1: \documentclass[aps,floats,superscriptaddress]{revtex4}
2: \usepackage{amsmath}
3: \usepackage{graphicx}
4:
5: \def\GroupeEquations#1{\begin{subequations} #1 \end{subequations}}
6: \def\moy#1{\left\langle #1 \right\rangle}
7: \def\ppint{P\int}
8: \def\Im{\hbox{Im}}
9: \def\Tr{\text{Tr}}
10: \def\sgn{\text{sgn}\,}
11: \def\parent#1{\left(#1\right)}
12: \def\fig#1#2{\includegraphics[height=#1]{#2}}
13:
14: \begin{document}
15:
16: \title{Quantum Fluctuations of a Nearly Critical Heisenberg Spin Glass}
17:
18: \author{A. Georges}
19: \affiliation{Laboratoire de Physique Th{\'e}orique, Ecole Normale
20: Sup{\'e}rieure, 24 Rue Lhomond 75005 Paris FRANCE}
21: \author{O. Parcollet}
22: \affiliation{Center for Material Theory,
23: Department of Physics and Astronomy, Rutgers University, Piscataway, NJ
24: 08854 USA}
25: \author{S. Sachdev}
26: \affiliation{Department of Physics, Yale University, New Haven, CT
27: 06520 USA}
28: \begin{abstract}
29: We describe the interplay of quantum and thermal fluctuations in
30: the infinite-range Heisenberg spin glass. This model is
31: generalized to $SU(N)$ symmetry, and we describe the phase
32: diagram as a function of the spin $S$ and the temperature $T$.
33: The model is solved in the large $N$ limit and certain universal
34: critical properties are shown to hold to all orders in $1/N$. For
35: large $S$, the ground state is a spin glass, but quantum effects
36: are crucial in determining the low $T$ thermodynamics: we find a
37: specific heat linear in $T$ and a local spectral density of spin
38: excitations, $\chi^{\prime\prime}_{loc} ( \omega ) \sim \omega$
39: for a spin glass state which is marginally stable to fluctuations
40: in the replicon modes. For small $S$, the spin-glass order is
41: fragile, and a spin-liquid state with $\chi^{\prime\prime}_{loc}
42: \sim \tanh (\omega/2T)$ dominates the properties over a
43: significant range of $T$ and $\omega$. We argue that the latter
44: state may be relevant in understanding the properties of
45: strongly-disordered transition metal and rare earth compounds.
46: \end{abstract}
47:
48: \maketitle
49:
50:
51: The study of intermetallic compounds of the transition metals and
52: rare earths has been a subject at the forefront of condensed
53: matter physics for some time now \cite{itp,coleman}.
54: A rich and complex variety of
55: behaviors is observed in low temperature electrical and magnetic
56: measurements, much of which lacks a comprehensive theoretical
57: description. The complexity arises from the dominant role played
58: by the local magnetic moments on the $d$ and $f$ orbitals and
59: their interactions with each other and the itinerant charge
60: carriers.
61:
62: It is convenient to begin our discussion in a phase with
63: well-established magnetic order, in which
64: each magnetic
65: moment is effectively static. This static moment could be
66: polarized in a regular manner (as in a commensurate
67: antiferromagnet or an incommensurate spin density wave), or point
68: in random directions (as in a spin glass state). In most realistic
69: systems, the magnetic moment is either quite small, or has
70: averaged to zero by dynamic quantum fluctuations: so it is
71: useful to consider mechanisms which reduce the magnetic moment,
72: and eventually cause it to vanish at a quantum phase transition to
73: some paramagnetic state. Two distinct routes to such a quantum
74: phase transition can be envisaged, and, we believe, the interplay
75: between them is at the heart of the complexity of the problem. In
76: the first route, originally discussed by Doniach \cite{Doniach}, the moment is
77: quenched by Kondo screening by the itinerant electrons: theories
78: of such quantum critical points have been proposed\cite{Hertz,Millis,SRO,AG}
79: in which the
80: predominant role of the itinerancy is to overdamp the collective
81: magnetic excitations. In the second route, the exchange
82: interactions between the moments play a more fundamental role:
83: a pair of spins interacting with an antiferromagnetic exchange
84: prefers to form a singlet valence bond, and the proliferation of
85: such singlets can destroy the magnetic order. Analytic theories
86: for such transitions has been made mainly for systems without
87: quenched disorder\cite{SachdevBook}.
88: Simple models of crossovers between these two
89: routes have also been presented\cite{SCS,GS,AMS}.
90:
91: This paper will present a detailed study of the second route
92: to destruction of magnetic order for the case of a strongly random
93: system with spin glass magnetic order. There are a number of
94: motivations for
95: focusing on random systems. First, randomness is
96: inevitably present in all materials, and it is clear that it
97: strongly perturbs the low temperature properties. Spin glass order
98: is present in a number of systems, while others appear to be in the
99: vicinity of such a state. Finally, a technical motivation is in
100: the structure of the mean-field theory we shall present: it builds
101: in important feedback effect between the inter-site magnetic
102: correlations and the single-site spin dynamics, and this is
103: crucial to all the non-trivial spin correlations we shall
104: describe. Such a feedback is absent in previous studies of the
105: magnetic quantum critical point, and it has been argued that this
106: is an important limitation for them\cite{AG,AMS}.
107: A different route to
108: incorporating these feedback effects has been taken in some recent
109: studies\cite{Si}; however, they discuss only the paramagnetic state of
110: their model, and the extent to which magnetically ordered states
111: preempt their results remains to be clarified.
112:
113: This paper is organized as follows : In Section \ref{outline}, we
114: present our spin glass model and give an outline of our results,
115: including the phase diagram. Section
116: \ref{ParamagneticPhases} is devoted to the nature of the
117: paramagnetic solutions, and more specifically to the quantum
118: critical regime. Section \ref{SpinGlassPhases} is devoted to the
119: spin glass phase and to the various regimes within this phase, as
120: a function of temperature and of the size of the spin. The
121: Appendices contain technical details and some additional results
122: on the quantum rotor and Ising spin glasses of
123: Ref.~\onlinecite{rsy}.
124:
125: \section{Model and outline of the results}\label{outline}
126:
127: The numerous recent studies of quantum
128: fluctuations in spin glasses \cite{RevueBhatt},
129: have focused either on infinite-range models of Ising
130: and rotor models \cite{millerhuse,rsy}
131: %(which don't have Berry phases in their quantum
132: %path integral because the ground state of each site, when
133: %decoupled from the others, is non-degenerate)
134: or models in low dimensions which flow to
135: strong disorder fixed points \cite{dsf,Pich,Motrunich}.
136: Here we shall continue the study of infinite range models,
137: but will consider a model of Heisenberg spins : in this case,
138: the path integral for each spin has a important Berry phase term
139: which imposes the spin
140: commutation relations.
141: %, and each site, when decoupled,has
142: % a $2S+1$ fold degeneracy.
143: As we will see, this leads to a great deal of
144: new physics\cite{SachdevYe} and non-trivial dynamic spin correlations
145: even in the spin glass state.
146: More specifically, we present a complete solution of
147: the quantum Heisenberg spin glass on a fully
148: connected lattice of $\cal N$ sites with strong Gaussian disorder,
149: both in the paramagnetic and the glassy phase,
150: when the spin symmetry group is extended from $SU(2)$ to $SU(N)$ and the
151: large-$N$ limit is taken.
152: In the limit of large connectivity, (dynamical) mean-field techniques
153: apply and the
154: model can be reduced to the study of a self-consistent single-site
155: problem, which is however
156: still highly non-trivial because of quantum effects.
157: The large-$N$ limit is instrumental in allowing for an explicit
158: solution. Nevertheless some of our results regarding the quantum
159: critical regime have been extended beyond the large-N limit.
160: In a recent publication \cite{NotreLettre}, we summarized the main
161: results of the present study. Here, we provide detailed derivations
162: and new results, such as a full discussion of the paramagnetic
163: phases and a discussion beyond large-$N$.
164:
165: The model considered in this paper is defined by the Hamiltonian:
166: \begin{equation}\label{DefSachdevYe}
167: H = \frac{1}{\sqrt{{\cal N}N}} \sum_{i<j} J_{ij}
168: \vec{S}_{i}\cdot \vec{S}_{j},
169: \end{equation}
170: where the magnetic exchange couplings $J_{ij}$ are independent,
171: quenched random variables distributed according to a Gaussian distribution
172: \begin{equation}\label{DefLoiJ}
173: P (J_{ij}) = \frac{1}{J\sqrt{2\pi}} e^{-J_{ij}^{2}/ (2J^{2})}
174: \end{equation}
175: As already pointed out by Bray and Moore \cite{BrayMoore},
176: after using the replica trick to
177: average over the disorder \cite{MezardParisiBook}, the mean-field (infinite
178: dimensional) limit maps the model onto a {\sl self-consistent single
179: site model} with the action (in imaginary time $\tau$, with $\beta$ the
180: inverse temperature) :
181: \begin{equation}\label{ActionEffective}
182: S_{eff} = S_{B} - \frac{J^{2}}{2N} \int_{0}^{\beta} d\tau d\tau'\,
183: Q^{ab} (\tau -\tau')
184: \overrightarrow{S}^{a} (\tau )\cdot \overrightarrow{S}^{b} (\tau')
185: \end{equation}
186: and the self-consistency condition
187: \begin{equation}\label{CondCoherence}
188: Q^{ab} (\tau -\tau') = \frac{1}{N^{2}}
189: \moy{
190: \overrightarrow{S}^{a} (\tau )\cdot \overrightarrow{S}^{b}(\tau')
191: }_{S_{eff}}
192: \end{equation}
193: where $a,b=1,\cdots,n$ denote the replica indices (the limit
194: $n\rightarrow 0$ has to be taken later) and $S_{B}$ is the Berry
195: phase of the spin \cite{SachdevYe}. Due to their time-dependence,
196: the solution of these mean-field equations remains a very
197: difficult problem for $N=2$, even in the paramagnetic phase.
198: Thus, in Ref.~\onlinecite{BrayMoore}, as well as in most
199: subsequent work \cite{kopec}, the {\it static approximation} was
200: used, neglecting the $\tau-$dependence of $Q^{ab}(\tau)$. This
201: approximation may be reasonable in some regimes but prevents a
202: study of the quantum equilibrium dynamics, and is particularly
203: inappropriate in the quantum-critical regime. However, this
204: imaginary time dynamics has been explicitly studied in a Quantum
205: Monte Carlo simulation {\sl in the paramagnetic phase} with spin
206: $S=1/2$ by Grempel and Rozenberg \cite{Grempel1}. Recently, we
207: introduced a large-$N$ solution of the mean field problem
208: \cite{NotreLettre}, in which the problem is exactly solvable and,
209: as explained below, the solution provides a good description of
210: the physics of the $N=2$ mean field model, to the extent the
211: latter is understood. More specifically, in the following we will
212: consider two different types of spin representations for the $SU
213: (N)$ spins :
214: \begin{itemize}
215: \item [a)] {\sl Bosonic} representations : the spin operator $S$ is
216: represented using Schwinger bosons $b$ by $S_{\alpha \beta}=
217: b^{\dagger}_{\alpha }b_{\beta } - S\delta_{\alpha \beta }$,
218: with the constraint $\sum_{\alpha}
219: b^{\dagger}_{\alpha }b_{\alpha }= SN$ ($0\leq S$). In
220: the language of Young tableaux, these representations are described by
221: one line of length $SN$. They
222: are a natural generalisation of an $SU (2)$ spin of
223: size $S$.
224: \item [b)] {\sl Fermionic} representations : the spin operator $S$ is
225: represented using Abrikosov fermions $f$ by $S_{\alpha \beta}=
226: f^{\dagger}_{\alpha }f_{\beta } - q_{0} \delta_{\alpha \beta }$,
227: with the constraint $\sum_{\alpha}
228: f^{\dagger}_{\alpha }f_{\alpha }= q_{0}N$ ($0\leq q_{0}\leq 1$). In
229: the language of Young tableaux, these representations are described by
230: one column of length $q_{0}N$.
231: Note that for $SU (2)$, only $S=0$ and $S=1/2$ can be represented in this
232: manner.
233: \end{itemize}
234: In the following, we refer to the model with bosonic (resp.
235: fermionic) representations
236: as the bosonic (resp. fermionic) model. In the
237: fermionic model, quantum fluctuations are so strong in large-$N$
238: that the spin glass ordering is destroyed \cite{SachdevYe},
239: contrary to the bosonic model, where a spin glass phase exists,
240: as explained below. The two models have different theoretical
241: interest : if one wants to concentrate on the quantum critical
242: regime, above the spin glass ordering temperature, one can use the
243: fermionic model (as e.g. in Ref.~\onlinecite{Slush}). However,
244: since we are interested in the spin glass phase itself, we will
245: now focus on the bosonic model. Nevertheless, our results on
246: the paramagnet will be valid for both cases with only slight
247: modifications explicitly quoted below.
248:
249:
250: In the $N\rightarrow \infty$ limit, the mean field self-consistent
251: model (\ref{ActionEffective}) reduces to
252: an integral equation for the Green's function of the boson
253: $G_{b}^{ab} (\tau) \equiv - \overline{\moy{T b^{a} (\tau)
254: b^{\dagger b} (0)}}$ where the bar denotes the average over disorder
255: and the brackets the thermal average \cite{SachdevYe} :
256: \begin{subequations}\label{EqBase}
257: \begin{align}
258: (G_{b}^{-1})^{ab} (i\nu_{n}) &=i\nu_{n}\delta_{ab} +
259: \lambda^{a}\delta_{ab} - \Sigma_{b}^{ab} (i\nu
260: _{n}) \\
261: \Sigma^{ab}_{b}(\tau ) &=\ J^{2} \bigl( G_{b}^{ab} (\tau)\bigr)^{2}
262: G_{b}^{ab} (-\tau) \\
263: G_{b}^{aa} (\tau=0^{-} ) &= - S
264: \end{align}
265: \end{subequations}
266: Similarly for the fermionic model, we have :
267: \begin{subequations}\label{EqBasefermions}
268: \begin{align}
269: (G_{f}^{-1})^{ab} (i\omega_{n}) &=i\omega_{n}\delta_{ab} +
270: \lambda^{a}\delta_{ab} - \Sigma_{f}^{ab} (i\omega
271: _{n}) \\
272: \Sigma^{ab}_{f}(\tau ) &=\ -J^{2} \bigl( G_{f}^{ab} (\tau)\bigr)^{2}
273: G_{f}^{ab} (-\tau) \\
274: G_{f}^{aa} (\tau=0^{-} ) &= q_{0}
275: \end{align}
276: \end{subequations}
277: In these equations, $\nu$ (resp. $\omega$) are the bosonic (resp.
278: fermionic) Matsubara frequencies and the inversion should be
279: taken with respect to the replica indices $a,b$. Note that our
280: conventions for the sign of the Green functions in this paper
281: slightly differ from those of Ref.~\onlinecite{SachdevYe}. Note
282: that, although the equations are written in term of $G$, the
283: physical quantity is the {\it local } spin susceptibility
284: $\chi_{loc} (\tau )=\moy{S (\tau )S (0)}$ which is given in the
285: large-$N$ limit by
286: \begin{equation}\label{DefChi}
287: \chi_{loc} (\tau )= G_{b}^{aa} (\tau)G_{b}^{aa} (-\tau )
288: \end{equation}
289: In many instances below (where we mostly focus on the bosonic case),
290: we shall drop the index $b$ in $G_b$.
291:
292: From both analytical and numerical analysis of these integral
293: equations, we have constructed the phase diagram displayed on Figure
294: \ref{DiagrammePhase}, as a function of the size of
295: the spin $S$ and the temperature $T$.
296: \begin{figure}[ht]
297: \[
298: \fig{10cm}{DiagrammePhase.eps}
299: \]
300: \caption{\label{DiagrammePhase}
301: \sl
302: Phase diagram of the mean field bosonic model. There is a spin glass
303: phase below the spin glass temperature $T_{sg}^{c}$, which is determined
304: with the marginality condition (See Section
305: \ref{SpinGlassPhases}). $T^{eq}_{sg}$ is the spin glass temperature as
306: determined with the stationarity criterion.}
307: \end{figure}
308: %
309: Let us give here a brief overview of the main features of this
310: phase diagram, which will be studied in great detail in the rest
311: of this paper. As is evident from Fig~\ref{DiagrammePhase}, it is
312: useful to divide the discussion into models with $S$ large and
313: $S$ small. Both regimes are accessible in the large $N$ limit,
314: where $S$ is effectively a continuous parameter taking all
315: positive values. For the physical case $N=2$, we will present
316: evidence later that at least $S=1/2$ is in the small $S$ regime
317: for the infinite-range model; moreover, we can expect that at least
318: some of the consequences of increased quantum fluctuations in
319: a realistic model with
320: finite-range interactions are mimicked by taking small $S$ values
321: in the large $N$ theory of the infinite-range model.
