$ Tau model which neglected mixing. In Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars, we present the changes in the uncalibrated synthetic colors which result when each of the model parameters is varied as described above. The color changes presented in this table are the average changes produced by parameter variations in the positive and negative directions. Since the VJHK color changes were almost always identical for colors on the Johnson-Glass and CIT/CTIO systems, the $`\mathrm{\Delta }`$(V–K), $`\mathrm{\Delta }`$(J–K) and $`\mathrm{\Delta }`$(H–K) values given in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars are applicable to either system.
If our uncertainty estimates are realistic, then it is clear from Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars that the variations in most colors are dominated by the uncertainties in one parameter. For the V–R, V–I, V–K, J–K and H–K colors, this parameter is T<sub>eff</sub>, with the exception being the V–R color of the cool dwarf, which appears to be especially metallicity-sensitive. The behavior of the U–V and B–V colors is more complicated. Gravity and metallicity uncertainties dominate U–V for the cool giant and the hot dwarf, but effective temperature uncertainties have the greatest influence on the U–V color of the cool dwarf. Uncertainties in each of the model parameters have a similar significance in determining the B–V color of the cool giant, while those in T<sub>eff</sub> dominate for the cool dwarf, and T<sub>eff</sub> and \[Fe/H\] uncertainties are most important in the hot dwarf. Overall, we conclude that estimating the formal uncertainties of the synthetic colors needs to be done on a star-by-star and color-by-color basis, which we have not attempted to do. However, it is encouraging to see that T<sub>eff</sub> uncertainties dominate so many of the color determinations, since our effective temperature scale is well-established by the angular diameter measurements discussed in Section 3.2.1.
## 4 The New Color-Temperature Relations and Bolometric Corrections
From the comparisons of the calibrated, 4 Gyr, solar-metallicity isochrone and the CMDs of M67, we conclude that the color calibrations that were derived in Section 3.2.2 and presented in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars generally put the synthetic colors onto the photometric systems of the observers but leave some of the optical colors of cool dwarfs too blue. These calibrations, coupled with the previously described improvements in the model atmospheres and synthetic spectra, have encouraged us to calculate a new grid of color-temperature relations and bolometric corrections. The bolometric corrections are calculated after calibrating the model colors, assuming BC<sub>V,⊙</sub> = –0.12 and M<sub>V,⊙</sub> = +4.84; when coupled with the calibrated color of our solar model, (V–K)<sub>CIT</sub> = 1.530, we derive M<sub>K,⊙</sub> = +3.31 and BC<sub>K,⊙</sub> = +1.41. However, keep in mind that the color calibrations have been derived from Population I stars, so the colors of the models having \[Fe/H\] $``$ –0.5 should be used with some degree of caution (but see Section 3.3).
In Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars, we present a grid of calibrated colors and K-band bolometric corrections for stars having 4000 K $``$ T<sub>eff</sub> $``$ 6500 K and 0.0 $``$ log g $``$ 4.5 at five metallicities between \[Fe/H\] = –3.0 and solar metallicity; the optical colors of the dwarf (log g = 4.5) models resulting from the use of the cool-dwarf color calibrations discussed in Section 3.2.2 are enclosed in parentheses, allowing the reader to decide which color calibrations to adopt for these models. We will now proceed to compare our new, theoretical color-temperature relations to empirical relations of field stars and to previous MARCS/SSG results.
### 4.1 Comparing Our Results to Other Empirical Color-Temperature Relations
In the following sections, we use the photometry and effective temperatures of the stars listed in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars to derive empirical color-temperature relations for field giants and field dwarfs. We then compare our models and our empirical, solar-metallicity color-temperature relations to the CT relations of field stars which have been derived by Blackwell & Lynas-Gray (1994; hereafter BLG (94)), Gratton et al. (1996; hereafter GCC (96)), Bessell (1979; hereafter B (79)), Bessell (1995; hereafter B (95)), Bessell et al. (1998; hereafter BCP (98)) and Bessell (1998; hereafter B (98)).
#### 4.1.1 The Color-Temperature Relations of Our Field Stars
The color and temperature data for our set of color-calibrating field stars are plotted in Figures 8 and 9, where we have split the sample into giants (log g $``$ 3.6) and dwarfs (log g $`>`$ 3.6), shown in the lower and upper panels of the figures, respectively. We have fit quadratic relations to the effective temperatures of the dwarfs and giants separately as a function of color to derive empirical, solar-metallicity color-temperature relations for each; the resulting CT relations of the giants and dwarfs are shown as bold, solid lines and bold, dotted lines, respectively, in these figures. The coefficients of the fits (and their 1$`\sigma `$ uncertainties) are given in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars. The calibrated colors of our M67 isochrone (4 Gyr, Z) are also plotted in Figures 8 and 9 as solid lines; the dashed lines in the upper panels of Figure 8 show the calibrated isochrone when the cool-dwarf color calibrations are applied. Comparisons with our other isochrones indicate that the CT relation of the solar-metallicity isochrones is not sensitive to age.
Although we have divided our field-star sample into giants and dwarfs, the empirical CT relations of the two groups only appear to differ significantly in (V–R)<sub>C</sub> and (V–I)<sub>C</sub>. In B–V and J–K, the CT relations of the dwarfs and giants are virtually identical, and it is only the color of the coolest dwarf that prevents the same from being true in V–K. The differences in the U–V color-temperature relations of the dwarfs and giants can largely be attributed to the manner in which we chose to perform the least-squares fitting. Therefore, we want to emphasize that our tabulation of separate CT relations for dwarfs and giants does not necessarily infer that the two have significantly different colors at a given T<sub>eff</sub>.
It is apparent that the calibrated isochrone matches the properties of the field giants very nicely in Figures 8 and 9 and, in fact, is usually indistinguishable from the empirical CT relation. However, using the color calibrations of Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars, the dwarf and giant portions of the isochrone diverge cooler than $``$5000 K in all colors; this same behavior is not always seen in the stellar data. This discrepancy at cool temperatures means that one of the following conditions must apply: 1) the field star effective temperatures of the cool dwarfs are systematically too hot; 2) the isochrones predict the wrong T<sub>eff</sub>/gravity relation for cool dwarfs; or 3) the cool dwarf temperatures and gravities are correct but the model atmosphere and/or synthetic spectrum calculations are wrong for these stars. Only the latter condition would lead to different color calibrations being required to put the colors of cool dwarfs and cool giants onto the photometric systems.
The previously-mentioned comparison between our isochrones and those of VandenBerg (1999) lead us to believe that the lower main-sequence of our 4 Gyr, solar-metallicity isochrone is essentially correct, so it would appear that the synthetic colors of the coolest dwarf models are bluer than the color calibrations of Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars would predict because either we have adopted incorrect effective temperatures for these stars or there is some error in the model atmosphere or synthetic spectrum calculations, such as a missing opacity source.
Unpublished work on CO band strengths in cool dwarfs suggests that part of the problem could lie with the BG89 temperatures derived for these stars. In addition, some IRFM work (Alonso et al. 1996) indicates that the effective temperatures adopted for our K dwarfs cooler than 5200 K may be too hot by $``$100 K and infers an even larger discrepancy (up to 400 K) for HR 8086. A recent angular diameter measurement (Pauls (1999)) also predicts a hotter T<sub>eff</sub> than we adopted for HR 1084. If our K-dwarf effective temperatures are too warm, this would imply that the cool-dwarf color calibrations should not be used because the errors lie in the model parameters adopted for the cool dwarfs and not in the calculation of their synthetic spectra. However, T<sub>eff</sub> errors of 100 K are not sufficiently large to make the three cool field dwarfs lie on the optical color calibrations of the other stars (see Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars). In addition, the similarities of the V–K and J–K colors of our two coolest dwarfs and those of the calibrated isochrone at the assumed effective temperatures argue that the T<sub>eff</sub> estimates are essentially correct but do not preclude temperature changes of the magnitude implied by the IRFM and angular diameter measurements. Thus, some combination of T<sub>eff</sub> errors and model uncertainties may conspire to produce the color effects seen here, and more angular diameter measurements of K dwarfs would be very helpful in sorting out these possibilities. Nevertheless, the use of the cool-dwarf color calibrations improves the agreement between the models and the empirical data so remarkably well in Figures 7 and 8 that a good case can be made for their adoption.
As discussed in the Section 3.3, the model B–V colors of giants are noticeably dependent on gravity at low temperatures. The agreement between the observed B–V colors and the corresponding calibrated, synthetic colors for both the sample of field stars and the cool M67 giants suggests that the gravity-temperature relation of the isochrone truly represents the field stars, even at cool temperatures. While this is certainly anticipated, the fact that it occurs indicates that the surface boundary conditions and mixing-length ratio used for the interior models are satisfactory and gives us confidence that these newly-constructed isochrones are reasonable descriptions of the stellar populations they are meant to represent in our evolutionary synthesis program.
#### 4.1.2 Other Field-Star Color-Temperature Relations
BLG (94) estimated effective temperatures for 80 stars using the IRFM and derived V–K, T<sub>eff</sub> relations for the single stars, known binaries and total sample which they studied. Because these three relations are very similar, we have adopted their single star result for our comparisons, after converting their V–K colors from the Johnson to the Johnson-Glass system using the color transformation given by Bessell & Brett (1988). We expect the BLG (94) effective temperature scale to be in good agreement with that adopted here for the field stars used to derive the color calibrations – for the 19 stars in common between BLG (94) and BG (89), BLG (94) report that the average difference in T<sub>eff</sub> is 0.02%. For later discussion, we note that the coolest dwarf included in BLG (94)’s sample is BS 1325, a K1 star, with T<sub>eff</sub> = 5163 K.
GCC (96) derived their CT relations from photometry of about 140 of the BG (89) and BLG (94) stars. They adjusted BG (89)’s IRFM T<sub>eff</sub> estimates to put them onto the BLG (94) scale and then fit polynomials to the effective temperatures as a function of color to determine CT relations in Johnson’s B–V, V–R<sup>1</sup><sup>1</sup>1Gratton (1998) has advised us that the a3 coefficient of the T<sub>eff</sub>, V–R relation of the bluer class III stars should be +85.49; it is given as –85.49 in Table 1 of GCC (96)., R–I, V–K and J–K colors. In each color, they derived three relations – one represents the dwarfs, and the other two apply to giants bluer and redder than some specific color. For these relations, we have used B (79)’s color transformations to convert the Johnson V–R and V–I colors to the Cousins system; Bessell & Brett (1988) again supplied the V–K and J–K transformations to the Johnson-Glass system.
BCP (98) combined the IRFM effective temperatures of BLG (94) and temperatures estimated from the angular diameter measurements of di Benedetto & Rabbia (1987), Dyck et al. (1996) and Perrin et al. (1998) to derive a polynomial relation between V–K and T<sub>eff</sub> for giants. The color-temperature relations provided by B (98) are presumably derived in exactly the same manner from this same group of stars. For cool dwarfs, BCP (98) also determined a CT relation, this time in V–I, from the IRFM effective temperatures of BLG (94) and Alonso et al. (1996).
B (79) merged his V–I, V–K color-color relation for field giants and the V–K, T<sub>eff</sub> relation of Ridgway et al. (1980) to come up with a V–I, T<sub>eff</sub> relation. He then combined this with V–I, color trends to define CT relations for giants in other colors. By assuming that the same V–I, T<sub>eff</sub> relation applied to giants and dwarfs with 4000 K $``$ T<sub>eff</sub> $``$ 6000 K, B (79) derived dwarf-star CT relations in a similar manner. For the coolest dwarfs, updated versions of B (79)’s V–R, T<sub>eff</sub> and V–I, T<sub>eff</sub> relations were given by B (95).
#### 4.1.3 Comparisons of the Color-Temperature Relations
We compare our calibrated, solar-metallicity color grid to empirical determinations of color-temperature relations of field stars in Figures 1014. For these comparisons, we break our grid up into giants & subgiants, which we equate with models having 0.0 $``$ log g $``$ 4.0, and dwarfs (log g = 4.5). In all of the figures in this section, the giant-star CT relations appear in the upper panels, and the lower panels give the dwarf relations. The calibrated colors of our solar-metallicity grid are shown as open circles (or as asterisks when the cool-dwarf color calibrations have been used), but to relieve crowding in the giant-star plots, we connect the model colors at a given T<sub>eff</sub> by a solid line and plot only the highest and lowest gravity models as open circles. Our empirical color-temperature relations (Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars) are shown as solid lines in Figures 1014; the CT relations taken from the literature are represented by the symbols indicated in the figures. In addition, our calibrated, solar-metallicity, 4 Gyr isochrone is shown as a dotted line; the dashed line in the dwarf-star panels of Figures 1012 is the calibrated isochrone which results when the cool-dwarf color calibrations have been used. Comparisons with our other solar-metallicity isochrones shows that this CT relation is insensitive to age, so it should agree closely with the field-star relations. We omit the U–V and H–K color-temperature relations from these comparisons because we are not aware of any recent determinations of field-star relations in these colors, which nevertheless are less well-calibrated than the others.
For the giant stars, our empirical CT relations are in excellent agreement with the field relations of BLG (94), GCC (96), BCP (98) and B (98), with the temperature differences at a given color usually much less than 100 K. The 4 Gyr, Z isochrone also matches the field-giant relations quite well, as would be expected from Figures 8 and 9. In fact, most of the (small) differences between our empirical CT relations and the isochrone are probably due to our decision to use quadratic relations to represent the field stars; a quadratic function may not be a good representation of the true color-temperature relationships, and such a fit often introduces a bit of excess curvature at the color extremes of the observational data. The B (79) color-temperature relations for giants are also seen to be in good agreement with the other empirical and theoretical, giant-star data in V–R and V–I, but his B–V, T<sub>eff</sub> relation is bluer than the others; we have not attempted to determine the reason for this discrepancy.
Overall, since the colors of the field giants lie within the bounds of our solar-metallicity grid at a given effective temperature and also show good agreement with the 4 Gyr isochrone, our calibrated models are accurately reproducing the colors of solar-metallicity, F–K field giants of a specific temperature and surface gravity. Unfortunately, the dwarf-star color-temperature relations do not show the same level of agreement as those of the giants, overlying one another for T<sub>eff</sub> $``$ 5000 K but diverging at cooler effective temperatures. Therefore, we will examine each of the CT relations of the dwarfs individually.
In B–V (Figure 10), the field stars from Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars produce a CT relation which is in good agreement with that of GCC (96). When using the color calibrations of Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars, our isochrone and grid models of cool dwarfs are bluer than the empirical relations at a given T<sub>eff</sub>, which is consistent with the comparison to M67 in Figure 7, indicating that these calibrated B–V colors are too blue. On the other hand, the agreement between these models and the CT relation of B (79) suggests that these grid and isochrone colors are essentially correct. Similar quandaries are posed by the color-temperature relations of the dwarfs in V–R and V–I (Figures 11 and 12). Here, the B (79) relations predict slightly bluer colors than the analogous isochrone and grid models at a given temperature, but B (95)’s updated data for the coolest dwarfs agrees with our calibrated isochrone and models quite well. Our empirical CT relations for dwarfs and the GCC (96) data, on the other hand, lie to the red of these models and isochrones, again showing the same kind of discrepancy seen in the M67 comparisons. Unfortunately, the BCP (98) V–I, T<sub>eff</sub> relation does not extend to cool enough temperatures to help resolve the situation. Supplementing the color calibrations of Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars with the (optical) cool-dwarf color calibrations, however, brings the solar-metallicity isochrone and models into agreement with all of the field-dwarf CT relations of GCC (96) instead.
In Figure 13, we also see that there is one empirical CT relation (BLG (94)) which indicates that our cool-dwarf models are essentially correct at a given T<sub>eff</sub>, while another (GCC (96)) predicts that they are too blue, although the magnitude of disagreement in this figure is much smaller than that seen in Figures 1012. Since we are not aware of any near-infrared photometry for lower-main-sequence stars in M67, we cannot use Figure 7 to bolster any of the V–K CT relations. The only color in which all of the theoretical and empirical data are in relative agreement for the dwarfs is J–K; this is shown in Figure 14.
We hesitate to emphasize the similarities between our empirical color-temperature relations and the others plotted in the dwarf-star panels of Figures 1014. Our V–K CT relation for field stars only differs from that of BLG (94) because they neglected to treat dwarfs and giants separately; if we do the same, these differences disappear. Even so, the coolest dwarf in the BLG (94) study was BS 1325, a K1 V star with T<sub>eff</sub> = 5163 K, which is hotter than the temperature at which the dwarf and giant CT relations begin to bifurcate, so it is somewhat misleading to extend BLG (94)’s relation to redder colors in the dwarf-star panel of Figure 13. Likewise, it is not surprising that our CT relations and GCC (96)’s relations are so compatible – GCC (96) used the BLG (94) and BG (89) data to define their relations, so the cool end of their dwarf relations are defined by the same cool dwarfs included in our sample. The poor agreement between the other empirical CT relations and the dwarf-star relations of B (79) is probably due to the manner in which B (79) derived his field-star relations. He assumed a similar V–I, T<sub>eff</sub> relation for cool dwarfs and cool giants, an assumption which Figure 8 suggests to be inappropriate.
Overall, then, if we assume that the effective temperatures and photometry that we have adopted for the field stars of Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars and the members of M67 are all correct, we can account for the similarities and the differences between our color-temperature relations and those taken from the literature. In this case, it appears that the CT relations of GCC (96) are the most representative of field dwarfs, while each of those which we have examined are probably equally reliable for the field giants (except B (79)’s B–V, T<sub>eff</sub> relation). This also produces a consistent picture in which both the field-star CT relations and the M67 photometry indicate that our model colors for cool dwarfs are too blue at a given effective temperature if only the color calibrations of Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars are employed; adopting the cool-dwarf color calibrations for T<sub>eff</sub> $`<`$ 5000 K then brings the models and the observational data into close agreement.
However, we can envision an alternate scenario in which the coolest three (or more) field dwarfs have been assigned effective temperatures which are too hot. By lowering T<sub>eff</sub> by $``$100 K for HR 8085 and HR 8832 and by $``$200 K for HR 8086, these stars would lie along the the same color calibrations as the other field stars in all colors except perhaps U–V and B–V. This would shift our empirical CT relations and those of GCC (96) to bluer colors for the cool dwarfs and thus resolve the disagreement with B (95)’s relations and the models calibrated using the Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars color calibrations. It would also leave V–R and V–I as the only two colors in which the dwarf-star and giant-star CT relations differ significantly, which is understandable, since the V, R and I bands contain many molecular features (e.g., TiO, CN) which are gravity-sensitive. However, the discrepancies between the calibrated, 4 Gyr, solar-metallicity isochrone (solid line in Figure 7) and the U–V and B–V photometry of the lower main-sequence of M67 would likely remain.
Since it is not clear to us which of these situations applies, we have presented the calibrated model colors for each case here. The essential pieces of information needed to disentangle these two possibilities are angular diameter measurements of K dwarfs; these would certainly aid us in testing the BG (89) temperatures of the cool dwarfs in our sample and confirming their CT relations. However, for the time-being, the small number of cool dwarfs with effective temperature estimates and the fact that such stars make only a small contribution to the integrated light of a galaxy have led us to adopt the calibration relations given in Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars for the synthetic colors of all of our evolutionary synthesis models.
### 4.2 Comparisons to Previous MARCS/SSG Color-Temperature Relations
In Figures 15 and 16, we compare our new, solar-metallicity CT relations to the previous MARCS/SSG results published by VB (85) and BG (89); the models of the former and latter are shown as open and filled circles, respectively. In each panel of these figures, we plot log g vs. color and use dotted lines to represent isotherms resulting from the uncalibrated colors of our models; the isotherms produced by our improved, calibrated colors (Table Improved Color–Temperature Relations and Bolometric Corrections for Cool Stars calibrations only) are shown as solid lines. Thus, differences between the older colors and the appropriate dotted line can be identified with improvements in the models and color measurements, and differences between the dotted and solid lines of the same temperature can be equated with the effects of calibrating the synthetic colors.
#### 4.2.1 The Optical Color-Temperature Relations
In the upper panel of Figure 15, we compare our CT relations and the B–V colors calculated by VB (85) using the filter-transmission-profiles of Matthews & Sandage (1963). The calibrated B–V colors are always redder than their previous counterparts, but the uncalibrated colors of the new models do not differ appreciably from those of VB (85) for T<sub>eff</sub> $``$ 5000 K; the main improvement in the B–V colors of these hotter stars is due to the color calibrations. At 4500 K, especially as log g decreases, the improved low-temperature opacity data used here becomes important, and this becomes the dominant improvement for cooler stars, although the color corrections inferred from the calibrations are still substantial.
The V–R colors of VB (85) and BG (89), measured with the R-band filter of Cousins (1980), are compared to our new, solar-metallicity color grid in the middle panel of Figure 15. Surprisingly, for effective temperatures hotter than about 5000 K, the colors of the older models are in good agreement with those of the newer, calibrated models; either the color differences produced by the filter-transmission-profiles adopted by each group are being almost exactly offset by the effects of the color calibrations which we have derived, or the newer opacities used here have moved the V–R colors of the models to the blue. For the cool giants, however, the calibrated V–R colors are significantly redder than those previously published.
For V–I, the new and old models are shown in the lower panel of Figure 15. This diagram is similar to the V–R plot in that it is difficult to disentangle the color changes produced by the use of different filter profiles (VB (85) and BG (89) use the I-band filter of Cousins 1980) and those resulting from improvements in the stellar modelling. It is apparent, however, that the previous MARCS/SSG V–I colors are too red at hotter temperatures (T<sub>eff</sub> $``$ 5500 K) and too blue at cooler temperatures.
#### 4.2.2 The Near-Infrared Color-Temperature Relations
Figure 16 (upper panel) shows that the synthetic V–K colors on the CIT/CTIO system are not altered appreciably by calibration, at least with respect to the color range of the models and the magnitude of the color differences which exist between the newer and older models. While the colors of our models and those of BG (89) are about the same at T<sub>eff</sub> = 4500 K, the newer models are redder at temperatures cooler than this and the older models are redder at hotter temperatures. Differences in the model atmospheres and synthetic spectra are responsible for these effects, since BG (89) used the same filter-transmission-profiles that we have.
The CIT/CTIO J–K colors shown in the lower panel of Figure 16 are all significantly bluer than those previously published. For the hotter models, improvements in MARCS and the much better treatment of line absorption in SSG are the major causes of these blueward shifts. In fact, the magnitude of the color differences between the older colors and the newer, uncalibrated colors is about the same at all temperatures. The color calibrations, however, while also making the model J–K colors bluer, grow in importance as T<sub>eff</sub> decreases; this effect begins to dominate the color shifts due to the model improvements at T<sub>eff</sub> $``$ 4500 K.
## 5 Conclusion
We have calculated colors and bolometric corrections for grids of stellar models having 4000 $``$ T<sub>eff</sub> $``$ 6500 K, 0.0 $``$ log g $``$ 4.5 and –3.0 $``$ \[Fe/H\] $``$ 0.0. The synthetic colors which we measure include Johnson U–V and B–V; Cousins V–R and V–I; Johnson-Glass V–K, J–K and H–K; and CIT/CTIO V–K, J–K, H–K and CO. These synthetic colors have been determined by convolving the most recent estimates of the filter-transmission-profiles of the respective photometric systems with newly-calculated synthetic spectra. The synthetic spectra are produced by the SSG spectral synthesis code, which has been updated significantly, especially in terms of its low-temperature opacity data and spectral line lists. The MARCS model atmospheres which are used by SSG have also been calculated using this new opacity data and improved opacity distribution functions.