322:
323: For large $S$, the ground state must clearly be a spin glass
324: (Fig~\ref{DiagrammePhase}). However, even for very large $S$, it is
325: necessary to consider quantum effects in understanding the low $T$
326: excitations and thermodynamics, and these have not been previously
327: described. In this paper (Sections~\ref{sec:largeS}
328: and~\ref{Thermo}) we will show that the local spin
329: susceptibility has a low energy density of states which increases
330: linearly with energy. At the same time, the specific heat also has
331: a linear dependence upon temperature. These results hold for
332: temperatures $T < J \sqrt{S}$, although characteristic excitations
333: have an energy of order $JS$ for $T < JS$; we will provide scaling
334: functions which determine the dynamic response functions at these
335: energies. At even higher $T$, there is a phase transition to a
336: paramagnet at $T \sim J S^2$. For large $S$ the static properties of this
337: phase transition
338: are well-described by a purely classical theory in which the
339: $\vec{S}$ in (\ref{DefSachdevYe}) are commuting vectors of length
340: $S$. Notice also that we indicate two critical temperatures,
341: $T_{sg}^c$ and $T_{sg}^{eq}$: as we discuss in Section
342: \ref{SpinGlassPhases}, these are a consequence of peculiarities in
343: the nature of replica symmetry breaking, where the dynamic
344: freezing into the spin glass phase ($T_{sg}^c$), happens
345: at a slightly higher temperature than the equilibrium transition
346: ($T_{sg}^{eq}$).
347:
348: For small $S$, we also find a spin glass phase
349: (Fig~\ref{DiagrammePhase}) at $T=0$, but the order vanishes at a
350: small $T$. Moreover, its excitations and finite $T$ properties
351: are very different from those at large $S$. These are now
352: dominated by signals of a ``spin-liquid'' state discussed in
353: Ref.~\onlinecite{SachdevYe} and described in
354: Section~\ref{SubSectSpinliquid}. In particular, we describe a
355: novel quantum-critical region in the paramagnet where
356: $\mbox{max}(\omega, T)$ is the characteristic energy scale, and
357: the local dynamic spin susceptibility obeys
358: $\chi^{\prime\prime}_{loc} (\omega ) \sim \tanh(\omega/2T)$. We
359: believe that aspects of this regime may be relevant to disordered
360: transition metal and rare earth compounds in regimes where
361: exchange interactions between the magnetic moments are playing a
362: dominant role. Completion of this picture requires an
363: understanding of the stability of the ``spin-liquid'' picture to
364: mobile charge carriers, and this has also been addressed in a
365: previous work \cite{Slush}.
366:
367:
368: \section{Paramagnetic phase}\label{ParamagneticPhases}
369:
370: Contrary to the classical case, the paramagnetic phase of quantum
371: spin-glass models is non trivial in mean field theory. An early
372: discussion of these solutions has been given in
373: Ref.~\onlinecite{SachdevYe}, but we present here a much more
374: complete description, and compare our results to the $N=2$ case,
375: when numerical results are available. Since in this section we
376: look for paramagnetic solutions, we will consider only {\sl
377: replica diagonal} solutions of (\ref{EqBase}) : $G^{ab}\propto
378: \delta_{ab}$. Two types of paramagnetic solutions have been
379: found, that we will now consider successively : the {\it
380: spin-liquid} solutions and the {\it local moment} solutions.
381:
382: \subsection{Spin-liquid solutions}\label{SubSectSpinliquid}
383:
384: \subsubsection{The large-$N$ limit}
385:
386: A low-frequency, long-time analysis of the integral equations
387: (\ref{EqBase}) reveals that, under the condition that
388: $\lambda-\Sigma(i0^+)$ vanishes at low temperature, a solution
389: can be found which displays a power law decay of the Green's
390: function at long time \cite{SachdevYe}: $G (\tau )\sim
391: 1/\sqrt{\tau}$. These solutions display a singularity in the
392: complex plane of frequencies, $z$, at $z =0$, with an amplitude
393: which can be parameterized by an angle $\theta$ as:
394: \begin{equation}\label{SingulariteSpinliquid}
395: G (z)\sim \frac{Ae^{-i\pi /4-i\theta }}{\sqrt{z}} \qquad \qquad
396: \text{ for } z\rightarrow 0,\ \Im z>0
397: \end{equation}
398: (Values of $z$ on the imaginary frequency axis at the Matsubara
399: frequencies will be denoted by $\nu_n$, while on the real axis
400: will be denoted, $\omega$.)
401: Thus, these solutions display a slow local spin dynamics :
402: $\Im \chi _{loc} (\omega )\propto \sgn \omega $ for
403: $\omega \rightarrow 0, T=0$. Moreover, the {\it local} susceptibility
404: $\chi_{loc} (T)\equiv \int_{0}^{\beta }\chi_{loc} (\tau ) \, d\tau $
405: diverges as $\chi_{loc}
406: (T) \sim \ln T/J$ at low temperature.
407: More precisely, one can find the thermal scaling function characterizing
408: the $\tau \rightarrow \infty, T\rightarrow 0 $ limit, as explained in
409: a previous paper \cite{Slush} :
410: \begin{equation}\label{FormeEchelle}
411: \chi_{loc} (\tau ,\beta )
412: \propto \left(\frac{\pi /\beta }{\sin \pi \tau /\beta }
413: \right) + \dots \qquad \qquad \qquad
414: J\,\chi_{loc}^{\prime\prime}(\omega,T) \propto
415: \tanh\frac{\omega}{2T}
416: \end{equation}
417: Note that in the paramagnetic phase of {\it quantum} models, the local
418: susceptibility $\chi_{loc} (T)$ (which is the response to a {\it
419: local} magnetic field) differs from the uniform susceptibility $\chi (T)$
420: (response to a constant magnetic field), contrary to {\it classical}
421: spin glass models where $\chi = \chi_{loc}$ \cite{MezardParisiBook} :
422: this is a consequence of
423: the commutation relations of the spin, as can be seen for example in
424: the high-temperature expansion in the $SU (2)$ model.
425: In this large-$N$ limit, it can be shown that $\chi (T) \ll
426: \chi_{loc} (T)$ for $T\rightarrow 0$ and numerical computations
427: indeed suggest that $\chi (T)\sim \mbox{const.}$ \cite{Slush}.
428:
429: Remarkably, the parameter $\theta$ which characterizes the {\sl
430: spectral asymmetry} \cite{KondoLong} of the spectral density at
431: low frequency can be explicitly related to the size $S$ of the
432: spin (which involves a priori an integral of the spectral density
433: over all frequencies). This is very similar to a kind of Friedel
434: sum rule applying to this problem, and indeed the derivation
435: follows a very similar route, based on the existence of a
436: Luttinger-Ward functional. (Interestingly enough, the ``boundary
437: term'' which usually vanishes in such derivations contributes
438: here a finite value). This derivation is presented in detail in
439: Appendix \ref{AppSpectralAsym}, where the following relation
440: between $\theta$ and $S$ is established:
441: \begin{equation}\label{ValueOfTheta}
442: \frac{\theta }{\pi } + \frac{\sin 2\theta }{4} =
443: \left\{
444: \begin{aligned}
445: \frac{1}{2} + S& \text{ \qquad in the bosonic model}\\
446: \frac{1}{2} - q_{0}& \text{\qquad in the fermionic model}
447: \end{aligned}
448: \right.
449: \end{equation}
450: This relation has important consequences for the physical properties of the
451: spin-liquid solutions. First, we note that the spectral density must
452: obey the positivity conditions: $\Im G_f(\omega+i0^+) < 0$ and
453: $\mbox{sgn}(\omega)\Im G_b(\omega+i0^+) < 0$.
454: Hence, in the fermionic case, $\theta$ must obey $-\frac{\pi }{4}\leq \theta \leq
455: \frac{\pi }{4}$. It is easily checked from (\ref{ValueOfTheta}) that
456: $\theta$ precisely describes this range of parameters as $q_0$ is varied
457: from $q_0=0$ to $q_0=1$, and that the $\theta(q_0)$ relation is unique.
458: This suggests that the spin-liquid solution
459: is an acceptable low-temperature solutions for the whole range of
460: $q_0$ in the fermionic case.
461: \begin{figure}[ht]
462: \[
463: \fig{6cm}{CompTheoNumSLB.eps}
464: \]
465: \caption{\label{PlotOfTheta}
466: \sl
467: $S$ as a function of $\theta$ for the bosonic model.
468: The solid line is given by relation
469: (\ref{ValueOfTheta}) and the points were obtained previously from a
470: numerical solution of the saddle-point equation at zero
471: temperature \cite{SachdevYe}.}
472: \end{figure}
473: %
474: In contrast, in the bosonic case, the plot in Fig.\ref{PlotOfTheta}
475: shows that (\ref{ValueOfTheta}) actually defines {\it two} values of
476: $\theta$ (in the allowed range $\frac{\pi }{4}\leq \theta \leq
477: \frac{3\pi }{4}$) for a given spin $S$ as long as $S<S_{max}\simeq
478: 0.052$, while no value of $\theta$ is found for $S>S_{max}$. This
479: implies that no paramagnetic solution of the spin-liquid type is found
480: {\it at zero-temperature} in the bosonic case as soon as $S>S_{max}$
481: (note that furthermore $S_{max}$ is very small). For $S<S_{max}$, such
482: solutions exist at zero temperature
483: (even though they are not the true ground-state,
484: see below) with the locally stable solution corresponding to the
485: smallest of the two values of $\theta$. However, even for $S>S_{max}$,
486: {\it at low ($T<J$) but finite temperature}, the spin liquid solutions
487: do exist in the bosonic model. By this, we mean that a
488: numerical computation in imaginary time gives a solution which
489: exhibits the scaling form (\ref{FormeEchelle}), for which a
490: unambiguous value of
491: the spectral asymmetry $\theta$ can be defined and computed
492: numerically.
493: At very low temperature, these solutions are unstable to the spin glass solution,
494: but above the spin glass temperature at low spin, they are relevant
495: in the quantum critical regime associated
496: with the quantum critical point at $S=0$.
497: We shall comment in more detail, at the end of the following section, on the
498: nature of the paramagnetic solutions found at low temperature for small
499: values of $S$, in the bosonic case.
500:
501: Another consequence of relation
502: (\ref{ValueOfTheta}) is that it allows to predict that these
503: spin-liquid solutions have a {\it non-zero extensive entropy} at
504: zero-temperature and to calculate the value of this entropy
505: analytically. The derivation of this result follows very closely a
506: similar analysis of the overscreened multichannel Kondo problem
507: in the large-N limit, performed in
508: Ref.~\onlinecite{KondoLong,KondoPRL} and only the main steps will
509: be repeated here. This can be done either in the bosonic model or
510: in the fermionic one, with slight modifications. Since the spin liquid
511: solutions are relevant at zero temperature only in the fermionic model,
512: we shall present the result in this case.
513: First, denoting by ${\cal S}$ the value of the
514: entropy per spin at zero temperature, one establishes the
515: following thermodynamic equality:
516: \begin{equation}
517: \frac{\partial{\cal S}}{\partial q_{0}} =
518: -\frac{{\partial\lambda}}{{\partial T}}|_{T=0}
519: \end{equation}
520: Then, a low temperature expansion is used which allows to relate the slope
521: of $\lambda(T)$ to
522: the spectral asymmetry parameter $\theta$ above, so that one
523: finally gets (in the fermionic case):
524: \begin{equation}
525: \frac{\partial{\cal S}}{\partial q_{0}} =
526: \ln
527: \frac{\sin(\pi/4-\theta)}{\sin(\theta+\pi/4)}
528: \end{equation}
529: The entropy is then obtained by integration over the size of the
530: spin, with the physically obvious boundary conditions
531: ${\cal S}(q_0=0)={\cal S}(q_0=1)=0$. The resulting
532: value of the entropy as a function of $q_0$ is plotted in
533: Fig.\ref{PlotEntropy}.
534: \begin{figure}[ht]
535: \[
536: \fig{6cm}{Entropie.eps}
537: \]
538: \caption{\label{PlotEntropy}
539: \sl
540: Entropy as a function of the size of the spin ($q_{0}$) in the
541: fermionic model.}
542: \end{figure}
543:
544:
545: Finally, we comment on the physical nature of the spin-liquid
546: paramagnetic solutions found in this section. These solutions
547: correspond to a partial screening of the local moment at each
548: site, due to the interaction with the other spins. As a result
549: the local susceptibility diverges logarithmically (much slower
550: than a Curie law), but an extensive entropy is still present at
551: $T=0$, indicating a degenerate state. From a local point of view,
552: the physics is somewhat similar to an overscreened Kondo system,
553: but here the gapless bath which quenches the spin is not external
554: but self-consistently generated by the other spins. We suspect
555: that the physics of this phase has to do with the degeneracy of
556: the (large-$N$ generalization) of the ``triplet'' state in which two
557: spins are bound whenever a strong ferromagnetic bond $J_{ij}$ is
558: encountered. In Section \ref{SpinLiquidN=2}, we show that this
559: spin-liquid regime is not a peculiarity of the large-$N$ limit but
560: indeed survives in the mean-field description of the quantum
561: critical regime of a $SU(2)$ quantum Heisenberg spin-glass. It
562: would be very valuable to gain a more direct understanding of
563: this gapless spin-liquid regime from a study of the problem for a
564: fixed configuration of bonds, before averaging over disorder.
565: This could be achieved numerically and is left for future studies.
566:
567: \subsubsection{Beyond the large-$N$ limit}
568: \label{SpinLiquidN=2}
569: This subsection will show how
570: recent renormalization group analysis of related models\cite{AMS,SBV,VBS}
571: imply that
572: the above spin-liquid solution applies to all orders in $1/N$. In
573: particular, the large $N$ solution with $\mbox{Im} \chi_{loc}
574: \propto \mbox{sgn}~ \omega$ for small $\omega$ acquires no
575: corrections to its functional form: the only changes are to the
576: non-universal proportionality constant.
577: All the discussion below will be in a paramagnetic
578: phase where it is sufficient to consider only a single replica,
579: and so we will drop replica indices in this subsection.
580:
581: We begin by rewriting (\ref{ActionEffective}) in the following
582: form\cite{AMS}
583: \begin{equation}
584: S_{eff} = S_B - \gamma_0 \int_0^{\beta} d\tau \overrightarrow{S} (\tau)
585: \cdot \overrightarrow{\phi} (\tau)
586: \label{ss1}
587: \end{equation}
588: where $\gamma_0$ is a coupling constant and
589: $\overrightarrow{\phi}$ is an annealed Gaussian random field
590: with $\langle \overrightarrow{\phi} (\tau) \cdot
591: \overrightarrow{\phi} (0) \rangle = 1/|\tau|^{2-\epsilon}$.
592: It is reasonable to expect that the spin correlations in the
593: quantum ensemble defined by (\ref{ss1}) decay with the power-law
594: $\langle \overrightarrow{S} (\tau) \cdot
595: \overrightarrow{S} (0) \rangle \sim 1/|\tau|^{\sigma}$, and we are
596: interested in determining the value of the exponent $\sigma$.
597: A simple extension\cite{SachdevYe} of the
598: solution discussed above implies that in the large $N$ limit
599: $\sigma = \epsilon$. Here we will argue that this equality is in fact
600: {\em exact for all $N$}. Now using the self-consistency condition
601: (\ref{CondCoherence}) we obtain $\sigma=\epsilon=1$, which then
602: implies $\mbox{Im} \chi_{loc}
603: \propto \mbox{sgn}~ \omega$.
604:
605: The field-theoretic renormalization group analysis of (\ref{ss1})
606: was discussed in Ref.~\onlinecite{VBS}, and we will highlight the
607: main results. The key observation is that renormalization of the
608: theory (\ref{ss1}) requires only a single wave-function
609: renormalization factor $Z$, and that there is no independent
610: renormalization of the coupling constant $\gamma_0$. This result
611: was established diagrammatically in Ref.~\onlinecite{VBS}, and we
612: will not reproduce the argument here. So if we renormalize the
613: spin by $\overrightarrow{S} = \sqrt{Z} \overrightarrow{S}_R$,
614: then the coupling constant renormalization is simply $\gamma_0 =
615: \mu^{\epsilon/2} \gamma /\sqrt{Z}$, where $\mu$ is a
616: renormalization scale. The renormalization constant is in general
617: a complicated function of $\gamma$, and was determined to
618: two-loop order in Ref.~\onlinecite{VBS}:
619: \begin{equation}
620: Z = 1 - \frac{2\gamma^2}{\epsilon} + \frac{\gamma^4}{\epsilon} +
621: \ldots
622: \label{ss2}
623: \end{equation}
624: in a minimal subtraction scheme. However, even though $Z$ is not
625: known exactly, the exponent $\sigma$ can be determined exactly.
626: Standard field-theoretical technology shows that the above
627: renormalizations imply the $\beta$-function
628: \begin{equation}
629: \beta (\gamma) = - \frac{\epsilon \gamma}{2} \left( 1 -
630: \frac{1}{2} \frac{\partial \ln Z}{\partial \ln \gamma}
631: \right)^{-1}.
632: \label{ss3}
633: \end{equation}
634: Furthermore, the exponent $\sigma$ is given by the value of
635: \begin{equation}
636: \sigma (\gamma) = \beta(\gamma) \frac{\partial \ln Z}{\partial
637: \gamma}
638: \label{ss4}
639: \end{equation}
640: at the fixed point $\gamma=\gamma^{\ast}$ where $\beta(\gamma)$
641: vanishes. Comparing (\ref{ss3}) and (\ref{ss4}) we see that
642: \begin{equation}
643: \beta (\gamma) = -(\epsilon - \sigma(\gamma)) \gamma/2.
644: \label{ss5}
645: \end{equation}
646: Clearly, a zero of the $\beta$ function must have $\sigma =
647: \epsilon$, and this establishes the required result.
648:
649: We note that similar examples of a critical exponent being valid to all
650: orders (in spite of a non-trivial $\beta$-function) can be found for
651: other models in the statistical mechanics of disordered systems
652: (see e.g \cite{forster,honko}).
653:
654: \subsection{``Local moment'' solutions}
655:
656: In a mean field model, one usually expects to find locally stable
657: (while possibly unstable to ordering) paramagnetic solutions of the
658: mean-field equations down to zero temperature. Hence, the absence
659: of solutions of the spin-liquid type for $S>S_{max}$ suggests that
660: a different kind of paramagnetic solution should exist for those
661: values of the spin. Indeed, we have found that the integral equations
662: (\ref{EqBase}) have another class of paramagnetic solutions
663: in the bosonic case.
664: These solutions actually exist for all values of the spin $S$ and down
665: to zero temperature. Hence, they coexist at low $S$ with the spin-liquid
666: solutions in some range of temperature.
667: Their physical nature is very different from the previous spin-liquid
668: solutions, and as discussed below they are not very physical solutions
669: when considered at low temperature. They are characterized
670: by a Green's function which does not decay at long times and obeys the
671: asymptotic behavior:
672: $G_{b} (\tau ) \simeq -S - e^{-J^{2}\beta S^{3}\tau}$. In contrast to
673: the spin-liquid case, $\lambda$ diverges for $T\rightarrow 0$ in this
674: regime:
675: \begin{equation}\label{NewParaLambdaLowT}
676: \lambda \sim \frac{J^{2} S^{3}}{T}
677: \end{equation}
678: Finding numerically these solutions of (\ref{EqBase}) requires
679: some care. We have used an algorithm in which
680: we solve (\ref{EqBase}) in imaginary time for $G(\tau)$, for a fixed
681: value of $r=\int_{0}^{\beta }G (\tau )d\tau$,
682: and then adjust the number of particles to $S$ by a dichotomy on $r$.
683: %
684: \begin{figure}[ht]
685: \[
686: \fig{6cm}{Ki.eps}
687: \]
688: \caption{\label{KiNewParaFig}
689: \sl
690: $\chi_{loc} (\tau )$ extracted from a numerical solution
691: of saddle point equation (\ref{EqBase}) in
692: imaginary time, for $S=1$, $J=1$.
693: The solid curve is low temperature ($J\beta =10$), the dashed curve is
694: high temperature ($J\beta =.1$); for intermediate temperatures, the
695: curves interpolate between the two.}
696: \end{figure}
697: %
698: The local susceptibility $\chi_{loc} (\tau )$ obtained in this manner
699: is displayed on Figure
700: \ref{KiNewParaFig}. At high temperature, we find $\chi_{loc} (T)=
701: \frac{S (S+1)}{T}$ as expected since the spin is
702: essentially free. At low temperature, we find another Curie law, with a
703: reduction of the Curie constant due to quantum fluctuations :
704: \begin{equation}\label{ReducedCurie}
705: \chi_{loc} (T) = \frac{S^{2}}{T} \qquad\qquad \hbox{ for }
706: T\rightarrow 0
707: \end{equation}
708: Hence, for these solutions, the effect of the interactions with the other
709: spins is not strong enough to result in a qualitatively different
710: screening regime, resulting merely in a reduction of the Curie constant.
711: This is analogous to an {\it underscreened} Kondo regime.
712:
713: These solutions are of a similar type than those found in in a
714: quantum Monte-Carlo simulation of the $SU (2)$ model
715: \cite{Grempel1}. There also, a reduction of the Curie constant
716: from $S (S+1)/3$ to $S^{2}/3$ was clearly observed. Thus,
717: contrary to one of the conclusions of Ref.~\onlinecite{Grempel1},
718: the large-$N$ limit correctly reproduces the paramagnetic local
719: moment solution found for the physical $N=2$ case. Moreover, a
720: numerical solution of the large-N integral equations
721: (\ref{EqBase}) for real frequencies can also be obtained both at
722: high and low temperature. The results are presented in Figure
723: \ref{NewParaOmega} for both $\rho_{b} (\omega )\equiv -1/\pi \Im
724: G_{b} (\omega )$ and $\chi_{loc}'' (\omega )\equiv 1/\pi \Im
725: \chi_{loc} (\omega )$. At high temperature, $\rho_{b}$ is centered
726: around $\lambda \sim T\ln ((S+1)/S)$ and $\chi_{loc}'' (\omega )$
727: is a simple peak. At low temperature, we find in $\chi_{loc}''
728: (\omega)$ a delta-function peak at zero frequency with weight
729: $S^{2}$, which is associated with longitudinal relaxation (see
730: Ref.~\onlinecite{Grempel1} for a discussion for $N=2$) and two
731: peaks, centered around $\pm \lambda \propto 1/T$ (Cf Eq.
732: (\ref{NewParaLambdaLowT})) with a constant width, which are
733: associated with transverse relaxation. Again, these results are
734: very similar to the conclusions reached in
735: Ref.~\onlinecite{Grempel1} from a fit of the imaginary-time data
736: for $N=2$, using a model $\chi^{''}(\omega)$. The only difference
737: is that the central peak is not broadened by thermal fluctuations
738: in the large-$N$ limit.
739: \begin{widetext}
740: \begin{figure}[ht]
741: \[
742: \includegraphics[height=10cm]{LocalMoment.eps}
743: \]
744: \caption{
745: \sl
746: \label{NewParaOmega}
747: Spectral densities for $G_{b}$ (left) and for the local
748: susceptibility $\chi_{loc}$ (right), for high temperature (top) and a
749: very low temperature (bottom). These results are
750: extracted from a numerical solution of saddle point equation (\ref{EqBase}) in
751: real frequencies.}
752: \end{figure}
753: \end{widetext}
754: Finally, since these solutions describe the formation of a
755: local moment at low
756: temperature, and that the onset of the quenching temperature
757: (at which the reduction of the Curie constant sets in) turns out to be
758: lower than the temperature where spin-glass ordering occurs (as shown in the
759: next section),
760: we consider these solutions as a mean-field artefact of the spin glass
761: ordering. Static limits of such solutions are actually known
762: to occur in the classical SK model. We also note that the internal
763: energy of these solutions have an unphysical divergence
764: as $T\rightarrow 0$.
765:
766: We close this section by noting that, for small values of the spin
767: $S$, a rather intricate pattern emerges for the stability and
768: coexistence of the two kinds of paramagnetic solutions described
769: above. We have studied this numerically in some detail but do not
770: report this here, since most of these phenomena occur below the
771: spin-glass ordering temperature anyway. The important features
772: have been displayed on Fig.\ref{DiagrammePhase}. We emphasize
773: again that the spin-liquid solutions are the relevant solutions
774: describing the whole quantum-critical regime. Even though they
775: are unstable at $T=0$ for $S>S_{max}$ they remain consistent
776: {\it finite temperature} solutions in a much wider range of values
777: of $S$ for $T<J$. At higher $S$ and above the spin-glass ordering
778: temperature, the paramagnetic phase behaves as a local moment,
779: with a Curie constant getting gradually reduced as $T$ is lowered.
780:
781: \section{The spin-glass phase}\label{SpinGlassPhases}
782:
783: In this section, we investigate spin-glass ordering in this model.
784: The first observation that we make (Sec.\ref{sg_susc}) is that the
785: spin-glass susceptibility (i.e the response to a spin-glass ordering field)
786: is actually of order $1/N$ in the large-N limit. This does not
787: preclude a spin-glass phase, but means that the transition is not
788: associated associated with a linear instability. Indeed, we shall
789: find explicit solutions in the ordered phase in the bosonic case,
790: while the fermionic case does not have a spin-glass phase at
791: $N=\infty$ (but does order as soon as $1/N$ corrections are
792: considered).
793:
794: \subsection{Spin-glass susceptibility}\label{sg_susc}
795:
796: Here, we derive an exact expression for the spin-glass
797: susceptibility, valid for arbitrary $N$ in this mean-field model.
798: The derivation is most conveniently performed using replicas. We
799: note that in the presence of spin-glass order, the correlation
800: function $\moy{\overrightarrow{S}^{a} (\tau )\cdot
801: \overrightarrow{S}^{b}(\tau')}$ acquires a non-zero value for
802: $a\neq b$, which is however static, i.e independent of
803: $\tau-\tau'$ (see e.g Ref.~\onlinecite{SachdevBook}). This crucial
804: point is due to the fact that different replicas are independent
805: of one another before averaging, so that for $a\neq b$:
806: \begin{equation}
807: \overline{\moy{\overrightarrow{S}^{a}(\tau )\cdot
808: \overrightarrow{S}^{b}(\tau')}} =
809: \overline{\moy{\overrightarrow{S}^{a} (\tau )}\cdot
810: \moy{\overrightarrow{S}^{b}(\tau')}} =
811: \overline{\moy{\overrightarrow{S}^{a} (0)}\cdot
812: \moy{\overrightarrow{S}^{b}(0)}}
813: \end{equation}
814: In the following, we shall denote by $q_{ab}$ the (normalized) off-diagonal
815: correlation function which is an order parameter for the
816: spin-glass phase:
817: \begin{equation}
818: q_{ab} \equiv \frac{1}{N^2} \sum_{a\neq b}
819: \overline{\moy{\overrightarrow{S}^{a} (\tau )\cdot
820: \overrightarrow{S}^{b}(\tau')}}
821: \end{equation}
822: We consider the stability of the paramagnetic phase to this type
823: of ordering, and introduce an ordering field $H_{ab}$ conjugate to
824: $q_{ab}$. This has two effects:
825:
826: i) It adds to the effective action (\ref{ActionEffective}) an
827: explicit term:
828: \begin{equation}
829: \delta S = \frac{1}{N}
830: \int d\tau\int d\tau' \sum_{a\neq b} H_{ab}\,
831: \vec{S}_a(\tau)\cdot\vec{S}_b(\tau')
832: \end{equation}
833: The normalization of $H_{ab}$ has been chosen in such a way
834: that the change in the total free-energy is of order $N$.
835:
836: ii) It modifies the value of the self-consistent field $Q^{ab}$.
837: The change of the off-diagonal component, to linear order, is
838: imposed by the self-consistency condition (\ref{CondCoherence})
839: to be: $\delta Q^{ab}=\delta q_{ab}$, with $\delta q_{ab}$ the
840: induced order parameter.
841:
842: One can then perform an expansion of the off-diagonal correlation
843: function up to linear order in both $H_{ab}$ and $\delta Q^{ab}$.
844: This yields:
845: \begin{equation}
846: \delta q_{ab} = \frac{1}{N}
847: \chi_{loc}^2\,\left(H_{ab}+J^2\delta Q^{ab}\right)
848: \end{equation}
849: in which $\chi_{loc}$ is the local susceptibility of the
850: paramagnetic phase. Since $\delta Q^{ab}=\delta q_{ab}$, this
851: finally yields the susceptibility to spin-glass ordering:
852: \begin{equation}
853: \label{ChiSG}
854: \chi_{sg} \equiv \frac{\delta q_{ab}}{H_{ab}} =
855: \frac{1}{N}\,\frac{\chi_{loc}^2}{1-(J\chi_{loc})^2/N}
856: \end{equation}
857: This formula has two important consequences. The susceptibility
858: to spin-glass ordering is of order $1/N$
859: and any spin-glass instability at $N=\infty$ must be associated
860: with a non-linear effect of higher order (i.e come from terms of
861: higher than quadratic order in $q_{ab}$ in the free energy).
862: Furthermore, (\ref{ChiSG}) shows that for finite $N$, a (linear)
863: instability into a spin-glass phase will occur when
864: $J\chi_{loc}(T)=\sqrt{N}$. Hence the fermionic model, for which
865: $\chi_{loc}$ diverges at low-T, will have a spin-glass instability
866: for arbitrary large but finite $N$. More precisely, since
867: the low-temperature behavior $J\chi_{loc} \sim \ln \frac{J}{T}$
868: has been shown above to hold for all N for the paramagnetic solution of
869: the fermionic case, we conclude from (\ref{ChiSG}) that the spin-glass
870: transition temperature depends on $N$ in that case as :
871: $T_c^{f}\sim J e^{-\sqrt{N}}$.
872:
873:
874:
875: \subsection{Spin glass solutions and the ``replicon'' problem}
876:
877: We now turn to the explicit construction of solutions of the
878: integral equations (\ref{EqBase}) with spin-glass ordering, in the
879: bosonic case. The same reasoning as above shows that the Green's
880: function $G^{ab} (\tau )$ does not depend on $\tau$ for $a\neq b$.
881: Thus the most general Ansatz for the Green's function $G^{ab}$ can be
882: written as :
883: \begin{equation}\label{AnsatzParisi}
884: G^{ab} (\tau ) =\left\{
885: \begin{aligned}
886: \widetilde{G} (\tau ) - g_{1} & \qquad \qquad (a=b)\\
887: - g_{ab} & \qquad \qquad (a\neq b)\\
888: \end{aligned}
889: \right.
890: \end{equation}
891: where $\widetilde{G} (\tau )$ is a function of the imaginary time,
892: $g_{ab}$ a constant matrix and $g_{1}$ a constant.
893: By definition, $g_{1}$ is fixed so that $\widetilde{G}$ is
894: regular at $T=0$, {\sl i.e. } $\widetilde{G} (\tau) \rightarrow 0$ as
895: $\tau \rightarrow \infty$.
896: In the following discussion, we will restrict ourselves to solutions
897: given by a Parisi Ansatz for $g_{ab}$ (a replica symmetry breaking scheme).
898: In the $n\rightarrow 0$ limit (where $n$ is the number of replica),
899: this matrix becomes a function $g (u)$ of a
900: continuous variable $u$ with $0\leq u\leq 1$ \cite{MezardParisiBook}.
901:
902: Our equations (\ref{EqBase}) involve the Green function $G$, whereas the
903: physical quantity is the susceptibility $\chi^{ab} (\tau )=G^{ab}
904: (\tau )G^{ab} (-\tau )$. The order parameter $q_{ab}$ widely introduced in
905: the spin glass literature \cite{MezardParisiBook} is
906: given here by $q_{ab} =g_{ab}^{2}$, or in the limit $n\rightarrow 0$ :
907: \begin{equation}
908: q (u) = g (u)^{2}
909: \end{equation}
910: The Edwards-Anderson parameter is $q_{EA} = q (1)=g (1)^{2}$
911: \cite{MezardParisiBook,SachdevBook}. Since at zero temperature,
912: in the long time limit, we find $\lim_{\tau \rightarrow \infty
913: }{G^{aa} (\tau )G^{aa} (-\tau )} = q_{EA}$ we find $g_{1} = g
914: (1)$, by definition of $g_{1}$. More precisely, we look for
915: solutions in which these two definitions of $q_{EA}$ coincide but
916: it is not really an assumption in our computation : if this
917: relation was violated, we would simply find for $\widetilde{G}$ a
918: non vanishing limit for $\tau \rightarrow \infty$.