We have found that the initial synthetic colors require small corrections to be put onto the observational systems. Here, we have assembled a set of field stars which have effective temperatures measured using the infrared-flux method (IRFM), and we have used photometry of these stars to derive the color calibrations needed to put the synthetic colors onto the photometric systems. The scale factors and zero points of these calibrations are gratifyingly close to unity and small, respectively, but the UBVRI colors of the coolest dwarfs (T<sub>eff</sub> $`<`$ 5000 K) may require a more substantial correction. Nevertheless, the specific color calibrations which we adopt are only applicable to models calculated with our versions of the MARCS and SSG computer codes. We encourage others to examine the need for such calibrations in their models as well.
Using the field-star data adopted in deriving the color calibrations, we have also found that the recent angular diameter measurements of Mozurkewich et al. (1991), Mozurkewich (1997) and Pauls et al. (1997) match the angular diameters predicted by Bell & Gustafsson (1989) very well, thus giving strong confirmation of the latter’s effective temperature scale, which we have adopted. Since most of the synthetic colors are much more sensitive to effective temperature than they are to the other model parameters, such as metallicity and surface gravity, this infers that the differences between the uncalibrated, synthetic colors and the field-star photometry are not caused by errors in the temperatures of the field-star models. The coolest field dwarfs may again prove an exception to this generalization, as none of their angular diameters have been measured and their effective temperatures are therefore more uncertain. Knowing that the IRFM-derived temperatures are accurate has also allowed us to determine empirical, color-temperature (CT) relations for these field F–K stars.
After calibrating our synthetic colors, our newly-calculated, 4 Gyr, Z isochrone is seen to be in good agreement with optical and near-infrared color-magnitude diagrams of M67 from the upper main-sequence through the tip of the red-giant branch. For giants and for dwarfs hotter than about 5000 K, our theoretical, solar-metallicity CT relations and solar-metallicity isochrones are also found to match the empirical, field-star color-temperature relations derived here and found elsewhere in the literature. For the coolest dwarfs (T<sub>eff</sub> $`<`$ 5000 K), there are some differences between our empirical color-temperature relations and the best-determined relations of others. The calibrated colors of the cool-dwarf models also remain a bit bluer than the field-star relations (at a given T<sub>eff</sub>) and the M67 main-sequence stars (at a given absolute magnitude). Because we have not been able to determine whether the synthetic, optical colors of the cool dwarf models are too blue because their adopted effective temperatures are too hot or the synthetic spectra are in error, we present alternate UBVRI colors for our models having T<sub>eff</sub> $`<`$ 5000 K and log g = 4.5. These colors are based upon separate color calibrations derived from only the cool field dwarfs and which greatly improve the agreement between the calibrated, synthetic colors and the empirical data. We leave it to the reader to decide which model colors they prefer.
This work would greatly benefit from further effective temperature estimates of cool dwarfs and measurements of the corresponding angular diameters needed to confirm them. It would also be quite useful to acquire additional photometry of the field stars used here to derive the color calibrations. Only a small subset of these stars, which includes no cool dwarfs, has been observed on the Cousins system, making the V–R and V–I color calibrations less well-determined than some of the others. In addition, only Johnson near-infrared photometry was available for the field stars. Color transformations had to be applied to this photometry before the Johnson-Glass and CIT/CTIO color calibrations could be derived; this also prohibited the determination of a similar calibration for the CIT/CTIO CO index.
Overall, the agreement between our theoretical CT relations and the empirical data and between our 4 Gyr, solar-metallicity isochrone and the photometry of M67 indicates that our synthetic spectra are generally providing a good representation of the spectral energy distributions of stars of the corresponding T<sub>eff</sub>, surface gravity and metallicity. This conclusion is further supported by our synthetic spectra of M giants, which are presented in a companion paper (Houdashelt et al. 2000). In a future paper (Houdashelt et al. 2001), we will use similar synthetic spectra and newly-constructed isochrones as part of a program to simulate the spectral energy distributions of early-type galaxies through evolutionary synthesis.
We would like to thank the National Science Foundation (Grant AST93-14931) and NASA (Grant NAG53028) for their support of this research. We also thank Ben Dorman for allowing us to use his isochrone-construction code and Mike Bessell for providing many helpful suggestions on the manuscript. MLH would like to express his gratitude to Rosie Wyse for providing support while this work was completed. The research has made use of the Simbad database, operated at CDS, Strasbourg, France. The NSO/Kitt Peak FTS data used here were produced by NSF/NOAO. |
no-problem/9911/hep-ph9911210.html | ar5iv | text | # POMERON: BEYOND THE STANDARD APPROACH 11footnote 1To appear in Proceedings of “XXIX International Symposium on Multiparticle Dynamics, 9-13 August 1999, Brown University, Providence, RI 02912, USA”.
## 1 Introduction
Experiments at HERA and at hadron colliders have reported and characterized a class of events incorporating a hard (high transverse momentum) partonic scattering while carrying the characteristic signature of diffraction, namely a leading (anti)proton and/or a large rapidity gap (region of pseudorapidity devoid of particles). The prevailing theoretical idea is that the exchange across the gap is the Pomeron , which in QCD is a color-singlet exchange of gluons and/or quarks with vacuum quantum numbers.
A question of intense theoretical debate is whether the Pomeron has a unique particle-like partonic structure. This question can be addressed experimentally by comparing the parton distribution functions (pdf’ s) of the (anti)proton measured in a variety of hard single diffraction dissociation processes as a function of $`\xi `$, $`t`$, $`Q^2`$ and $`x`$ (or $`\beta x/\xi `$), where $`t`$ is the 4-momentum transfer and $`\xi `$ the fractional momentum loss of the (anti)proton. The gluon and quark pdf’ s can be sorted out by studying processes with different sensitivity to the gluon and quark components of the Pomeron.
The proton diffractive pdf’s have been measured in diffractive deep inelastic scattering (DDIS) at HERA by both the H1 and ZEUS Collaborations. The experiments measure directly the $`F_2`$ diffractive structure function, $`F_2^{D(3)}(\xi ,Q^2,\beta )=_{t=1}^{t_{min}}F_2^{D(4)}(t,\xi ,Q^2,\beta )𝑑t`$. The variable $`\xi `$ is related to the rapidity gap by $`\mathrm{\Delta }y=\mathrm{ln}\frac{1}{\xi }`$. The gluon diffractive $`pdf`$ was determined by H1 from a QCD analysis of the $`Q^2`$ evolution of $`F_2^{D(3)}`$. All HERA hard diffraction results are generally consistent with the parton densities obtained from DDIS. However, assuming factorization, calculations based on these $`pdf`$’s predict rates $`10`$ times larger than the measured $`W`$ and dijet production rates at the Tevatron . The suppression of Tevatron relative to HERA diffractive rates represents a breakdown of factorization.
The magnitude of the factorization breakdown is in general agreement with predictions based on the renormalized Pomeron flux model , in which the Pomeron flux (see next section) is viewed as a rapidity gap probability density and is normalized by scaling it to its integral over the available phase space in $`(\xi ,t)`$. In addition to its success in predicting soft and hard diffraction rates, the flux renormalization model describes differential soft diffraction cross sections remarkably well. However, the $`\beta `$-dependence of the Pomeron/diffractive $`pdf`$’s is not specified by the model and had to be introduced “by hand”. In this paper, the diffractive $`pdf`$’s are derived from the non-diffractive using a parton model approach to hard diffraction based on the the concept of a normalized rapidity gap probability.
## 2 Clues from soft physics
The Pomeron was introduced in Regge theory to account for the high energy behavior of the elastic, diffractive and total hadronic cross sections. In terms of the Pomeron trajectory, $`\alpha (t)=1+ϵ+\alpha ^{}t`$, the $`pp`$ cross sections can be written as
$$\sigma _T\left(s\right)=\beta _{IPpp}^2\left(0\right)\left(\frac{s}{s_0}\right)^{\alpha \left(0\right)1}\left(\frac{s}{s_0}\right)^ϵ$$
(1)
$$\frac{d\sigma _{el}}{dt}=\frac{\beta _{IPpp}^4\left(t\right)}{16\pi }\left(\frac{s}{s_0}\right)^{2\left[\alpha \left(t\right)1\right]}\mathrm{exp}\left[\left(b_{}+2\alpha ^{}\mathrm{ln}s\right)t\right]$$
(2)
$$\frac{d^2\sigma _{sdd}}{d\xi dt}=\underset{\mathrm{Pomeron}\mathrm{flux}\frac{1}{\xi ^{1+2ϵ}}}{\underset{}{\frac{\beta _{IPpp}^2\left(t\right)}{16\pi }\frac{1}{\xi ^{2\alpha \left(t\right)1}}}}\underset{\sigma _T^{IPp}\left(\xi s\right)^ϵ}{\underset{}{\beta _{IPpp}\left(0\right)g\left(t\right)\left(\frac{\xi s}{s_0}\right)^{\alpha \left(0\right)1}}}$$
(3)
The Regge approach has been successful in predicting the three salient features of high energy behavior, namely (i) the rise of the total cross secions with energy , (ii) the shrinking of the forward elastic scattering peak , and (iii) the shape of the $`M^2`$ dependence of SDD . However, these features are also present in a parton model approach , as outlined below.
### 2.1 Rise of total cross sections
The $`pp`$ total cross section is basically proportional to the number of partons in the proton, integrated down to $`x=s_{}/s`$, where $`s_{}`$ is the energy scale for soft physics. The latter is of $`𝒪(M_T^2)`$, where $`M_T1`$ GeV is the average transverse mass of the particles in the final state. Expressing the parton density as a power law in $`1/x`$, which is an appropriate parameterization for the small $`x`$ region responsible for the cross section rise at high energies, we obtain
$$\sigma _T_{\left(s_{}/s\right)}^1\frac{dx}{x^{1+n}}\left(\frac{s}{s_{}}\right)^n$$
(4)
The parameter $`n`$ is identified with the $`ϵ=\alpha (0)1`$ of the Pomeron trajectory.
### 2.2 Shrinking of forward elastic peak
In Regge theory, the slope of the forward elastic peak increases as $`\mathrm{ln}s`$, while in the parton picture one would naively expect the slope to follow the increase of the total cross section:
$$\frac{d\sigma ^{el}}{dt}\underset{\mathrm{Regge}}{\underset{}{\mathrm{exp}\left[\left(b_{}+2\alpha ^{}\mathrm{ln}s\right)t\right]}}\underset{\mathrm{Parton}\mathrm{model}}{\underset{}{\mathrm{exp}\left[\left(b_{}+cs^n\right)t\right]}}$$
(5)
Using $`n=ϵ=0.104`$ , we note that from $`\sqrt{s}=20`$ to 1800 GeV the terms $`\mathrm{ln}s`$ and $`s^n`$ increase by factors of 2.50 and 2.55, respectively. Thus, the Regge and parton model expressions for the slope are experimentally indistinguishable.
### 2.3 $`M^2`$ dependence of diffraction dissociation
The diffractive mass is related to $`\xi `$ by $`M^2\xi s`$. In the Regge picture, the $`t=0`$ single diffractive cross section <sup>2</sup><sup>2</sup>2In the rest of this paper we use $`t=0`$ cross sections for simplicity, but the results we obtain generally apply also to the cross sections integrated over $`t`$. is given by (see Eq. 3)
$$\mathrm{Regge}:\frac{d\sigma _{sdd}}{dM^2}\frac{s^{2ϵ}}{\left(M^2\right)^{1+ϵ}}$$
(6)
However, experimentally the cross section is found to be independent of $`s`$ :
$$\mathrm{Experiment}:\frac{d\sigma _{sdd}}{dM^2}\frac{1}{\left(M^2\right)^{1+ϵ}}$$
(7)
In terms of $`\xi =\frac{M^2}{s}`$, we have $`\frac{d\sigma _{sdd}^{\mathrm{exp}}}{d\xi }\frac{1}{s^ϵ}\frac{1}{\xi ^{1+ϵ}}`$, which can be written as
$$\frac{d\sigma _{sdd}}{d\xi }\underset{}{\frac{1}{s^{2ϵ}}\frac{1}{\xi ^{1+2ϵ}}}\times \left(\xi s\right)^ϵ$$
(8)
Recalling that $`\xi `$ is related to the rapidity gap and $`\xi s=M^2`$ is the s-value of the diffractive sub-system, and noting that $`_{(s_{}/s)}^1\frac{1}{s^{2ϵ}}\frac{d\xi }{\xi ^{1+2ϵ}}=\mathrm{constant}`$, Eq. 8 may be viewed as representing the product of the total cross section at the sub-system energy multiplied by a normalized rapidity gap probability. This experimentally established scaling behavior is used in the next section as a clue in deriving the diffractive $`F_2`$ structure function from the non-diffractive in the parton model approach to diffraction.
## 3 Diffractive deep inelastic scattering
The inclusive and diffractive DIS cross sections are proportional to the corresponding $`F_2`$ structure functions of the proton,
$$\frac{d^2\sigma }{dxdQ^2}\frac{F_2^h(x,Q^2)}{x};\frac{d^3\sigma }{d\xi dxdQ^2}\frac{F_2^{D\left(3\right)}(\xi ,x,Q^2)}{x}$$
(9)
where the superscripts $`h`$ and $`D(3)`$ indicate a hard inclusive structure (at the scale $`Q^2`$) and a 3-variable diffractive structure (integrated over $`t`$). The latter depends not only on the hard scale $`Q^2`$, but also on the soft scale $`M_T^2`$, which is the relevant one for the formation of the gap.
The only marker of the rapidity gap is the variable $`\xi `$. We therefore postulate that the rapidity gap probability is proportional to the soft parton density at $`\xi `$ and write the DDIS cross section as
$$\frac{d^3\sigma }{d\xi dxdQ^2}\frac{F_2^h(x,Q^2)}{x}\times \frac{F_2^s\left(\xi \right)}{\xi }\xi \mathrm{norm}$$
(10)
where the symbolic notation “$`\xi \mathrm{norm}`$” is used to indicate that the $`\xi `$ probability is normalized. Since $`x=\beta \xi `$, the normalization over all available $`\xi `$ values involves not only $`F_2^s`$ but also $`F_2^h`$, breaking down factorization. It is therefore prudent to write the DDIS cross section in terms of $`\beta `$ instead of $`x`$, so that the dependence of $`F_2^h`$ on $`\xi `$ is shown explicitely:
$$\frac{d^3\sigma }{d\xi d\beta dQ^2}\frac{F_2^h(\beta \xi ,Q^2)}{\beta }\times \frac{F_2^s\left(\xi \right)}{\xi }\xi \mathrm{norm}$$
(11)
The term to the right of the $``$ sign represents the $`F_2^{D(3)}(\xi ,\beta ,Q^2)`$ structure function. Guided by the scaling behaviour of Eq. 8, we now factorize $`F_2^{D(3)}(\xi ,\beta ,Q^2)`$ into $`F_2^h(\beta ,Q^2)`$, the sub-energy cross section, times a normalized gap probability:
$$F_2^{D\left(3\right)}(\xi ,\beta ,Q^2)=P_{gap}(\xi ,\beta ,Q^2)\times F_2^h(\beta ,Q^2)$$
(12)
The gap probability is given by
$$P_{gap}(\xi ,\beta ,Q^2)=\frac{F_2^h(\beta \xi ,Q^2)}{\beta }\times \frac{F_2^s\left(\xi \right)}{\xi }\times N(s,\beta ,Q^2)$$
(13)
where $`N(s,\beta ,Q^2)`$ is the normalization factor,
$$N(s,\beta ,Q^2)=f_q/_{\xi _{min}=Q^2/\beta s}^1\frac{F_2^h(\beta \xi ,Q^2)}{\beta }\times \frac{F_2^s\left(\xi \right)}{\xi }𝑑\xi $$
(14)
in which $`f_q`$ denotes the quark fraction of the hard structure (since only quarks participate in DIS).
### 3.1 Comparison with data
At small $`x`$ and small $`\xi `$, the structure functions $`F_2^h`$ and $`F_2^s`$ are represented well by $`F_2^h(x,Q^2)=A^h/x^{ϵ^h(Q^2)}`$ and $`F_2^s(\xi )=A^s/\xi ^{ϵ^s}`$. An analytic evaluation of Eqs. 14 and 12 yields
$$N(s,\beta ,Q^2)=f_q/\left[\frac{A^h}{\beta ^{1+ϵ^h}}\frac{A^s}{ϵ^h+ϵ^s}\left(\frac{\beta s}{Q^2}\right)^{ϵ^h+ϵ^s}\right]$$
(15)
$$F_2^{D\left(3\right)}(\xi ,\beta ,Q^2)=\frac{1}{\xi ^{1+ϵ^h+ϵ^s}}\times f_q\left(ϵ^h+ϵ^s\right)\left(\frac{Q^2}{\beta s}\right)^{ϵ^h+ϵ^s}\times \frac{A^h}{\beta ^{ϵ^h}}$$
(16)
Figure 1 shows a comparison of the $`\beta `$ distribution of H1 data at $`\xi 0.01`$ and $`Q^2=45`$ GeV<sup>2</sup> with the prediction of Eq. 16. The following parameters were used in the calculation: $`\sqrt{s}=280`$ GeV, $`ϵ^s=0.1`$ , $`Q^2=45`$ GeV<sup>2</sup>, $`ϵ^h(Q^2=45)=0.3`$, $`f_q=0.4`$, $`\xi =0.01`$ and $`A^h=0.2`$, evaluated from $`F_2(Q^2=50,x=0.00133)=1.46\frac{A^h}{x^{0.3}}`$. The agreement between theory and experiment may be considered excellent, particularly since no free parameters were used in the calculation.
Figure 1
## References |
no-problem/9911/quant-ph9911072.html | ar5iv | text | # Use of Composite Rotations to Correct Systematic Errors in NMR Quantum Computation
## I Introduction
Quantum computers are information processing devices which operate by—and exploit—the laws of quantum mechanics, giving them the potential to tackle problems which are intractable using classical computers . Although there has been a great deal of interest in the theory of quantum computation, actually building a general purpose quantum computer has proved extremely difficult. The multiparticle coherent states on which quantum computers rely are extremely vulnerable to the effects of errors, and so considerable effort has been expended on tackling random errors introduced by decoherence, including methods of error correction and fault tolerant computation . Comparatively little effort, however, has been directed at the issue of systematic errors, arising from reproducible imperfections in the apparatus used to implement quantum computations.
It makes sense to address systematic errors, however, because many of them can be eliminated relatively easily. It is convenient to visualise one spin quantum states as three-dimensional vectors, where the $`x`$, $`y`$, and $`z`$ components are given by decomposition of the state into a linear combination of Pauli spin matrices. In this picture, unitary operations are interpreted as rotations of the Bloch vector on a sphere. The sensitivity of the final state to rotational imperfections can be much reduced by replacing single rotations with composed rotations. Ordinarily replacing one erroneous operation by several erroneous operations will increase the overall error, but this isn’t necessarily the case as rotations are non-linear, and the non-linearity opens the possibility of errors cancelling one another.
It is worth stressing the distinction between improved experimental technique, quantum error correcting codes , and the use of composite pulses. Improved experimental technique minimises conditions which lead to data errors. Error correcting codes diagnose errors (however obliquely) and use this knowledge of the errors to explicitly reverse them. Composite pulses prevent errors in the first place by reducing the impact on the system of conditions which cause systematic data errors, without actually eliminating these conditions. Composite pulses cannot, however, correct random errors, as they rely on the reproducible nature of systematic errors to correct them.
## II Systematic errors in NMR quantum computers
In the last few years nuclear magnetic resonance (NMR) techniques have been used to implement small quantum computers . Several simple quantum algorithms have been implemented using NMR, such as Deutsch’s algorithm and Grover’s algorithm . The comparative ease with which such experiments have been performed reflects, in part, the fact that more conventional NMR experiments, used throughout the molecular sciences, themselves rely on the generation and manipulation of multi-spin coherent states, and so techniques for handling them are highly developed.
Despite this pre-existing experimental sophistication, NMR quantum computers are nonetheless subject to systematic errors. Implementing complex quantum algorithms require a large network of logic gates, which within an NMR implementation entails even longer cascades of pulses. In this case small systematic errors in the pulses (which can be largely ignored in conventional NMR experiments) accumulate and become significant.
Interactions between individual spins and between spins and the environment are mediated by RF pulses, which are applications of an RF field with phase $`\varphi `$ (in the rotating frame) for some duration $`\tau `$. In the ideal case, the pulse will drive the Bloch vector through an angle $`\theta `$ about an axis orthogonal to the $`z`$-axis and at an angle $`\varphi `$ to the $`x`$-axis. The rotation angle, $`\theta `$, depends on the rotation rate induced by the RF field, usually written $`\omega _1`$, and the duration of the pulse, $`\tau `$. In practice the RF field is not ideal, and this leads to two important classes of systematic errors, usually referred to as pulse length errors and off-resonance effects.
Pulse length errors arise either when the length of the pulse is set incorrectly, or (more commonly) when the RF field strength deviates from its assumed value, so that the rotation angle achieved deviates from its theoretical value. This effect is most commonly observed within NMR as a result of inhomogeneity in the applied RF field; in this case it is impossible for all the spins within the sample to experience the same rotation angle.
Off resonance effects arise from the use of a single RF source to excite transitions in two or more spins which have different resonance frequencies. This is done for a variety of practical reasons. For example, in conventional NMR experiments, which seek to analyse molecular systems, there will be a large number of different transition frequencies, whose exact values are unknown at the start of the experiment. Thus the only practical approach is to use a single RF source, with sufficient power that it can excite transitions across a wide range of frequencies. In quantum computation experiments the transition frequencies are known beforehand, but for a variety of reasons the use of a single RF source for each type of atomic nucleus remains the most practical approach (different sources *are* used for different nuclei, such as <sup>1</sup>H and <sup>13</sup>C).
Composite pulses are widely used in NMR to minimize the sensitivity of the system to pulse-length and off-resonance errors, by replacing direct rotations with composite rotations which are less sensitive to such effects. However, conventional composite pulse sequences are rarely appropriate for quantum computation because they usually incorporate assumptions about the initial state of the spins or introduce phase errors. These conventional sequences were initially derived by considering the trajectories of Bloch vectors during a pulse, and this approach is only useful if the starting point of the Bloch vector is known.
QC requires fully-compensating (type A) composite pulse sequences . Fortuitously, a number of such sequences were developed fifteen years ago by Tycko . While these pulse sequences do not offer quite the same degree of compensation as is found with some more conventional sequences, the compensation which does occur is effective whatever the initial position of the Bloch vector. As such sequences are rarely (if ever) required within conventional NMR experiments, these sequences have previously found no experimental use. They are, however, ideally suited to quantum computations.
## III Example: NMR quantum counting
Systematic errors are an issue with many types of quantum algorithms, but we will focus on a counting algorithm which can involve particularly long pulse trains. Counting the number of items matching a search criterion is a well-known computer science problem, and an efficient quantum counting algorithm based on a variation of Grover’s quantum search has been implemented using NMR techniques.
Consider a match function $`f(x)`$ which maps $`n`$-bit binary strings to a single output bit. In general there will be $`N=2^n`$ possible input values, of which $`k`$ will yield a match, $`f(x)`$=1. The object of the quantum counting algorithm is to estimate $`k`$ by estimating an eigenvalue of the Grover iterate $`G=HU_0H^1U_{\overline{f}}`$, where $`H`$ represents an $`n`$-bit Hadamard transform, $`U_0`$ maps $`|0`$ to $`|0`$, and $`U_{\overline{f}}`$ maps $`|x`$ to $`(1)^{f(x)+1}|x`$. The algorithm involves applying $`G`$ repeatedly and observing the variation in signal intensity. The signal intensity is modulated in $`r`$, the number of applications of G, with frequency proportional to $`k`$, the number of matches. A more detailed explanation of this algorithm is found in reference .