919:
920: Among the various possible replica symmetry breaking schemes
921: \cite{MezardParisiBook}, we will now focus on {\it one-step}
922: solutions, since we have not found any other, either two-steps
923: or with continuous replica symmetry breaking. In this case, the
924: function $g(u)$ is piecewise constant : $g (u)= \overline{g}$ for
925: $0<u<x$ and $g(u)=g$ for $x\leq u\leq 1$. In the following, we
926: will refer to $x$ as the {\sl breakpoint}. According to
927: (\ref{EqBase}), the self-energy $\Sigma$ has the form :
928: \begin{equation}\label{AnsatzParisiSigma}
929: \Sigma^{ab} (\tau ) = \left\{
930: \begin{aligned}
931: \widetilde{\Sigma} (\tau ) - J^{2} g^{3} & \qquad \qquad (a=b)\\
932: -J^{2} g_{ab}^{3} & \qquad \qquad (a\neq b)\\
933: \end{aligned}
934: \right.
935: \end{equation}
936: with $\widetilde{\Sigma }$ is given by (\ref{EqSG_DefSigTilde}).
937: Using the standard formulas to invert the Parisi matrices
938: in the limit $n\rightarrow 0$ \cite{MezardParisiManifold}, we find :
939: \begin{align}\label{Intermed1}
940: \widetilde{G}^{-1} (i\nu_{n}) = i\nu_{n} + \lambda
941: -\widetilde{\Sigma}(i\nu_{n}) \\
942: \label{Intermed2} J^{2} g^{2} \widetilde{G}(i\nu_n = 0)^{2} = 1 +
943: J^{2} \beta x g^{3} \widetilde{G} (i \nu_n = 0)
944: \end{align}
945: At this stage, it is useful to introduce here a new parameter $\Theta$
946: defined by :
947: \begin{equation}\label{DefTheta}
948: \widetilde{G} (i \nu_n ) = -\frac{\Theta}{Jg}
949: \end{equation}
950: We then eliminate $\lambda$ in (\ref{Intermed1}) and $\widetilde{G} (0)$
951: in (\ref{Intermed2}) and obtain finally a closed set of equations for
952: $\widetilde{G}$ and $g$ :
953: \begin{subequations}\label{EqSG}
954: \begin{equation}\label{EqSG_Dyson}
955: \left( \widetilde{G} (i\nu_{n})\right)^{-1} = i\nu_{n} - \frac{Jg}{\Theta } -
956: \left(\widetilde{\Sigma} (i\nu_{n}) - \widetilde{\Sigma} (0) \right)
957: \end{equation}
958: \begin{equation}\label{EqSG_DefSigTilde}
959: \widetilde{\Sigma} (\tau ) = J^{2} \left(
960: \widetilde{G}^{2} (\tau )\widetilde{G} (-\tau ) -
961: 2 g \widetilde{G} (\tau ) \widetilde{G} (-\tau) - g
962: \widetilde{G}^{2} (\tau ) + 2 g^{2} \widetilde{G} (\tau )
963: + g^{2} \widetilde{G} (-\tau )\right)
964: \end{equation}
965: \begin{equation}\label{EqSG_NumberParticle}
966: \widetilde{G} (\tau =0^{-}) = - (S - g)
967: \end{equation}
968: \begin{equation}\label{EqSG_x}
969: \beta x = \frac{1}{J g^{2}} \left(\frac{1}{\Theta} - \Theta \right)
970: \end{equation}
971: \end{subequations}
972: %
973: The crucial observation at this point is that these
974: saddle-point equations possess a one parameter family of solutions,
975: parametrised by $\Theta $ or equivalently by the breakpoint $x$.
976: This phenomenon already occur in other models which have
977: a {\sl one-step} replica symmetry breaking solution \cite{GiamLedou}.
978: The determination of the breakpoint turns out to be the most difficult
979: question of this analysis.
980: Two possible criteria are :
981: \begin{enumerate}
982: \item To minimize the free energy ${\mathcal F}(x)$ as a function of $x$, as would be
983: required by the thermodynamics. In the
984: following, we will refer to this as the {\sl equilibrium
985: criterion}. This criterion has been used in a previous attempt to
986: understand this spin glass phase \cite{kopec}.
987: \item To impose a vanishing lowest eigenvalue of the fluctuation matrix
988: in the replica space. We will refer to this as
989: the {\sl marginality} or {\sl replicon } criterion.
990: Although it is not really justified up to now, we
991: will argue that it is the correct choice.
992: \end{enumerate}
993: This problem is not due to the quantum aspect of our model :
994: it already appears similarly in some classical spin glass
995: models, in the $p$-spin model for example.
996: In this classical model, the study of the dynamics shows the existence
997: of a dynamical transition $T^{dyn}$ above the static spin glass
998: temperature $T^{eq}$ given by the static solution of the mean field
999: model. It turns out that, in this classical model, the
1000: replicon criterion has been proven to give
1001: the same transition temperature $T^{c}=T^{dyn}$.
1002: Moreover, it has been shown \cite{CugliandoloPspinQ,LeticiaPspin}
1003: that the same
1004: phenomenon occurs in some {\sl quantum } version of the p-spin model.
1005: Thus, using this condition, it is possible in some sense to mimic the
1006: dynamics by simply solving a static problem, although this is not
1007: fully understood at present.
1008:
1009: In the present model, the two criteria give a coherent solution
1010: but with totally different spectra of {\it equilibrium}
1011: fluctuations : the equilibrium criterion leads to a gap in $\chi''
1012: (\omega )$ whereas the replicon criterion is the only one which
1013: give a {\sl gapless } $\chi''(\omega )$ (a similar observation was
1014: made in Ref.~\onlinecite{GiamLedou} in a one dimensional quantum
1015: model with disorder). We believe that in this quantum Heisenberg
1016: spin glass the replicon criterion provides us with the correct
1017: physical solution ($T_{c}$), contrary to the equilibrium
1018: solution, which gives the static transition temperature
1019: ($T_{eq}$). However this claim cannot be proved in the present
1020: context : in particular, the static solution does give a full
1021: solution of (\ref{EqBase}). A study of the true Hamiltonian
1022: dynamics in real time and finite temperature of this quantum
1023: problem is necessary for a deeper understanding of this question,
1024: but this is beyond the scope of this paper. Let us now examine
1025: the two criteria separately in more details.
1026:
1027: \subsubsection{The replicon criterion}
1028: \label{sec:replicon}
1029: To apply the replicon criterion, we need to study the
1030: fluctuations of the free energy in the replica space around the one-step
1031: solution.
1032: In the large-$N$ limit, the free energy is given by the expression :
1033: \begin{equation}\label{FormuleF1}
1034: {\mathcal F}[G^{ab}, \lambda ] = \frac{1}{\beta } \sum_{n} \Tr
1035: \ln \Bigl(i\nu_{n} + \lambda - \Sigma^{ab} (i\nu_{n}) \Bigr) +
1036: \frac{3J^{2}}{4} \sum_{ab} \int_{0}^{\beta} d\tau \, \left[
1037: G^{ab} (\tau ) G^{ab} (-\tau) \right]^2 - \lambda S
1038: \end{equation}
1039: Under infinitesimal variation $\delta g_{ab}$ for $a\neq b$,
1040: the variation of the free energy is (up to second order)
1041: \begin{equation}\label{DefMatrixFluctuations}
1042: \delta {\mathcal F}=\sum_{a>b\atop c>d} M_{ab,cd}\delta
1043: g_{ab}\delta g_{cd}.
1044: \end{equation}
1045: Strictly speaking, as this is a quantum problem,
1046: we have to simultaneously consider the
1047: variation of the diagonal component, $\delta\widetilde{G}(\tau)$ in
1048: (\ref{AnsatzParisi}) in a study of the fluctuations. In a spin
1049: glass phase, there is indeed a coupling between $\delta g_{ab}$
1050: and $\delta \widetilde{G} (\tau)$ which modifies the fluctuation
1051: eigenvalues. Fortunately however, as we show in Appendix
1052: \ref{AppReplicon}, this coupling does not modify the eigenvalue
1053: $e_{1}$ and our main result (\ref{ThetaMarginality}) below, and so
1054: we will neglect $\delta \widetilde{G} (\tau)$ here.
1055: The diagonalization of the $n (n-1)/2 \times n (n-1)/2 $ matrix $M$ is
1056: briefly explained in Appendix \ref{AppReplicon} and gives three eigenvalues
1057: \begin{eqnarray}\label{EigenvaluesFluct}
1058: \nonumber
1059: &e_{1} = 3\beta J^{2}g^{2} (1-3\Theta^{2})\\
1060: &e_{2} = \frac{3\beta J^{2}g^{2}}{\Theta^{2}}\left(\Theta^{2} -3
1061: +3\beta Jg^{2}\Theta (1+\Theta) \right)\\
1062: \nonumber
1063: &e_{3} = 6\beta J^{2}g^{2}\left(3\beta J g^{2} \Theta -1 \right)
1064: \end{eqnarray}
1065:
1066: A first consequence of this analysis is that replica symmetric
1067: solutions are unstable, since from Eq. (\ref{EqSG_x}) they correspond
1068: to $\theta =1$ and then $e_{1}<0$.
1069: Hence, these solutions will not be considered in the following discussion.
1070:
1071: A full solution of Eqs.(\ref{EigenvaluesFluct}) is required to show the
1072: positivity of $e_{2},e_{3}$, but we immediately see that $e_{1}=0$ for
1073: \begin{equation}\label{ThetaMarginality}
1074: \Theta_{R}= \frac{1}{\sqrt{3}}
1075: \end{equation}
1076:
1077: Quite remarkably, we will see below in Section~\ref{sec:largeS}
1078: and Appendix~\ref{app:gapless} that precisely the same
1079: value of $\Theta$ is selected by a criterion which is seemingly
1080: entirely independent. We will study the dynamic spectral functions
1081: in the spin-glass phase, as defined by $\widetilde{G}(\tau)$,
1082: and show that their associated spectral densities are non-zero
1083: as $|\omega| \rightarrow 0$ only for the value of $\Theta$ in
1084: (\ref{ThetaMarginality}). So marginal stability in replica
1085: space appears to be connected to a gapless quantum excitation
1086: spectrum. We may intuitively understand this as due to the
1087: availability of many low energy states when the system first
1088: freezes, but a better understand should emerge from a real-time
1089: analysis.
1090:
1091: \subsubsection{The equilibrium criterion}
1092: To apply the equilibrium criterion, we start from the expression
1093: of the free energy ${\mathcal F}$ and solve for $x$ :
1094: \begin{equation}\label{Critere1}
1095: \frac{d{\mathcal F} (x)}{dx}=0
1096: \end{equation}
1097: The computation of the {\sl total} derivative (\ref{Critere1})
1098: reduces to $\partial_{x}{\mathcal F} (x)|_{\widetilde{G},\lambda}$
1099: because the saddle-point equations (\ref{EqBase}) {\sl for finite
1100: $n$} are equivalent to
1101: \begin{equation}\label{PropF1}
1102: \frac{\partial {\mathcal F}}{\partial G^{ab}} =\frac{\partial
1103: F}{\partial\lambda }= 0
1104: \end{equation}
1105: as can be checked by an explicit calculation. Computing the
1106: logarithm in (\ref{FormuleF1}) (using Appendix II of
1107: Ref.~\onlinecite{MezardParisiManifold}) and taking the derivative
1108: leads to :
1109: \begin{equation}
1110: \frac{3}{4} J^{2} g^{4} - \frac{2}{(\beta x)^{2}} \ln \bigl(-J g
1111: \widetilde{G} (i\nu_n = 0)) = - \frac{ g}{\beta x \widetilde{G}
1112: (i\nu_n = 0) }
1113: \end{equation}
1114: and finally to a equation for $\Theta$ :
1115: \begin{equation}\label{ThetaStationnarity}
1116: 2 \ln \Theta + \frac{1}{4\Theta^{2}} + \frac{1}{2} -
1117: \frac{3\Theta^{2} }{4}=0
1118: \end{equation}
1119: This equation has two solutions : the replica symmetric one
1120: $\Theta=1$ (unstable, as explained above),
1121: and a non trivial one $\Theta =\Theta_{eq}\approx 0.4421\dots$.
1122: Contrary to the previous solution, we will
1123: see in Section~\ref{sec:largeS} that $\Im \widetilde{G}$ has a gap
1124: for this value of $\Theta$.
1125:
1126: \subsection{The phase diagram}
1127:
1128: Once $\Theta $ has been determined, the equations (\ref{EqSG}) can be
1129: solved either numerically (both in
1130: imaginary time and in real frequency) or analytically in the
1131: $S\rightarrow \infty$ limit. In the following, we
1132: will mainly restrict ourselves to $\Theta =\Theta_{R}$ since we believe that
1133: it is the correct solution. However all calculations have been redone for
1134: $\Theta =\Theta_{eq}$ with related results.
1135:
1136: \subsubsection{Numerical solution}
1137: First, the critical temperature, is obtained from the numerical solution of
1138: Eqs.(\ref{EqSG}) in imaginary time : the spin glass order
1139: parameter $q (T) = g^{2} (T)$ and the breakpoint $x (T)$ are displayed
1140: in Figure \ref{TcFig} as a function of the temperature.
1141: %
1142: \begin{figure}[ht]
1143: \[
1144: \fig{6cm}{Tc.eps}
1145: \]
1146: \caption{\label{TcFig} \sl The Edwards-Anderson parameter $q_{EA}$ and the
1147: breakpoint $x$ as a function of the temperature $T$ for a fixed value
1148: of the size of the spin $S=1$ ($J=1$). The transition to the
1149: paramagnet is given by the condition $x=1$.}
1150: \end{figure}
1151: %
1152: $x$ increases linearly with $T$ from 0 at $T=0$ (there is no
1153: replica symmetry breaking at zero temperature) and $T_{sg}$ is determined
1154: by the condition $x (T_{sg})=1$, since we must have $0\leq x\leq 1$ by
1155: definition \cite{MezardParisiBook}.
1156: Hence there is a discontinuity in $q$ at the transition, but we will
1157: show below that the transition is second order.
1158: A careful numerical study shows that the transition is always driven
1159: by $x=1$ for all values of $S$ and produces the critical temperature
1160: displayed on Figure \ref{DiagrammePhase}. The computation is similar
1161: for the critical temperature $T^{c}_{sg}$ given by the ``replicon''
1162: criterion and for the critical temperature $T^{eq}_{sg}$ given by
1163: the equilibrium criterion. Moreover,
1164: we find that $T^{c}_{sg}$, the {\sl dynamic} transition temperature, is {\bf
1165: higher} than $T^{eq}_{sg}$, the {\sl static} transition temperature :
1166: this is required by our physical interpretation of the two solutions
1167: but it was not obvious a priori from the integral equations solved.
1168:
1169: \subsubsection{The large-$S$ limit and spectral densities}
1170: \label{sec:largeS}
1171:
1172: Further analytical insight into the spin glass phase itself can be obtained by
1173: considering various large-$S$ limits, which differ by the manner in which
1174: temperatures and frequencies are scaled with $S$ (See
1175: Fig~\ref{DiagrammePhase}).
1176:
1177: In a first simple large-$S$ limit, we take $T$ large enough so
1178: that $T/JS^2$ is of order unity. This is the simple classical
1179: limit in which we can neglect all non-zero Matsubara frequencies,
1180: and (\ref{DefSachdevYe}) reduces to the classical problem in which
1181: $\vec{S}$ are commuting vectors of length $S$. The equations
1182: (\ref{EqBase}) are analytically solvable and we can obtain a
1183: closed form expression for the critical temperature at which
1184: spin-glass order vanishes:
1185: \begin{equation}\label{TsgClass}
1186: T^{c}_{sg} \sim \frac{2}{3\sqrt{3}} JS^{2}.
1187: \end{equation}
1188:
1189: A second, more sophisticated limit, valid at lower temperatures (well within
1190: the spin-glass phase) is when we examine $\omega$ and $T$ of order
1191: $JS$. It is therefore useful to define the variables
1192: $\overline{\omega} =\omega /(JS)$ and $\overline{T}=T/(JS)$, which remain of order
1193: unity at large $S$.
1194: With this scaling,
1195: the integral equations (\ref{EqSG}) reduce to independent quartic equations for
1196: each frequency.
1197: More precisely, if we make the following Ansatz for the Green's
1198: function
1199: \begin{equation}\label{DefScalingLimitLargeS}
1200: \widetilde{G} (\omega ,T) = \frac{1}{JS} g_{1} (\overline
1201: {\omega} ,\overline{T}) + \frac{1}{JS^{2}} g_{2}
1202: (\overline{\omega} ,\overline{T}) + \dots,
1203: \end{equation}
1204: then to leading order in $1/S$, (\ref{EqSG}) reduce to
1205: \begin{eqnarray}\label{EqQuartic}
1206: g (T) &=& S - \int_{-\infty }^{\infty} \rho_{1} (\overline{\omega}
1207: ) d\overline{\omega} \nonumber \\
1208: g_{1} (\overline{\omega} )^{-1} &=& \overline{\omega} -
1209: \frac{1}{\Theta }- 3\Theta - 2g_{1} (\overline{\omega} ) -
1210: \overline{g_{1} (-\overline{\omega})}
1211: \end{eqnarray}
1212: where $\rho_{1} =- \Im g_{1}/\pi $ as usual. Eliminating the
1213: frequency $-\overline{\omega} $, we find a quartic equation for
1214: $g_{1} (\overline{\omega} )$. We do not explicitly display the far
1215: more complicated equation for the subleading term $g_2$.
1216:
1217: The solution of the quartic equation for $\Theta=\Theta_R$
1218: is presented on Figure
1219: \ref{FigRho} together with a numerical solution of the full integral
1220: equation for $S=5$.
1221: From the solution of the quartic equation we find that $\rho_1$
1222: vanishes linearly frequency at low frequencies; indeed, we find the analytic expansion
1223: \begin{equation}
1224: g_1 (\overline{\omega}) = - \frac{1}{\sqrt{3}} - \frac{(1+i)}{2}
1225: \overline{\omega} + \frac{(2-3i)\sqrt{3}}{4} \overline{\omega}^2
1226: \label{ss10}
1227: \end{equation}
1228: at low frequencies. We expect that the full Green's function in
1229: (\ref{DefScalingLimitLargeS}) also has a similar low frequency
1230: expansion, although
1231: this has not been proved. It is not difficult to show that the linear
1232: low frequency
1233: spectral density holds at higher orders in the $1/S$ expansion;
1234: moreover, our numerical
1235: results, shown in Fig~\ref{FigRho}, also clearly indicate a linear
1236: behavior at small $\omega$. At dominant order in the present large $S$ theory, the spin susceptibility is given by :
1237: \begin{equation}\label{KiInSpinGlass}
1238: \chi'' (\omega ) = - \pi \int_{0}^{\omega } \, dx \rho_{1} (x)
1239: \rho_{1} (x-\omega )
1240: + g\pi \left(\rho_{1} (\omega ) - \rho_{1} (-\omega ) \right) + \pi
1241: g^{2} \beta \omega
1242: \delta (\omega )
1243: \end{equation}
1244: and this is also shown in Figure \ref{FigRho}.
1245: \begin{figure}[ht]
1246: \[
1247: \fig{5cm}{Rho.eps} \kern 2cm \fig{6cm}{Kiom.eps}
1248: \]
1249: \caption{\label{FigRho} \sl
1250: {\sl a)} $\rho_{1} (\omega )$ at $T\approx 0$ for $S=5$ and
1251: $\Theta=\Theta_R$ : the solid line
1252: is the numerical
1253: solution for the integral equation (\ref{EqSG}), the dashed line is
1254: the solution of the quartic equation (\ref{DefScalingLimitLargeS}) for
1255: $g_{1}$.
1256: {\sl b)} $\chi'' (\omega )/\omega $ at $T\approx 0$ from the quartic
1257: equation.}
1258: \end{figure}
1259:
1260: It is important to realize that the deceptively simple structure
1261: in (\ref{ss10}) relies on the special value $\Theta=\Theta_R =
1262: 1/\sqrt{3}$ determined by the entirely different replicon
1263: argument in Section~\ref{sec:replicon}. For arbitrary values of
1264: $\Theta$ we either find no physically sensible solution of the
1265: large $S$ quartic equation (this is the case at the replica
1266: symmetric value $\Theta = 1$ where the spectral density does not
1267: satisfy the required positivity criteria) or a solution with a
1268: spectral gap. In the latter case, the solution for $g_1$ is real
1269: for small real $\overline{\omega}$, and there is an onset in the
1270: imaginary part $\sim (\overline{\omega} -
1271: \overline{\omega}_c)^{1/2}$ above some critical frequency
1272: $\overline{\omega}_c$. The solution for $\Theta=\Theta_{eq}$ is of
1273: the second type: it has a finite energy gap, but does not violate
1274: any spectral positivity criteria.
1275:
1276: This subsection has so far identified two distinct large $S$ regimes.
1277: In the regime $T \sim JS^2$ we have purely classical behavior (the non-zero
1278: Matsubara frequencies can be neglected for static properties) and
1279: a phase transition at a critical temperature in (\ref{TsgClass}) where
1280: the spin-glass order vanishes. At lower temperatures, $T \sim JS$, we are well
1281: within the spin-glass phase, and the semiclassical dynamics is described
1282: by the solution of a quartic equation defined by (\ref{EqQuartic}).
1283: As we noted in Fig~\ref{DiagrammePhase}, there is a third ``quantum''
1284: regime at even lower temperatures, $T \sim J \sqrt{S}$, and this
1285: becomes evident in a study of the thermodynamic properties presented
1286: in the following section.
1287:
1288: \subsection{Thermodynamics}\label{Thermo}
1289:
1290: We now turn to the internal energy $U$ and the specific heat $C$.
1291: By computing the average of the Hamiltonian in the $N\rightarrow
1292: \infty $ limit, we find that $U$ is given by :
1293: \begin{equation}\label{FormulaU}
1294: U (T) = -\frac{J^{2}}{2}\int_{0}^{\beta } d \tau \left[
1295: G^{ab}(\tau ) G^{ab} (-\tau ) \right]^2
1296: \end{equation}
1297: Using (\ref{AnsatzParisi}) and the one-step replica symmetry
1298: breaking Ansatz in the spin glass phase, we find :
1299: \begin{equation}\label{FormuleDeveloppeeU}
1300: U (T) = \left\{
1301: \begin{aligned}
1302: -&\frac{J^{2}}{2}\int_{0}^{\beta } G(\tau )^{2}G(-\tau )^{2}d\tau
1303: & \qquad \text{ in the paramagnetic phase} \\
1304: -&\frac{J^{2}}{2}\int_{0}^{\beta }
1305: \bigl (\widetilde{G}(\tau) -g \bigr )^{2} \bigl
1306: (\widetilde{G}(-\tau )-g \bigr )^{2}d\tau - \frac{J^{2}}{2} \beta
1307: (x-1)g^{4}
1308: & \qquad \text{ in the spin glass phase} \\
1309: \end{aligned}
1310: \right.
1311: \end{equation}
1312:
1313: A numerical computation of the internal energy $U (T)$ and the
1314: specific heat $C (T)$ is displayed in Figure \ref{CetU} for
1315: $\Theta=\Theta_R$.
1316: \begin{figure}[ht]
1317: \[
1318: \fig{8cm}{CetU.eps}
1319: \]
1320: \caption{\label{CetU} \sl The specific heat $C (T)$ and the
1321: internal energy $U (T)$ vs. the temperature $T$, from a numerical
1322: solution of Eqs. (\ref{EqSG}) for $S=5$ and $\Theta=\Theta_R$. }
1323: \end{figure}
1324: The condition for the phase transition between the two phases is
1325: that the breakpoint in the Parisi function reaches its limiting
1326: value $x=1$. In this limit the equations determining the
1327: parameters in the spin glass phase, (\ref{EqSG}), transform
1328: continuously to those for the paramagnet. As the equations are
1329: believed to have a unique solution, this implies that there is no
1330: discontinuity in the internal energy at the transition, indicating
1331: its second-order nature. This is confirmed by the numerical
1332: solution displayed on Figure \ref{CetU}.
1333:
1334: Moreover, in the large-$S$ limit defined above, we can perform a
1335: low temperature expansion of the internal energy. Inserting
1336: (\ref{DefScalingLimitLargeS}) and (\ref{EqQuartic}) into
1337: (\ref{FormulaU}), and keeping the dominant terms at large $S$ we
1338: find
1339: \begin{eqnarray}
1340: U(T) &=& - \frac{J}{2} \left( \frac{1}{\Theta} + 3 \Theta \right)
1341: g^2 -J^2 g^2 \frac{1}{\beta} \sum_{\nu_n} \left[ \widetilde{G} (
1342: i \nu_n ) \widetilde{G} ( - i \nu_n) + 2 \widetilde{G}^2 (i \nu_n
1343: )
1344: \right] + \ldots \nonumber \\
1345: &=& - \frac{JS^2 }{2} \left( \frac{1}{\Theta} + 3 \Theta \right)
1346: + J S \left( \frac{1}{\Theta} + 3 \Theta \right) \overline{T}
1347: \sum_{\overline{\nu}_n} g_1 ( i \overline{\nu}_n) -J S
1348: \overline{T} \sum_{\overline{\nu}_n} \left[ g_1 (i
1349: \overline{\nu}_n ) g_1 ( - i \overline{\nu}_n) + 2 g_1^2 ( i
1350: \overline{\nu}_n ) \right] + {\cal O} (JS^0) \label{ssU}
1351: \end{eqnarray}
1352: where $\overline{\nu}_n = \nu_n / (JS)$. Clearly, this result
1353: indicates that the leading term in $U(T)$ is a
1354: temperature-independent constant of order $JS^2$, followed by a
1355: term of order $JS$ whose coefficient is a function only of
1356: $(T/JS)$. Evaluation of the latter function at low $T$ for
1357: $\Theta=\Theta_R$ yields a curious accident: the gapless
1358: structure of the spectral functions suggests that the low $T$
1359: expansion should depend only on even powers of $T/JS$, but it is
1360: not difficult to show using (\ref{ss10}) that the coefficient of
1361: the term of order $S(T/JS)^2$ vanishes. The first non-vanishing,
1362: $T$-dependent term among those shown explicitly in (\ref{ssU})
1363: turns out to be order $JS (T/JS)^4$. To obtain the true low $T$
1364: behavior we need to expand (\ref{ssU}) to one higher-order in
1365: $1/S$, and this requires use of the second term, $g_2$, in
1366: (\ref{DefScalingLimitLargeS}). We do not expect any cancellation
1367: of the term of order $(T/JS)^2$ at this point, and so the low $T$
1368: expansion for $U$ looks like
1369: \begin{equation}\label{DevEnergy}
1370: U(T) = U(0) + a S (T/JS)^{4} + b (T/JS)^{2} +\dots
1371: \end{equation}
1372: Rather than numerically evaluating the values $a$ and $b$, we
1373: will be satisfied by the full numerical solution of (\ref{EqSG}),
1374: followed by the evaluation of (\ref{FormuleDeveloppeeU}). The
1375: results are shown in Fig~\ref{CetU} and are consistent with
1376: (\ref{DevEnergy}). The structure of the expansion in
1377: (\ref{DevEnergy}) suggests that these results are valid for $T < J
1378: \sqrt{S}$, where the specific heat depends linearly on the
1379: temperature. Although the present discussion has been carried out
1380: for large $S$, we expect, and this is supported by our numerical
1381: results, that the linear $T$ dependence of the specific heat holds
1382: even for small $S$ as $T \rightarrow 0$.
1383:
1384: In Appendix~\ref{rotor} we describe the computation of the
1385: specific heat of the quantum rotor and Ising spin glasses
1386: considered in Ref.~\onlinecite{rsy}. As noted in the introduction,
1387: these models are simpler because they do not have quantum Berry
1388: phases in their effective action. Further, at low orders in their
1389: Landau theory, the solution for the spin glass phase is
1390: replica-symmetric. However, understanding the true $T \rightarrow
1391: 0$ behavior requires inclusion of higher-order, ``dangerously
1392: irrelevant'' terms which induce replica symmetry breaking; this
1393: is carried out in Appendix~\ref{rotor}, and we find that these
1394: quantum spin glasses also have a linear specific heat at low $T$.
1395:
1396: \section{Conclusion}\label{conclusion}
1397: We believe that the results of this paper provide a reasonably
1398: complete understanding of the infinite-range quantum Heisenberg
1399: spin glass. While there have been a large number of previous
1400: studies of quantum spin glasses of Ising spins and rotors
1401: (including models with $(p>2)$-spin interactions), none of these
1402: models contain quantum Berry phases in their effective actions,
1403: as is the case with the Heisenberg model.
1404: They have strong consequences: the spin-liquid solution of
1405: Section~\ref{SubSectSpinliquid} and its spectral density
1406: (\ref{FormeEchelle}) are novel properties of the Heisenberg
1407: model. There is an intricate interplay in stability between this
1408: spin-liquid state and the state with spin-glass order at low $T$
1409: which we have also described. At sufficiently low $T$, the
1410: spin-glass order always appears, and we have also described the
1411: thermodynamic properties of this state.
1412:
1413: An important issue not resolved in our analysis is the origin of
1414: the marginal stability criterion in the fluctuation eigenvalues
1415: in replica space. We imposed this criterion in a rather {\em ad
1416: hoc} manner, and found that it was the unique case under which the
1417: quantum excitation spectrum was gapless. Ultimately, the
1418: selection criterion for the spin glass state has to be a dynamic
1419: one, and this requires an analysis of the approach to equilibrium
1420: in real-time dynamics. Such an analysis was not carried out here,
1421: and is an important direction for future research.
1422:
1423: Another interesting open problem is to extend the study of
1424: (\ref{DefSachdevYe}) to cases where $J_{ij}$ has a non-zero
1425: average value. This will allow for ground states with other types
1426: of magnetic order, ferromagnetic and antiferromagnetic, and their
1427: competition with the spin glass state should be of some
1428: experimental interest. Interesting transitions in the
1429: paramagnetic states from the spin liquid state discussed also
1430: appear possible.
1431:
1432: We have already mentioned a recent study \cite{Slush} of the
1433: quenching of the spin liquid state by mobile charge carriers into
1434: a disordered Fermi liquid. Combining this with models just
1435: mentioned, with a non-zero average $J_{ij}$, should lead to
1436: results of direct physical interest in the heavy fermion and
1437: cuprate series of compounds.
1438:
1439: \begin{acknowledgements}
1440: We thank G. Biroli, L. Cugliandolo, D. Grempel, P. Le Doussal,
1441: M. Rozenberg for useful discussions.
1442: S.S. was supported by US NSF Grant No DMR 96--23181.
1443: O.P is supported by the Center of Material Theory, Rutgers University, NJ,USA.
1444: \end{acknowledgements}
1445:
1446: \begingroup
1447: \appendix
1448:
1449: \section{Computation of the spectral asymmetry}\label{AppSpectralAsym}
1450: This appendix is devoted to the derivation of Eqs
1451: (\ref{ValueOfTheta}).
1452: We will consider
1453: hereafter the fermionic case (the bosonic one is very similar).