Figure 1 shows experimental results for a search over a one qubit search space, using a two-qubit quantum computer based on the two <sup>1</sup>H nuclei of cytosine in solution in $`\mathrm{D}_2\mathrm{O}`$ . The NMR spectrometer used a <sup>1</sup>H operating frequency of 500 MHz. On a one qubit search space, there are three possible results corresponding to four possible functions $`f`$: no inputs match (in which case $`f=f_{00}`$), the first input matches ($`f=f_{01}`$), the second input matches ($`f=f_{10}`$), or both inputs match ($`f=f_{11}`$). For simplicity, only the cases $`f_{00}`$ and $`f_{11}`$ are shown.
The observed signal loss with $`r`$ could arise from a number of sources. Clearly one possibility is the effects of decoherence, but the observed decay rate is too rapid to be so simply explained, and patterns in the signal loss (this is especially clear for $`f_{00}`$, where data points lie alternately above and below a smooth curve) suggest a more complex explanation. Repeating the experiments on spectrometers with higher operating frequencies markedly degrades the results, as shown in figure 2, which shows data from computations implemented on 600 and 750 MHz spectrometers. Once more, only the $`f_{00}`$ and $`f_{11}`$ cases are shown. These results clearly demonstrate that the use of expensive high field magnets does not always give better results. An unwanted beat frequency and high frequency chatter have been introduced, and the signal decays more rapidly. These effects are more severe at 750 MHz than at 600 MHz.
The observed field dependence strongly suggests that these errors arise from off-resonance effects. These arise because only a single RF transmitter is used to excite both <sup>1</sup>H nuclei, and it cannot be on-resonance with both sets of transitions. Instead the transmitter is placed halfway between the two resonance frequencies, resulting in equal and opposite off resonance terms. Two factors are responsible for the poor results at higher frequencies. Firstly, the frequency offsets, $`\pm \delta \nu `$, are proportional to the resonance frequency, and thus are larger at high frequencies. Secondly, the RF field strength is typically weaker at higher frequencies. These combine to make the off-resonance effects much more serious. This diagnosis can be confirmed by including off-resonance effects in simulations of the NMR experiment. After this correction is made, the theoretical results, shown in figure 3, agree strikingly well with the experimental results. Even the high frequency ’noise’ on the peaks is reproduced.
Since our simulations confirm that the major errors in NMR quantum counting arise from off-resonance effects in the $`90_y^{}`$ pulses, we chose to use Tycko’s $`385_y^{}\mathrm{\hspace{0.17em}320}_y^{}\mathrm{\hspace{0.17em}25}_y^{}`$ composite $`90_y^{}`$ sequence, which offers good compensation for off-resonance errors while remaining fairly insensitive to pulse-length errors. Substituting this sequence for the $`90_y^{}`$ pulses produced a significant improvement in the experimental results at 750 MHz, as shown in figure 4. The low frequency beating almost disappears, and the high frequency noise is much reduced.
When using composite pulses, there is a trade-off between the cancellation of errors and the extra manipulation of the system necessary to induce this cancellation. While there exist sequences which in theory compensate for an extraordinary range of errors, these sequences involve prohibitively long cascades of pulses. In practice, three-pulse sequences seem to provide a good balance between insensitivity and simplicity. The hazards of over-manipulation can be seen when the counting experiment is repeated using composite $`180^{}`$ pulses as well as $`90^{}`$ ones, shown in figure 5. The composite $`180_x^{}`$ pulse used was $`90_x^{}\mathrm{\hspace{0.17em}225}_x^{}\mathrm{\hspace{0.17em}315}_x^{}`$ . Although simulations predict that the results should be slightly better when both $`180^{}`$ and $`90^{}`$ compensated pulses are used, the results for $`f_{11}`$ are in fact slightly worse. The errors introduced by the extra manipulation outweigh the small gain achieved by the use of more sophisticated pulse sequences.
## IV Composite pulses for arbitrary rotation angles
Discussion of composite pulses in the NMR literature has focussed almost exclusively on $`90^{}`$ and $`180^{}`$ pulses, for the simple reason that these are the rotation angles most commonly used in conventional pulse sequences. While these composite pulses can be very useful for quantum computation, as shown above, it would be preferable to have composite pulses for other rotation angles as well. For example, a $`45^{}`$ pulse could be used in the implementation of a true Hadamard gate , which might be applied many times in a single network.
We use the method of coherent averaging to generate an analytic expression for composite pulses similar to Tycko’s $`90^{}`$ pulse but with arbitrary rotation angle. The basic idea of the method is to split the Hamiltonian of the system into intended and error components,
$$H(t)=H_0(t)+V,$$
(1)
where the “ideal” Hamiltonian (written in the rotating frame and using product operator notation) is
$$H_0\left(t\right)=\omega _1\left(I_x\mathrm{cos}\varphi +I_y\mathrm{sin}\varphi \right)$$
(2)
and the error term is
$$V=\delta \omega I_z$$
(3)
($`\delta \omega `$ is the resonance offset angular frequency). We then seek to minimise the propagator of the error component by expanding it as a power series . Knowing $`V`$, we can rewrite the propagator as a product,
$`U(t)`$ $`=`$ $`U_0(t)U_V(t)`$ (4)
$`=`$ $`T\mathrm{exp}\left(i{\displaystyle _0^\tau }𝑑tH_0(t)\right)T\mathrm{exp}\left(i{\displaystyle _0^\tau }𝑑t\stackrel{~}{V}(t)\right)`$ (5)
where $`T`$ is the Dyson time-ordering operator and
$$\stackrel{~}{V}(t)=U_0(t)^1VU_0(t).$$
(6)
Since $`U_V`$ varies rather slowly with time, it can be expanded as a power series in $`\stackrel{~}{V}`$:
$$U_V(\tau )=\mathrm{exp}\left(i\tau \left(V^{(0)}+V^{(1)}+\mathrm{}\right)\right)$$
(7)
using the Magnus expansion
$`V^{(0)}`$ $`=`$ $`{\displaystyle \frac{1}{\tau }}{\displaystyle _0^\tau }𝑑t\stackrel{~}{V}(t)`$ (8)
$`V^{(1)}`$ $`=`$ $`{\displaystyle \frac{i}{2\tau }}{\displaystyle _0^\tau }𝑑t_1{\displaystyle _0^{t_1}}𝑑t_2[\stackrel{~}{V}(t_1),\stackrel{~}{V}(t_2)]`$ (9)
etc. Complete expressions for the higher order terms are given in . Our objective is to find a pulse sequence which satisfies the requirement that it perform an ideal rotation under ideal conditions,
$$T\underset{i=1}{\overset{m}{}}U_i(\tau _i)=U_0(\tau )$$
(10)
while minimising $`V^{(0)}`$. We do not consider higher order terms as using more equations would require more variables in the solution and we wish to restrict the length of our composite sequence to three pulses. Here $`U_0(\tau )`$ is the propagator of an ideal uncomposed pulse and the $`U_i(\tau _i)`$ are the constituent propagators of the ideal composite pulse. Ordinarily, a numerical search for a solution must be conducted for each set of target values, $`\theta `$ and $`\varphi `$, but an analytic solution exists if the axes of rotation are taken to be $`\varphi _1=\varphi `$, $`\varphi _2=\pi +\varphi `$, and $`\varphi _3=\varphi `$. While this is nominally a loss of generality, the solutions we find are perfectly compensated to first order, and so no more general solution would be better to first order. The solutions also have reasonable properties with respect to RF field inhomogeneity.
With these phases, it follows trivially from equation (10) that
$$\theta _2=\theta _1+\theta _3\theta $$
(11)
Working out $`V^{(0)}`$ with these values of $`\varphi `$ and $`\theta _2`$ and setting $`V_y^{(0)}`$ and $`V_z^{(0)}`$ to zero (it happens that $`V_x^{(0)}`$ is already zero) gives a restriction on $`\theta _3`$,
$$\theta _3=\pm \theta _1+2n\pi n=0,\pm 1,\pm 2,\mathrm{}$$
(12)
Taking the positive result, $`\theta _3=\theta _1+2n\pi `$, and back-substituting gives two solution families,
$`\theta _1`$ $`=`$ $`\pm \mathrm{arccos}\left({\displaystyle \frac{\left(1\mathrm{cos}\theta \right)^2\pm \mathrm{sin}\theta \sqrt{\left(1\mathrm{cos}\theta \right)\left(7+\mathrm{cos}\theta \right)}}{4\left(1\mathrm{cos}\theta \right)}}\right)+2n\pi `$ (13)
$`\theta _1`$ $`=`$ $`\pm \mathrm{arcsin}\left({\displaystyle \frac{\mathrm{sin}\theta \left(1+\mathrm{cos}\theta \sqrt{\left(1+\mathrm{cos}\theta \right)\left(7+\mathrm{cos}\theta \right)}\right)}{4\left(1+\mathrm{cos}\theta \right)}}\right)+2n\pi `$ (14)
These solutions are equal in the region $`0\theta _{\text{target}}\pi `$ for the positive root and a consistent choice of signs. (This is easily checked by working out the Taylor expansions about 0, and less easily checked by working out the difference between equations (13) and (14) using an identity for $`\mathrm{arcsin}\alpha \mathrm{arccos}\beta `$.)
Although we have not explicitly considered the effects of RF field inhomogeneity when determining the $`\varphi _i`$ and $`\theta _1`$, it happens they are trivial for the rotation axes we have chosen. Using $`V_{\text{rf}}=\delta \omega _1\left(I_x\mathrm{cos}\varphi \left(t\right)+I_y\mathrm{sin}\varphi \left(t\right)\right)`$, where $`\delta \omega _1`$ is a measure of the field inhomogeneity, for $`V`$ and working out $`V^{(0)}`$ with the $`\varphi _i`$ specified above, we get
$`V_x^{(0)}`$ $`=`$ $`\delta \omega _1`$ (15)
$`V_y^{(0)}`$ $`=`$ $`0`$ (16)
$`V_z^{(0)}`$ $`=`$ $`0`$ (17)
Therefore, equations (13) and (14) can be used exclusively to determine $`\theta _1`$. We follow Tycko and choose $`n=1`$ so that $`2\pi \theta _14\pi `$. The final rotation angles are
$`\theta _1`$ $`=`$ $`\mathrm{arccos}\left({\displaystyle \frac{\left(1\mathrm{cos}\theta \right)^2+\sqrt{\left(1\mathrm{cos}\theta \right)\left(7+\mathrm{cos}\theta \right)}\mathrm{sin}\theta }{4\left(1\mathrm{cos}\theta \right)}}\right)+2\pi `$ (18)
$`\theta _2`$ $`=`$ $`\theta _1+\theta _3\theta `$ (19)
$`\theta _3`$ $`=`$ $`\theta _12\pi `$ (20)
Within this family, the quality of the pulses is remarkably insensitive to to actual details of the pulse sequence. As long as equation (11) is strictly satisfied, any sequence with angles vaguely like those given by equation (20) will produce broad-band rotations.
Figure 6 shows fidelity plots for Tycko’s original $`90^{}`$ composite pulse, and a new pulse derived using our analytic method. For comparison purposes, fidelity plots for the equivalent uncompensated pulses are included as well. The fidelity was evaluated using quaternions as described in reference .
The operation of the sequence whose fidelity is plotted in Figure 6(d) is illustrated in Figure 7. The resonance offset was chosen to be fairly large so that the mechanism of the composite pulse is clear.
The composite pulses we have derived have good properties in both quantum computation and spectroscopic contexts, but it may be possible to produce pulses with properties that are even more favourable for quantum computation if we stop using spectroscopy as our implicit model. Consider equation (10) again. Although it seems like an obvious condition that the behaviour of the pulse should be ideal in the absence of errors, this is misleading. We would never want to use a composite pulse in a situation where all pulses were on-resonance, and so there is no *a priori* reason to design pulses which perform well in the event of this non-existent ideality. The resonance offsets for quantum computation will always be known exactly, so we would like to be able to design pulse sequences which perform best as these known offsets rather than at the zero offset. This is a problem we are currently considering.
## V Conclusions
Composite pulses show great promise for reducing data errors in NMR quantum computers. More generally, any implementation of a quantum computer must be concerned on some level with rotations on the Bloch sphere, and so composite pulse techniques may find very broad application in quantum computing. Composite pulses are not, however, a panacea, and some caution must be exercised in their use.
## ACKNOWLEDGMENTS
JAJ is a Royal Society University Research Fellow. HKC thanks NSERC (Canada) and the TMR programme (EU) for their financial assistance. We thank Starlab (Riverland NV) for their support. The OCMS is supported by the UK EPSRC, BBSRC, and MRC. |
no-problem/9911/cond-mat9911006.html | ar5iv | text | # Instantons in the working memory: implications for schizophrenia.
## I INTRODUCTION
Long-term memory is thought to be stored in biochemical modulation of the inter-neuron synaptic connections. As each neuron forms about $`10^4`$ synapses such a mechanism of memory seems to be very effective, allowing potentially the storage of $`>10^4`$ bits per neuron. However, although the synaptic modifications persist for a long time, it may take as long as minutes to form them. Since the external environment operates at much shorter time scales, the synaptic plasticity is virtually useless for the short-term needs of survival. Such needs are satisfied by the working memory (WM), which is stored in the state of neuronal activity, rather than in modification of synaptic conductances (Miyashita, 1988; Miyashita and Chang, 1988; Sakai and Miyashita, 1991; Funahashi et al., 1989; Goldman-Rakic et al., 1990; Fuster, 1995).
WM is believed to be formed by recurrent neural circuits (Wilson and Cowan, 1972; Amit and Tsodyks, 1991a, b; Goldman-Rakic, 1995). A recurrent neural network can have two stable states (attractors), characterized by low and high firing frequencies. External inputs produce transitions between these states. After transition the network maintain high or low firing frequency during the delay period. Such a bipositional “switch” can therefore store one bit of information.
During the last few years it has become clear that NMDA receptor is critically involved in the mechanism of WM. It is evident from the impairments of the capacity to perform the delayed response task produced by the receptor blockade (Krystal et al., 1994; Adler et al., 1998, Pontecorvo et al., 1991; Cole et al., 1993; Aura et al., 1999). Similarly the injection of the NMDA receptor antagonists brings about weakening of the delayed activity demonstrated in the electrophysiological studies (Javitt et al., 1996, Dudkin et al., 1997a). At the same time intracortical perfusion with glutamate receptor agonists both improves the performance in the delayed response task and increases the duration of WM information storage (Dudkin et al., 1997a and b). Such evidence is especially interesting since administration of NMDA receptor antagonists (PCP or ketamine) reproduces many of the symptoms of schizophrenia including the deficits in WM.
Out of many unusual properties of NMDA channel two may seem to be essential for the memory storage purposes: the non-linearity of its current-voltage characteristics and high affinity to the neurotransmitter, resulting in long lasting excitatory post synaptic current (EPSC). The former have been recently implicated in being crucial for WM (Lisman et al., 1998). It was shown that by carefully balancing NMDA, AMPA, and GABA synaptic currents one can produce an N-shaped synaptic current-voltage characteristic, which complemented by the recurrent neural circuitry results in the bistability. This approach implies therefore that a single neuron can be bistable and maintain a stable membrane state corresponding to high or low firing frequencies (Camperi and Wang, 1998). Such a bistability therefore is extremely fragile and can be easily destroyed by disturbing the balance between NMDA, GABA or AMPA conductances. This may occur in the experiments with the NMDA antagonist.
The single neuron bistability should be contrasted to a more conventional paradigm, in which the bistability originates solely from the network feedback. It is based on the non-linear relation between the synaptic currents entering the neuron and its average firing rate (Amit and Tsodyks, 1991a, b; Amit and Brunel, 1997). Such a mechanism is not sensitive to the membrane voltage and therefore can be realized using e.g. AMPA receptor. Thus it may seem that this mechanism is ruled out by the NMDA antagonist experiments.
In this paper we study how the properties of the synaptic receptor can influence the behavior of the conventional bistable network. We therefore disregard the phenomena associated with the non-linearity of synaptic current-voltage relation (Lisman et al., 1998). In effect we study the deviations from the mean-field picture suggested earlier (see e. g. Amit and Tsodyks, 1991a, b). The main result of this paper concerns the question of stability of the high frequency attractor. This state is locally stable in the mean-field picture, i. e. small synaptic and other noises cannot kick the system out of the basin of attraction surrounding the state. However this state is not guaranteed to be globally stable. After a certain period of time a large fluctuation of the synaptic currents makes the system reach the edge of the attraction basin and the persistent memory state decay into the low frequency state. We therefore find how long can the working memory be maintained, subject to the influence of noises.
Before the network reaches the edge of attraction basin it performs an excursion into the region of parameters which is rarely visited in usual circumstances. This justifies the use of term “instanton” to name such an excursion, emphasizing the analogy with the particle traveling in the classically forbidden region in quantum mechanics. Such analogy have been used before in application to the perception of ambiguous signals by Bialek and DeWeese, 1995. The present study however is based on microscopic picture, deriving the decay times from the synaptic properties.
In the most trivial scenario the network can shut itself down by not releasing neurotransmitter in a certain fraction of synapses of all neurons. Thus all the neurons cross the border of attraction basin simultaneously. This mechanism of memory state decay is shown to be ineffective. Instead the network chooses to cross the border by small groups of neurons, taking advantage of large combinatorial space spanned by such groups. The cooperative nature of decay in this problem is analogous to some examples of tunneling of macroscopic objects in condensed matter physics (Lee and Larkin, 1978; Levitov et al., 1995).
The decay time of the memory state evaluated below increases with the size of the network growing (see Section III). This is a natural result since in a large network the relative strength of noise is small according to the central limit theorem. Therefore if one is given the minimum time during which the information has to be stored, there is the minimum size of the network that can perform this task. We calculate therefore the minimum number of neurons necessary to store one bit of information in a recurrent network. This number weakly depends on the storage time and for the majority of realistic cases is equal to 5-15.
Two complimentary measures of WM stability, the memory decay time and the minimum number of neurons, necessary to store one bit, are shown below to depend on the synaptic channel properties. The former grows exponentially with increasing duration of EPSC, which is unusually large for the NMDA channel due to high affinity to glutamate. This allows to interpret the NMDA channel blocking experiments in the framework of conventional network bistability.
In recent study Wang, 1999 considered a similar problem for the network subjected to the influence of external noises, disregarding the effects of finiteness of the probability of neurotrasmitter release. In our work we analyze the effects of synaptic failures on the stability of the persistent state. We therefore consider the internal sources of noise. Our study is therefore complimentary to Wang, 1999.
Analytical calculations presented below are confirmed by computer simulations. For the individual neurons we use the modified leaky integrate-and-fire model due to Stevens and Zador, 1998, which is shown to reproduce correctly the timing of spikes in vitro. Such realistic neurons may have firing frequencies within the range 15-30Hz for purely excitatory network in the absence of inhibitory inputs (Section II). This solves the high firing frequency problem (see e.g. Amit and Tsodyks, 1991a). The resolution of the problem is based on the unusual property of the Stevens and Zador neuron, which in contrast to the simple leaky integrator has two time scales. These are the membrane time constant and characteristic time during which the time-constant changes. The latter, being much longer than the former, determines the minimum firing frequency in the recurrent network, making it consistent to the physiologically observed values (see Section II for more detail).
## II DEFINITION OF THE MODEL AND ITS MEAN-FIELD SOLUTION
In this section we first examine the properties of single neuron, define our network model, and, finally solve the network approximately, using the mean-field approximation.
The Stevens-Zador (SZ) model of neuron is an extension of the standard leaky integrator (see e.g. Tuckwell, 1998). It is shown to accurately predict the spike timings in layer 2/3 cells of rat sensory neocortex. The membrane potential $`V`$ satisfies the leaky integrator equation with time-varying resting potential $`E`$ and integration time $`\tau `$:
$$\dot{V}=\frac{E\left(t\right)V}{\tau \left(t\right)}+I(t).$$
(1)
Here $`t`$ is the time elapsed since the last spike generated, and the input current $`I(t)`$ is measured in volts per second. When the membrane voltage reaches the threshold voltage $`\theta `$ the neuron emits a spike and the voltage is reset to $`V_{reset}`$.
This model is quite general to describe many types of neurons, differing only by the functions $`E(t)`$, $`\tau (t)`$, and the parameters $`\theta `$ and $`V_{reset}`$. For the pyramidal cells in cortical layer 2/3 of rats the functions can be fitted by
$$E\left(t\right)=E_0\mathrm{\Delta }E\left[1\mathrm{exp}\left(\alpha t/\tau _0\right)\right]$$
(2)
and
$$\tau \left(t\right)=\tau _0\left[1\mathrm{exp}\left(\alpha t/\tau _0\right)\right].$$
(3)
The parameters of the model for these cells have the following numerical values: $`E_0=32`$mV, $`\mathrm{\Delta }E=28`$mV, $`\alpha =0.3`$, $`\tau _0=10`$msec, $`\theta =22`$mV, and $`V_{reset}=32`$mV (Stevens and Zador, 1998). The resting potential and integration time with these parameters are shown in Figure 1. <sup>*</sup><sup>*</sup>*Since the time constant is zero at $`t=0`$ to resolve the singularity an implicit Runge-Kutta scheme should be used in numerical integration of (1).
In the next step we calculate the transduction function $`f=T(I)`$, relating the average external current to the average firing frequency. This function is evaluated in Appendix A and is shown in Figure 2. The closed hand expression for this function cannot be obtained. Some approximate asymptotic expressions can be found however. For large frequencies ($`f\alpha /\tau _0=30Hz`$) it is approximately given by the linear function (dashed line in Figure 2):
$$T(I)g\left(II_0\right),$$
(4)
where
$$g=\frac{\alpha }{1+\alpha }\frac{1}{\left|E_0\theta \right|},$$
(5)
and
$$I_0=\frac{\mathrm{\Delta }E}{\tau _0}+\frac{\left|E_0\theta \right|}{2\tau _0}\frac{1+\alpha }{1+2\alpha }$$
(6)
For small frequencies ($`f\alpha /\tau _0`$) we obtain:
$$T(I)\frac{\alpha }{\tau _0}\frac{1}{\mathrm{ln}\left[\left(\left|E_0\theta \right|\right)/\left(\left|E_0\theta \right|+\mathrm{\Delta }E\tau _0I\right)\right]}$$
(7)
The neuron therefore starts firing significantly when current exceeds the critical value
$$I^{}=\frac{\left|E_0\theta \right|+\mathrm{\Delta }E}{\tau _0}.$$
(8)
We ignore spontaneous activity in our consideration assuming all firing frequencies below $`5Hz`$ to be zero. We also disregard the effects related to refractory period since they are irrelevant at frequencies $`1050Hz`$.
We consider the network consisting of $`N`$ SZ neurons, establishing all-to-all connections. After each neuron emits a spike a EPSC is generated in the input currents of all cells with probability $`\kappa `$. This is intended to simulate the finiteness of probability of the neurotransmitter release in synapse. The total input current of $`n`$-th neuron is therefore
$$I_n(t)=\underset{k=1}{\overset{N}{}}s_kb_{n,s_k}j(tt_{s_k})+I_n^{ext}(t),$$
(9)
where $`k`$ enumerates the neurons making synapses on the $`n`$-th neuron (all neurons in the network), $`t_{s_k}`$ is the time of spike number $`s_k`$ emitted by cell $`k`$, and $`I_n^{ext}`$ is the external current. $`b`$ is a boolean variable equal to $`1`$ with probability $`\kappa `$ (in our computer simulations always equal to 0.3). It is the presence of this variable that distinguishes our approach from Wang, 1999. We chose the EPSC represented by $`j(t)`$ to be (see Amit and Brunel, 1997)
$$j=j_0\mathrm{exp}(t/\tau _{_{\mathrm{EPSC}}})H(t),$$
(10)
where $`H(t)=1`$, if $`t>0`$, and $`H(t)=0`$ otherwise. As evident from (10) $`\tau _{_{\mathrm{EPSC}}}`$ is the duration of EPSC. It is therefore the central variable in our consideration.