1454: At zero temperature, the number of particles is given by
1455: \begin{equation}\label{NbrePart}
1456: q_{0} =i \int_{-\infty }^{\infty } \frac{d\omega}{2\pi } \,\,
1457: G_{f}^{F}(\omega )e^{i \omega 0^{+}}
1458: =i \ppint_{-\infty }^{\infty } \frac{d\omega}{2\pi } \,\,
1459: \partial_{\omega } \ln G_{f}^{F} (\omega ) e^{i \omega 0^{+}}
1460: - i \ppint_{-\infty }^{\infty } \frac{d\omega}{2\pi } \,\,
1461: G_{f}^{F} (\omega ) \partial_{\omega }\Sigma_{f}^{F} (\omega)
1462: e^{i \omega 0^{+}}
1463: \end{equation}
1464: where $G^{F}$ is the Green function with Feynman prescription on the
1465: real axis, and the symmetric principal part is defined by
1466: \begin{equation}\label{DefPP}
1467: \ppint_{-\infty}^{\infty} = \lim_{\eta \rightarrow 0 }{
1468: \int_{-\infty}^{-\eta} + \int_{\eta }^{\infty }}
1469: \end{equation}
1470: Using the relation between $G^{F}$ and the {\it retarded Green
1471: function} $G^{R}$, we find for the first term :
1472: \begin{equation}\label{Terme1}
1473: i \ppint_{-\infty }^{\infty } \frac{d\omega}{2\pi } \,\,
1474: \partial_{\omega } \ln G_{f}^{F} (\omega ) e^{i \omega 0^{+}}
1475: = \frac{\arg G_{f}^{R} (0^{-}) - \arg G_{f}^{R} (-\infty)}{\pi} +
1476: i \ppint_{-\infty }^{\infty } \frac{d\omega}{2\pi } \,\,
1477: \partial_{\omega } \ln G_{f}^{R} (\omega ) e^{i \omega 0^{+}}
1478: \end{equation}
1479: The arguments can be extracted from the low-energy and the high-energy
1480: behavior of the Green function, which leads to $\arg G_{f}^{R}
1481: (0^{-}) = -3 \pi /4 -\theta$ and $\arg G_{f}^{R} (-\infty) = -\pi $
1482: respectively. The integral on the right of (\ref{Terme1}) can be
1483: easily evaluated : we close the contour of integration, avoiding the
1484: singularity at $\omega =0$ and use the analyticity of the retarded
1485: Green function in the upper half plane, in which it has no zeros nor
1486: poles.
1487: We find finally :
1488: \begin{equation}\label{Finalq01}
1489: q_{0} = \frac{1}{2} - \frac{\theta }{\pi} -
1490: i \ppint_{-\infty }^{\infty } \frac{d\omega}{2\pi } \,\,
1491: G_{f}^{F} (\omega ) \partial_{\omega }\Sigma_{f}^{F} (\omega)
1492: e^{i \omega 0^{+}}
1493: \end{equation}
1494:
1495: The problem is now reduced to the computation of the integral in
1496: (\ref{Finalq01}) as function of $\theta$, which turn out to be
1497: the most difficult point. An analogous computation was performed
1498: in the overscreened regime of a large-$N$ description of Kondo
1499: effects \cite{KondoPRL,KondoLong}, but it turns out to be more
1500: complex here. Proceeding along the lines of
1501: Ref.~\onlinecite{KondoPRL,KondoLong}, we note the existence of
1502: Luttinger-Ward functional
1503: $\Phi_{LW} = \int dt \,G^{2} (t)G^{2} (-t)$
1504: which has two properties : first we have $\Sigma^{R} (\omega ) =
1505: \delta \Phi_{LW} /\delta G^{R} (\omega )$; second $\Phi_{LW }$ is invariant
1506: in the transformation $G (\omega )\rightarrow G (\omega +\epsilon)$.
1507: From this, we could naively think that the integral of
1508: (\ref{Finalq01}) vanishes. However, it is not possible to find a
1509: regularization for the integral for which we could use the invariance
1510: of the Luttinger-Ward functional and a more careful analysis shows that :
1511: \begin{equation}\label{Anomalie}
1512: i \ppint_{-\infty }^{\infty } \frac{d\omega}{2\pi } \,\,
1513: G_{f}^{F} (\omega ) \partial_{\omega }\Sigma_{f}^{F} (\omega)
1514: = \frac{\sin 2\theta }{4}
1515: \end{equation}
1516:
1517: To obtain this result, we introduce the following parametrisation of
1518: the singularity at $\omega =0$ :
1519: \begin{equation}\label{DefC+C-}
1520: \rho (\omega ) \sim
1521: \left\{
1522: \begin{aligned}
1523: \frac{C_{+}}{\sqrt{\omega}} &\text{\qquad \qquad for $\omega >0$}\\
1524: \frac{C_{-}}{\sqrt{|\omega|}} &\text{\qquad \qquad for $\omega <0$}
1525: \end{aligned}
1526: \right.
1527: \end{equation}
1528: The principle of the computation is very simple : we compute
1529: explicitly the integral with a regulator $\eta>0$ and
1530: then perform the limit $\eta \rightarrow 0$.
1531: Going to the real axis, we find :
1532: \begin{equation}\label{ExprSigF}
1533: \Sigma^{F}(\omega ) = -
1534: \int_{\substack{\\ \\
1535: \omega_{1}>0\\
1536: \omega_{2}>0\\
1537: \omega_{3}<0\\
1538: }
1539: \substack{\\ \\ \\ \text{ ou }\\ \\}
1540: \substack{\\ \\
1541: \omega_{1}<0\\
1542: \omega_{2}<0\\
1543: \omega_{3}>0\\
1544: }
1545: }
1546: %
1547: d \omega_{1} d \omega_{2} d \omega_{3} \,\,
1548: \frac{\rho (\omega_{1})\rho (\omega_{2})\rho (\omega_{3})}{
1549: \omega_{1} + \omega_{2}-\omega_{3} -\omega -i0^{+}\sgn \omega_{1}}
1550: \end{equation}
1551: Using the notations $a = \overline{a} + i\epsilon_{a}$, $b =
1552: \overline{b} + i\epsilon_{b}$, $\epsilon_{a/b} = \pm 0^{+}$ and
1553: $\psi_{\eta }(x) = \Theta \left(|x| - \eta \right)$
1554: ($\overline{a}$ are $\overline{b} $ real and $\Theta$ is the
1555: Heaviside function), we obtain (using the definition of the
1556: principal part (\ref{DefPP}))
1557: \begin{align}\label{Utile1}
1558: \phi_{\eta } (a,b) = & \ppint_{-\infty}^{\infty} \frac{dz}{\left(z-a
1559: \right)^{2} (z-b)}\\
1560: \nonumber
1561: = & \frac{1}{(a-b)^{2}}
1562: \left(
1563: \ln \left| \frac{(\eta +b) (\eta -a)}{(\eta -b) (\eta +a)} \right|
1564: + i\pi \psi_{\eta } (b) \sgn \epsilon_{b} - i\pi \psi_{\eta }(a)
1565: \sgn \epsilon_{a}
1566: \right)
1567: + \frac{1}{a-b}\left(\frac{1}{\eta -a} + \frac{1}{\eta +a} \right)
1568: \end{align}
1569: Using the spectral representation for $G^{F}$ and (\ref{ExprSigF}),
1570: we find :
1571: \begin{equation}\label{R1}
1572: {\cal I} = -
1573: \int_{\Delta_{1}\cup \Delta_{2}}^{}\prod_{k=0}^{3} d \omega_{k} \,\,
1574: \rho (\omega_{0})\rho (\omega_{1})\rho (\omega_{2})\rho (\omega_{3})
1575: \phi_{\eta} (\omega_{1} +\omega_{2}-\omega_{3}-i\epsilon_{1}\sgn
1576: \omega_{1}, \omega_{0} - i\epsilon_{0}\sgn \omega_{0})
1577: \end{equation}
1578: with an explicit integration over $\omega $ with (\ref{Utile1}).
1579: In this expression, the integration domains are defined as :
1580: \begin{equation}\label{Domaines}
1581: \Delta_{1} =
1582: \left\{\begin{aligned}
1583: \omega_{0}&<0\\
1584: \omega_{1}&>0\\
1585: \omega_{2}&>0\\
1586: \omega_{3}&<0
1587: \end{aligned}\right\}
1588: \bigcup
1589: \left\{\begin{aligned}
1590: \omega_{0}&>0\\
1591: \omega_{1}&<0\\
1592: \omega_{2}&<0\\
1593: \omega_{3}&>0
1594: \end{aligned}\right\}
1595: \qquad \qquad
1596: \Delta_{2} =
1597: \left\{\begin{aligned}
1598: \omega_{0}&>0\\
1599: \omega_{1}&>0\\
1600: \omega_{2}&>0\\
1601: \omega_{3}&<0
1602: \end{aligned}\right\}
1603: \bigcup
1604: \left\{\begin{aligned}
1605: \omega_{0}&<0\\
1606: \omega_{1}&<0\\
1607: \omega_{2}&<0\\
1608: \omega_{3}&>0
1609: \end{aligned}\right\}
1610: \end{equation}
1611: Since $\phi_{\eta } (-a,-b) = - \phi_{\eta } (a,b)$,
1612: a simple change of variable leads to :
1613: \begin{align}\label{R2}
1614: \nonumber
1615: {\cal I} &=
1616: -\int_{x_{i}>0}
1617: \Biggl[
1618: \Bigl(
1619: \rho(\omega_{1})\rho(\omega_{2})\check\rho(\omega_{3})\check\rho(\omega_{0})-
1620: \check\rho(\omega_{1})\check\rho(\omega_{2})\rho(\omega_{3})\rho(\omega_{0})
1621: \Bigr)
1622: \phi_{\eta} (x_{1}+x_{2}+x_{3} - i\epsilon_{1},-x_{0}+i\epsilon_{0})\\
1623: & + \Bigl(
1624: \rho(\omega_{1})\rho(\omega_{2})\check\rho(\omega_{3})\rho(\omega_{0})-
1625: \check\rho(\omega_{1})\check\rho(\omega_{2})\rho(\omega_{3})
1626: \check\rho(\omega_{0})
1627: \Bigr)
1628: \phi_{\eta} (x_{1}+x_{2}+x_{3} - i\epsilon_{1},x_{0}-i\epsilon_{0})
1629: \Biggr]
1630: \end{align}
1631: with $\check \rho (\omega ) = \rho (-\omega )$.
1632: To take the limit $\eta \rightarrow 0$, we use the new variables
1633: $x_{i}=\eta u_{i}$ and the behavior of $\rho (x)$ for $x\rightarrow
1634: 0$, parametrised according to (\ref{DefC+C-}).
1635: The first integral in (\ref{R2}) vanishes at dominant order in
1636: $\eta$ (this term is proportional to $C_{+}^{2}C_{-}^{2}
1637: -C_{-}^{2}C_{+}^{2}=0 $), but the second integral gives :
1638: \begin{equation}\label{R3}
1639: {\cal I}= \int_{u_{i}>0} \frac{C_{+}^{3}C_{-} -
1640: C_{-}^{3}C_{+}}{\sqrt{u_{0}u_{1}u_{2}u_{3}}}
1641: \phi_{\eta=1} (u_{1}+ u_{2} + u_{3} -i\epsilon_{1},u_{0}-i\epsilon_{0})
1642: \end{equation}
1643: Using $x=u_{0}$, $y=u_{1}+u_{2}+u_{3}$ and polar coordinates in
1644: $\sqrt{u_{i}}$, we find
1645: ${\cal I} = 2\pi (C_{+}^{3}C_{-} - C_{-}^{3}C_{+}) {\cal I}_{2 }$
1646: with
1647: \begin{multline}\label{R4}
1648: {\cal I}_{2} = \int_{0}^{\infty }\!\!\!\! \int_{0}^{\infty }
1649: \frac{dx}{\sqrt{x}} \sqrt{y}dy
1650: \left[ \frac{1}{(x-y+i\epsilon )^{2}}
1651: \left(
1652: \ln \left| \frac{1+x}{1-x}\frac{1-y}{1+y} \right|
1653: + i \pi \left(\psi_{1} (y) - \psi_{1} (x) \right)
1654: \right) + \right.\\
1655: \left. \frac{1}{y-x-i\epsilon }
1656: \left(\frac{1}{1-y+i\epsilon_{1}} + \frac{1}{1+y-i\epsilon_{1}} \right)
1657: \right]
1658: \end{multline}
1659: After an integration by parts on $y$ and using %:
1660: %\begin{equation}\label{R5}
1661: %{\cal I}_{2} = -\frac{1}{2} \int_{0}^{\infty }\!\!\!\! \int_{0}^{\infty } \frac{dx}{\sqrt{x}}
1662: %\frac{dy}{\sqrt{y}} \frac{1}{x-y+i\epsilon}
1663: %\left( \ln \left|\frac{1+x}{1-x}\right| + \ln \left|\frac{1-y}{1+y}\right|+
1664: %i\pi \parent{\psi_{1} (y) - \psi_{1} (x)}
1665: %\right)
1666: %= -2 i\pi \int_{0}^{1} \frac{dx}{x} \ln \left|\frac{1+x}{1-x} \right|
1667: %\end{equation}
1668: %Since
1669: \begin{equation}\label{Utile2}
1670: \int_{0}^{1} \frac{dx}{x} \ln \left|\frac{1+x}{1-x} \right| = \frac{\pi^{2}}{4}
1671: \end{equation}
1672: we find
1673: \begin{equation}\label{R7}
1674: i\ppint_{-\infty }^{\infty } \frac{d\omega}{2\pi} \,\,
1675: G_{f}^{F} (\omega ) \partial_{\omega }\Sigma_{f}^{F} (\omega) =
1676: \frac{\pi^{3}}{2}(C_{+}^{3}C_{-} - C_{-}^{3}C_{+})
1677: \end{equation}
1678: Finally, a analogous computation can be performed in the bosonic case,
1679: leading in both cases to :
1680: \begin{equation}\label{FinalThetaPre}
1681: i\ppint_{-\infty }^{\infty} \frac{d\omega}{2\pi} \,\,
1682: G^{F} (\omega ) \partial_{\omega }\Sigma^{F} (\omega) =
1683: \frac{\sin 2\theta }{4}
1684: \end{equation}
1685: (in this expression, $-\pi \leq\theta \leq \pi $).
1686: These expressions have been shown to agree perfectly with numerical
1687: computations in
1688: imaginary time for the fermionic case and on the real axis at zero temperature
1689: in the bosonic case.
1690:
1691: Let us note finally that we can guess the result if we admit a priori
1692: that the integral is given
1693: by an homogeneous polynomial of degree 4 : due to the particle-hole
1694: symmetry (in the fermionic case : $f\leftrightarrow f^{\dagger}$, the
1695: result can be expressed as a function of $C^{4}_{+}-C^{4}_{-}$ and
1696: $C_{+}^{3}C_{-} - C_{-}^{3}C_{+}$. The first term is rejected since it
1697: leads to a singularity at $\theta =\pm \pi /4$. The proportionality
1698: coefficient is fixed by imposing $\theta =\pi /4$ for $q_{0}=0$.
1699:
1700:
1701: \section{The marginality criterion}\label{AppReplicon}
1702: \subsection{Diagonalization of the fluctuation matrix}
1703: First, we diagonalize the fluctuation matrix $M$ in the
1704: replica space defined by Eq. (\ref{DefMatrixFluctuations}).
1705: A priori $M$ is a $n (n-1)/2\times n (n-1)/2$ matrix. However, we have
1706: taken for $g_{ab}$ the simple one step replica symmetry breaking
1707: Ansatz on $g_{ab}$, {\it i.e.} the $n\times n $ matrix splits into
1708: $n/m\times n/m$ blocks : $g_{ab}=g$ if $\lfloor a/m \rfloor = \lfloor
1709: b/m \rfloor $, 0 otherwise.
1710: Thus $M$ splits into $n/m$ identical $m (m-1)/2\times m (m-1)/2$
1711: blocks ($M_{ab,cd}$ does not vanish if and only if all indices are in
1712: the same block
1713: $\lfloor a/m \rfloor = \lfloor b/m \rfloor =\lfloor c/m \rfloor = \lfloor
1714: d/m \rfloor $).
1715: Hence the diagonalization is to be performed only on one block (we set
1716: $1\leq a,b,c,d \leq m$), which elements are given by (with $a,b,c,d$
1717: distinct replica indices) :
1718: \begin{align}\label{AppRepValueM}
1719: M_{ab,ab}&=A \equiv 3 \beta J^{2}g^{2} \left[1-3
1720: \beta^{2}J^{2}g^{2}\left (g^{2} +\Bigl(g + \frac{\Theta }{\beta Jg}
1721: \Bigr)^{2}\right) \right] \\
1722: M_{ab,ac}&=B\equiv
1723: -9\beta ^{3}J^{4}g^{5} \left(2g +\frac{\Theta }{\beta J g} \right)\\
1724: M_{ab,cd}&=C\equiv -18\beta^{3}J^{4}g^{6}
1725: \end{align}
1726: This matrix has already been diagonalized in
1727: Ref.~\onlinecite{DeAlmeida} and its eigenvalues are given by :
1728: \begin{align}\label{AppRepValeurde_e}
1729: e_{1} &= A-2B+C \\
1730: e_{2} &= A +2 (m-2)B + (m-2) (m-3)C/2\\
1731: e_{3} &= A + (m-4) B - (m-3)C
1732: \end{align}
1733: Moreover, the degeneracies of $e_{1}$, $e_{2}$ and $e_{3}$ are
1734: $n (m-3)/2$, $n/m$ and $n (m-1)/m$ respectively.
1735: Using (\ref{AppRepValueM}) and (\ref{AppRepValeurde_e}), we find
1736: finally the result of the text (\ref{EigenvaluesFluct}).
1737:
1738: The above calculation has entirely ignored perturbations,
1739: $\delta \widetilde{G}(\tau)$
1740: in the
1741: diagonal elements of (\ref{AnsatzParisi}). Including these greatly
1742: complicates the analysis, but a simple observation will suffice
1743: for our purposes. Our main attention is on the cross-coupling
1744: between $\delta \widetilde{G}(\tau)$ and the $\delta g_{ab}$. A
1745: simple consequence of the block-diagonal structure of the $g_{ab}$
1746: in the mean-field solution is that this cross-coupling has the
1747: form
1748: \begin{equation}
1749: \delta {\cal F} \sim \sum_{a>b, \lfloor a/m \rfloor = \lfloor
1750: b/m \rfloor} \int_0^{\beta} d \tau \delta \widetilde{G}
1751: (\tau) \delta g_{ab}.
1752: \label{ssapp}
1753: \end{equation}
1754: Now we can expand $\delta g_{ab}$ in terms of the eigenvectors
1755: associated with (\ref{AppRepValeurde_e}), which were computed in
1756: Ref.~\onlinecite{DeAlmeida}. The key observation is that after
1757: the sum over $a,b$ in (\ref{ssapp}), the cross-terms
1758: corresponding to all the eigenvectors associated with $e_{1}$
1759: vanish. Consequently these eigenvectors remain eigenvectors even
1760: upon including $\delta \widetilde{G} (\tau)$, and the eigenvalue
1761: $e_1$ remains unchanged. A similar argument shows that the
1762: eigenvalue $e_3$ also remains unchanged, and only the
1763: eigenvectors associated with $e_2$ are modified non-trivially by
1764: the coupling to $\delta \widetilde{G} (\tau)$.
1765:
1766: \subsection{The replicon solution is gapless}
1767: \label{app:gapless}
1768:
1769: Let us assume that there is no gap in the boson spectral density and
1770: more precisely that for small $\omega$ :
1771: \begin{equation}\label{AppGaplessAnsatz}
1772: \widetilde{G} (\omega ) = \frac{\Theta }{Jg} + (a+ib)\omega^{\alpha}
1773: \Theta (\omega ) + (a'+ib')|\omega|^{\alpha}\Theta (-\omega ) + o
1774: (|\omega|^{\alpha})
1775: \end{equation}
1776: where $\Theta $ is the Heaviside function, $a,b,a',b'$ are real
1777: constants, and $\alpha >0$.
1778: Then from (\ref{EqSG_DefSigTilde}) we obtain for $\omega >0$ :
1779: \begin{align}
1780: \Im \left(\widetilde{\Sigma } (\omega )\right) &= J^{2}g^{2}
1781: \Im \left( 2 \widetilde{G} (\omega ) + \overline{\widetilde{G}
1782: (-\omega )} \right) + \dots \\
1783: \widetilde{\Sigma } (\omega ) &= c+ \left(d+ i (2b-b')J^{2}g^{2}
1784: \right) \omega^{\alpha}
1785: \end{align}
1786: where $c$ and $d$ are real constants.
1787: The other terms are subdominant in the limit
1788: $\omega \rightarrow 0$ as can be seen using a spectral representation.
1789: We then expand (\ref{EqSG_Dyson}) to second order and obtain
1790: at first order $\lambda =c +\frac{Jg}{\Theta }$ and for the imaginary
1791: part at second order :
1792: \begin{equation}
1793: b - (2b-b')\Theta^{2} = b' - (2b'-b)\Theta^{2}=0
1794: \end{equation}
1795: which leads to ($\Theta =1$ is excluded since $b>0$ and $b'<0$)
1796: \begin{equation}
1797: \Theta^{2}=\frac{1}{3}
1798: \end{equation}
1799: Thus the value of $\Theta $ given by the replicon
1800: condition is {\sl the only one} that leads to a gapless bosonic
1801: spectral density.
1802:
1803: \section{Free energy of quantum rotor and Ising spin glasses}
1804: \label{rotor}
1805:
1806: Quantum spin glasses of quantum rotors and Ising spins were
1807: studied extensively in Ref.~\onlinecite{rsy}. However, while the
1808: paramagnetic phase and the vicinity of the quantum-critical point
1809: were fairly completely described, the $T \rightarrow 0$
1810: thermodynamics within the spin glass phases were only studied in
1811: the replica-symmetric solution. A proper understanding of this
1812: low $T$ limit requires consideration of replica symmetry
1813: breaking, and we will provide that here. We will find, as in the
1814: more complex Heisenberg spin model considered in the body of the
1815: paper, that the specific heat is linear in $T$ at low $T$.
1816:
1817: As we are restricting our attention to mean-field theory, we can
1818: neglect the spatial dependence of all degrees of freedom.
1819: Further, we will also restrict ourselves to the Ising case, and
1820: the generalization to the multi-component rotor case is
1821: immediate. As discussed in Ref.~\onlinecite{rsy}, the effective
1822: action of the quantum Ising spin glass is expressed in terms of
1823: the order parameter functional
1824: \begin{equation}
1825: Q^{ab} (\tau, \tau^{\prime} ) = \langle \sigma^a (\tau) \sigma^b
1826: (\tau^{\prime} ) \rangle \label{rsy1}
1827: \end{equation}
1828: where $\sigma^a$ is the Ising spin in replica $a$. The important
1829: low-order terms in the free energy density are
1830: \begin{eqnarray}
1831: {\mathcal F} &=& \frac{1}{\kappa} \int d \tau \sum_a \left. \left[
1832: \frac{\partial}{\partial \tau_1} \frac{\partial}{\partial \tau_2}
1833: + r \right] Q^{aa} (\tau_1, \tau_2 ) \right|_{\tau_1 = \tau_2 =
1834: \tau} -\frac{\kappa}{3} \int d \tau_1 d \tau_2 d \tau_3 \sum_{abc}
1835: Q^{ab} (\tau_1 , \tau_2 ) Q^{bc} (\tau_2, \tau_3) Q^{ca} (\tau_3,
1836: \tau_1 ) \nonumber \\
1837: &~&~~~~~~~+ \frac{u}{2} \int d \tau \sum_a Q^{aa} (\tau,\tau)
1838: Q^{aa} (\tau,\tau) - \frac{y}{6} \int d \tau_1 d \tau_2 \sum_{ab}
1839: \left[ Q^{ab} (\tau_1 , \tau_2 ) \right]^4. \label{rsy2}
1840: \end{eqnarray}
1841: Here $r$ is the parameter which tunes across the spin glass
1842: transition, and $\kappa$, $y$ measure the strength of various
1843: non-linearities. The analysis of thermodynamic properties in the
1844: spin-glass phase in Ref.~\onlinecite{rsy} was carried out with a
1845: vanishing coefficient of the quartic term, $y=0$: in this case
1846: the order parameter has replica symmetry, and it was found that
1847: the specific heat $\sim T^3$ as $T \rightarrow 0$. Here, we will
1848: extend the solution to small $y \neq 0$, and show that the
1849: solution with broken replica symmetry has a linear specific heat.
1850:
1851: Time-translational symmetry requires that the mean-field solution
1852: take the form
1853: \begin{equation}
1854: Q^{ab} (\tau_1 , \tau_2 ) = \frac{1}{\beta} \sum_{\nu_n} Q^{ab}
1855: (i \nu_n) e^{i \nu_n (\tau_1 - \tau_2 )}. \label{rsy3}
1856: \end{equation}
1857: As in (\ref{AnsatzParisi}), we choose the following Ansatz for
1858: $Q^{ab}$:
1859: \begin{equation}
1860: Q^{ab} (i \nu_n) = \left\{ \begin{array}{ccc} D(i \nu_n) + \beta
1861: q_{EA}
1862: & \qquad & a=b \\
1863: \beta q_{ab} & \qquad & a \neq b
1864: \end{array}
1865: \right. \label{rsy4}
1866: \end{equation}
1867: where the off-diagonal terms, $q_{ab}$, are time-independent and
1868: characterized by the Parisi function $q(u)$, and $q(1) \equiv
1869: q_{EA}$. We have included an additive factor of $\beta q_{EA}$ in
1870: the diagonal term for convenience, and without loss of
1871: generality: as in the discussion below (\ref{AnsatzParisi}), we
1872: will find that this ensures that at $T=0$ the solution for
1873: $D(\tau)$ vanishes as $\tau \rightarrow \infty$. Also, the
1874: diagonal components $q_{aa}$ do not appear in the above, and we
1875: are therefore free to choose them as $q^{aa} = 0$. Here, and in
1876: the remainder of this appendix we are assuming that $r$ is
1877: sufficiently negative so that the system has a spin glass ground
1878: state; for larger $r$, the ground state is a paramagnet \cite{rsy}
1879: with $q_{EA}=q_{ab} =0$ whose properties are not addressed here.
1880:
1881: We now need to insert (\ref{rsy4}) into (\ref{rsy2}) and find the
1882: saddle-point with respect to variations in the functions $q(u)$
1883: and $D ( i \nu_n)$. This is, in principle, a straightforward
1884: exercise, but the computations are somewhat lengthy.
1885:
1886: We first identify just the terms that depend upon $q^{ab}$; these
1887: have the form
1888: \begin{equation}
1889: {\mathcal F} = - R_1 \mbox{Tr} q^2 - \frac{R_2}{3} \mbox{Tr} q^3
1890: - \frac{R_3}{6} \sum_{ab} q_{ab}^4 + \ldots, \label{rsy5}
1891: \end{equation}
1892: where
1893: \begin{eqnarray}
1894: R_1 &=& \beta \kappa (D(0) + \beta q_{EA}) \nonumber \\
1895: R_2 &=& \kappa \beta^2 \nonumber \\
1896: R_3 &=& \beta y. \label{rsy6}
1897: \end{eqnarray}
1898: We first address the problem of determining the saddle-point of
1899: (\ref{rsy5}) with respect to variations in $q(u)$. Fortunately,
1900: this problem has been completely solved in the classical spin
1901: glass literature \cite{HertzBook}, and we can directly borrow the
1902: results: the function $q(u)$ increases linearly as a function of
1903: $u$ for $0 < u < 1 - (1- 4 R_1 R_3/R_2^2)^{1/2}$, where it
1904: saturates until $u=1$ at the constant value
1905: \begin{equation}
1906: q_{EA} = \frac{R_2 - (R_2^2 - 4 R_1 R_3)^{1/2}}{2 R_3}.
1907: \label{rsy7}
1908: \end{equation}
1909: Combining (\ref{rsy7}) with (\ref{rsy5}), we obtain the simple
1910: result
1911: \begin{equation}
1912: q_{EA}^2 = - \kappa D(0)/y \label{rsy8}
1913: \end{equation}
1914:
1915: Next, we consider the variation of ${\mathcal F}$ in (\ref{rsy2})
1916: with respect to $D(i \nu_n)$. This is most easily done for $\nu_n
1917: \neq 0$, for which we obtain the following saddle-point equation
1918: \begin{eqnarray}
1919: && \frac{1}{\kappa} ( \nu_n^2 + r) - \kappa D^2 (i \nu_n) + u
1920: \left[ \frac{1}{\beta} \sum_{\nu^{\prime}_n} D( i \nu^{\prime}_n )
1921: +
1922: q_{EA} \right] - 2 y q_{EA}^2 D(-i \nu_n) \nonumber \\
1923: &&~~- \frac{2 y q_{EA}}{\beta} \sum_{\nu^{\prime}_n} D(i
1924: \nu^{\prime}_n) D(- i\nu_n - i\nu^{\prime}_n) - \frac{2 y}{3
1925: \beta^2} \sum_{\nu^{\prime}_n, \nu^{\prime\prime}_n}
1926: D(i\nu^{\prime}_n) D(i\nu^{\prime\prime}_n) D(- i\nu_n -
1927: i\nu^{\prime}_n-i\nu^{\prime\prime}_n)=0. \label{rsy10}
1928: \end{eqnarray}
1929: Upon consideration of the saddle point equation for $D(0)$ one
1930: initially finds a number of additional term associated with the
1931: coupling of $D(0)$ to the $q_{ab}$. However, our parameterization
1932: in (\ref{rsy4}) was chosen judiciously, and has the feature that
1933: all these additional terms vanish upon using (\ref{rsy8}); so,
1934: the result (\ref{rsy10}) applies {\em also} for $\nu_n = 0$.
1935:
1936: Let us also note the complete expression for the free energy
1937: density, obtained by inserting (\ref{rsy4}) and (\ref{rsy10})
1938: into (\ref{rsy2}):
1939: \begin{eqnarray}
1940: {\mathcal F}/n &=& \frac{q_{EA} r}{\kappa} + \frac{y^2 q_{EA}^5}{5
1941: \kappa} + \frac{1}{\beta \kappa} \sum_{\nu_n} ( \nu_n^2 + r) D(i
1942: \nu_n) - \frac{\kappa}{3 \beta} \sum_{nu_n} D^3 (i \nu_n) +
1943: \frac{u}{2} \left[ \frac{1}{\beta} \sum_{\nu_n} D( i\nu_n ) +
1944: q_{EA} \right]^2 \nonumber \\
1945: &&~~- \frac{y q_{EA}^2}{\beta} \sum_{\nu_n} D(i\nu_n) D(-i\nu_n)
1946: - \frac{2 y q_{EA}}{3 \beta^2} \sum_{\nu_n, \nu^{\prime}_n}
1947: D(i\nu_n) D(i\nu^{\prime}_n) D(- i\nu_n - i\nu^{\prime}_n)
1948: \nonumber
1949: \\ &&~~~~~ - \frac{ y}{6 \beta^3} \sum_{\nu_n, \nu^{\prime}_n,
1950: \nu^{\prime\prime}_n} D(i\nu_n) D(i\nu^{\prime}_n)
1951: D(i\nu^{\prime\prime}_n) D(- i\nu_n -
1952: i\nu^{\prime}_n-i\nu^{\prime\prime}_n). \label{rsy11}
1953: \end{eqnarray}
1954:
1955: We are now left with the task of solving the saddle-point
1956: equations (\ref{rsy8}) and (\ref{rsy10}) for $q_{EA}$ and
1957: $D(i\nu_n)$, and inserting the result in (\ref{rsy11}). This is
1958: clearly a daunting task, and we will be satisfied in describing
1959: the $T \rightarrow 0$ limit to first order in $y$. This is
1960: similar in spirit to the large $S$ expansion of
1961: Sections~\ref{sec:largeS}, \ref{Thermo}, and we expect that
1962: higher-order corrections in $y$ will not modify the nature of the
1963: low $T$ limit.
1964:
1965: First, we consider the case $y=0$. Here a complete analytical
1966: solution is possible, and was presented in Ref.~\onlinecite{rsy}.
1967: We have at $y=0$:
1968: \begin{eqnarray}
1969: q_{EA}^{0} &=& \frac{1}{\beta \kappa} \sum_{\nu_n} |\nu_n| -
1970: \frac{r}{\kappa u} \nonumber \\
1971: D^0 (i\nu_n) &=& - \frac{|\nu_n|}{\kappa} \nonumber \\
1972: {\mathcal F}^0 (T)/n &=& \frac{2}{3 \beta \kappa^2} \sum_{\nu_n}
1973: |\nu_n|^3 - \frac{r^2}{2\kappa^2 u}\nonumber \\
1974: &=& {\mathcal F}^0 (0)/n - \frac{4 \pi^3 T^4}{45 \kappa^2}.
1975: \label{rsy12}
1976: \end{eqnarray}
1977: We observe that the free energy density behaves as $T^4$, while
1978: the specific heat $\sim T^3$ as $T \rightarrow 0$. Notice also
1979: that positivity of $q_{EA}^{0}$ requires an upper bound on $r$,
1980: which we have assumed to hold.
1981:
1982: Before considering explicit corrections in powers of $y$, we make
1983: an observation that is valid to all orders in $y$. The solution
1984: for $D(i\nu_n)$ in (\ref{rsy12}), when analytically continued to
1985: real frequencies, $\omega$, has an imaginary part which vanishes
1986: linearly in $\omega$ at small $\omega$. We now show that this
1987: conclusion holds to all orders in $y$; the constraint
1988: (\ref{rsy8}) will play a key role in establishing this result.