If the number of neurons in the network is large its dynimics is well described by the mean-field approximation. How large the number of neurons should be is discussed in the next Section. For the purposes of mean-field treatment it is sufficient to use the average firing frequency $`\overline{f}`$ and the average input current $`\overline{I}`$ to describe the network completely. Thus the network dynamics can be approximated by one “effective” neuron receiving the average input current and firing at the average frequency. Assume that this hypothetical neuron emits spikes at frequency $`f(t)`$ (point 1 in Figure 2). Due to the network feedback this results in the input current equal to the average of Eq. (9), displaced in time by the average duration of EPSC, i.e.
$$I(t+\tau _{_{\mathrm{EPSC}}})\overline{I}=N\kappa j_0\tau _{_{\mathrm{EPSC}}}\overline{f}(t)+\overline{I}^{ext}.$$
(11)
This corresponds to the transition between points 1 and 2 in Figure 2. At last the transition between the input current and the firing frequency is accomplished by the transduction function $`T(I)`$ (points 2 and 3). The delay due to this transition is of the order of somatic membrane time constant ($`10`$msec) and is negligible compaed to $`\tau _{_{\mathrm{EPSC}}}100`$msec. We therefore obtain the equation on the firing frequency $`\overline{f}(t+\tau _{_{\mathrm{EPSC}}})=T\left(\overline{I}(\overline{f}(t))\right)`$ (Wilson and Cowan, 1972), or using the Taylor expansion
$$\tau _{_{\mathrm{EPSC}}}\dot{\overline{f}}=T(\overline{I}(\overline{f}))\overline{f}.$$
(12)
To obtain the steady state solutions we set the time derivatives in (12) to zero
$$T(I)=I^1(I).$$
(13)
Here $`I^1(I)`$ is the function inverse to (11). It is shown in Figure 2 by the straight solid line. This equation has three solutions, two of which are stable. They are marked in by letters $`A`$ and $`B`$ in the Figure.
Point $`B`$ represents the high frequency attractor. Frequencies obtained in SZ model neurons are not too high. They are of the order of $`\alpha /\tau _0`$, i.e. in the range $`1530`$ Hz. This coinsides with the range of frequencies observed in the delayed activity experiments (Miyashita, 1988; Miyashita and Chang, 1988; Sakai and Miyashita, 1991; Funahashi et al., 1989; Goldman-Rakic et al., 1990; Fuster, 1995). The reason for the relatively low firing frequency rate is as follows. For the leaky integrator model the characteristic firing rates in the recurrent network are of the order of $`1/\tau `$, i. e. are in the range $`50100`$Hz. For the SZ neuron $`\tau `$ is even smaller (see Figure 1b), and therefore it seems that the firing rates should be larger than for leaky intergator. This however is no true, since for the SZ neuron, in contrast to the leaky integrator, there is the second time scale. It is the characteristic time of variation of the time constant $`\tau _0/\alpha 3050`$ msec (see Figure 1b). Since the time constant itself is very short it becomes irrelevant for the spike generation purposes at low firing rates and the second time scale determines the characteristic frequency. It is therefore in the range $`1530`$Hz.
Finally we would like to discuss the stability of attactor $`B`$. It can be locally stable, i.e. small noises cannot produce the transition from state $`B`$ to state $`A`$. The condition for this follows from the linerized near equilibrium Eq. (12):
$$\nu =\frac{dT}{d\overline{I}}\frac{d\overline{I}}{d\overline{f}}<1$$
(14)
Very rarely, however, a large fluctuation of noise can occur, that kicks the system out of the basin of attraction of state $`B`$. It is therefore never globally stable. This is the topic of the next Section.
## III BEYOND MEAN-FIELD APPROXIMATION
Computer simulations show that our network can successfully generate delayed activity response. An example is shown in Figure 3a where a short pulse of external current (dashed line) produced transition to the high frequency state. This state is well described by the mean-field treatment given in the previous section. This is true, however, only for the networks containing a large number of neurons $`N`$. If the size of the network is smaller than some critical number $`N^{}`$ the following phenomenon is observed (see Figure 3b). The fluctuations of the current reach the edge of the attraction basin (dotted line) and the network abruptly shuts down, jumping from high frequency to the low frequency state. The quantitative treatment of these events is the subject of this section.
Similar decay processes have been observed by Funahashi et al., 1989 in prefrontal cortex. One of the examples of such error trials is shown in Figure 4.
Although the decay is abrupt, the moment at which it occurs is not reprodusable from experiment to experiment. It is of interest therefore to study the distribution of the time interval between the initiation of the delayed activity and the moment of its decay. Our simulations and the arguments given in Appendix B show that the decay can be considered a Poisson process. The decay times have therefore an exponential distribution (see Figure 5).
Since failure to maintain the persistent activity entails the loss of the memory and incorrect performance in the delayed responce task, at least in monkey’s prefrontal cortex (Funahashi et al., 1989), one can use the conclusion about Poisson distribution to interpret some psychophysical data. In some experiments on rats performing the binary delayed matching to position tasks deterioration of WM is observed as a function of delay time (Cole et al., 1993). The detereoration was characterized by “forgetting” curve, with matching accuracy decreasing from about 100% at zero delay to approximately 70% at 30 second delay. The performance of the animal approaches the regime of random guessing with 50% of correct responses in this binary task (Figure 6). Our prediction for the shape of the “forgetting” curve that follows from the Poisson distribution of delayed activity times is
$$Correct=(1+\mathrm{exp}(t/\overline{t}))/2\times 100\%.$$
(15)
This prediction is used to fit the experimental data in Figure 6.
This Figure also shows the effect of competitive NMDA antagonist CPP. Application of the antagonist reduces the average memory retention time $`\overline{t}`$ from about 40 seconds to 15 seconds. The presence of non-compatitive antagonists impairs performance even at zero delay (Cole et al., 1993; Pontecorvoet al., 1991). Non-competitive antagonists have therefore an effect on the components of animal behavior different from WM. When the delayed component of “forgetting” curve is extracted it can be well fitted by an expression containing exponential similar to (15). Such fits also show that the average WM storage time $`\overline{t}`$ decreases with application of NMDA antagonist.
Our calculation in Appendix B show that the average memory storage time is given by
$$\overline{t}\tau _{_{\mathrm{EPSC}}}\mathrm{exp}\left[\kappa f\tau _{_{\mathrm{EPSC}}}N^2\left(\mathrm{\Delta }I/I\right)^3\right].$$
(16)
Here $`\mathrm{\Delta }I`$ is the distance to the edge of the attraction basin from the stable state and $`I`$ is the average feedback current. This result holds if $`f\tau _{_{\mathrm{EPSC}}}1`$. Because $`\overline{t}`$ is of the order of tens of seconds and $`\tau _{_{\mathrm{EPSC}}}`$ is approximately $`100`$ msec the exponential in (16) is of the order of $`10^2`$-$`10^3`$ for the realistic cases.
There are therefore two ways how synaptic receptor blockade can affect $`\overline{t}`$. First, the attenuation of the EPSC decreases the average firing frequency $`f`$. Second, it moves the system closer to the edge of the attraction basin, reducing $`\mathrm{\Delta }I`$. Both factors increase the effect of noise onto the system, decreasing the average memory storage time. This is manifested by Eq. (16).
Another consequence of the formula is the importance of NMDA receptor for the WM storage. It is based on the large affinity of the receptor to glutamate, leading to long EPSC ($`\tau _{_{\mathrm{NMDA}}}100`$ msec), compared for example to AMPA receptor ($`\tau _{_{\mathrm{AMPA}}}15`$ msec). Eq. (16) implies that if AMPA receptor is used in the bistable neural net and all other parameters ($`\kappa `$, $`f`$, $`N`$, and $`\mathrm{\Delta }I/I`$) are kept the the same, the memory storage time is equal to $`\tau _{\mathrm{AMPA}}15`$ msec. Thus it is not suprising that NMDA receptor is chosen by evolution as a mediator of WM and the highest density of the receptor is observed in places involved into the WM storage, i.e. in prefrontal cortex (Cotman et al., 1987).
We now would like to illustrate what processes lead to the decay time given by (16). Consider a simple network consisting of three neurons (Figure 7). Assume that the ratio $`\mathrm{\Delta }I/I`$ for this network is equal to 1/3. This implies that the neurons have to lose only 1/3 of their recurrent input current due to a fluctuation to stop firing. This can be accomplished by various means. Our research shows that the most effective fluctuation is as follows. Due to probabilistic nature of synaptic transmition some of the synapses release glutamate when spike arrives onto the presinaptic terminal (full circles in Figure 7) some fail to do so (open circle). It is easy to see that if the same synapse fails to release neurotransmitter in responce to any spike arriving during the time interval $`\tau _{_{\mathrm{EPSC}}}`$ the reverberations of current terminate. Indeed, the failing synapse (open circle) deprives the neuron of 1/3 of its current. This is just enough to put the input current into the neuron below the threshold. The neuron therefore stops firing. This deprives the entire network, consisting of three neurons, of 1/3 of its feedback current. Therefore the delayed activity in this network terminates.
In the the most reasonable alternative mechanism of decay the average mean-field current would reach the edge of the attraction basin i.e. would be reduced by 1/3. This can be accomplished by shutting down three synapses in three different neurons instead of one. Such mechanism is therefore less effective than the proposed above. Quantitatively the difference is manifested in reducing the exponent of the factor containg $`\mathrm{\Delta }I/I<1`$ in Eq. (16) from 3 to 2. This bring about an increase in the average storage time. The mean-field mechanism is therefore less restrictive than the proposed one and is disregarded in this paper.
Since an increase of the number of neurons in the network $`N`$ dramatically influences the memory storage time according to Eq. (16), another characteristic of the reliability of WM circuit is the minimum number of neurons $`N^{}`$ necessary to store one bit of information during time $`\overline{t}`$. We first study this quantity computationally as follows. We run the network simulation many times with the same values of $`N`$ and $`\tau _{_{\mathrm{EPSC}}}`$. We then determine the average decay time $`\overline{t}`$. Having done this we decrease or increase $`N`$ depending on wheater $`\overline{t}`$ is larger or smaller than a given value (20 sec in all our simulations). This process converges to the number of neurons necessary to sustain the delayed activity $`N^{}`$ for the given value of $`\tau _{_{\mathrm{EPSC}}}`$. The process is then repeated for different values of synaptic time-constant. However, the network feedback is always renormalized so that the firing frequency stays the same, close to the physiologically feasible value $`\overline{f}16Hz`$. This corresponds to the attractor state shown in Figure (2). The resulting dependence of $`N^{}`$ versus $`\tau _{_{\mathrm{EPSC}}}`$ is shown in Figure 8 by markers.
There are two sets of computational results in Figure 8. The higher values of $`N^{}`$ are obtained for the network with no noise in the external inputs (dots in Figure 8). The only source of noise in such a network is therefore the unreliability of synaptic connections. It appears for this case that no delayed activity can exist for $`\tau _{_{\mathrm{EPSC}}}`$ below 37 msec. The lower values of $`N^{}`$ are obtained for the case of white noise added to the external input (triangles). The amplitute of white noise is $`10\%`$ of the total input current and the correlation time is 1 msec. The limiting value of synaptic time constant for which the dalayed activity is not possible is much smaller for this case ($`5`$ msec).
This result may seem counter-intuitive. Having added the external noise we increased the viability of the high frequency state, reducing the effects of the internal synaptic noise. This however is not so suprising if one takes the neuronal synchrony into account. Synchrony of neuronal firing results in oscillations in the average input current (see inset in Figure 2a). Such oscillations periodically bring the system closer to the edge of the attraction basin, creating additional opprotunities for decay. Thus the syncronous network is less stable than the asynchronous one. External noice attenuates neuronal synchrony, smearing the ascillations of the average input current. Thus the system with noise in the external inputs should have a larger decay time and smaller critical number of neurons $`N^{}`$. This idea is discussed quantitatively below in this Section. This is similar to the stabilisation of the mean-field solutions in the networks of pulse-coupled oscillators (Abbott and Vreeswijk, 1993).
Figure 9 shows the results of similar calculations for the attractor state with a higher average firing frequency ($`\overline{f}=28`$Hz). This network shows higher reliability and smaller values of $`N^{}`$ for both synchronous and asyncronous (10% noise added) regimes. This is consistent with Eq. (16). The results of analytical calculations (solid lines) show satisfactory agreement with the computer modeling (markers). The rest of the Section is dedicated to the discussion of different aspects of the analytical calculations and their results. A more thorough treatment of the problem can be found in Appendix B.
To evaluate the minimum number of neurons in the closed hand form we solve Eq. (16) for $`N`$
$$N^{}N_1^{}\frac{1}{\sqrt{\kappa f\tau _{_{\mathrm{EPSC}}}}}\left(\frac{I}{\mathrm{\Delta }I}\right)^{3/2}\sqrt{\mathrm{log}\left(\frac{\overline{t}}{\tau _{_{\mathrm{EPSC}}}}\right)}$$
(17)
This equation is valid for $`f\tau _{_{\mathrm{EPSC}}}1`$. In the opposite case $`f\tau _{_{\mathrm{EPSC}}}1`$ it has to be amended. To obtain the correct expression in the latter case the following considerations should be taken into account.
### A Fluctuations of the average mean-field current.
It is obvious that when the mean-field attractor becomes less and less stable in the local sense, i.e. the dimensionless feedback coefficient $`\nu <1`$ \[Eq.(14)\] approaches unity, the global stability should also suffer. Indeed, if the system is weakly locally stable the fluctuations in the average current are large. This should facilitate global instability. The facilitation can be accounted for by noticing that the transition from high frequency state $`B`$ to the low frequency one $`A`$ is most probable when the average current is low. Hence to obtain the most realistic probability of transition and correct values for $`\overline{t}`$ and $`N^{}`$ one has to decrease the values of average current and distance to the edge of the attraction basin by
$$\begin{array}{c}II\delta I,\hfill \\ \\ \mathrm{\Delta }I\mathrm{\Delta }I\delta I,\hfill \end{array}$$
(18)
where the standard deviation of the average current $`\delta I`$ is calculated in Appendix B
$$\delta I=\frac{I}{N\sqrt{2(1\nu )\kappa f\tau _{_{\mathrm{EPSC}}}}}.$$
(19)
Here $`I`$ is the average current before the shift. This correction decreases the ratio $`\mathrm{\Delta }I/I`$, decreasing $`\overline{t}`$ \[see Eq. (16)\]. This implies the reduced reliability of the network due to the average current fluctuations.
### B Large combinatorial space covered by small groups of neurons attempting to cross the attraction basin edge.
In the simple example with three neurons, in principle, each of them can contribute to the process of decay. If the network is larger the number of potentially dangerous groups grows as a binomial coefficient $`C_N^n`$, where $`n=N\mathrm{\Delta }I/IN`$ is the size of the dangerous group. The correction to $`N^{}`$ can be expressed as follows (Appendix B):
$$N^{}=\sqrt{\left(N_1^{}\right)^2+\left(N_2^{}\right)^2}+N_2^{}.$$
(20)
Here $`N_2^{}`$ is the combinatorial correction
$$N_2^{}=\left(\frac{I}{\mathrm{\Delta }I}\right)^3\frac{\varphi \left(\mathrm{\Delta }I/I\right)+\left(\mathrm{\Delta }I/I\right)\mathrm{ln}A}{\kappa f\tau _{_{\mathrm{EPSC}}}}.$$
(21)
Here
$$\begin{array}{c}A=\frac{I}{\mathrm{\Delta }I\sqrt{4\pi N^{}\kappa f\tau _{_{\mathrm{EPSC}}}}},\hfill \\ \\ \varphi (x)=x\mathrm{ln}\left[1/x\right]+(1x)\mathrm{ln}\left[1/(1x)\right].\hfill \end{array}$$
(22)
Since the combinatorial contribution $`N_2^{}`$ is proportional to $`1/f\tau _{_{\mathrm{EPSC}}}`$ it is negligible compared to $`N_1^{}1/\sqrt{f\tau _{_{\mathrm{EPSC}}}}`$ at large values of $`\tau _{_{\mathrm{EPSC}}}`$. Therefore $`N^{}N_1^{}`$ as claimed by Eq. (17) in this limit. On the other hand if $`\tau _{_{\mathrm{EPSC}}}`$ is small the main contribution to the critical number of neurons $`N^{}`$ comes from $`N_2^{}`$.
The right hand sides of Eqs. (17) and 21 contains dependence on $`N^{}`$ through the shift in distance to the attraction edge $`\delta I`$ and the coefficient $`A`$. This dependence is weak however since the former represents a very small correction and the latter depends on $`N^{}`$ only logarithmically. Nevertheless to generate a numerically precise prediction we iterate Eq. (17), (20), and (17) until a consistent value of $`N^{}`$ is reached. The results are shown by lower solid lines in Figures 8 and 9. They are in a good agreement with numerical simulations in which the synchrony of firing was suppressed by external noise.
### C Neuronal synchrony.
The synchronizaton of neuronal firing can be critical for the stability of the high frequency attractor. Consider the following simple example (Wang, 1999). Consider the network consisting on only one neuron. Assume that the external input current to the neuron exceeds the firing threshold $`I^{}`$ at certain moment and the neuron starts emiting spikes. Assume that the external current is then decreased to some value below the threshold. The only possibility for the neuron to keep firing is to receive enough feedback current from itself, so that the total input current is above threshfold. Suppose then for illustration purposes that the synapse that the neuron forms on itself (autapse) is infinitely reliable, i.e. every spike arriving on the presynaptic terminal brings about the release of neurotransmitter. It is easy to see that if the duration of EPSC is much shorter than the interspike interval the persistent activity is impossible. Indeed, if EPSC is short, at the moment, when the new spike has to be generated there is almost no residual EPSC. The input current at that moment is almost entirely due to external sources, i.e. is below the threshold. The new spike generation is therefore impossible. We conclude that the persistent activity is impossible if EPSC is short, i. e. for $`f\tau _{_{\mathrm{EPSC}}}1`$. Note, that this conclusion is made assuming no noise in the system.
The natural question is why this conclusion is relevant to the network consisting of many neurons? Our computer modeling shows that in the large network the neuronal firing pattern is highly sychronized on the scales of the order of $`T=1/f`$. In essence all neurons fire simultaneously and periodically with the period $`T`$. Therefore they behave as one neuron. Hence the delayed activity in such network is impossible if $`f\tau _{_{\mathrm{EPSC}}}1`$ even in the absence of synatic noise. Thus the synchrony facilitates the decay of the delayed activity brought about by the noise. This is consistent with the results of computer simulations presented if Figures 8 and 9.
As it was mentioned above synchrony facilitates the instability by producing the oscillations in the average current and therefore by bringing the edge of the attraction basin closer. It therefore effectively decreases $`\mathrm{\Delta }I`$. The magnitude of this decrease is estimated in Appendix B to be
$$\begin{array}{c}\delta I_{syn}\frac{2I}{\pi f\tau _{_{\mathrm{EPSC}}}}\hfill \\ \\ \times \frac{f\tau _{_{\mathrm{EPSC}}}e^{1/f\tau _{_{\mathrm{EPSC}}}}\left(f\tau _{_{\mathrm{EPSC}}}+1\right)}{1e^{1/f\tau _{_{\mathrm{EPSC}}}}}\hfill \end{array}$$
(23)
and
$$\mathrm{\Delta }I\mathrm{\Delta }I\delta I_{syn}.$$
(24)
Here $`I`$ is the average current given by (18). This correction, which is the amplitude of the oscillations of the average current due to synchrony, is in good agreement with computer modeling (see insert in in Figure 3a). The synchrony does not change the average current in the network significantly. Therefore the latter should not be shifted as in (18).
These corrections imply that if the original value of $`\mathrm{\Delta }I`$ is equal to the sum $`\delta I+\delta I_{syn}`$ \[see Eq. (19)\], the effective distance to the edge of attraction basin vanishes. The delayed activity cannot be sustained under such a condition. This occurs at small values of $`\tau _{_{\mathrm{EPSC}}}`$ and determines the positions of vertical asymptotes (dashed lines) in Figures 8 and 9. This gives a quantitative meaning to the argument of impossibility of stable delayed activity in synchronous network at small values of synaptic time constant given above (see also Wang, 1999). When the synchrony is diminished by external noise, the cut off value of $`\tau _{_{\mathrm{EPSC}}}`$ is determined only by $`\mathrm{\Delta }I=\delta I`$ and is therefore much smaller. It is about 5 msec in our computer similations.
Synchrony also affects the values of the critical number of neurons for large and small $`f\tau _{_{\mathrm{EPSC}}}`$ ($`N_1^{}`$ and $`N_2^{}`$ respectively)
$$N_{1syn}^{}=\left(\frac{I}{\mathrm{\Delta }I}\right)^{3/2}\frac{1}{f\tau _{_{\mathrm{EPSC}}}}\sqrt{\frac{2\mathrm{ln}\left(t/\tau _{_{\mathrm{EPSC}}}\right)}{\kappa \left(e^{2/f\tau _{_{\mathrm{EPSC}}}}1\right)}},$$
(25)
$$N_{2syn}^{}=\left(\frac{I}{\mathrm{\Delta }I}\right)^3\frac{\varphi \left(\mathrm{\Delta }I/I\right)+\left(\mathrm{\Delta }I/I\right)\mathrm{ln}A}{e^{2/f\tau _{_{\mathrm{EPSC}}}}1}\frac{1}{\kappa f^2\tau _{_{\mathrm{EPSC}}}^2}$$
(26)
These equations are derived in Appendix B. To obtain the value of critical number of neurons $`N^{}`$ Eq. (20) should be used. Let us compare the latter equations to (17) and (21). In the limit $`f1`$ the expressions for $`N_1^{}`$ and $`N_{1syn}^{}`$ converge to the same asymptote $`N_1^{}N_{1syn}^{}`$, whereas both $`N_2^{}`$ and $`N_{2syn}^{}`$ go to zero.
Since again, as in asynchronous case, both $`N_{1syn}^{}`$ and $`N_{2syn}^{}`$ depend on the quantity that we are looking for $`N^{}`$ (however, again, very weakly), an iterative procedure should be used to determine the consistent value of $`N^{}`$. Having applied this iterative procedure we obtain the ipper solid lines in Figures 8 and 9. We therefore obtain an excellent agreement with the results of the numerical study. In addition we calculate the cut-off $`\tau _{_{\mathrm{EPSC}}}`$, below which the delayed activity is not possible for the synchronous case. For the attractor with $`\overline{f}16`$ Hz the cut-off value is $`37`$ msec, while for the higher frequency attractor ($`\overline{f}28`$ Hz) the value is $`21`$ msec. These values are shown in Figures 8 and 9 by dashed lines. We therefore conclude that the delayed activity mediated by AMPA receptor is impossible in the synchronous case.
## IV DISCUSSION
In this work we derive the relationship between the dynamic properties of the synaptic receptor channels and the stability of the delayed activity. We conclude that the decay of the latter is a Poisson process with the average decay time exponentially depending on the time constant of EPSC. Our quantitative conclusion applied to AMPA receptor, having a short EPSC, implies that it is incapable of sustaining the persistent activity in case of synchronization of firing in the network. For the case of asynchronous network one needs a large number of neurons $`>30`$ to store one bit of information with AMPA receptor. On the other hand NMDA receptor seems to stay away from these problems, providing reliable quantum of information storage with about 15 neurons for both synchronized and asynchronous case. We therefore suggest an explanation to the obvious from experiments high significance of the NMDA channel for WM.
One can conclude from our study that if the time constant of NMDA channel EPSC is further increased, the WM can be stored for much longer time. Assume that $`\tau _{_{\mathrm{EPSC}}}`$ is increased by a factor of 2, for instance, by genetic enhancement (Tang et al., 1999). If the network connectivity, firing frequency, and average currents stay the same as in the wild type, the exponential in Eq. (16) is increased by a factor of $`10^3`$. This implies that the working memory can be stored by such an animal for days, instead of minutes. Alternatively the the number of neurons responsible for the storage of quantum of information can be decreased by a factor $`0.7`$ keeping the storage time the same. This implies the higher storage capacity of the brain of mutant animals.
We predict that with normal NMDA receptor the number of neurons able to store one bit of information for $`40`$ sec is about 15. Should our theory be applicable to rats performing delayed matching to position task (Cole et al., 1993) the conclusion would be that the recurrent circuit responsible for this task contains $`<`$ 15 neurons. Of course this would imply that other sources of loss of memory, such as distraction, are not present.
The use of very simple model network allowed us to look into the nature of the global instability of delayed activity. The principal result of this paper is that the unreliability of synaptic conductance provides the most effective channel for the delayed activity decay. We propose the optimum fluctuation of the synaptic noises leading to the loss of WM. Decay rates due to such a fluctuation agree well with the results of the numerical study (Section III). Thus any theory or computer simulation that does not take the unreliability of synapses into account does not reproduce the phenomenon exactly.
We also studied the decay process in the presence of noise in the afferent inputs. The study suggests that the effect of noise on the WM storage reliability is not monotonic. Addition of small white noise ($`<10`$% of the total external current) increases the reliability, by destroying synchronization. It is therefore beneficial for the WM storage. Further increase of noise ($`>12`$%) destroys WM, producing transitions between the low and high frequency states. We conclude therefore that there is an optimum amount of the afferent noise, which on one hand smoothes the synchrony out and on the other hand does not produce the decay of delayed activity itself. Further work is needed to study the nature of influence of the external noise in the case of unreliable synapses.
If the afferent noise is not too large the neurons fire in synchrony. The natural consequences of synchrony are the decrease of the coefficient of variation of the interspike interval for single neuron and the increase of the crosscorrelations between neurons. The latter prediction is consistent with findings of some multielectrode studies in monkey prefrontal cortex (see Dudkin et al., 1997a). On the other hand sinchrony may also be relevant to the phenomenon of temporal binding in the striate cortex (Engel et al., 1999; Roskies, 1999). We argue therefore that temporal binding with the precision of dosens of millisconds can be accomplished by formation of recurrent neural networks.
We also suggest the solution to the high firing frequency problem by using a more precise model of the spike generation mechanism, i.e. the leaky integrator model with varying in time resting potential and integration time. The minimum firing frequency for the recurrent network based on such model is determined by the rate of variation of the potential and time-constant and is within the range of physiologically observed values. Further experimental work is needed to see if the model is applicable to other types of neurons, such as inhibitory cells.
In conclusion we have studied the stability of delayed activity in the recurrent neural network subjected to the influence of noise. We conclude that the global stability of the persistent activity is affected by properties of synaptic receptor channel. NMDA channel, having a long EPSC duration time, is a reliable mediator of the delayed responce. On the other hand AMPA receptor is much less reliable, and for the case of synchronized firing in principle cannot be used to sustain responce. Effect of the NMDA channel blockade on the WM task performance is discussed.
## V ACKNOLEDGEMENTS
The author is grateful to Thomas Albright, Paul Tiesinga, and Tony Zador for discussions and numerous helpful suggestions. This work was supported by the Alfred P. Sloan Foundation.
## A Single-neuron transduction function
In this Appendix we calculate the single-neuron transduction function. The first step is to find the membrane voltage as a function of time. From Eq. (1) using the variation of integration constant we obtain
$$\begin{array}{c}V(t)=V_{reset}e^{S(t)}\hfill \\ \\ +_{t_0}^t𝑑t^{}e^{S(t^{})S(t)}\left[I(t^{})+E(t^{})/\tau (t^{})\right].\hfill \end{array}$$
(A1)
Here
$$S(t)=_{t_0}^t𝑑t^{}/\tau (t^{}),$$
(A2)
and we introduced the refractory period both for generality and to resolve the peculiarity at $`t=0`$. For constant current and functions given by (2) and (3) the expression for voltage can be further simplified:
$$\begin{array}{c}V(t)=V_{reset}e^{S(t)}+E_0\left[1e^{S(t)}\right]\hfill \\ \\ +\left(I\frac{\mathrm{\Delta }E}{\tau _0}\right)\frac{\tau _0}{\alpha }\frac{L_{1/\alpha }\left(e^{\alpha t/\tau _0}\right)e^{S(t)}}{\left(e^{\alpha t_0/\tau _0}1\right)^{1/\alpha }},\hfill \end{array}$$
(A3)
where
$$S(t)=\frac{1}{\alpha }\mathrm{ln}\left(\frac{e^{\alpha t/\tau _0}1}{e^{\alpha t_0/\tau _0}1}\right)$$
(A4)
and $`L_n(x)`$ is defined for integer $`n`$ by
$$\begin{array}{c}L_0(x)=\mathrm{ln}(x),\hfill \\ \\ L_n(x)=\frac{1}{n}(x1)^nL_{n1}(x),\hfill \end{array}$$
(A5)
and is obtained for fractional $`n`$ by analytical continuation. For example
$$\begin{array}{c}L_1(x)=x1\mathrm{ln}(x),\hfill \\ \\ L_2(x)=(x1)^2/2x+1\mathrm{ln}(x),\hfill \\ \\ L_3(x)=(x1)^3/3(x1)^2/2+x1\mathrm{ln}(x),\hfill \\ \\ \mathrm{}\hfill \end{array}$$
(A6)
Solving the equation $`V(t)=\theta `$ produces the interspike interval $`t`$ and frequency $`f=1/t`$. The solution cannot be done in the closed form. Some asymptotes can be calculated however. The calculation depends on the value of parameter $`\xi =\alpha t/\tau _0`$.
i) $`\xi =\alpha t/\tau _01`$.
In this limit
$$e^{S(t)}\frac{\left(e^{\alpha t_0/\tau _0}1\right)^{1/\alpha }}{\xi +\xi ^2/2},$$
(A7)
is very small and can be neglected everywhere, except for when multiplied by large factor $`L_{1/\alpha }`$. The latter product
$$\frac{L_{1/\alpha }\left(e^{\alpha t/\tau _0}\right)e^{S(t)}}{\left(e^{\alpha t_0/\tau _0}1\right)^{1/\alpha }}\frac{\alpha }{1+\alpha }\left(\xi +\frac{\xi ^2}{2}\right)\frac{\alpha }{1+2\alpha }\xi ^2.$$
(A8)
The equation on the interspike interval is
$$\theta =E_0+(\tau _0I\mathrm{\Delta }E)\left[\frac{\xi }{1+\alpha }\frac{\xi ^2}{2}\frac{1}{(1+\alpha )(1+2\alpha )}\right].$$
(A9)
Solving this quadratic equation with respect to $`\xi `$ we obtain the asymptote given by Eq. (4).
ii) $`\xi 1`$.
In this case the solution can be found directly from Eq. (1) by assuming $`\dot{V}=0`$. Solving the resulting algebraic equation for $`t`$ we obtain Eq. (7) above.
## B Decay of the high frequency attractor
### 1 Asynchronous case: derivation of $`N_1^{}`$ and $`N_2^{}`$
The average and standard deviation of the input current of each neuron, given by Eqs. (9) and (10) are
$$\overline{I}=N\kappa f\tau _{_{\mathrm{EPSC}}}j_0$$
(B1)
and
$$\overline{\delta I_n^2}=\overline{(I_n\overline{I})^2}=N\kappa f\tau _{_{\mathrm{EPSC}}}j_0^2/2.$$
(B2)
The distribution of $`I_n`$ is therefore, according to the central limit theorem
$$p(I_n)=\frac{1}{\sqrt{2\pi \overline{\delta I_n^2}}}\mathrm{exp}\left[\frac{(I_n\overline{I})^2}{2\overline{\delta I_n^2}}\right].$$
(B3)
The probability that the input current is below threshold $`I^{}`$, i.e. the neuron does not fire, is given by the error function derived from distribution (B3)
$$P=_{\mathrm{}}^I^{}𝑑I_np(I_n)Ae^{\kappa f\tau _{_{\mathrm{EPSC}}}N\mathrm{\Delta }I^2/\overline{I}^2},$$
(B4)
where $`A=\overline{I}/\mathrm{\Delta }I\sqrt{4\pi \kappa f\tau _{_{\mathrm{EPSC}}}N}`$, and $`\mathrm{\Delta }I=\overline{I}I^{}`$. In derivation of (B4) we used the asymptotic expression for the error function
$$_{\mathrm{}}^y𝑑x\frac{e^{x^2/2\sigma ^2}}{\sigma \sqrt{2\pi }}\frac{\sigma }{|y|\sqrt{2\pi }}e^{y^2/2\sigma ^2},$$
(B5)
when $`y\sigma `$.
The reverberations of current will be impossible if $`n=N\mathrm{\Delta }I/\overline{I}`$ neurons are below threshold. When this occurs the feedback current to each neuron is reduced with respect to the average current by $`\mathrm{\Delta }I`$, i.e. is below threshold, and further delayed activity is impossible. The probability of this event is given by the binomial distribution:
$$p_0C_N^nP^n.$$
(B6)
Here the binomial coefficient $`C_N^n=N!/(Nn)!n!`$ accounts for the large number of groups of neurons that can contribute to the decay.
Since the input currents stay approximately constant during time interval $`\tau _{_{\mathrm{EPSC}}}`$, we brake the time axis into windows with the duration $`\mathrm{\Delta }t\tau _{_{\mathrm{EPSC}}}`$. Denote by $`p`$ the average probability of decay of the high frequency activity during such a little window. Assume that $`k`$ windows have been passed since the delayed activity commenced. The probability that the decay occurs during the $`k`$-th time window, i.e. exactly between $`t=k\mathrm{\Delta }t`$ and $`t=(k+1)\mathrm{\Delta }t`$, is
$$\rho (t)\mathrm{\Delta }t=(1p)^kp.$$
(B7)
Here $`(1p)^k`$ is the probability that the decay did not happen during either of $`k`$ early time intervals. After simple manipulations with this expression we obtain for the density of probability of decay as a function of time
$$\rho (t)=e^{pt/\mathrm{\Delta }t}/p\mathrm{\Delta }t.$$
(B8)
This is a Poisson distribution and decay is a Poisson process. The reason for this is that the system retains the values of input currents during the time interval $`\mathrm{\Delta }t\tau _{_{\mathrm{EPSC}}}100`$ msec. Therefore all processes separated by longer times are independent. Since the decay time is of the order of $`10`$-$`100`$ seconds, the attempts of system to decay at various times can be considered independent. We finally notice that since the values of current are preserved on the scales of $`\tau _{_{\mathrm{EPSC}}}`$ we can conclude that
$$pp_0$$
(B9)
given by Eq. (B6). The average decay time can then be estimated using (B8)
$$\frac{\overline{t}}{\tau _{_{\mathrm{EPSC}}}}p_0^1.$$
(B10)
In the limit $`f\tau _{_{\mathrm{EPSC}}}1`$ the minimum size of the network necessary to maintain the delayed activity is small. We can assume therefore $`N`$ to be small in this limit. The same assumption can be made for $`n<N`$. Thus for an effective network (using the neurons sparingly) we can assume the combinatorial term in Eq. (B6) to be close to unity. Since in the expression for $`p_0C_N^nA^n\mathrm{exp}\left(n\kappa f\tau _{_{\mathrm{EPSC}}}N\mathrm{\Delta }I^2/\overline{I}^2\right)`$ it is also compensated by the small prefactor of the exponential $`A^n<1`$ \[see Eq. (B4)\], we can assume $`C_N^nA^n1`$ for the effective network. Therefore in the limit $`f\tau _{_{\mathrm{EPSC}}}1`$ the main dependence of the parameters of the model is concentrated in the exponential factor. This set of assumptions in combination with Eq. (B10) leads to the expression (16) in the main text. If more precision is needed Eq. (B10) can be used directly.
To find the critical number of neurons $`N^{}`$ we calculate logarithm of both sides of Eq. (B10). We obtain
$$\mathrm{ln}\left(\overline{t}/\tau _{_{\mathrm{EPSC}}}\right)\mathrm{ln}C_N^{}^nn\mathrm{ln}p_0.$$
(B11)
Using asymptotic expression for the logarithm of the binomial coefficient, which can be derived from the Stirling’s formula,
$$\begin{array}{c}\mathrm{ln}C_N^nN\varphi \left(n/N\right);N,n1;\hfill \\ \\ \varphi (x)=x\mathrm{ln}\left[1/x\right]+(1x)\mathrm{ln}\left[1/(1x)\right],\hfill \end{array}$$
(B12)
we derive the quadratic equation for $`N^{}`$
$$\begin{array}{c}\left(\frac{\mathrm{\Delta }I}{I}\right)^3\kappa f\tau _{_{\mathrm{EPSC}}}\left(N^{}\right)^2\hfill \\ \\ N^{}\left[\varphi \left(\frac{\mathrm{\Delta }I}{I}\right)+\frac{\mathrm{\Delta }I}{I}A\right]\mathrm{ln}\frac{\overline{t}}{\tau _{_{\mathrm{EPSC}}}}0.\hfill \end{array}$$
(B13)
Solving this equation we obtain the set of equations (17), (20), and (21) in the main text.
### 2 Synchronous firing case
In this subsection we assume that all the neurons in the network fire simultaneously. The important quantities which will be studied are the time dependence of the averaged over network current $`\overline{I}(t)`$ and the fluctuations of the input current into single neuron $`\overline{\delta j_n^2}`$. The former is responsible for the shift in the distance to the edge of the attraction basin (Eq. (24)), the latter determines the average decay time, as follows from the previous subsection.
Assume that the neuron fired at $`t=0`$. The average number of EPSC’s arriving to the postsynaptic terminal is $`\kappa N`$ due to finitness probability of the synaptic vesicle release $`\kappa `$ and all-to-all topology of the connections. The average current at times $`0<t<T=1/f`$ is contributed by the spikes at $`t=0`$ and by all previous spikes at times $`Tk`$, $`k=1,2,\mathrm{}`$:
$$\begin{array}{c}\overline{I}(t)=I_0e^{t/\tau _{_{\mathrm{EPSC}}}}+I_0e^{(t+T)/\tau _{_{\mathrm{EPSC}}}}+\mathrm{}\hfill \\ \\ =I_0e^{t/\tau _{_{\mathrm{EPSC}}}}/(1e^{1/f\tau _{_{\mathrm{EPSC}}}}),\hfill \end{array}$$
(B14)
where we introduced the notation $`I_0=\kappa Nj_0`$. The average over time value of this averaged over neurons current is $`<\overline{I}>=I_0f\tau _{_{\mathrm{EPSC}}}`$, where by angular brackets we denote the time average: $`<A>=_0^TA(t)𝑑t/T`$. The average current $`\overline{I}`$ therefore experiences oscillations between $`\overline{I}_{max}=\overline{I}(0)`$ and $`\overline{I}_{min}=\overline{I}(T)`$. The minimum value of $`\overline{I}`$ is below the average current $`<\overline{I}>`$. This minimum value, according to the logic presented in the main text, determines the shift of the distance to the edge of the attraction basin. This shift is therefore $`<\overline{I}>\overline{I}_{min}`$. Our computer modeling however shows that the actual amplitude of the current oscillations is consistently about $`60`$% of the value predicted by this argument. This was tested at various values of parameters. The explanation of this $`60`$% factor is as follows. In case if all the neurons fire simultaneously the shape of the dependence of the average current versus time is saw-like. However the spikes do not fire absolutely simultaneously. The uncertainty in the spiking time is of the order of $`T`$. This uncertainty, which is intrinsic to the recurrent synaptic noises, and therefore is difficult to calculate exactly, smears out the saw-like dependence of the average current of time (see inset in Figure 3a). Approximately this smearing can be accounted for by dumping the higher harmonics of the saw-like dependence. When only principal harmonic remains, the amplitude of the saw-like curve is reduced by a factor $`2/pi0.6`$. This is consistent with the numerical result. We therefore conclude that the shift of $`\mathrm{\Delta }I`$ is
$$\delta I_{syn}0.6\left[<\overline{I}>\overline{I}(T)\right],$$
(B15)
what leads to Eq. (23) in the text.
We now turn to the calculation of the standard deviation of the input current into one neuron. Similar to (B14), using the central limit theorem we obtain
$$\delta I_n^2(t)=\underset{k=0}{\overset{\mathrm{}}{}}j_0^2\kappa Ne^{(2t+2Tk)/\tau _{_{\mathrm{EPSC}}}}$$
(B16)
This quantity is most important at $`tT`$ when the average current reaches its minimum and the neuron stops firing with maximum probability. Performing the summation we therefore obtain
$$\delta I_n^2(T)=\frac{\kappa I^2}{(\kappa f\tau _{_{\mathrm{EPSC}}})^2N}\frac{1}{e^{2/f\tau _{_{\mathrm{EPSC}}}}1}.$$
(B17)
Substituting this value into Eq. (B4) in the previous subsection and repeating the subsequent derivation we obtain Eqs. (25) and (26) in the main text.
### 3 Fluctuations of the average current
In this subsection we calculate the fluctuations of the average netwotk current. We consider asynchronous case for simplicity. The conclusions are perfectly good for synchronous case, for the reasons that will become clear later in this subsection, and agree well with computer similations.
The average current satisfies linarized equation similar to linearized version of Eq. (12)
$$\tau _{_{\mathrm{EPSC}}}\mathrm{\Delta }\dot{\overline{I}}(t)=(\nu 1)\mathrm{\Delta }\overline{I}(t)+\xi (t).$$
(B18)
Here $`\mathrm{\Delta }\overline{I}(t)`$ is the deviation of the averaged over network current from the equilibrium value and $`\xi (t)`$ is the noise. The unitless network feedback coefficient $`\nu <1`$ is defined by (14). As evident from this equation $`\mathrm{\Delta }\overline{I}(t)`$ has a slow time constant $`\tau _{_{\mathrm{EPSC}}}/(1\nu )`$ \[since $`1/(1\nu )>5`$ in our simulations\]. On the other hand the noise $`\xi (t)`$ is determined by synapses and has a correlation time that is relatively small ($`\tau _{_{\mathrm{EPSC}}}`$).
The correlation function of noise in the average current can be found from Eq. (9)
$$\mathrm{\Xi }(t)\xi (t)\xi (0)=\frac{\overline{\delta I_n^2}}{N}\mathrm{exp}(|t|/\tau _{_{\mathrm{EPSC}}}).$$
(B19)
Here angular brackets imply averaging over time and the value of $`\mathrm{\Xi }(t=0)`$ follows from Eq. (B2) and the central limit theorem (dispersion of the average is equal to the dispersion of each of the homogenious constituents $`I_n`$ divided by the number of elements $`N`$). We conclude therefore that
$$\mathrm{\Xi }(t=0)=\frac{I^2}{2f\tau _{_{\mathrm{EPSC}}}N^2}.$$
(B20)
The correlation function of $`\mathrm{\Delta }\overline{I}`$ can then be easily found from Eq. (B18) using the Fourier transform. Define $`C(t)\mathrm{\Delta }\overline{I}(t)\mathrm{\Delta }\overline{I}(0)`$. Then
$$C(\omega )=\left|\mathrm{\Delta }\overline{I}(\omega )\right|^2=\frac{\mathrm{\Xi }(\omega )}{(1\nu )^2+\left(\omega \tau _{_{\mathrm{EPSC}}}\right)^2}$$
(B21)
Since
$$\mathrm{\Xi }(\omega )=\frac{2i\tau _{_{\mathrm{EPSC}}}\mathrm{\Xi }(t=0)}{1+\left(\omega \tau _{_{\mathrm{EPSC}}}\right)^2},$$
(B22)
the expression for $`C(t)`$ is readily obtained by inverting the Fourier transform (B21). In the limit $`1\nu 1`$, which holds in our simulations the answer is
$$C(t)=\frac{\mathrm{\Xi }(t=0)}{1\nu }\mathrm{exp}\left[(1\nu )|t|/\tau _{_{\mathrm{EPSC}}}\right].$$
(B23)
The value of $`C`$ taken at $`t=0`$ determines the standard deviation of the average current (19).
The fluctuations described by the correlation function (B23) have a large correlation time compared to the firing frequency. We conclude therefore that synchrony should not affect the long range component of the correlation function. |
no-problem/9911/physics9911017.html | ar5iv | text | # The RR interval spectrum, the ECG signal and aliasing
## I Introduction
The R-R interval spectral analysis is usually based on heart rate data collected in two ways. In one method the data are collected by analog to digital conversion of the ECG signal and computer evaluation of the R-R intervals from the ECG signal. In the second method, devices are used whose output is the R-R interval alone. The advantage of the first method is the control of accuracy and flexibility of the evaluations. The second method has the advantage of storing smaller amount of data, and it can be easily used on-line.
In the first method, usually the number of collected data (sampled ECG signal) is of two to three orders of magnitude larger than the R-R interval data. Thus if only R-R interval is analyzed a large amount of data is unused. In this paper we are trying to take advantage of the ECG sampled signal and to derive new information in addition to the conventional R-R interval analysis ,,,, .
The ECG signal spectrum is bounded below the frequency f<sub>B</sub> by using an electronic filter and sampled at rate larger than 2f<sub>B</sub>, thus excluding aliasing from spectral analysis. A similar procedure cannot be applied to the R-R interval spectral analysis, and in this case an aliasing is possible. One of our main efforts in this paper is devoted to the problem of how to detect aliasing in the R-R interval spectral analysis.
In order to get an insight, we performed an experiment, in which the ECG signal of one of the authors (A.G) was detected while the breathing rate was larger than half the heart rate. A constant breathing rate for a time exceeding 5 minutes was monitored with good accuracy using a special breathing procedure with a metronome. The results show distinctively a very sharp peak in the spectral analysis of the ECG signal and corresponding (diffused) aliasing peaks in the R-R interval spectral analysis.
The spectral analysis of the ECG signal was performed with the standard FFT procedures. The spectral analysis of the R-R intervals was performed with several techniques in order to take into consideration that the data were unevenly sampled. This is presented in section 2. In section 3 we discuss the possibility of aliasing in the spectral analysis of the R-R intervals. In section 4 we compare power estimations of ECG’s and R-R intervals of 3 experiments. In section 5 we analyze the results. In section 6 summary and conclusions are presented.
## II Spectral Analysis of Unevenly Sampled Data
The methods of spectral analysis are well developed for evenly sampled data ,. The R-R interval data are unevenly sampled in time. In most cases an analysis is performed with respect to beat numbers which are evenly spaced. We will below justify this method using least squares principles. But as was recently indicated by Laguna et Al. , the resampling of data is causing the appearance of additional harmonics. They recommend to use a method developed by N. R. Lomb . The errors of resampling the beats can, to large extent, be overcome by using a cubic spline interpolation. In this work we are suggesting a new method of treating unevenly sampled data, which, unexpectedly, gave good results beyond the Nyquist frequency.
### A Analysis according to beat numbers
Let us assume that the RR intervals are given at unevenly sampled times $`t_n`$, with the values $`s\left(t_n\right),`$ where $`n`$ is the beat number, $`n=1\mathrm{}N`$. Let us divide the interval $`[t_1,t_N]`$ into equal subintervals
$$\mathrm{\Delta }\tau =\frac{t_Nt_1}{N1},$$
(1)
and let us generate in the interval $`[t_1,t_N]`$ evenly sampled times:
$$\tau _n=(n1)\mathrm{\Delta }\tau +t_1.$$
(2)
We will use the discrete time Fourier transform (DFT) for a basis formed from the evenly sampled times $`\tau _n`$. We will assume that
$$s(t_n)=\frac{1}{N}\underset{k=1}{\overset{N}{}}S_k\mathrm{exp}\left(i\omega _k\tau _n\right),\omega _k=2\pi \left(k1\right)/\left(N\mathrm{\Delta }\tau \right).$$
(3)
The coefficients $`S_k`$ will be determined by minimizing the expression
$$\sigma =\underset{n=1}{\overset{N}{}}\left\{\left[s(t_n)\frac{1}{N}\underset{k=1}{\overset{N}{}}S_k\mathrm{exp}\left(i\omega _k\tau _n\right)\right]\left[s(t_n)\frac{1}{N}\underset{k=1}{\overset{N}{}}S_k^{}\mathrm{exp}\left(i\omega _k\tau _n\right)\right]\right\}$$
(4)
with the result
$$S_k=\underset{n=1}{\overset{N}{}}s\left(t_n\right)\mathrm{exp}\left(i\omega _k\tau _n\right).$$
(5)
Eqns. 5 and 3 can be handled easily with standard FFT programs. This is the usual procedure which is adopted in most of the papers dealing with R-R interval analysis.
### B Other Methods
FFT can be applied more efficiently if the unevenly sampled data are interpolated at evenly spaced intervals of Eq. 2. The cubic spline interpolation is one of the good ways to do it.
The Lomb method was extensively analyzed in ref. . We give here only the formulae in the form of the Lomb normalized periodogram
$$P_X\left(\omega _k\right)=\frac{1}{2\sigma ^2}\left\{\frac{\left[_{n=1}^N\left[s\left(t_n\right)\overline{s}\right]\mathrm{cos}\left(\omega _k\left(t_n\tau \right)\right)\right]^2}{_{n=1}^N\mathrm{cos}^2\left(\omega _k\left(t_n\tau \right)\right)}+\frac{\left[_{n=1}^N\left[s\left(t_n\right)\overline{s}\right]\mathrm{sin}\left(\omega _k\left(t_n\tau \right)\right)\right]^2}{_{n=1}^N\mathrm{sin}^2\left(\omega _k\left(t_n\tau \right)\right)}\right\}$$
(6)
where $`\overline{s}`$ and $`\sigma ^{2\text{ }}`$are the mean and variance of the data and the value of $`\tau `$ is defined as
$$\mathrm{tan}\left(2\omega _k\tau \right)=\frac{_{n=1}^N\mathrm{sin}\left(2\omega _kt_n\right)}{_{n=1}^N\mathrm{cos}\left(2\omega _kt_n\right)}$$
(7)
### C Non-Uniform Discrete Fourier Transform (NUDFT)
We present here a new method of treating unevenly spaced events which we call the ”non-uniform discrete Fourier transform” (NUDFT).
Let us assume that $`s\left(\tau _n\right)`$ are the exact values of the signal at the points given by Eq. 2. The corresponding DFT is
$$S_k=\underset{n=1}{\overset{N}{}}s\left(\tau _n\right)\mathrm{exp}\left(i\omega _k\tau _n\right).$$
(8)
Our aim is to find a good approximation to this expression in terms of the unevenly sampled signal $`s(t_n)`$.
We start with the Euler summation formula
$$\underset{n=1}{\overset{N}{}}f\left(\tau _n\right)=\frac{1}{\mathrm{\Delta }\tau }_{\tau _1}^{\tau _N}f\left(\tau \right)𝑑\tau +\frac{1}{2}\left[f\left(\tau _1\right)+f\left(\tau _N\right)\right]+\frac{\mathrm{\Delta }\tau }{12}\left[f^{}\left(\tau _N\right)f^{}\left(\tau _1\right)\right]+O(\mathrm{\Delta }\tau ^2)$$
(9)
and make the following decomposition of the integral on the right hand side of Eq.9:
$$_{\tau _1}^{\tau _N}f\left(\tau \right)𝑑\tau =_{t_1}^{t_2}f\left(\tau \right)𝑑\tau +_{t_2}^{t_3}f\left(\tau \right)𝑑\tau +\mathrm{}+_{t_{N1}}^{t_N}f\left(\tau \right)𝑑\tau $$
(10)
and approximate each of the integrals on the right hand side with the trapezoidal rule:
$$_{\tau _1}^{\tau _N}f\left(\tau \right)𝑑\tau =\frac{1}{2}\left[f\left(t_1\right)+f\left(t_2\right)\right]\left(t_2t_1\right)+\mathrm{}+\frac{1}{2}\left[f\left(t_{N1}\right)+f\left(t_N\right)\right]\left(t_Nt_{N1}\right)+O\left(\mathrm{\Delta }\tau \right)$$
(11)
From Eqs. 9 and 11 we obtain:
$`{\displaystyle \underset{n=1}{\overset{N}{}}}f\left(\tau _n\right)={\displaystyle \frac{1}{2\mathrm{\Delta }\tau }}\left\{\left[f\left(t_1\right)+f\left(t_2\right)\right]\left(t_2t_1\right)+\mathrm{}+\left[f\left(t_{N1}\right)+f\left(t_N\right)\right]\left(t_Nt_{N1}\right)\right\}`$
$$+\frac{1}{2}\left[f\left(t_1\right)+f\left(t_N\right)\right]+O\left(\mathrm{\Delta }\tau \right).$$
(12)
When the $`t_n`$ are equally spaced Eq. 12 becomes an identity with the $`O\left(\mathrm{\Delta }\tau \right)=0`$, therefore it seems to us that Eq. 12 is satisfied with an higher accuracy than just $`O\left(\mathrm{\Delta }\tau \right).`$ Eq. 12 can be applied to approximate Eq. 8 with the substitution
$$f\left(t_n\right)=s\left(t_n\right)\mathrm{exp}\left(i\omega _kt_n\right),$$
(13)
and the final result, the approximation to Eq. 8, after rearranging the terms, becomes:
$$S_k=\underset{n=1}{\overset{N}{}}c_ns\left(t_n\right)\mathrm{exp}\left(i\omega _kt_n\right)+O\left(\mathrm{\Delta }\tau \right),$$
(14)
where
$$\begin{array}{c}c_1=\frac{\mathrm{\Delta }\tau +t_2t_1}{2\mathrm{\Delta }\tau },\\ c_2=\frac{t_3t_1}{2\mathrm{\Delta }\tau },\\ \mathrm{}\\ c_{N1}=\frac{t_Nt_{N2}}{2\mathrm{\Delta }\tau },\\ c_N=\frac{\mathrm{\Delta }\tau +t_Nt_{N1}}{2\mathrm{\Delta }\tau },\end{array}$$
(15)
with the inverse formula
$$s(\tau _n)=\frac{1}{N}\underset{k=1}{\overset{N}{}}S_k\mathrm{exp}\left(i\omega _k\tau _n\right)+O\left(\mathrm{\Delta }\tau \right),\omega _k=2\pi \left(k1\right)/\left(N\mathrm{\Delta }\tau \right),$$
(16)
which is an interpolation formula for $`s\left(t_n\right)`$ at the evenly spaced points $`\tau _1\mathrm{}\tau _N`$.
## III <br>Aliasing
Aliasing is a result of undersampling and is a well known phenomenon. In ref. aliasing was looked upon from the point of view of symmetry. It is an example of wrong symmetry, and as such should be given more attention. It is the outcome of an incomplete basis. It was found in ref. , that for evenly sampled data with a sampling rate $`f_S`$, the spectral amplitude $`S(f)`$ evaluated with FFT, has the following symmetry properties
$$\left|S\left(f\right)\right|=\left|S\left(f\pm f_S\right)\right|=\left|S\left(f\pm f_S\right)\right|=\left|S\left(\pm f\pm nf_S\right)\right|,$$
(17)
where $`f`$ is the frequency and n is an arbitrary integer.
In order to avoid the aliasing symmetry of Eq. 17, the frequencies should be bounded by the Nyquist frequency (denoted here by $`f_B`$) according to
$$f_B=\frac{f_S}{2}.$$
(18)
The ECG signal was sampled with sampling rate 250 Hz, and an electronic filter was applied, which have eliminated practically all frequencies above 32 Hz, thus aliasing can not occur at frequencies below 125 Hz or even below 32 Hz. The R-R intervals were calculated directly from the ECG signal. The sampling rate for R-R intervals can be defined only for evenly sampled data, for the methods which interpolates the unevenly sampled data, or one can consider the average sampling rate from Eq. 1, in both cases
$$\overline{f}_S=1/\mathrm{\Delta }\tau =2f_N,$$
(19)
where $`f_N`$ is the Nyquist frequency for the R-R intervals. As the ECG signal contains frequencies much grater than $`f_N`$, and the R-R intervals are derived from the ECG signal, one can not be sure that the spectral analysis of the R-R intervals is free from aliasing. As a matter of fact there are indications of aliasing in some rare cases. One way to identify aliasing is to change the sampling rate and follow the changes in the spectrum. Unfortunately, for the R-R intervals, one can not speak about a definite sampling rate, but rather can consider a distribution of sampling rates. The changes in sampling rate required to observe aliasing are of the same order as the fluctuations in the sampling rate. Therefore in practice it is almost impossible to observe consistent changes in the spectrum slightly changing the heart rate.
Other possibility of detecting aliasing is by comparing the heart rate spectrum with the ECG signal spectrum. Marked differences below the Nyquist frequency for the power distribution of the RR intervals compared to the ECG signal power distribution in the same range may indicate aliasing. But we do not have yet a sound basis to treat this problem.
We have devised an experiment which definitely demonstrates the aliasing in the R-R intervals spectrum. To the best of our knowledge this is the first experiment in which one can exactly know the correct frequency above the Nyquist frequency and can follow the development of the aliasing, which appears to be diffused to great extent because the symmetry of Eq. 17 is represented not by one sampling rate but by a distribution of sampling rates, as the R-R interval is unevenly sampled.
Below we describe 3 experiments. One of them was devised to demonstrate aliasing and the other two for learning about the relations between the R-R interval spectrum and the spectrum of the ECG signal.
## IV Three Experiments
We present below results of three experiments. In the first experiment the ECG signal was collected in a normal resting state. The aim of this experiment was to compare the ECG spectrum with the R-R intervals spectrum. In the second experiment very slow breathing was monitored at a rate of 0.04 Hz. Again the ECG and R-R interval spectra were compared. In the third experiment very fast breathing was accurately monitored at the rate of 74/min and 84/min. These respiratory rates were above half of the heart rates thus allowing to observe in detail the development of aliasing.
### A The First Experiment
In this experiment (linked with the names of Zahi and Ori, where the second is one of the authors: O.G) which was done in normal, resting conditions, we compare the power estimation of the R-R interval and the ECG signal, from which the R-R interval was obtained. The ECG signal was sampled at a rate of 250 Hz. Stable intervals of 7 minutes duration were chosen for analysis.
In Fig. 1a the power distribution of the ECG signal of Zahi is depicted. The attenuation of the power with increasing frequency above 12 Hz is due to the action of an electronic filter. Above 32 Hz the contribution is practically zero. The average heart rate was 0.97 Hz. The above results were zoomed to the interval \[0-12\] Hz in Fig. 1b. One can see distinctively the peak around the average heart rate and the higher harmonics of this peak. The second harmonic is missing, but the third, fourth, fifth and sixth are distinctively visible, higher harmonics became more and more smeared and indistinguishable above the sixth harmonic. One should also note the large difference in power in the heart rate range, below the Nyquist frequency of 0.49Hz, which is much smaller compared to the peak around the average heart rate (0.97 Hz).
The power distribution of the RR intervals in the range \[0-0.5\] Hz was computed according to the methods discussed in section 2 and are presented in Figs. 2a (DFT, beat number analysis), 2b (Spline interpolation), 2c (NUDFT). For comparison also the power distribution of the ECG signal in the above range is presented in Fig. 2d.
The results of Figs. 2a, 2b, and 2c are quite similar, but the spline interpolation (Fig. 2b) and the NUDFT (Fig. 2c) are practically identical. The three graphs show the structure commonly found in the power estimation analysis of RR intervals, namely the existence of the ”high frequency” (HF), ”low frequency” (LF) and the ”very low frequency” (VLF) peaks. The ECG spectrum shows qualitatively the same structure (but not a quantitative agreement), except that the ECG spectrum is highly suppressed below 0.04 Hz, in the VLF region, indicating a possibility of aliasing in this region in the RR analysis.
In Figs. 3 and 4a-4d the results of Ori are presented. The conclusions are similar to those of Zahi, except that in the ECG spectrum both VLF and LF peaks are missing, indicating the possibility of aliasing in these regions for the RR analysis. Also in the ECG spectrum of Ofek, Fig. 5b the VLF and LF, present in Fig. 5a, are missing. VLF is missing in J.C.’s ECG spectrum (see Figs. 6a-6b).
### B The Second Experiment
In this experiment (linked again with the name Ori) we have checked the ECG spectrum near the VLF region, as the VLF was absent in the ECG spectrum for the resting state in the first experiment. The question was whether such a result persists in all ECG spectra. Therefore we have probed the VLF region by monitoring very prolonged breathing with a rate of 0.04 Hz. For the spectrum of RR intervals we found that the DFT, Spline interpolation and NUDFT gives similar results, and again NUDFT was practically identical to the spline interpolation. Therefore we present only the results of NUDFT, which are presented in Fig. 7a. For comparison the spectrum of the ECG signal is given in Fig. 7b. In Fig. 7a one can see a very clean pattern of a peak at 0.04 Hz and its higher harmonics. In Fig. 7b one can see a similar but somewhat diffused pattern. Thus this experiment indicates that similar respiratory patterns exists in both the RR as well as in the ECG signal.
### C The Third Experiment
In this experiment (linked to the name Alex, who is one of the authors: A.G) very fast breathing was accurately monitored at the rate of 74/min and 84/min respectively. These rates were well above half of the average heart rate thus allowing to observe in detail the development of aliasing. In Fig. 8 the ECG spectrum is dominated by the very high and narrow peak at the frequency $`f_1=1.234Hz`$, also its higher harmonics can be distinctively seen. The frequency $`f_1`$ is just the breathing frequency 74/min. In the same figure one can also see the diffused peaks near the average heart rate frequency of 1.636 Hz and its higher harmonics. One should observe aliasing at about $`1.636`$ $`f_1=0.402Hz`$. Indeed one can see diffused peaks around that frequency in Fig. 9a, which displays the power estimation of the RR intervals using the NUDFT (which below the Nyquist rate is similar to the spline interpolation). The width of this region can be estimated by noting that the RR intervals have different instantaneous sampling rates which are equal to the inverse of the RR interval time. In Fig. 10 we have calculated the distribution of the sampling rates by dividing the frequency region into 100 beans. We have shifted that distribution by subtracting $`f_1`$. As one can see the results are confined approximately to the region 0.32-0.47 Hz. Indeed the aliasing peaks of Fig. 9a appear in this region. The pictures below the Nyquist frequency are very similar for the DFT, NUDFT, the spline interpolation and the Lomb method (Fig. 9b) with a similar aliasing behavior.
In principle the NUDFT and the Lomb methods should not be used above the Nyquist frequency. Surprisingly enough we have found that both methods have a sharp peak at $`f_1`$, as can be seen in Figs. 9a and 9b. Both methods do not have the aliasing symmetry of the DFT as given by Eq. 17, therefore the results are not symmetric with respect to the Nyquist frequency (half the sampling rate), as it is satisfied, for example, in the case of the spline interpolation. We have found an exact result at $`f_1`$ and a diffused aliasing around 0.4 Hz. It is interesting to note that both methods give almost the same result below and above the Nyquist frequency. One can interpret the appearance of the sharp peak at $`f_1`$ as a result of a partial destruction of aliasing symmetry due to uneven samplings.
Similar results for the breathing frequency 84/min are presented in Figs. 11-12.
## V Summary and Conclusions
The ECG signal spectrum is bounded below the Nyquist frequency f<sub>B</sub> by using an electronic filter and sampled at rate larger than 2f<sub>B</sub>, thus excluding aliasing from spectral analysis. A similar procedure cannot be applied to the RR interval spectral analysis, and in this case an aliasing is possible. One of our main efforts in this paper was devoted to the problem of how to detect aliasing in the R-R interval spectral analysis.
In order to get insight into this problem three experiments have been analyzed. In the first experiment the ECG signal was collected in a normal resting state. The aim of this experiment was to compare the ECG spectrum with the R-R interval spectrum. In the second experiment very slow breathing was monitored at a rate of 0.04 Hz. Again the ECG and R-R interval spectra were compared. In the third experiment very fast breathing was accurately monitored at the rate of 74/min and 84/min respectively. These respiratory rates were above half of the heart rates thus allowing to observe in detail the development of aliasing.
The experiments which were described above led us to the following conclusions:
1. The spectral analysis of the ECG signal is more sensitive and accurate compared to the R-R interval spectral analysis and is free from aliasing. Still in the present stage it contains too much information to be of practical use. Efforts should be made to understand what will be the best way to extract information (not related to the heart condition alone as in the standard analysis of ECG) about the external influences on the heart signal.
2. We have conducted an experiment which gave a clear insight about the mechanism of aliasing in the R-R interval spectrum. The very sharp peak in the spectrum of the ECG signal, which came as the result of enforced quick breathing, reappeared as a diffused signal in the RR spectrum. The extension of the diffuseness agrees with the extension of the sampling rates of unevenly sampled data..
3. The VLF peak observed in the R-R interval spectrum is usually missing in the ECG spectrum. This lead us to suspect that the VLF observed in the RR spectrum has its origin in aliasing.
4. In some cases the LF peak does not show up in the ECG spectrum. This led us to suspect that part of the LF peak is of aliasing origin.
5. Unlike in electronic devices, it is very difficult to devise procedures to detect aliasing in humans. In electronic devices aliasing can be easily detected by changing the sampling rate. In humans the fluctuations of the heart rate are of the same order as the required changes in the sampling rates. It will be an important task to develop a proper procedure for detecting aliasing in humans.
6. We have developed a new technique for spectral analysis for unevenly sampled data called non-uniform discrete Fourier transform (NUDFT). When employed to the RR data, below the Nyquist frequency, it gave similar results as those obtained by interpolating the data with a cubic spline. Above the Nyquist frequency, the correct peak in the spectrum was detected with great accuracy. A similar result was obtained with the recently rediscovered Lomb method. We interpret this unexpected result by a partial destruction of aliasing symmetry in both methods. More efforts should be made in order to understand the anti-aliasing properties of the above methods.
7. We consider aliasing to be a wrong symmetry, resulting from the use of an incomplete basis, which has intrinsic symmetries inconsistent with the properties of the signal. Aliasing can be partially removed by reducing the symmetry of the basis.
Figure Captions
* Figure 1. The relative power of the ECG signal of Zahi, a) in the spectral range of 0-36 Hz, b) in the spectral range of 0-12 Hz.
* Figure 2. The relative power computed (from the ECG signal of Zahi) by four different methods, in the spectral range of 0-0.5 Hz, a) by DFT, b) by spline interpolation of the RR data, by NUDFT, d) from the ECG signal.
* Figure 3. The relative power of the ECG signal of Ori.
* Figure 4. The relative power computed (from the ECG signal of Ori) by four different methods, in the spectral range of 0-0.52 Hz. a) by DFT, b) by spline interpolation of the RR data, c) by NUDFT, d) from the ECG signal.
* Figure 5. The relative power computed (from the ECG signal of Ofek) by two different methods, in the spectral range of 0-0.6 Hz, a) by spline interpolation of the RR data, b) from the ECG signal.
* Figure 6. The relative power computed (from the ECG signal of J.C.) by two different methods, in the spectral range of 0-0.46 Hz, a) by spline interpolation of the RR data, b) from the ECG signal.
* Figure 7. The relative power computed (from the ECG signal of Ori with breathing rate of 0.04 Hz) by two different methods, in the spectral range of 0-0.62 Hz, a) by NUDFT, b) from the ECG signal.
* Figure 8. The relative power of the ECG signal of Alex with a breathing rate of 1.234 Hz.
* Figure 9. The relative power computed (from the ECG signal of Alex with a breathing rate of 1.234 Hz) by two different methods, in the spectral range of 0-1.5 Hz, a) by NUDFT, b) from the ECG signal.
* Figure 10. A 100 bin histogram of the heart rates of Alex which are subtracted by the breathing rate of 1.234 Hz.
* Figure 11. The relative power of the ECG signal of Alex with a breathing rate of 1.404 Hz.
* Figure 12. The relative power computed (from the ECG signal of Alex with a breathing rate of 1.404 Hz) by two different methods, in the spectral range of 0-1.6 Hz, a) by NUDFT, b) from the ECG signal. |
no-problem/9911/cond-mat9911137.html | ar5iv | text | # Commensurability, excitation gap and topology in quantum many-particle systems on a periodic lattice
## Abstract
Combined with Laughlin’s argument on the quantized Hall conductivity, Lieb-Schultz-Mattis argument is extended to quantum many-particle systems (including quantum spin systems) with a conserved particle number, on a periodic lattice in arbitrary dimensions. Regardless of dimensionality, interaction strength and particle statistics (bose/fermi), a finite excitation gap is possible only when the particle number per unit cell of the groundstate is an integer.
Strongly interacting quantum many-particle systems is one of the central topics of theoretical physics. The Renormalization Group (RG) is an important concept in studying such problems . According to the RG picture, low-energy, long-distance behavior of a quantum many-particle system is governed by RG fixed points. A quantum critical system, which has gapless excitation spectrum, is renormalized into a critical RG fixed point. However, in general, a critical RG fixed point allows some relevant perturbations. When the perturbations are added, the system is driven away from the critical fixed point. Such system generally has a finite excitation gap. Then it is expected that a gapless quantum critical system is unstable and only achieved by an appropriate fine-tuning of the Hamiltonian, which makes the relevant perturbations vanish.
In reality, there are rather many quantum critical system with gapless excitation spectrum, obtained without any apparent fine-tuning. Thus, one may naturally ask a question: is there some mechanism which protects a gapless critical system? There is one well-known mechanism of such kind: a gapless Nambu-Goldstone mode exists when the system has a spontaneously broken continuous symmetry. In fact, it describes variety of gapless excitations in quantum many-body systems, like spin waves and phonons, etc. However, it does not exhaust all the (stable) quantum critical systems. On the other hand, few universal mechanism other than the Nambu-Goldstone theorem are known .
In this letter, we argue that an incommensurability is another universal mechanism which protects the gapless excitation spectrum in quantum many-particle systems. We will show the following statement:
> In a quantum many-particle system defined on a periodic lattice, with an exactly conserved particle number, a finite excitation gap is possible only if the particle number per unit cell of the ground state is an integer.
This condition to have a finite gap may be called as the “commensurability condition”. When the particle number per unit cell of the lattice is $`\nu =p/q`$ where $`p`$ and $`q`$ are coprimes, a gapful groundstate must spontaneously break the translation symmetry, so that the unit cell of the groundstate is enlarged by factor of $`q`$. In the case of quantum spin system with the conserved total magnetization $`_jS_j^z`$, mapping of the spin system to an interacting boson system gives $`Sm`$ particles per spin where $`S`$ is the spin quantum number, $`m`$ is the average magnetization per spin.
An incommensurate filling corresponds to the limit of large $`q`$, giving large ground state degeneracy if there is a finite excitation gap. This usually implies the spectrum is actually gapless. That the incommensurate filling gives a gapless spectrum is empirically recognized more or less. In fact, all trivial ground states (e.g. completely dimerized state) with an excitation gap, as well as the less trivial Valence-Bond-Solid states , satisfy the commensurability condition. However, it is not trivial whether the incommensurability generally guarantees gapless excitations in the presence of interaction and quantum fluctuation.
If a gapful phase is adiabatically connected to those trivial states within the Hamiltonians conserving the particle number and periodicity, it must also satisfy the same commensurability condition. This is because an infinitesimal modification of the Hamiltonian does not mix states with different particle number; thus the particle number per unit cell in the groundstate is unchanged during the adiabatic change, if the gap always remains open. This explains the stability of the particle density $`\nu `$ in such cases (e.g. in the “strong coupling” approach to the magnetization plateau ). However, this argument does not exclude the possibility of a gapful phase, which is not adiabatically connected to any trivial state, not obeying the commensurability condition.
In this letter, based on a topological argument, we will show that the commensurability condition is a robust non-perturbative constraint. We consider general quantum many particle systems on a lattice with a periodic structure with any strength of interaction, in $`D`$-dimensions. For simplicity, here we assume that there is a single species of particles. We assume that the number of particles is conserved (i.e. commutes with the Hamiltonian.) Let us call the direction of an arbitrary chosen primitive vector $`\stackrel{}{a}`$ of the lattice as $`x`$-direction. We impose the periodic boundary condition in $`x`$-direction with length $`L`$ measured in unit of $`|\stackrel{}{a}|`$. Defining the translation operator $`T_x`$ which translates the system by $`\stackrel{}{a}`$, the periodic boundary condition is represented as $`T_{x}^{}{}_{}{}^{L}=1`$, and the Hamiltonian invariant under the translation $`T_x`$. The “cross section” of the lattice is defined so that the whole lattice is spanned (without overlap) by translation of the cross section by $`T_x`$. We denote the number of unit cells contained in the cross section by $`C`$; the total volume (ie. number of unit cells) of the system is given by $`CL`$.
In one dimension (including ladders etc.), the proposed statement was already shown in Refs. by generalizing the Lieb-Schultz-Mattis (LSM) argument . Therefore the remaining problem is to understand higher dimensions. Applying the LSM argument to the higher dimensions $`D>1`$ meets a difficulty. The energy of the variational state is bounded only by $`O(C/L)`$, which is generally not small in the thermodynamic limit; in an isotropic (in size) system, $`CL^{D1}`$.
Affleck discussed some application of the LSM argument to $`D>1`$. While in Ref. only spin systems at zero magnetization were considered, it is straightforward to extend the discussion in Ref. to quantum many-particle systems with general particle density (quantum spin systems with general magnetization), as was done in one dimension . Unfortunately, the strong anisotropy limit $`C/L0`$ is necessary to apply the LSM argument as in Ref. . He argued it plausible that the conclusion is still valid for an isotropic system where $`CL^{D1}`$. However, the strong anisotropy limit makes the system essentially one-dimensional. Thus one might suspect that the LSM argument does not give a useful information on higher-dimensional system which is isotropic in size. Below we will argue that the same conclusion holds, without relying on the strong anisotropy limit.
As in Ref. , we impose the periodic boundary condition for $`x`$-direction, and require $`C`$ to be mutually prime with $`q`$ but not the anisotropy condition $`CL`$. The boundary conditions for other than $`x`$-direction can be either open or periodic if they are uniform in $`x`$-direction. The particles may or may not have a real electric charge. Here we introduce a fictitious charge $`e`$ for each particle, which couples to an externally given (fictitious) electromagnetic field.
Because of the periodic boundary condition in the $`x`$-direction, the system may be regarded as a ring. Following Laughlin’s discussion of the Quantum Hall Effect (QHE), we consider a magnetic flux $`\mathrm{\Phi }`$ piercing through the ring. In a simplest gauge choice. the magnetic flux can be represented by the uniform vector potential $`A_x=\mathrm{\Phi }/L`$ in the $`x`$-direction, Now let us adiabatically increase the magnetic flux $`\mathrm{\Phi }`$ by a unit flux quantum $`\mathrm{\Phi }_0=hc/e`$, when the system is in the groundstate $`|\mathrm{\Psi }_0`$ at $`\mathrm{\Phi }=0`$. The groundstate $`|\mathrm{\Psi }_0`$ is chosen (when the groundstates are degenerate) so that it is also an eigenstate of $`T_x`$. This is always possible, at least in a finite size system, because we assumed periodic lattice structure and periodic boundary condition in $`x`$-direction; the Hamiltonian commutes with $`T_x=e^{iP_x}`$. Here $`P_x`$ is the $`x`$-component of the total (crystal) momentum. The groundstate is thus also an eigenstate of the momentum with an eigenvalue $`P_x^0`$:
$$P_x|\mathrm{\Psi }_0=P_x^0|\mathrm{\Psi }_0,$$
(1)
During the adiabatic process, the Hamiltonian is only modified by the uniform vector potential $`A_x=\mathrm{\Phi }/L`$ in the above gauge choice. Then the Hamiltonian always commutes with $`T_x`$. When the magnetic flux reaches the unit flux quantum, the original groundstate evolves into some state $`|\mathrm{\Psi }_0^{}`$. The Hamiltonian $`H(\mathrm{\Phi })`$ generally depends on the flux $`\mathrm{\Phi }`$ through the vector potential, reflecting the Aharanov-Bohm (AB) effect. However, when the AB flux $`\mathrm{\Phi }`$ is equal to the unit flux quantum, there is no AB effect and the energy spectrum is identical to the zero flux case. In fact, the vector potential is eliminated by the large gauge transformation
$$U=\mathrm{exp}[\frac{2\pi i}{L}\underset{\stackrel{}{r}}{}xn_\stackrel{}{r}],$$
(2)
where $`n_\stackrel{}{r}`$ is the particle number operator at site $`\stackrel{}{r}`$, and $`x`$ is the $`x`$-coordinate of $`\stackrel{}{r}`$. Namely, the Hamiltonian with the unit flux quantum goes back to the original one by the large gauge transformation as $`UH(\mathrm{\Phi }_0)U^1=H(0)`$. By the same transformation, the adiabatic evolution of the groundstate $`|\mathrm{\Psi }_0^{}`$ is mapped to $`U|\mathrm{\Psi }_0^{}`$. Thus, after the adiabatic procedure and the large gauge transformation, we get back to the original Hamiltonian but the groundstate $`|\mathrm{\Psi }_0`$ is changed to $`U|\mathrm{\Psi }_0^{}`$.
On the other hand, in the presence of a finite excitation gap, the ground state $`|\mathrm{\Psi }_0`$ can only be transformed into itself or, possibly, into another one of degenerate groundstates during the adiabatic process . As already explained, the reason why the LSM argument has been applied only to one dimension (or to the strong anisotropic limit) is that, the energy expectation value of the variational state is bounded only by $`O(C/L)`$, which is generally not small. Thus one can not say $`U|\mathrm{\Psi }_0`$ is always a low-energy state in $`D>1`$ dimensions. However, applying the adiabatic argument, we are able to claim that, if the system has a finite excitation gap, the outcome of the adiabatic evolution $`U|\mathrm{\Psi }_0^{}`$ should be one of the groundstates .
In the case of QHE, an implicit assumption of uniqueness of the groundstate led to an integer quantization of Hall conductivity. However, as pointed out by Tao and Wu , it is possible that the ground states are degenerate, and that is what needed in fractional QHE. Therefore we have to check whether $`U|\mathrm{\Psi }_0^{}`$ is identical to $`|\mathrm{\Psi }_0`$ or not.
Here let us recall that, $`U|\mathrm{\Psi }_0`$, which is similar to $`U|\mathrm{\Psi }_0^{}`$, is nothing but the variational state constructed in the LSM argument and its generalizations . Now we see a rather close relation between LSM and Laughlin’s arguments. While our state $`U|\mathrm{\Psi }_0^{}`$ is not identical to the LSM state $`U|\mathrm{\Psi }_0`$, we can still invoke the LSM orthogonality argument used for $`U|\mathrm{\Psi }_0`$.
We have chosen the original groundstate as an eigenstate of $`P_x`$ as in eq. (1). Since the Hamiltonian always commutes with $`P_x`$ during the adiabatic process, the eigenvalue of $`P_x`$ is unchanged. This can be seen easily from perturbation theory on an infinitesimal increase of the AB flux; the infinitesimal modification of the Hamiltonian commutes with $`P_x`$ and it does not mix states with different eigenvalues of $`P_x`$. Thus, $`|\mathrm{\Psi }_0^{}`$ belongs to the same eigenvalue $`P_x^0`$ as in eq. (1). Now, after the gauge transformation, $`|\mathrm{\Psi }_0^{}`$ is transformed to $`U|\mathrm{\Psi }_0^{}`$, which may belong to a different eigenvalue. By using the identity
$$U^1T_xU=T_x\mathrm{exp}[2\pi i\underset{\stackrel{}{r}}{}\frac{n_\stackrel{}{r}}{L}]$$
(3)
we see that $`U|\mathrm{\Psi }_0^{}`$ is an eigenstate of $`P_x`$ with $`P_x=P_x^0+2\pi \nu C`$. Thus, if choose $`C`$ to be mutually prime with $`q`$ ($`\nu =p/q`$), $`U|\mathrm{\Psi }_0^{}`$ is orthogonal to $`|\mathrm{\Psi }_0^{}`$ and $`|\mathrm{\Psi }_0`$, because it belongs to a different eigenvalue of $`P_x`$. By repeating the same procedure several times, we can conclude that there are at least $`q`$ degenerate ground states. Thus we have shown that the similar conclusion to the one-dimensional case holds in any dimensions, without relying on the strong anisotropy limit. The optimistic view taken in Ref. is actually justified, as far as the anisotropy condition is concerned. It is valid for arbitrary strong interaction, and is even independent of the particle statistics (bose/fermi). The ground state degeneracy is a robust, non-perturbative property related to the topology of the gauge field and the symmetries of the system.
An unsatisfactory point still remaining is that we have to take $`C`$ to be mutually prime with $`q`$. If $`C`$ is an integral multiple of $`q`$, nothing can be said on the degeneracy, and a unique ground state with an excitation gap is possible in principle. This is also related to the fact that the present argument is not yet mathematically rigorous for the thermodynamic limit. (Compare to Ref. .) However, if we assume that the groundstate degeneracy does depend on whether $`C`$ is an integral multiple of $`q`$ or not in a large enough system, it suggests some long-range structure of period $`q`$. Then it is naturally expected that the ground states have the $`q`$-fold degeneracy anyway, for a sufficiently large system. In addition, the present argument can be applied to many different boundary conditions, because there are various possible choices of the primitive vector $`\stackrel{}{a}`$ and the corresponding cross section. The degeneracy looks less artificial in the light of this fact.
The ground states in a finite system would be actually split by exponentially small energy due to tunneling effects. In a finite size system, the $`q`$ (near-)groundstates $`|\mathrm{\Psi }_n`$ ($`n=0,1,\mathrm{},q1`$) are eigenstates of the momentum $`P_x`$: $`P_x|\mathrm{\Psi }_n=(2\pi n/q)|\mathrm{\Psi }_n`$. However, in the thermodynamic limit, the physical groundstates are given by $`|\stackrel{~}{\mathrm{\Psi }}_j`$’s, which are defined as $`|\stackrel{~}{\mathrm{\Psi }}_j=_ne^{iP_xj}|\mathrm{\Psi }_n`$. These physical ground states are connected by the translation operator:
$$T_x|\stackrel{~}{\mathrm{\Psi }}_j=|\stackrel{~}{\mathrm{\Psi }}_{(j+1)\mathrm{mod}n}.$$
(4)
Thus the translation symmetry (to $`x`$-direction) is spontaneously broken, and the periodicity of the ground state is an integral multiple of $`q`$. This concludes the derivation of the proposed statement.
We note that, Lee and Shankar had derived a similar statement for the limited case of hard-core models in two dimensions. However, their argument relies on a certain field-theory mapping, and looks less reliable compared to ours. In fact, their statement appears to be too strong: they state that there must be a Charge Density Wave (CDW) order if the system with a fractional filling $`\nu <1`$ has a gap. This has a rather simple counterexample: the spontaneously dimerized ground state of a $`S=1/2`$ Heisenberg magnet at zero magnetization, which corresponds to a hard-core boson system with $`\nu =1/2`$, has no long-range Néel-type (namely, CDW) order. On the other hand, the spontaneously dimerized state does break the translational symmetry, and is consistent with our conclusion.
The ground state degeneracy of $`S=1/2`$ quantum spin systems has been discussed in several different contexts, for example in Ref. . In particular, there has been a lot of discussions on the possibility of the exotic spin-liquid state called the Resonating Valence Bond (RVB) state. While various possibilities were considered under the name of RVB, here we refer to the proposals of a disordered groundstate with a finite excitation gap, but without any apparent breaking of the translation symmetry. If there is really such a spin-liquid state, it appears contradictory to our result. This might be the reason why there is no established example of such an RVB groundstate, despite intensive search in various $`S=1/2`$ spin systems with odd number of spins per unit cell.
However, in spite of its uniform appearance, some degeneracy is argued to exist in the RVB state, under the periodic boundary condition. In Refs. , this degeneracy is argued to be unphysical. For the degeneracy to be unphysical, it must be that no physical operators distinguish the groundstates with the spontaneously broken translation symmetry. While we are not sure it is possible, we do not rule out such a possibility. In any case, the (near-) degeneracy of the finite size system concluded from the present argument seems consistent with Refs. .
The ground-state degeneracy in a gapped phase required by the present argument, is related to the commensurability. Intuitively, this is quite natural; the particles can be locked into a stable groundstate only when it can have a commensurate structure with the lattice. Such an intuition is, however, more or less based on some trivial states which can be easily imagined in minds. Nevertheless, the commensurability condition turned out to be essential even in the presence of arbitrarily strong interaction and quantum fluctuation, because of the topological mechanism discussed in the present letter.
In generic cases, a finite excitation gap would be only possible at special commensurate (rational) filling. In case of charged particles, such a gapped phase includes Mott and band insulators. On the other hand, gapless phases at generic filling includes superfluid, Fermi liquid and possibly other conducting phases. In case of quantum spin systems, the gapful phases are related to the magnetization plateau at quantized magnetization. In any dimensions, a magnetization plateau with a finite excitation gap is possible only if the commensurability condition is satisfied: $`n(Sm)=\text{integer}`$ where $`n`$ is the number of spins in the unit cell of the ground state, similarly to the one-dimensional case . In fact, several magnetization plateaus reported in $`D>1`$ dimensions (examples include ) satisfies this quantization condition. However, a plateau in a magnetization curve appears if there is no gapless excitation which changes the total magnetization. As already mentioned in Ref. , the LSM argument (and the present argument) does not directly guarantee gapless excitations of such kind. Thus, it may be possible to have an “exceptional” magnetization plateau which does not obey the above simple quantization condition, if there are gapless excitations only with the same total magnetization as the groundstate. In one dimension, all plateaus should obey the quantization condition as far as the general Abelian bosonization treatment is valid. (However, see for a possible exceptional “plateau” at $`m=0`$, and for a discussion in the doped case.) In higher dimensions, the situation is less clear, while certainly many plateaus satisfy the simple quantization condition. An exceptional plateau at $`m=0`$ might be realized in Kagomé lattice , which is argued to have singlet gapless excitations.
Finally, we note that our commensurability condition has obvious generalizations to the spinful electron systems, Kondo lattices, and other multi-species particle systems in arbitrary dimensions.
I would like to thank Ian Affleck, Hal Tasaki and Masanori Yamanaka for many stimulating discussions over several years. I also thank Subir Sachdev for pointing out some relevant references. This work is supported by a Grant-in-Aid from Ministry of Education, Science, Sports and Culture of Japan. |
no-problem/9911/cond-mat9911228.html | ar5iv | text | # Dynamics of Kinks: Nucleation, Diffusion and Annihilation
## Abstract
We investigate the nucleation, annihilation, and dynamics of kinks in a classical $`(1+1)`$-dimensional $`\varphi ^4`$ field theory at finite temperature. From large scale Langevin simulations, we establish that the nucleation rate is proportional to the square of the equilibrium density of kinks. We identify two annihilation time scales: one due to kink-antikink pair recombination after nucleation, the other from non-recombinant annihilation. We introduce a mesoscopic model of diffusing kinks based on “paired” and “survivor” kinks/antikinks. Analytical predictions for the dynamical time scales, as well as the corresponding length scales, are in good agreement with the simulations.
preprint: LA-UR 99-1618
2
Many extended systems have localized coherent structures that maintain their identity as they move, interact and are buffeted by local fluctuations. The statistical mechanics of these objects has diverse applications, e.g., in condensed matter physics , biology , and particle physics . The model to be studied here is a kink-bearing $`\varphi ^4`$ field theory in $`(1+1)`$ dimensions, popular because its properties are representative of those found in many applications. Static equilibrium quantities of this theory, such as the kink density and spatial correlation functions, are now well understood and recent work has shown that theory and simulations are in good agreement . However, dynamical processes, both close to and far out of equilibrium, are much less well understood. Questions include: What is the nucleation rate of kink-antikink pairs? How is an equilibrium population maintained? How do these dynamical processes depend on the temperature and damping? These questions, among others, are the subject of this Letter.
We introduce and analyze below a simple model of kink diffusion and annihilation that predicts the nucleation rate and provides a picture of the physical situation, including the existence of multiple time and length scales. We also carry out high resolution numerical simulations. As one consequence of our work, we are able to settle a recent controversy as to whether the nucleation rate of kinks in an overdamped system is proportional to $`\mathrm{exp}(2E_k\beta )`$ or $`\mathrm{exp}(3E_k\beta )`$ in favor of the first result ($`E_k`$ is the kink energy and $`\beta =1/k_BT`$).
We consider the dynamics of the $`\varphi ^4`$ field obeying the following dimensionless Langevin equation :
$$_{tt}^2\varphi =_{xx}^2\varphi +\varphi (1\varphi ^2)\eta _t\varphi +\xi (x,t),$$
(1)
with the fluctuation-dissipation relation enforced by $`<\xi (x,t)\xi (x^{},t^{})>=2\eta \beta ^1\delta (xx^{})\delta (tt^{})`$. We perform simulations on lattices typically of $`10^6`$ sites, using a finite difference algorithm that has second-order convergence to the continuum . Typical values of the grid spacing and time step are $`\mathrm{\Delta }x=0.4`$ and $`\mathrm{\Delta }t=0.01`$.
At zero temperature, the static kink solution centered at $`x=x_0`$ is $`\varphi _k(x)=k(xx_0)`$ where $`k(x)=\mathrm{tanh}(x/\sqrt{2})`$; the corresponding antikink solution is $`\varphi _a(x)=k(xx_0)`$. Because there are only two potential minima, kinks alternate with antikinks on the spatial lattice. Imposing periodic boundary conditions constrains the number of kinks and antikinks to be equal. During the time evolution, we identify kinks and antikinks individually and follow the “lifeline” of each kink or antikink (Fig. 1).
Equilibrium properties of one-dimensional systems, such as the free energy density and the correlation function $`<\varphi (0)\varphi (x)>`$, can be calculated using the transfer integral method . The calculation is exact, although one typically must evaluate eigenvalues of the resulting Schrödinger equation numerically. When the on-site potential has the double-well form, as is the case here, one part of the free energy density at low temperature can be interpreted as due to kinks, forming a dilute gas with density : $`\rho _k\sqrt{E_k\beta }\mathrm{exp}(E_k\beta )`$. This WKB approximation is consistent with recent simulations at $`\beta >6`$, where unambiguous identification of kinks is possible .
An equilibrium density of kinks is maintained by a dynamical balance of nucleation and annihilation of kink-antikink pairs (Fig. 1). The dependence of the nucleation rate $`\mathrm{\Gamma }`$ on temperature and damping, however, is not directly calculable from the transfer integral; nor are unambiguous results for symmetric potentials available from saddle-point calculations . While analogy with the Kramers’ problem suggests $`\mathrm{\Gamma }\mathrm{exp}(2\beta E_k)`$ , the relationship $`\mathrm{\Gamma }\mathrm{exp}(3\beta E_k)`$ has also been suggested . Our direct counting of nucleation events establishes that their rate is proportional to the square of the equilibrium density, that is $`\mathrm{\Gamma }\mathrm{exp}(2\beta E_k)`$ (Fig. 2). Below we show how this relation can be understood from a mesoscopic model of diffusing kinks with paired nucleation.
At equilibrium, the nucleation rate is related to the mean kink lifetime $`\tau `$ by $`\rho _k=\mathrm{\Gamma }\tau `$. Previous attempts to evaluate $`\mathrm{\Gamma }`$ numerically have proceeded by counting the number of kinks $`n_k(t)`$ and assuming exponential decay of $`n_k(t+\tau )n_k(t)`$. Unfortunately this approach provides no information on the underlying processes, and yields incorrect results if kinks are not properly identified on the lattice. In particular, results that appeared to support $`\mathrm{\Gamma }\mathrm{exp}(3\beta E_k)`$ were performed at temperatures too high for accurate computation of $`n_k(t+\tau )n_k(t)`$ .
Because we identify individual nucleation events and follow individual kink lifelines, we can distinguish “paired” kinks (whose partner antikink is still alive) from “survivor” kinks (whose partner has been killed). We also distinguish and measure the contributions to the annihilation rate from the recombinant and various non-recombinant mechanisms (Fig. 3). The most frequent annihilation event is recombination of a recently-nucleated pair (designated I in Fig. 3) . However, the “survivor” kinks that remain after a non-recombinant annihilation event (II or III) have a longer mean lifetime.
At finite temperature, the mean-squared displacement of an isolated kink is given by $`<𝐗_t^2>=2Dt`$. The diffusivity $`D`$ can be estimated by using the zero-temperature kink as an ansatz in the equation of motion (1), yielding $`D(E_k\beta \eta )^1`$ , where $`E_k=\sqrt{8/9}`$ for a static kink. Corrections to $`D`$, arising because of fluctuations in the kink shape, are proportional to $`\beta ^2`$ and subdominant in the temperature range considered here.
Our numerical observations, in particular that kink-antikink collisions at moderate to large damping always result in annihilation, motivate us to introduce the following mesoscopic model of kink dynamics: (i) kink-antikink pairs are nucleated at random times and positions with initial separation $`b\rho _k^1`$; (ii) once born, kinks and antikinks diffuse independently with diffusivity $`D`$; (iii) kinks and antikinks annihilate on collision. The separation between a kink and its partner performs Brownian motion with diffusivity $`2D`$. Thus, if only recombinate annihilation (I in Fig. 3) were allowed, the time $`𝐭_0`$ between nucleation and annihilation would have the density $`\frac{\mathrm{d}}{\mathrm{d}t}𝒫[𝐭_0<t]=bt^{\frac{3}{2}}(8\pi D)^{\frac{1}{2}}\mathrm{exp}(\frac{b^2}{8Dt})`$ .
To analyse our model, we use the following approximation for non-recombinant annihilation: as long as both members of a pair are alive, there is a constant probability $`\mu `$ per unit time of a member being struck and “killed” by an outsider, i.e. of an event II or III. Thus, to each pair we assign a killing time $`𝐭_\mu `$, distributed according to $`𝒫[𝐭_\mu >t]=\mathrm{exp}(\mu t)`$. Non-recombinant annihilation happens with probability $`𝒫[𝐭_\mu <𝐭_0]=1\mathrm{exp}(b\nu )`$, where $`\nu ^2=\frac{\mu }{2D}`$ . The killing rate $`\mu `$ depends on the density of kinks; we estimate it as follows. A new-born pair finds itself in a domain between an existing kink and antikink of typical length $`1/(2\rho _k)`$. The mean time for a diffusing particle to exit the region is proportional to $`(2D\rho _k^2)^1`$. Therefore, let
$$\mu =2D\alpha ^2\rho _k^2.$$
(2)
The value of the dimensionless factor was obtained from numerical measurements of length and timescales (see Figures 4 and 6 below): we estimate $`\alpha 8`$.
Let $`R(t)=\frac{\mathrm{d}}{\mathrm{d}t}𝒫[𝐭_0<t|𝐭_0<𝐭_\mu ]`$. Then
$$R(t)=N(b)\mathrm{exp}(\frac{b^2}{8Dt})t^{\frac{3}{2}}\mathrm{exp}(\mu t),$$
(3)
where $`N(b)=b(8\pi D)^{\frac{1}{2}}\mathrm{exp}(\nu b)`$. In Fig. 4 we plot (3) and a histogram of values of $`𝐭_0`$ obtained from a large numerical solution of (1). The behavior $`R(t)t^{\frac{3}{2}}`$ is characteristic of Brownian excursions .
Our mesoscopic model has two timescales :
$`\tau _0`$ $`=`$ $`<𝐭_0|𝐭_0<𝐭_\mu >={\displaystyle \frac{b}{2\sqrt{\mu D}}},`$ (4)
$`\tau _\mu `$ $`=`$ $`<𝐭_\mu |𝐭_\mu <𝐭_0>={\displaystyle \frac{1}{\mu }}\left(1{\displaystyle \frac{\nu b}{2}}{\displaystyle \frac{1}{\mathrm{e}^{\nu b}1}}\right).`$ (5)
The mean recombination time (4) depends on $`b`$; in contrast $`\tau _\mu `$ has a non-zero limit for $`\nu b0`$: $`\tau _\mu 1/(2\mu )`$. With the approximation that a “survivor” kink/antikink has the same probability per unit time, $`\mu `$, of collision and death, the mean lifetime of a kink or antikink is given by
$$\tau =\tau _0\mathrm{e}^{\nu b}+\tau _\mu (1\mathrm{e}^{\nu b})+\frac{1}{2\mu }(1\mathrm{e}^{\nu b}).$$
(6)
As $`\rho b0`$, $`\tau (3/2)b(2\mu D)^{1/2}=(3/4)b(\alpha D\rho )^1`$. In the same limit, combining (2) and (6) yields the prediction that the nucleation rate is proportional to the square of the equilibrium density:
$$\mathrm{\Gamma }=\rho _k/\tau =\frac{4}{3b}D\alpha \rho _k^2,$$
(7)
The relation $`\mathrm{\Gamma }\rho _k^2`$ is also found in the discrete-space Ising model . In contrast, the nucleation rate is proportional to $`\rho _k^3`$ in systems where nucleation does not occur in pairs . The latter scaling was incorrectly predicted for the $`\varphi ^4`$ system , from an estimate of the annihilation rate that does not take into account paired nucleation. (In the $`\varphi ^4`$ system one does, however, find that the rate of survivor-survivor annihilation events – IV in Fig. 3 – is proportional to $`\rho _k^3`$.) In the $`\varphi ^4`$ SPDE, the parameters $`D`$, $`\mathrm{\Gamma }`$ and $`b`$ have in general a weak (non-exponential) temperature dependence. The lengthscale $`b`$ is of the same order as the width of an isolated kink.
In Fig. 5 we plot the measured nucleation rate versus the damping coefficient at fixed temperature. The nucleation rate is proportional to $`\eta ^1`$ for $`\eta 1`$ \[in agreement with (7)\] and appears to plateau for $`\eta 0`$. At low damping, however, direct measurement of the nucleation rate is problematic because kink-antikink collisions may result in single or multiple bounces rather than immediate annihilation .
We now turn to the lengthscales in the system. A histogram of distances between neighboring kinks and antikinks is well-approximated by an exponential with characteristic length $`(2\rho _k)^1`$. This simple form results from the cancellation of the tendency of paired kinks to be closer together than $`(2\rho _k)^1`$ with the opposite tendency of survivor kinks. In Fig. 6 we plot $`f(x)=`$(number of occurrences of separation $`(x,x+\mathrm{d}x)`$)$`/(L\mathrm{d}x)`$. We also construct the histogram for the separations of only paired kinks and antikinks. The dashed curve is the probability of being at $`x`$, averaged over the lifetime, for a Brownian motion killed at $`x=0`$ and at rate $`\mu `$ :
$$l(x)=N\mathrm{exp}(\nu x)=N\mathrm{exp}(\alpha \rho _kx).$$
(8)
The classification of kinks into paired kinks and survivors, with the approximation that kinks have a constant probability (2) per unit time of non-recombinant annihilation, allows us to construct a macroscopic rate theory for the two densities $`n_p(t)`$ (paired kinks) and $`n_s(t)`$ (survivor kinks). The equation for $`n_p(t)`$ has a positive term due to nucleation and a negative term inversely proportional to the lifetime of pairs, (4). The terms in the equation for $`n_s(t)`$ correspond to processes III and IV in Fig. 3. Note that process II does not change the number of survivor kinks. We obtain
$`\dot{n}_p=`$ $`\mathrm{\Gamma }2b^1\alpha D(n_s+n_p)n_p`$ (9)
$`\dot{n}_s=`$ $`D\alpha ^2(n_s+n_p)^2(n_p2n_s).`$ (10)
The steady state solution of (10) gives the relationship (7) between $`\mathrm{\Gamma }`$ and the equilibrium kink density. Nonequilibrium dynamics are also correctly described: if $`\mathrm{\Gamma }=0`$, the paired density quickly decays and, for late times, $`\dot{n}_sn_s^3`$, in agreement with an exact result for the survival probability in the diffusion-limited reaction A$`+`$A$``$0 . While not exact, (10) illustrate that at least two coupled equations are necessary to capture the two timescales in the dynamics: no single rate equation can suffice.
We have benefited from discussions with Kalvis Jansons, Eli Ben-Naim and Vincent Hakim. Computations were performed at the National Energy Research Scientific Computing Center (NERSC), Lawrence Berkeley National Laboratory. |
no-problem/9911/astro-ph9911144.html | ar5iv | text | # A Multi-Component Measurement of the Cosmic Ray Composition Between 𝟏𝟎^𝟏𝟕eV and 𝟏𝟎^𝟏𝟖eV
## Abstract
The average mass composition of cosmic rays with primary energies between $`10^{17}`$eV and $`10^{18}`$eV has been studied using a hybrid detector consisting of the High Resolution Fly’s Eye (HiRes) prototype and the MIA muon array. Measurements have been made of the change in the depth of shower maximum, $`X_{max}`$, and in the change in the muon density at a fixed core location, $`\rho _\mu (600m)`$, as a function of energy. The composition has also been evaluated in terms of the combination of $`X_{max}`$ and $`\rho _\mu (600m)`$. The results show that the composition is changing from a heavy to lighter mix as the energy increases.
The source of cosmic rays with particle energies above $`10^{14}`$ eV is still unknown. Models of origin, acceleration, and propagation must be evaluated in light of the observed energy spectrum and chemical composition of the cosmic ray flux arriving at the earth. The cosmic ray energy spectrum generally follows a simple power law over many decades of energy. This might lead one to believe that cosmic rays of all energies share the same source. However, there are two detectable breaks in this otherwise smooth spectrum. At an energy of about $`10^{15}`$ eV the spectrum softens. At energies above $`10^{18}`$ eV it hardens again. These two features, known as the “knee” and “ankle”, suggest that the source of cosmic rays or propagation effects might be changing in these regions. Observations of the mass composition as a function of energy may provide a path to further understanding.
Several experiments have attempted to determine the mean cosmic ray composition through the knee region of the spectrum. While the results are not in complete agreement, there is some consensus for a composition becoming heavier at energies above the knee , a result consistent with charge-dependent acceleration theories or rigidity-dependent escape models. In the region above the knee, the Fly’s Eye experiment has reported a changing composition from a heavy mix around $`10^{17}`$eV to a proton dominated flux around $`10^{19}`$eV . Muon data from the AGASA experiment show broad agreement with this trend if the data are interpreted using the same hadronic interaction model as in the Fly’s Eye analysis .
In this experiment, our measurements are unique in that two normally independent detection techniques are employed simultaneously in the measurement of various aspects of extensive air shower(EAS). We use a hybrid detector consisting of the prototype High Resolution Fly’s Eye (HiRes) air fluorescence detector and the Michigan Muon Array (MIA). We have undertaken to independently measure parameters reflecting the average cosmic ray nuclear composition at energies above $`10^{17}`$ eV. The detectors are located in the western desert of Utah, USA at $`112^{}\text{W}\text{longitude}`$ and $`40^{}\text{N}\text{latitude}`$. The HiRes detector is situated atop Little Granite Mountain at a vertical atmospheric depth of $`850`$ g/cm<sup>2</sup>. It overlooks the CASA-MIA arrays some 3.3 km to the northeast. The surface arrays are some 150 m below the fluorescence detector at an atmospheric depth of 870 g/cm<sup>2</sup>.
The HiRes prototype has been described in detail elsewhere . It views the night sky with an array of 14 optical reflecting telescopes. They image the EAS as it progresses through the detection volume. Nitrogen fluorescence light (300–400 nm) is emitted at an atmospheric depth $`X`$ in proportion to the number of charged particles in the EAS at that depth, $`S(X)`$. Part of this shower development profile (at least 250 g/cm<sup>2</sup> long) can be determined by measuring the light flux arriving at the detector. Assuming $`S(X)`$ to be the Gaisser-Hillas shower developement function and correcting for Cherenkov light contamination and atmospheric sccattering effects one can measure the primary particle energy $`E`$, and the depth at which the shower reaches maximum size, $`X_{max}`$.
MIA , consisting of over 2500 m<sup>2</sup> of active area distributed in 16 patches of 64 scintillation counters, measures EAS muon arrival times with a precision of 4 ns and records all hits occurring within 4 $`\mu \text{s}`$ of the system trigger. MIA records only the identification and firing time of the counters participating in a given event. The average efficiency of MIA counters for detecting minimum ionizing particles was $`93\%`$ when they were buried, and the average threshold energy for vertical muons is 850 MeV. MIA determines the muon density via the pattern of hit counters observed in the shower . An estimate of the muon density at 600 m from the core, $`\rho _\mu (600m)`$, is then determined by a fit.
It is expected that changes in the mean mass composition of the cosmic ray flux as a function of $`E`$ will be manifested as changes in the mean values of two measurable quantities $`X_{max}`$ and $`\rho _\mu (600m)`$. To indicate those changes, a rate of change of $`X_{max}`$ with $`\mathrm{log}E`$, called the elongation rate, $`\alpha `$, has been introduced. Similarly for muons, we define a power law index for $`\overline{\rho _\mu (600m)}`$ as a function of $`E`$, called the “$`\mu `$ content index”, $`\beta `$, in this study. Hence:
$`\alpha ={\displaystyle \frac{d\overline{X}_{max}}{d\mathrm{log}E}}\mathrm{and}\beta ={\displaystyle \frac{\mathrm{d}\mathrm{log}\overline{\rho _\mu (600\mathrm{m})}}{\mathrm{d}\mathrm{log}\mathrm{E}}}.`$ (1)
Assuming that a shower initiated by a nucleus of mass number $`A`$ and energy $`E`$ is a superposition of $`A`$ subshowers each with energy $`E/A`$, $`\overline{X}_{max}\alpha _0\mathrm{log}\left(E/A\right)`$ and $`\overline{\rho _\mu (600m)}A\left(E/A\right)^{\beta _0}`$ where $`\alpha _0`$ and $`\beta _0`$ are for a pure beam of primary nuclei of mass $`A`$. The values of $`\alpha _0`$ and $`\beta _0`$ are dependent of the hadronic interaction model, but we find them largely independent of $`A`$ in our simulations described below. Therefore, any deviation of our observed elongation rate, $`\alpha `$ and $`\mu `$ content index, $`\beta `$, from those for pure composition imply a changing composition, i.e.
$$\frac{d\overline{\mathrm{log}A}}{d\mathrm{log}E}=\frac{\alpha \alpha _0}{\alpha _0}=\frac{\beta \beta _0}{1\beta _0}.$$
(2)
Since the superposition model is not fully realistic, a more reliable comparison between the data and predictions is based on detailed simulation of shower development described below.
HiRes/MIA coincident data were collected on clear moonless nights between Aug. 23, 1993 and Aug. 24, 1996. The total coincident exposure time was 2878 hours corresponding to a duty cycle of 10.2%. 4034 coincident events were observed. The shower trajectory for each event was obtained in an iterative procedure using the information from both HiRes and MIA . HiRes uses its spatial pixel patterns to find the plane in space containing the the detector and the shower axis (the shower-detector plane or SDP), and the time development to find the distance of closest approach $`R_p`$. MIA helps to constrain the HiRes fits using its muon arrival time distribution to provide the shower direction within the SDP. The accuracy of the shower axis determination depends on the number of observed muons, the HiRes angular track length, and the core distances from MIA and HiRes. Using its density measurement, MIA also helps to determine the shower core location in the SDP. 2491 events are reconstructed via this procedure. Monte Carlo studies show that over the full energy range studied, the median shower direction error is $`0.85^{}`$ with a median core location error of 45 m.
We have imposed cuts on the data to remove the poorly reconstructed events and maintain good resolution. The cuts are as follows: triggered HiRes pixels should subtend an angle of at least $`20^{}`$ and view a depth range of at least 250 g/cm<sup>2</sup>; shower maximum should be bracketed by measurements; the estimated error in shower maximum should be less than 50 g/cm<sup>2</sup>; the reduced $`\chi ^2`$ of the shower profile fit should be less than 10; to reduce the influence of direct Cerenkov light, all pixels should view the shower axis at angles larger than $`10^{}`$; and the reconstructed shower core should be less than 2000 m from the center of MIA. These cuts leave a sample of 891 events. When determining the $`\rho _\mu (600m)`$ from muon data we also require that: the shower core lie between 300 m and 1000 m from MIA; and the number of hit MIA counters should be less than 700 to avoid the saturation of the $`\mu `$ counters. This more restricted sample contains 573 events.
The HiRes data are shown in FIG. 1. We show bands to represent the statistical and systematic uncertainties in the depth of maximum. The measured elongation rate is $`93.0\pm 8.5\pm (10.5)`$ g/cm<sup>2</sup>/decade over the energy range from $`10^{17}`$ to $`10^{18.1}`$ eV. The MIA data are shown in FIG. 2. The data show a $`\mu `$ content index of $`0.73\pm 0.03\pm (0.02)`$/decade in the same energy range as for the elongation rate. Numbers in the brackets provide the systematic error based on the analysis described below.
The systematic errors on the muon local density $`\rho _\mu (600m)`$ result from the uncertainties in the absolute efficiencies of the MIA counters over time. The average efficiency is 80.7% with an RMS of 4.7% over the 16 patches during the time the data were taken. This is the only significant systematic uncertainty associated with $`\rho _\mu (600m)`$. For $`X_{max}`$, we have considered systematic errors in the atmospheric transmission of light and in the production of Cerenkov light. These are related since atmospherically scattered Cerenkov light can masquerade as fluorescence light if not accounted for properly. For atmospheric scattering, there was uncertainty in the aerosol content in the air from night to night. The uncertainty, equivalent to one standard deviation about the mean, is expressed as a range of possible horizontal extinction lengths for aerosol scattering at 350 nm (taken as 11 km to 17 km based on measurements using Xenon flashers) and a range of scale heights for the vertical distribution of aerosol density above the mixing layer (taken as 0.6 km to 1.8 km). For Cerenkov light production, we have systematically varied the angular scale for the Cerenkov emission around the shower axis. The emission angle distribution is related to the angular distribution of the shower particles and the intrinsic Cerenkov emission angle. Both vary with depth in the atmosphere, and at ground level we take the distribution as an exponential function of the angle from the shower axis, with a scale of $`4.0\pm 0.3^{}`$. Those uncertainties are shown by the shaded areas in FIG 1 & 2.
Another systematic error, that in the determination of energy, is considered to be independent of energy and will therefore have no effect on the values of the measured elongation rate and $`\mu `$ content index. It will however effect the normalization of the mean depth of maximum or muon size at a particular energy. We estimate that the systematic error in energy is no larger than 25%, made up of an uncertainty in the nitrogen fluorescence efficiency of no more than 20% and a calibration systematic of less than 5%.
Also shown in FIG. 1 and 2 are Monte Carlo simulation results. These full shower simulations have been performed using the CORSIKA package , employing QGSJET and SIBYLL hadronic interaction models. We have generated showers of fixed energies and fixed zenith angles in order to parameterize the shape of the shower development profile and the muon content at any energy between $`3\times 10^{16}`$eV and $`5\times 10^{18}`$eV and at any zenith angle out to $`80^{}`$. We then pass those showers through a realistic simulation of the detector and generate data to be analyzed by the reconstruction software used for real data. A $`E^3`$ differential spectrum is assumed. The minimum energy is well below the HiRes/MIA threshold to allow for a study of any threshold effects. Distributions of energy, impact parameter and zenith angle are well predicted by the simulation .
After reconstruction, and after applying the same quality cuts as we apply to the real data, we find that a pure proton flux and the QGSJET model gives an elongation rate of $`\alpha _0=58.5\pm 1.3`$ g/cm<sup>2</sup>/decade and a $`\mu `$ content index of $`\beta _0=0.83\pm 0.01`$/decade over the range from $`10^{17}`$ to $`10^{18}`$eV. For a pure iron composition and the QGSJET model we find corresponding values of $`\alpha _0=60.9\pm 1.1`$ g/cm<sup>2</sup>/decade and the same $`\beta _0=0.83\pm 0.01`$ g/cm<sup>2</sup>/decade as for protons. Results from SIBYLL show similar elongation rates, but have the $`X_{max}`$ approximately 25 g/cm<sup>2</sup> deeper than QGSJET. SIBYLL also predicts significantly fewer muons at 600 m for both proton and iron showers. The effect of any triggering and reconstruction biases is very small for $`X_{max}`$, as can be seen in FIG.1 by comparing these reconstructed data (dots) with the “input” (lines) directly from CORSIKA. The application of well chosen cuts has resulted in a bias-free measurement of the elongation rate. However, for muon density measurement, reconstruction effects change the index by 8%. The “input” for the simulation is 0.88 (true QGSJET prediction) for both proton and iron showers. We suspect that the presence of an asymmetry in core distance error can result in a small overestimate of the muon density. It is possible that this effect changes with shower energy.
Data and simulations are clearly inconsistent with each other in both the elongation rate and the $`\mu `$ content index. It leads support to the hypothesis that the cosmic ray composition is changing towards a lighter mix of nuclei from $`10^{17}`$ to $`10^{18}`$eV. HiRes and MIA reach the same conclusion by using different experimental techniques and measuring different physics variables. Substituting the measured and simulated values of $`\alpha `$ and $`\beta `$ in (2) shows that the result from HiRes and MIA are consistent, with an implied change in $`\mathrm{\Delta }\overline{\mathrm{log}A}`$ of about -0.58 over one decade of energy.
There is a problem with the absolute density of muons at 600 m, with respect to the model predictions. As seen in FIG. 2, the data show values of $`\rho _\mu `$ at the lower energies which are larger than pure iron showers. The comparison with SIBYLL predictions (not shown) is even worse. The reason for this discrepancy is not clear. We have searched for a possible experimental reason for a systematic overestimation of the muon size, and have not found an affect of sufficient magnitude. A change in the energy scale by the maximum estimated systematic error of 25% world lessen the problem but not completely remove it. A shift of 40% is required, which is well beyond the systematic error. such a shift would also produce large discrepancies in the R<sub>p</sub> distribution between data and Monte Carlo. We conclude that the model is deficient in muon production. Evidence for this has also been found at lower EAS energies . The local density is consistent with AGASA experiment, e.g. $`\rho _\mu (600m)=0.24\pm 0.02\pm (0.02)`$/m<sup>2</sup> for this work and $`0.25`$/m<sup>2</sup> for AGASA at $`3\times 10^{17}`$eV. Since models with very different predictions of $`\rho _\mu (600m)`$ have very similar $`\mu `$ content indices, we believe that the $`\mu `$ content indices are more reliable than the muon normalization.
We can take full advantage of the hybrid nature of our experiment by constructing a parameter which is a combination of the $`X_{max}`$ and $`\rho _\mu (600m)`$ measurements. Based on the simulation with fully considered shower development and detector responses, a new parameter can be determined by first finding a line in $`X_{max}`$-$`\mathrm{log}\rho _\mu (600m)`$ space which efficiently separates simulated proton and iron showers (FIG. 3(a)). The dimension perpendicular to this line, called $`m`$, can be defined as $`m=\mathrm{cos}\gamma \mathrm{log}\rho _\mu (600m)`$-$`\mathrm{sin}\gamma X_{max}+m_0`$, where $`\gamma `$ refers to the angle between the line and the log$`\rho _\mu (600m)`$ axis and $`m_0`$ is an arbitrary parameter to make $`m>0`$ for iron and $`m<0`$ for proton. The value of $`m`$ for a given shower indicates the degree of similarity between the data and either a pure iron or a pure proton event. FIG. 3(a) shows that the separation between proton and iron showers is larger in terms of $`m`$ than it is for either $`X_{max}`$ or $`\rho _\mu (600m)`$, thus the fluctuations from event to event may be useful as an indication of the composition.
We plot the distribution of $`m`$ for three different energy ranges in FIG. 3. We compare this for the data (solid lines) with the estimates for simulated fluxes of pure protons (dashed lines) and pure iron (dotted lines) under the QGSJET assumption with full detector simulation and reconstruction. All the distributions are normalized. It is obvious that neither pure proton nor pure iron can account for the data for energies under $`3\times 10^{17}`$ eV. The data are highly peaked at $`m=0`$, implying a mixed composition around $`10^{17}`$ eV. The proportion of proton like events increases with energy. Note that due to the discrepancy regarding the muon normalization between data and simulation, we have shifted the log$`\rho _\mu (600m)`$ values by adding 0.17 to the predictions. This makes the muon results consistent with the $`X_{max}`$ results in terms of the normalization.
Thus FIG. 3 shows a change in the mean value of $`m`$ indicating a lightening of the mean cosmic ray mass. In the three energy ranges indicated, $`m`$ has values of $`0.061\pm 0.015`$, $`0.148\pm 0.036`$ and $`0.42\pm 0.25`$. For comparison, $`m`$ for pure proton changes from -0.23$`\pm `$0.01 to -0.46$`\pm `$0.02 while $`m`$ for pure iron is relatively stable around 0.2.
We conclude that the HiRes-MIA hybrid experiment confirms the Fly’s Eye experiment result that the primary composition changes towards a lighter mix of nuclei from $`10^{17}`$ to $`10^{18}`$eV. This confirmation is nontrivial not only because of its unique combination of the simultaneous observation of shower longitudinal development and muon density on the ground, but also because of its significantly improved $`X_{max}`$ and energy resolution. While the conclusion regarding the primary composition depends on the interaction model used, this study shows that the elongation rate is relatively stable with respect to choice of models. No modern interaction model has produced an $`\alpha _0`$ much larger than 60 g/cm<sup>2</sup>/decade.
A change from a heavy to a light composition in this energy region may indicate the increasing abundance of extra-galactic cosmic rays. Indeed, the Flys’s Eye experiment reports a change in the spectral index near $`5\times 10^{18}`$eV which can be interpreted in this light. A number of new experiments, such as HiRes, the Pierre Auger Project and the Telescope Array, could address this issue. However, the source of the lower energy heavy composition remains a mystery. In that regard we note that there is a need to explore the energy region between $`10^{16}`$ to $`10^{17}`$ eV in order to connect our results with the measurements performed below $`10^{16}`$ eV. A measurement of the composition in this region may be crucial for the understanding of the sources of cosmic rays above the “knee”. |
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.