1989: Let us write $D(\omega) = D(0) + i D_1 \omega + \ldots$ for small
1990: $\omega$, where $D(0)$ and $D_1$ are some real constants.
1991: Inserting this in (\ref{rsy10}) and evaluating it at $T=0$ for
1992: small $\omega$, we note that the last two terms in (\ref{rsy10})
1993: have imaginary parts which vanish as $\omega^3$ and $\omega^5$.
1994: Keeping only the leading $\omega$ dependence of the imaginary
1995: part, we obtain the simple expression
1996: \begin{equation}
1997: -2 i \kappa D(0) D_1 \omega - 2 y q_{EA}^2 i D_1 \omega = 0.
1998: \label{rsy12a}
1999: \end{equation}
2000: From (\ref{rsy8}) we see that this condition is satisfied, and so
2001: $D_1$ can be non-zero.
2002:
2003: Now, we consider explicit first-order corrections in $y$: we will
2004: see that this leads to terms in the thermodynamics which vanish
2005: more slowly as $T \rightarrow 0$. We can easily use (\ref{rsy8})
2006: and (\ref{rsy10}) to determine the corrections to $D(i\nu_n)$ and
2007: $q_{EA}$ to linear order in $y$; however, these are not needed
2008: here as the shift in the free energy due to such corrections will
2009: only appear at order $y^2$, because the free energy is at a saddle
2010: point. Indeed, to obtain the free energy correct to first order in
2011: $y$, we need only insert (\ref{rsy12}) into (\ref{rsy11}). It is
2012: then quite easy to see that the free energy will have a term of
2013: order $y T^2$, and that the coefficient of this term will be
2014: non-universal and dependent upon the nature of the high energy
2015: cutoff. The required term comes from the $T$ dependence of
2016: $q_{EA}^0$, which from (\ref{rsy12}) is seen to be
2017: \begin{equation}
2018: q_{EA}^0 (T) = q_{EA}^0 (0) - \frac{\pi T^2}{3
2019: \kappa}.\label{rsy13}
2020: \end{equation}
2021: Upon inserting (\ref{rsy13}) and (\ref{rsy12}) into (\ref{rsy11})
2022: we will now obtain numerous terms in which the above $T^2$ term
2023: multiplies $T$-independent, cutoff-dependent terms coming from
2024: the upper bounds in the summations over the $D (i\nu_n)$. As the
2025: relative values of these contributions will depend upon the
2026: nature of the cutoff, there is no general reason for them to
2027: cancel against each other. Hence we obtain a $T^2$ contribution
2028: to ${\cal F}$ and a linear $T$ term in the low $T$ specific heat.
2029:
2030: \endgroup
2031:
2032: %\nocite{*}
2033: %\bibliographystyle{revtex}
2034: %\bibliography{SG}
2035:
2036: \begin{thebibliography}{10}
2037: \providecommand*{\bibinfo}[2]{#2}
2038: \providecommand*{\eprint}[1]{#1}
2039: \providecommand*{\url}[1]{#1}
2040: \bibitem{itp}
2041: \bibinfo{journal}{J. Phys.: Condens. Matt.} \bibinfo{volume}{\textbf{8}},
2042: \bibinfo{pages}{9675} (\bibinfo{date}{1996}), conference on Non-Fermi Liquid
2043: Behavior in Metals.
2044: \bibitem{coleman}
2045: \bibinfo{author}{P.~Coleman}, \bibinfo{journal}{Physica B}
2046: \bibinfo{volume}{\textbf{259-261}}, \bibinfo{pages}{353}
2047: (\bibinfo{date}{1999}).
2048: \bibitem{Doniach}
2049: \bibinfo{author}{S.~Doniach}, \bibinfo{journal}{Physica B}
2050: \bibinfo{volume}{\textbf{91}}, \bibinfo{pages}{213} (\bibinfo{date}{1977}).
2051: \bibitem{Hertz}
2052: \bibinfo{author}{J.~Hertz}, \bibinfo{journal}{Phys. Rev. B}
2053: \bibinfo{volume}{\textbf{14}}, \bibinfo{pages}{1165} (\bibinfo{date}{1976}).
2054: \bibitem{Millis}
2055: \bibinfo{author}{A.~Millis}, \bibinfo{journal}{Phys. Rev. B}
2056: \bibinfo{volume}{\textbf{48}}, \bibinfo{pages}{7183} (\bibinfo{date}{1993}).
2057: \bibitem{SRO}
2058: \bibinfo{author}{S.~Sachdev}, \bibinfo{author}{N.~Read}, and
2059: \bibinfo{author}{R.~Oppermann}, \bibinfo{journal}{Phys. Rev. B}
2060: \bibinfo{volume}{\textbf{52}}, \bibinfo{pages}{10286} (\bibinfo{date}{1995}).
2061: \bibitem{AG}
2062: \bibinfo{author}{A.~Sengupta} and \bibinfo{author}{A.~Georges},
2063: \bibinfo{journal}{Phys. Rev. B} \bibinfo{volume}{\textbf{52}},
2064: \bibinfo{pages}{10295} (\bibinfo{date}{1995}).
2065: \bibitem{SachdevBook}
2066: \bibinfo{author}{S.~Sachdev}, \bibinfo{title}{\emph{Quantum phase transition}}
2067: (\bibinfo{publisher}{Cambridge University Press}, \bibinfo{year}{1999}).
2068: \bibitem{SCS}
2069: \bibinfo{author}{S.~Sachdev}, \bibinfo{author}{A.~Chubukov}, and
2070: \bibinfo{author}{A.~Sokol}, \bibinfo{journal}{Phys. Rev. B}
2071: \bibinfo{volume}{\textbf{51}}, \bibinfo{pages}{14874} (\bibinfo{date}{1995}).
2072: \bibitem{GS}
2073: \bibinfo{author}{S.~Sachdev} and \bibinfo{author}{A.~Georges},
2074: \bibinfo{journal}{Phys. Rev. B} \bibinfo{volume}{\textbf{52}},
2075: \bibinfo{pages}{9520} (\bibinfo{date}{1995}).
2076: \bibitem{AMS}
2077: \bibinfo{author}{A.~Sengupta}, \bibinfo{journal}{Phys. Rev. B}
2078: \bibinfo{volume}{\textbf{61}}, \bibinfo{pages}{4041} (\bibinfo{date}{2000}).
2079: \bibitem{Si}
2080: \bibinfo{author}{J.~Smith} and \bibinfo{author}{Q.~Si}, \bibinfo{journal}{Phys.
2081: Rev. B} \bibinfo{volume}{\textbf{61}}, \bibinfo{pages}{5184}
2082: (\bibinfo{date}{2000}).
2083: \bibitem{rsy}
2084: \bibinfo{author}{N.~Read}, \bibinfo{author}{S.~Sachdev}, and
2085: \bibinfo{author}{J.~Ye}, \bibinfo{journal}{Phys. Rev. B}
2086: \bibinfo{volume}{\textbf{52}}, \bibinfo{pages}{6411} (\bibinfo{date}{1995}).
2087: \bibitem{RevueBhatt}
2088: \bibinfo{author}{R.~Bhatt} (\bibinfo{date}{1997}), in "Spin Glasses and Random
2089: Fields", A.P. Young eds. World Scientific.
2090: \bibitem{millerhuse}
2091: \bibinfo{author}{J.~Miller} and \bibinfo{author}{D.A.Huse},
2092: \bibinfo{journal}{Phys. Rev. Lett.} \bibinfo{volume}{\textbf{70}},
2093: \bibinfo{pages}{3147} (\bibinfo{date}{1993}).
2094: \bibitem{dsf}
2095: \bibinfo{author}{D.~Fisher}, \bibinfo{journal}{Phys. Rev. B}
2096: \bibinfo{volume}{\textbf{51}}, \bibinfo{pages}{6411} (\bibinfo{date}{1995}).
2097: \bibitem{Pich}
2098: \bibinfo{author}{C.~Pich} and \bibinfo{author}{et~al.}, \bibinfo{journal}{Phys.
2099: Rev. Lett.} \bibinfo{volume}{\textbf{81}}, \bibinfo{pages}{5916}
2100: (\bibinfo{date}{1998}).
2101: \bibitem{Motrunich}
2102: \bibinfo{author}{O.~Motrunich}, \bibinfo{author}{S.~Mau},
2103: \bibinfo{author}{D.~Huse}, and \bibinfo{author}{D.~Fisher},
2104: \bibinfo{journal}{Phys. Rev. B} \bibinfo{volume}{\textbf{61}},
2105: \bibinfo{pages}{1160} (\bibinfo{date}{2000}), \eprint{cond-mat/9906322}.
2106: \bibitem{SachdevYe}
2107: \bibinfo{author}{S.~Sachdev} and \bibinfo{author}{J.~Ye},
2108: \bibinfo{journal}{Phys. Rev. Lett} \bibinfo{volume}{\textbf{70}},
2109: \bibinfo{pages}{3339} (\bibinfo{date}{1993}).
2110: \bibitem{NotreLettre}
2111: \bibinfo{author}{A.~Georges}, \bibinfo{author}{O.~Parcollet}, and
2112: \bibinfo{author}{S.~Sachdev}, \bibinfo{journal}{Phys. Rev. Lett.}
2113: \bibinfo{volume}{\textbf{85}}, \bibinfo{pages}{840} (\bibinfo{date}{2000}).
2114: \bibitem{BrayMoore}
2115: \bibinfo{author}{A.~Bray} and \bibinfo{author}{M.~Moore}, \bibinfo{journal}{J.
2116: Phys. C : Solid. St. Phys} \bibinfo{volume}{\textbf{13}},
2117: \bibinfo{pages}{L655} (\bibinfo{date}{1980}).
2118: \bibitem{MezardParisiBook}
2119: \bibinfo{author}{M.~M{\'e}zard}, \bibinfo{author}{G.~Parisi}, and
2120: \bibinfo{author}{M.~Virasoro}, \bibinfo{title}{\emph{Spin Glass Theory and
2121: beyond}} (\bibinfo{publisher}{World Scientific}, \bibinfo{year}{1987}).
2122: \bibitem{kopec}
2123: \bibinfo{author}{T.~{Kope\' c}}, \bibinfo{journal}{Phys. Rev. B}
2124: \bibinfo{volume}{\textbf{52}}, \bibinfo{pages}{9590} (\bibinfo{date}{1995}).
2125: \bibitem{Grempel1}
2126: \bibinfo{author}{D.~Grempel} and \bibinfo{author}{M.~Rozenberg},
2127: \bibinfo{journal}{Phys. Rev. Lett.} \bibinfo{volume}{\textbf{80}},
2128: \bibinfo{pages}{389} (\bibinfo{date}{1998}).
2129: \bibitem{Slush}
2130: \bibinfo{author}{O.~Parcollet} and \bibinfo{author}{A.~Georges},
2131: \bibinfo{journal}{Phys. Rev. B} \bibinfo{volume}{\textbf{59}},
2132: \bibinfo{pages}{5341} (\bibinfo{date}{1999}), in this paper only the
2133: fermionic model is considered but the results apply also to the paramagnet in
2134: the bosonic model (with minor changes).
2135: \bibitem{KondoLong}
2136: \bibinfo{author}{O.~Parcollet}, \bibinfo{author}{A.~Georges},
2137: \bibinfo{author}{G.~Kotliar}, and \bibinfo{author}{A.~Sengupta},
2138: \bibinfo{journal}{Phys. Rev. B} \bibinfo{volume}{\textbf{58}},
2139: \bibinfo{pages}{3794} (\bibinfo{date}{1998}).
2140: \bibitem{KondoPRL}
2141: \bibinfo{author}{O.~Parcollet} and \bibinfo{author}{A.~Georges},
2142: \bibinfo{journal}{Phys. Rev. Lett.} \bibinfo{volume}{\textbf{79}},
2143: \bibinfo{pages}{4665} (\bibinfo{date}{1997}).
2144: \bibitem{SBV}
2145: \bibinfo{author}{S.~Sachdev}, \bibinfo{author}{C.~Buragohain}, and
2146: \bibinfo{author}{M.~Vojta}, \bibinfo{journal}{Science}
2147: \bibinfo{volume}{\textbf{286}}, \bibinfo{pages}{2479} (\bibinfo{date}{1999}).
2148: \bibitem{VBS}
2149: \bibinfo{author}{M.~Vojta}, \bibinfo{author}{C.~Buragohain}, and
2150: \bibinfo{author}{S.~Sachdev}, \bibinfo{journal}{Phys. Rev. B}
2151: \bibinfo{volume}{\textbf{61}}, \bibinfo{pages}{15152} (\bibinfo{date}{2000}).
2152: \bibitem{forster}
2153: \bibinfo{author}{D.~Forster}, \bibinfo{author}{D.~Nelson}, and
2154: \bibinfo{author}{M.~Stephen}, \bibinfo{journal}{Phys. Rev. A}
2155: \bibinfo{volume}{\textbf{16}}, \bibinfo{pages}{732} (\bibinfo{date}{1977}).
2156: \bibitem{honko}
2157: \bibinfo{author}{J.~Honkonen} and \bibinfo{author}{E.~Karjalainen},
2158: \bibinfo{journal}{J. Phys. A} \bibinfo{volume}{\textbf{21}},
2159: \bibinfo{pages}{4217} (\bibinfo{date}{1988}), for a review see also : J. P
2160: Bouchaud and A. Georges {\sl Phys. Reports} {\bf 195} 127 (1990).
2161: \bibitem{MezardParisiManifold}
2162: \bibinfo{author}{M.~M{\'e}zard} and \bibinfo{author}{G.~Parisi},
2163: \bibinfo{journal}{J. Phys. I} \bibinfo{volume}{\textbf{1}},
2164: \bibinfo{pages}{809} (\bibinfo{date}{1991}).
2165: \bibitem{GiamLedou}
2166: \bibinfo{author}{T.~Giamarchi} and \bibinfo{author}{P.~{Le Doussal}},
2167: \bibinfo{journal}{Physical Review B} \bibinfo{volume}{\textbf{53}},
2168: \bibinfo{pages}{15206} (\bibinfo{date}{1996}).
2169: \bibitem{CugliandoloPspinQ}
2170: \bibinfo{author}{L.~Cugliandolo}, \bibinfo{author}{C.~{Da Silva Santos}}, and
2171: \bibinfo{author}{D.~Grempel}, \bibinfo{journal}{Phys. Rev. Lett.}
2172: \bibinfo{volume}{\textbf{85}}, \bibinfo{pages}{2589} (\bibinfo{date}{2000}).
2173: \bibitem{LeticiaPspin}
2174: \bibinfo{author}{L.~Cugliandolo} and \bibinfo{author}{G.~Lozano},
2175: \bibinfo{journal}{Phys. Rev. B} \bibinfo{volume}{\textbf{59}},
2176: \bibinfo{pages}{915} (\bibinfo{date}{1999}).
2177: \bibitem{DeAlmeida}
2178: \bibinfo{author}{A.~{de~Almeida}} and \bibinfo{author}{D.~Thouless},
2179: \bibinfo{journal}{J. Phys. A} \bibinfo{volume}{\textbf{11}},
2180: \bibinfo{pages}{983} (\bibinfo{date}{1978}), reprinted in
2181: \cite{MezardParisiBook}.
2182: \bibitem{HertzBook}
2183: \bibinfo{author}{K.~Fischer} and \bibinfo{author}{J.~Hertz},
2184: \bibinfo{title}{\emph{Spin Glasses}} (\bibinfo{publisher}{Cambridge
2185: University Press, Cambridge}, \bibinfo{year}{1993}).
2186: \bibitem{Brooke}
2187: \bibinfo{author}{J.~Brooke} and \bibinfo{author}{al.},
2188: \bibinfo{journal}{Science} \bibinfo{volume}{\textbf{284}},
2189: \bibinfo{pages}{779} (\bibinfo{date}{1999}).
2190: \bibitem{CugliandoloKurchan}
2191: \bibinfo{author}{L.~F. Cugliandolo} and \bibinfo{author}{J.~Kurchan},
2192: \bibinfo{journal}{Phys. Rev. Lett.} \bibinfo{volume}{\textbf{71}},
2193: \bibinfo{pages}{173} (\bibinfo{date}{1993}).
2194:
2195: \end{thebibliography}
2196:
2197: \end{document}
2198: