1:
2: \documentstyle[12pt,graphicx,aps,epsf]{revtex}
3:
4: \newcommand {\be} {\begin{equation}}
5: \newcommand {\bea} {\begin{eqnarray} \nonumber }
6: \newcommand {\ee} {\end{equation}}
7: \newcommand {\eea} {\end{eqnarray}}
8:
9: \begin{document}
10: \parskip=0cm
11: \noindent
12:
13: \title {On the Effects of Changing the Boundary Conditions on the
14: Ground State of Ising Spin Glasses}
15: \author{
16: Enzo Marinari and Giorgio Parisi\\
17: \small Dipartimento di Fisica, INFM and INFN, Universit\`a di Roma
18: {\em La Sapienza},\\
19: \small P. A. Moro 2, 00185 Roma (Italy)\\
20: \small e-mail:
21: {\tt Enzo.Marinari@roma1.infn.it,
22: Giorgio.Parisi@roma1.infn.it}}
23:
24: \date{April 2000}
25:
26: \maketitle
27:
28: \begin{abstract}
29: We compute and analyze couples of ground states of $3D$ spin glass
30: systems with the same quenched noise but periodic and anti-periodic
31: boundary conditions for different lattice sizes. We discuss the
32: possible different behaviors of the system, we analyze the average
33: link overlap, the probability distribution of window overlaps (among
34: ground states computed with different boundary conditions) and the
35: spatial overlap and link overlap correlation functions. We establish
36: that the picture based on Replica Symmetry Breaking correctly
37: describes the behavior of $3D$ Spin Glasses.
38: \end{abstract}
39:
40: %\smallskip
41:
42: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
43: \section{Introduction}
44:
45: Understanding the structure of the states of three dimensional ($3D$)
46: Ising spin glasses at finite temperature is a very interesting
47: problem. Long time ago a mean field theory for spin glasses has been
48: constructed and successfully compared with the numerical properties of
49: a long range model, the Sherrington-Kirkpatrick model, where mean
50: field theory is exact by definition \cite{RSB}.
51:
52: In the mean field approach one finds that the free energy landscape is
53: strongly corrugated, and it is characterized by the presence of many
54: local minima, which correspond to spin configurations very different
55: one from the other. There are many low free energy local minima which
56: have a total free energy which differs from the ground state free
57: energy of an amount which is of order $1$; these minima contribute to
58: the probability distribution of the spins also for arbitrary large
59: volumes. In some sense one can say that in the infinite volume there
60: are many equilibrium states (for a better qualification of the
61: terminology see \cite{MAPARIRUZU}): these states are locally different
62: one from the other. For historical reasons this picture goes under
63: the name of {\em Replica Symmetry Breaking} (RSB).
64:
65: Since analytic progress on finite dimensional system is hard, the
66: study of $3D$ systems has mainly been based on numerical simulations.
67: Most of the numerical simulations have investigated systems of size up
68: to $V\equiv L^3=16^{3}$ and temperature values greater or equal than
69: half of the critical temperature ($T_{c}$): the results are in very
70: good agreement with the RSB picture. Using the best techniques that
71: are available to us today it is impossible to thermalize in a
72: reasonable amount of time systems of this size at much lower
73: temperature: because of that at today the very low temperature region
74: cannot be explored using Monte Carlo simulations. On the other end a
75: different class of algorithms exists which allow the determination of
76: the ground state at zero temperature for systems of comparable size
77: \cite{RIEGER,MARTIN,YOUNG,MIDDLETON}: it is clear that it would be
78: very interesting to compare the results of the RSB approach with the
79: numerical results at zero temperature.
80:
81: We face however a difficulty: in the RSB approach one computes
82: quantities at finite temperature in the infinite volume limit, while
83: if we look first for the ground state in a finite volume and only
84: later we send the volume to infinity we are exchanging the order of
85: the two limits. One can assume in a tentative way that the two limits
86: can be exchanged, in the sense that in the zero temperature limit the
87: probability distributions of the free energies of different
88: equilibrium states becomes the probability distributions of the local
89: minima of the energy: this has been the point of view of the authors
90: that have employed this technique in the last months, it is not
91: incompatible with any of the findings obtained to date, and we will do
92: the same in the following.
93:
94: Of course at zero temperature the ground state (i.e. the global
95: minimum of the energy) of a finite volume system with quenched random
96: couplings selected under a continuous distribution is unique, so that
97: interesting information can be extracted for example if we consider
98: the effects of an external perturbation.
99:
100: Different choices are possible. In a recent, very interesting paper,
101: Palassini and Young \cite{PALYOU} have shown that it is possible to
102: get useful information about the nature of the low $T$ phase of $3D$
103: Ising spin glasses with Gaussian distributed couplings by studying the
104: behavior of the ground state after changing the boundary conditions of
105: an $L^{3}$ lattice system from periodic ($PBC$) to anti-periodic
106: ($ABC$). They analyze ground states ($GS$) obtained with the same
107: realization of the quenched disorder and different boundary conditions
108: (or with disorder realizations that differ only in some locations) for
109: $L\le 10 $. In this paper we perform a more detailed analysis of the
110: same systems by considering a larger set of observables. Here we use
111: a modified genetic algorithm which is more efficient than the original
112: algorithm and that will be described in details elsewhere \cite{MAPA}.
113: This modified algorithm has allowed us to study larger systems ($L\le
114: 14$) than in former work. We arrive at the conclusion that the whole
115: set of data strongly suggests that the picture based on Replica
116: Symmetry Breaking is correct in describing the behavior of $3D$ Spin
117: Glasses.
118:
119: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
120: \section{Some Theoretical Considerations}
121:
122: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
123: \subsection{Four Possibilities\protect\label{FOURP}}
124:
125: As we have already discussed in the introduction we consider a $3D$
126: Ising spin glass with quenched couplings assigned under a Gaussian
127: probability distribution with zero expectation value, and we study the
128: behavior of the ground state after changing the boundary conditions
129: ($BC$) of a simple cubic lattice system of size $V=L^3$ from periodic
130: ($P$) to anti-periodic ($AP$). When we do such a change the ground
131: state is usually modified and in the new ground state some spins are
132: reversed. Let us call $\sigma(i)$ and $\tau(i)$ the spins in the
133: ground states with periodic and anti-periodic boundary conditions
134: respectively. The local overlap on site $i$ is defined as
135:
136: \begin{equation}
137: q(i)\equiv\sigma(i)\ \tau(i)\ .
138: \end{equation}
139: The link overlap is defined on the links and it is given
140: by
141:
142: \begin{equation}
143: q_{l}(i,\mu) \equiv q(i)\ q(i+\mu)\ ,
144: \end{equation}
145: where $\mu$ is the direction of the link, that can take $D$ positive
146: and $D$ negative values. If two spin configurations differ by a
147: global reverse of the spins their link overlap is identically equal to
148: one (while their overlap is equal to $-1$). We will consider in the
149: following the case where we flip the boundary conditions in the
150: direction $x$, while we leave them unchanged in directions $y$ and
151: $z$.
152:
153: In a first approximation (neglecting the points at the boundaries, see
154: next subsection) we can define the interface among the flipped and the
155: non flipped region as the sets of link where $q_{l}(i,\mu)=-1$. We
156: are interested in finding out the geometrical properties of this
157: interface in the infinite volume limit. Of course the probability of
158: finding the interface on a random link is given by
159:
160: \begin{equation}
161: \rho \equiv \frac12 (1-q_{l})\ ,
162: \end{equation}
163: where $q_{l}$ is the disorder expectation value of $q_{l}(i,\mu)$,
164: averaged over sites $i$ and directions $\mu$.
165:
166: Here we discuss a few possible situations:
167:
168: \begin{enumerate}
169:
170: \item
171: The interface is confined in a region of width $L^{z}$ (with $z<1$):
172: inside this region the interface may have overhangs. The interface
173: density goes to zero as $L^{-\alpha}$ with $\alpha\ge 1-z$. This is
174: what happens in ferromagnetic models, both with ordered and disordered
175: Hamiltonians, where $\alpha=1$, al least in the ordered case. We will
176: see immediately that this possibility is completely excluded by the
177: data.
178:
179: \item
180: The wandering exponent $z$ is equal to one and the interface density
181: goes to zero as $L^{-\alpha}$. We also assume than in the large volume
182: limit the interface is a fractal object with fractal dimension
183: $D_{s}=D-\alpha$ (where the space dimension $D$ is equal to 3). More
184: precisely we assume that if we define a continuous coordinate (in the
185: interval $[0-1]$) as $\frac{i}{L}$, in the infinite volume limit the
186: interface becomes a fractal object defined on the continuum
187: characterized by a fractal dimension $D_{s}$. In other words we are
188: assuming that the interface is not a multi-fractal (i.e. that it can
189: be characterized by a single fractal dimensions as usually happens for
190: fractal objects one defines on a lattice in statistical mechanics),
191: and so the relation $D_{s}=D-\alpha$ is consistent.
192:
193: \item
194: The exponent $\alpha$ is equal to zero and the density $\rho$ goes to
195: a non zero value (i.e. $q_{l}$ does not go to $1$ in the large volume
196: limit). Here the interface is space filling and in the above
197: described continuum limit, the interface is a dense set of measure
198: $1$. We expect that the probability that the interface does not
199: intersect a region $\cal R$, whose size is proportional to the system
200: size, goes to zero when the volume goes to infinity. On the contrary
201: in the previous case such a probability would be a non trivial
202: function of $\cal R$. As we shall see later, this possibility is the
203: one realized in the case of RSB.
204:
205: \item
206: There is a last possibility, which is unusual, and we mention for
207: completeness (although its description will take a disproportionate
208: amount of space): this is when the wandering exponent $z$ is equal to
209: one and $\alpha$ is greater than zero, but in continuum limit the
210: interface does not become a fractal object of dimension
211: $D_{s}=D-\alpha$, but it becomes a dense set. This happens for
212: example if we consider a set of parallel planes at distance
213: proportional to $L^{\alpha}$ or a set of isolated points at distance
214: $L^{\frac{\alpha}{D}}$. In both cases the density goes to zero as
215: $L^{-\alpha}$, but the resulting object is {\em not} a fractal in the
216: infinite volume limit. A similar case happens in three dimensions if
217: we take a random walk of length ${\cal L}=L^{3-\alpha}$ with $\alpha
218: <1$. The system is a fractal of dimension $D_{s}$ up to a distance
219: $\xi(L)$, while it looks homogeneous at larger distances.
220:
221: In this case the density-density correlation function will scale as
222:
223: \begin{equation}
224: C(x,L) \equiv \left. \overline{ \rho(x) \rho(0) }\right|_L
225: \propto L^{-\alpha} x^{-C_{s}}
226: \label{CON}
227: \end{equation}
228: in the region $1 \ll x \ll \xi(L)$, where $C_{s}=D-D_{s}$ is the
229: co-dimension of the fractal (the underlying random walk in our case).
230: We will denote by the upper bar the average over the quenched
231: disorder. On the contrary in the region $\xi(L) \ll x$ the
232: correlation function will scale as
233:
234: \begin{equation}
235: C(x,L) \propto L^{-2 \alpha}\ .
236: \label{DIS}
237: \end{equation}
238: The condition that at large distances the term in equation (\ref{DIS})
239: dominates over the one in equation (\ref{CON}) implies that
240: $C_{s}>\alpha$. Moreover the two formulas must be consistent in the
241: crossover region where $x\approx\xi(L)$. Imposing this condition one
242: finds that $ L^{-\alpha}
243: \xi(L)^{-C_{s}} \propto L^{-2 \alpha}$ and therefore
244: we obtain that
245:
246: \begin{equation}
247: \xi(L) \propto
248: L^{\omega}\ , \ \ \ \omega = {\alpha \over C_{s}} <1 \ .
249: \end{equation}
250: This case is anomalous in the sense that usual scaling is not valid
251: and there is a crossover length which increases as a fractional power
252: of the side of the system. The usual scaling law (valid in the region
253: $x=O(L)$)
254:
255: \begin{equation}
256: C(x,L) =L^{-2 \alpha} f(x L^{-1})
257: \end{equation}
258: is not satisfied and it is replaced by the unusual relation
259:
260: \begin{equation}
261: C(x,L) =L^{- \alpha -C_{s}} f(x
262: L^{-1}) + A L^{-2 \alpha}g(x L^{-1})\ .
263: \end{equation}
264: Only in the case $\alpha=0$ we are in a familiar situation: in this
265: case the crossover length $\xi$ does not diverge for large $L$ and we
266: recover the third case of our list.
267:
268: \end{enumerate}
269:
270: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
271: \subsection{The Effect of Changing Boundaries}
272:
273: In the replica approach one finds that at zero temperature there are
274: many different local minima of the Hamiltonian such that even in the
275: limit where $L$ goes to infinity the difference among the energy of
276: the ground state and the energy of such minima is of order $1$. It is
277: crucial that the among these minima there are some whose site overlap
278: $q$ and link overlap $q_l$ with the ground state remains different
279: from $1$ in large volume limit. It is also important to assume that
280: for large volumes there exists a function $f$ such that
281: $q_{l}=f(q^{2})$, in other words that for fixed overlap $q$ the link
282: overlap $q_{l}$ does not fluctuate when the volume goes to infinity.
283:
284: When in a $3D$ spin glass the boundary conditions in the $x$ direction
285: are flipped from positive to negative it is known that the total
286: energy changes by an amount $\Delta E$, which increases as a power of
287: $L$. In a ferromagnet, where the ground state is unique, the ground
288: state obtained under anti-periodic condition would be locally similar
289: to the ground state obtained under periodic boundary conditions. In
290: other words in any region of side $M$, if the region does not
291: intersect the interface, which is just a flat surface, the ground
292: state spin configuration obtained under anti-periodic boundary
293: conditions will be equal (or equal to the reverse) to the ground state
294: obtained under periodic boundary conditions.
295:
296: A similar conclusion holds in the case of a spin glass behaving
297: according to the predictions of the droplet model \cite{DROPLET},
298: where the boundary may be a corrugated surface: in this case we expect
299: that the link overlap among the ground states obtained with $PBC$ and
300: $ABC$ tends to one in the infinite volume limit. On the contrary in
301: the case of a RSB like behavior the ground state obtained under $ABC$
302: can be locally similar to one of the low energy excited states: in
303: this case the link overlap among the two ground states can tend to a
304: value {\em different} from one as $L\to\infty$.
305:
306: That an excited state may be selected as new ground state when
307: changing the boundary conditions is strongly suggested by the fact
308: that the energy difference among ground states obtained with periodic
309: and anti-periodic boundary conditions increases with $L$, while the
310: difference among the ground state energy and the excited state energy
311: at fixed boundary conditions remains fixed.
312:
313: The nature of the excited state that is selected when we change the
314: boundary conditions cannot easily derived using simple arguments. It
315: is reasonable to assume that in three dimensions, where $\Delta E$
316: which increases as a power of $L$, the new ground state will be as
317: different as possible from the old ground state: one expects the
318: overlap $q$ among the two ground states to vanish in this limit.
319:
320: The interest in changing the boundary conditions is also due to the
321: fact that this is also a convenient method to to study the properties
322: of the excited states. Other methods can be used to clarify similar
323: questions, but we will not use them here.
324:
325: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
326: \subsection{Gauge Invariance}
327:
328: In order to change the boundary conditions in the $x$ direction we can
329: change the sign of all the couplings connecting the plane $x=x_{0}$
330: with the plane $x=x_{0}+1$. The choice of $x_{0}$ does not matter:
331: the ground state energy does not depend on $x_{0}$, and the ground
332: state obtained after a different choice of the reversed plane
333: (e.g. $x=x_{1}$) can be obtained starting from the ground state where
334: anti-periodic boundary conditions have been enforced at $x_{0}$, by
335: flipping all the spins in the interval $ x_{0} < x \le x_{1}$.
336:
337: In other words going from periodic to anti-periodic boundary
338: conditions is a global change that locally is not visible. At this
339: end it is convenient to define quantities which are gauge invariant,
340: in the sense that they do not change when the plane where the
341: anti-periodic boundary conditions are imposed.
342:
343: Let us consider a very simple example: the correct definition of
344: $q_l(i,+\hat{x})$ is
345:
346: \begin{eqnarray} \nonumber
347: q_l(i,+\hat{x})&\equiv&q(i)\ q(i+\hat{x}) \ \ \ \
348: \mbox{for} \ \ \ x\ne x_{0},
349: \\ q_l(i,+\hat{x})&\equiv&-q(i)\ q(i+\hat{x}) \ \ \ \ \mbox{for} \
350: \ \ x=x_{0}\ ,
351: \end{eqnarray}
352: where $x$ is the first component of the three dimensional vector $i$.
353:
354: A more complex case is the definition of the window overlap
355: \cite{WINDOW} in a box of side $M$ which intersects the plane
356: $x=x_{0}$. Here the overlap $q_{M}$ can be defined as
357:
358: \begin{equation}
359: \sum_{i \in {\cal A}}q(i) -
360: \sum_{i \in {\cal B}}q(i)
361: \end{equation}
362: where $\cal A$ and $\cal B$ are the set of points of the box which are
363: respectively on the right and on the left of the plane $x=x_{0}$.
364:
365: The advantage of this prescription (which can be easily generalized to
366: more complex situations) is that in absence of a magnetic field
367: nothing depends on the actual position of the plane where the boundary
368: conditions are enforced: the fact that the reversal of the spins is
369: imposed at a given value of $x$ is immaterial. The presence of this
370: translational invariance is very effective from a practical point of
371: view: the expectation value of $q_{l}$ does not depend on the point,
372: so that data can be taken on the whole lattice. The possibility of
373: doing measurements on the whole lattice and not only far from the
374: point where the boundary conditions are changed reduces the
375: statistical errors in a substantial way.
376:
377: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
378: \section{A First Look to the Numerical Results}
379:
380: We have computed $16000$ couples of ground states on three dimensional
381: lattices of linear size $L=4$, $73740$ for $L=6$, $8532$ for $L=8$,
382: $2434$ for $L=10$, $910$ for $L=12$ and $670$ for $L=14$: for each
383: sample of the quenched disorder (Gaussian couplings) we have computed the two ground states
384: under periodic and anti-periodic boundary conditions, using a modified
385: genetic algorithm which will be described elsewhere \cite{MAPA}.
386:
387: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
388: \subsection{The Average Link Overlap}
389:
390: The introduction of anti-periodic boundary conditions breaks the
391: discrete rotational invariance of the simple cubic lattice, so that
392: the expectation value of $q_{l}(i,\hat{x})$, i.e. of the link overlap
393: in the $x$ direction, can be different from that of the link overlap
394: in the other two directions. In the ferromagnetic case (without
395: disorder) the change of the boundary conditions produces an interface
396: in one of the $y-z$ planes. We will call {\em perpendicular link
397: overlap} $q_{P}$ the expectation value of $q_{l}(i,\hat{x})$ averaged
398: over the lattice sites, since it is perpendicular to the $y-z$
399: interface, while we will call {\em transverse link overlap} $q_{T}$
400: the average over sites of the quantity $\frac12 \left(
401: q_{l}(i,\hat{y}) + q_{l}(i,\hat{z}) \right)$.
402:
403: In the ferromagnetic case (without disorder) one finds that
404:
405: \begin{equation}
406: 1-q_{P}=\frac{1}{L}, \ \ \ \ 1-q_{T}=0\ .
407: \end{equation}
408: We show in figure (\ref{qlink}) the numerical data for $1-q_{T}$ and
409: $1-q_{P}$ in our $3D$ spin glass. It is clear that the difference
410: $q_{T}-q_{P}$ goes to zero with $\frac{1}{L}$: it can be very well
411: fitted as $a L^{-b}$ with $b\simeq 2.5$ and a normalized $\chi^2$ of
412: order one.
413:
414: It is also evident that $1-q_{P}$ is with a very good approximation a
415: linear function of $L^{-1}$ in the same way it would be in the
416: ferromagnetic case, with the difference that here the extrapolation in
417: the $L\to\infty$ limit is clearly different from zero.
418:
419: A closer look shows that $1-q_{P}$ is not exactly a linear function of
420: $L^{-1}$, and it can be very well fitted (with a normalized $\chi^{2}$
421: of order one) as a second order polynomial in $L^{-1}$. The same
422: polynomial fit works very well also for $1-q_{T}$. The two fits
423: extrapolate to the same value (in the limit given by the statistical
424: error). In the same picture we show also the data for
425: $1-q_{P}-\frac1L$, i.e. the value we measure for $1-q_{P}$ in our spin
426: glass minus its {\em a priori} lower bound, which coincides with the
427: value in the ferromagnetic case, and we note it has a very weak
428: dependence over $L$.
429:
430: \begin{figure}
431: \centering
432: \includegraphics[width=0.5\textwidth,angle=270]{fig1.ps}
433: \caption[a]{$1-q_{T}$, $1-q_{P}$ and $1-q_{P}-\frac{1}{L}$ versus
434: $\frac{1}{L}$ for $L$ $=$ $4$, $6$, $8$, $10$, $12$, $14$. We plot
435: three polynomial best fits (of second degree in $\frac{1}{L}$) to
436: these three quantities and a simple power best fit to $1-q_{T}$.
437: \protect\label{qlink} }
438: \end{figure}
439:
440: A power law fit to a zero asymptotic value (i.e. to the form $A
441: L^{-\alpha}$) of $1-q_{P}$ does not give acceptable results, while the
442: data for $1-q_{T}$ can be fitted by a pure power law, with $\alpha
443: \simeq 0.27$. A power law fits also works (even if with a higher value
444: of $\chi^2$) for $1-q_{l}(L)\equiv 1- \frac23 q_{T} -\frac13 q_{P}$
445: with a similar value of $\alpha$.
446:
447: It clear that it is not acceptable to fit the values of the two link
448: overlaps with two different functional forms. Our conclusion is that
449: the data do not support the droplet model prediction that $q_{l}$ is
450: zero in the large $L$ asymptotic limit, and they hint for
451:
452: \begin{equation}
453: \lim_{L\to\infty}\left(1 - q_{l}\right)
454: = 0.245 \pm 0.015\ .
455: \label{E-TWOFOURFIVE}
456: \end{equation}
457: This estimate of the errors relays on the polynomial fit in powers of
458: $\frac{1}{L}$. Considering a third order polynomial for $q_{T}$ does
459: not change the situation. If we fit $1-q_{T}$ with a four parameter
460: function of the form $A + B L^{-\alpha} + C L ^{-2\alpha}$ we find an
461: extrapolated value of $A = q_{l} = 0.20 \pm 0.06$, which is consistent
462: with the previous estimate: it is clear that in providing a complete
463: estimate of the error over the extrapolated value one has to include
464: the systematic uncertainty due to the possibility of using different
465: functional forms for the extrapolating function. On the other end
466: corrections proportional to a power of $\frac{1}{L}$ arise naturally,
467: as it will be seen in the section on the correlations functions.
468:
469: The analysis of $q_{T}$ and $q_{P}$ shows that the RSB picture gives a
470: consistent and satisfactory explanation of the numerical data. We
471: will see in the rest of this note that this conclusion is supported by
472: the analysis of many other quantities.
473:
474: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
475: \subsection{The Hole Distribution}
476:
477: It is interesting to consider a region of shape $\cal R$ and to define
478: the probability $P_1({\cal R},L)$ that a box of such a shape does {\em
479: not} intersect the interface. In the case of a box of shape $2\times
480: 1 \times 1$ it is clear that $P_1({\cal R},L)$ coincides with
481: $\frac{1}{2}(1+q_{l})$.
482:
483: Following Palassini and Young \cite{PALYOU} we define $1-P_1(M,L)$ as
484: the probability that the interface does hit a cubic box of side $M$,
485: i.e. we take ${\cal R} =M\times M \times M$. In figure (\ref{cubi}) we
486: show our numerical data for various values of $L$ and $M$. It is
487: possible to fit these data (if we restrict ourself in the region $M\le
488: \frac{L}{2}$, in order to have comparable finite volume effects for
489: the different $L$ values) by a simple power of $L$ (i.e. by the form
490: $A L^{-\alpha}$): however the power $\alpha$ of the best fits strongly
491: depends on $M$ (it is given by $0.32$, $0.23$ and $0.14$ for $M$ $=$
492: $2$, $3$ and $4$ respectively), while in a meaningful fit we would
493: expect to find a $M$-independent exponent.
494:
495: Second order polynomial fits in powers of $\frac{1}{L}$ work well in
496: the region $M\le \frac{L}{2}$: they give for the extrapolated
497: $\left(1-P_1\left(M,\infty\right)\right)$ the values $0.35$, $0.56$
498: and $0.65$ for $M$ $=$ $2$, $3$ and $4$ respectively. The data are
499: consistent with the possibility that $1-P_1(M,\infty)$ goes to $1$
500: when $M\to\infty$. Indeed the extrapolated data reported before can
501: be fitted as $1-P_1(M,\infty) = 1 - A M^{-\gamma}$ with $\gamma$ close
502: to one.
503:
504: It is clear that also in this case the data do not favor the kind of
505: behavior predicted by the droplet model, where $1-P_1(M,\infty) =0$: they
506: are far better consistent with the possibility that $1-P_1(M,\infty)$ is a
507: non trivial function of $M$, which goes to $1$ (and it is not zero) when
508: $M$ goes to infinity.
509:
510: \begin{figure}
511: \centering
512: \includegraphics[width=0.5\textwidth,angle=270]{fig2.ps}
513: \caption[a]{ $1-P_1(M,L)$ versus $\frac{M}{L}$ for $M=2, ...,
514: \min\left(L,8\right)$, and $L$ $=$ $4$, $6$, $8$, $10$, $12$, $14$.
515: \protect\label{cubi} }
516: \end{figure}
517:
518: This conclusion is strengthened if we look for example to the plot of
519: $P_1(2,L)-P_1(3,L)$ and $P_1(3,L)-P_1(4,L)$: the data for those two
520: quantities are shown in figure (\ref{differenze}). The droplet model
521: suggestion that these quantities go to zero as a power of $L$ when $L$
522: goes to infinity does not seem very convincing.
523:
524: \begin{figure}
525: \centering
526: \includegraphics[width=0.5\textwidth,angle=270]{fig3.ps}
527: \caption[a] {The quantities $P_1(2,L)-P_1(3,L)$ and
528: $P_1(3,L)-P_1(4,L)$ versus $\frac{1}{L}$ for $L$ $=$ $4$, $6$, $8$,
529: $10$, $12$, $14$. \protect\label{differenze} }
530: \end{figure}
531:
532: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
533: \section{Scaling Behavior}
534:
535: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
536: \subsection{The Hole Distribution Again}
537:
538: As we have discussed in the introduction if in the infinite volume
539: limit the interface becomes a fractal, characterized by an unique
540: fractal dimension $D_{s}$, and the density of the interface goes to
541: zero as $L^{-\alpha}$, with $\alpha=D-D_{s}$, the probability of
542: finding a hole in a box of size $\cal R$ goes to a non trivial
543: function of $\cal R$ if we keep constant the size of the box in units
544: of $L$. In the opposite case the interface would be become a space
545: filling dense object. If we consider boxes of side $M$ the previous
546: argument shows that the probability that the box does intersect the
547: interface $1-P_1(M,L)$ should be a function of $\frac{M}{L}$. A
548: glance to figure (\ref{cubi}) shows that our numerical data do not
549: exhibit such scaling: the interface is not a fractal characterized by
550: a simple fractal dimension. On the contrary if we plot
551: $L^{-1.3}\ln(P_1(M,L))$ versus $\frac{M}{L}$ (see figure
552: (\ref{cubibis}) we find a reasonable scaling behavior, indicating
553: again that $1-P_1(M,L)$ goes exponentially to $1$ when $L$ goes to
554: infinity at fixed $\frac{M}{L}$.
555:
556: The same conclusion holds very clearly if we consider the case of
557: boxes of size $L \times L \times 1$, i.e. planes, oriented in the
558: $y-z$ directions, i.e. parallel to the interface. The probability
559: that such a plane does {\em not} intersect the interface $P_{p}(L)$
560: would tend to $1$ in the situation (1) described in section
561: (\ref{FOURP}), while it would tend to a non zero number in situation
562: (2), The numerical data (in figure (\ref{piani})) decrease very
563: rapidly as a function of $L$ (they can be reasonably fitted as
564: $aL^{-\delta}$ with $\delta$ of order two, but a better fit has the
565: form $a\exp(-bL^{c})$: the asymptotic limit is zero in both cases).
566: The first scenario we have presented in section (\ref{FOURP}) i.e. the
567: one of an exponent $z<1$, would imply that this probability goes to
568: one asymptotically: this is certainly not the case.
569:
570:
571: This set of data shows that the interface is space filling on a scale
572: proportional to the side $L$ of the system. Of course one could
573: suggest that the numerical data are affected by {\em very} strong
574: corrections to scaling: however it is not clear why this should happen
575: (the scaling in figure (\ref{cubibis}) is very good). In order to see if
576: we are able to detect any trace of these corrections it is convenient
577: to consider other quantities.
578:
579: \begin{figure}
580: \centering
581: \includegraphics[width=0.5\textwidth,angle=270]{fig_pml_log.ps}
582: \caption[a]{$L^{-1.3}\ln(P_1(M,L))$ versus $\frac{M}{L}$ for
583: $M=2,..., \min\left(L,8\right)$ and $L$ $=$ $4$, $6$, $8$, $10$, $12$, $14$.
584: \protect\label{cubibis}}
585: \end{figure}
586:
587: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
588: \subsection{Other Scaling Laws}\label{other}
589:
590: If we go back to our boxes of side $M$ we can define the window
591: overlap \cite{WINDOW} as the value of the overlap $q$ restricted to
592: one of these boxes. We can define $P_{M,L}(|q|)$ as the probability
593: distribution of the absolute value of such window overlap: the
594: probability is symmetric, so that we can consider only non negative
595: values of $q$. The probability that the interface does not intersect
596: the cube of side $M$ is simply given by $P_{M,L}(1)$.
597:
598: In the case where the two ground states obtained under different
599: boundary conditions are as different as possible the quantity
600:
601: \begin{equation}
602: q^{2}_{M,L} \equiv \int dq\ q^{2}\ P_{M,L}(q)
603: \end{equation}
604: should go zero at fixed $\frac{M}{L}$ at large $M$. A simple possible
605: scaling behavior is that for large $M$ and $L$
606:
607: \begin{equation}
608: q^{2}_{M,L} \simeq L^{-\delta}f(ML^{-1})\ .
609: \end{equation}
610: We show the numerical data for $q^{2}_{M,L}L^{\delta}$ versus
611: $ML^{-1}$ figure (\ref{q2cubi}) (with $\delta=.32$). The data for $L>4$
612: show a beautiful and very accurate scaling behavior.
613:
614: \begin{figure}
615: \centering
616: \includegraphics[width=0.5\textwidth,angle=270]{fig4.ps}
617: \caption[a]{$P_{p}(L)$ versus $L^{-1}$ for
618: $L$ $=$ $4$, $6$, $8$, $10$, $12$, $14$.
619: \protect\label{piani} }
620: \end{figure}
621:
622: The data in figures (\ref{cubibis}) and (\ref{q2cubi}) strongly
623: suggest that asymptotic scaling laws are valid with good accuracy
624: already for the lattice and block sizes that we have considered:
625: sub-leading corrections are small, and they do not seem to be of
626: exceptionally large size. Moreover we find that for windows that
627: occupy a finite fraction of the entire system in the infinite volume
628: limit the window overlap distribution seems to become a delta function
629: $\delta(q)$.
630: \begin{figure}
631: \centering
632: \includegraphics[width=0.5\textwidth,angle=270]{fig5.ps}
633: \caption[a]{ $q^{2}_{M,L}L^{\delta}$ versus
634: $ML^{-1}$ for $M=2, ..., \min\left(L,8\right)$ and
635: $L$ $=$ $4$, $6$, $8$, $10$, $12$, $14$.
636: \protect\label{q2cubi} }
637: \end{figure}
638:
639: Up to now all of our numerical evidence seems in perfect agreement
640: with the RSB predictions, and the first two possibilities we had
641: discussed in section (\ref{FOURP}) are excluded. In order to exclude the
642: more exotic last possibility, and to present further evidence for the
643: correctness of the RSB picture, it is convenient to study the
644: correlation functions, both of $q_{l}$ and of $q$. This will be done
645: in the next section.
646:
647: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
648: \section{Correlation Functions}
649:
650: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
651: \subsection {The Link Overlap Correlation Functions}
652:
653: We will present here a detailed analysis of the correlation functions:
654: this is important since they carry a large amount of information that
655: is crucial to distinguish among the different possible behaviors.
656:
657:
658: As we have already discussed if we consider the case where there is an
659: interface in the $y-z$ plane we must distinguish among correlations in
660: the transverse and in the perpendicular directions. Moreover in the
661: case of vector like quantities like $q_{l}$ we must distinguish among
662: correlations in the direction of the link and correlations in the
663: directions perpendicular to that of the link. For simplicity here we
664: will consider only one correlation, i.e. the transverse link overlap
665: correlation in the direction perpendicular to the link. More
666: precisely the link overlap correlation function we consider is defined
667: as
668:
669: \begin{equation}
670: G(d,L) \equiv
671: {1 \over 2 L^{3}}
672: \sum_{x,y,z}\overline{q_{l}(x,y,z;\hat{y})\ q_{l}(x,y,z+d;\hat{y})+
673: q_{l}(x,y,z;\hat{z})\ q_{l}(x,y+d,z;\hat{z})}\ .
674: \end{equation}
675: With this definition the correlation functions are periodic functions
676: of $d$ with period $L$, and they are symmetric functions of $d$, so
677: that only the case $0\le d\le \frac{L}{2}$ is interesting.
678:
679: \begin{figure}
680: \centering
681: \includegraphics[width=0.5\textwidth,angle=270]{fig_corlink.ps}
682: \caption[a]{The correlation functions $G(d,L)$ versus $w\equiv
683: \left(\frac{1}{d}+\frac{1}{L-d}\right)^{\lambda}$ with $\lambda=1.5$
684: for $d> 0$ and $L$ $=$ $4$, $6$, $8$, $10$, $12$, $14$. The solid
685: curves are smooth interpolations to the data point, and they turn out
686: to be remarkably linear and the lines corresponding to different sizes
687: are nearly parallel.
688: \protect\label{G}}
689: \end{figure}
690:
691: We show in figure (\ref{G}) the link overlap correlation functions at
692: distance greater than $0$ for different $L$ values. We plot them as a
693: function of the variable
694:
695: \begin{equation}
696: w\equiv \left(\frac{1}{d}+\frac{1}{L-d}\right)^{\lambda} \ ,
697: \end{equation}
698: with $\lambda=1.5$. The reasons for this choice of the dependent
699: variable will be clearer later: right now we just notice that the
700: points look remarkably linear.
701:
702: If we want to concentrate our attention on the $d$ dependence of the
703: correlation function (neglecting a possible constant value at large distance)
704: we can consider the quantity
705:
706: \begin{equation}
707: \Delta G(d+\frac12,L) \equiv G(d,L)-G(d+1,L)\ .
708: \end{equation}
709: Simple scaling implies that the function
710:
711: \begin{equation}
712: f(w,L)\equiv\Delta G(d+\frac12,L) (d^{-1}+(L-d)^{-1})^{1+\lambda}\ ,
713: \end{equation}
714: where $w=dL^{-1}$, should be independent from $L$ for large $L$. The
715: numerical data are shown in figure (\ref{DeltaG}). A good scaling is
716: obtained with the choice of the value $\lambda =1.5$.
717:
718: We expect that the correlation function can be fitted at large $L$ and
719: $d$ as
720:
721: \begin{equation}
722: G(d,L) = B(L)
723: \left(d^{-1}+(L-d)^{-1}\right)^{-\lambda}
724: + A(L)^{2} \ .
725: \label{fitG}
726: \end{equation}
727: The form that we have taken is the simplest one which enforces the
728: periodicity. Of course different forms are possible.
729: In order to consider also the $d$-independent part of the correlations
730: functions it is convenient to fit the whole set of data shown in
731: figure (\ref{G}) using equation (\ref{fitG}), where for simplicity we
732: fix $\lambda=1.5$. The fits are good (they would also be good for
733: similar values of $\lambda$) so that the whole data can be
734: reconstructed by the knowledge of the parameters $A(L)$ and $B(L)$,
735: which are shown in figure (\ref{AB}).
736:
737: \begin{figure}
738: \centering
739: \includegraphics[width=0.5\textwidth,angle=270]{fig_dsca.ps}
740: \caption[a]{The quantity $f(w,L)=\Delta G(d+\frac12,L)
741: (d^{-1}+(L-d)^{-1})^{1+\lambda}$ with $\lambda=1.5$ as a function of
742: $d$. Here $L$ $=$ $4$, $6$, $8$, $10$, $12$, $14$.
743: \protect\label{DeltaG}}
744: \end{figure}
745:
746: \begin{figure}
747: \centering
748: \includegraphics[width=0.5\textwidth,angle=270]{fig_ab2.ps}
749: \caption[a]{The two fit parameters $A(L)$ and $B(L)$ as function of
750: $L$ for $L$ $=$ $4$, $6$, $8$, $10$, $12$, $14$. We show both a a
751: power fit and a polynomial fit in $\frac{1}{L}$.
752: \protect\label{AB}}
753: \end{figure}
754:
755: While $A(L)$ displays some dependence on $L$, there is no hint that
756: the quantity $B(L)$ should go to zero when $L\to \infty$. Certainly
757: both of them do not go to zero as $L^{-\alpha}$ with $\alpha
758: \approx.3$ as expected in the case of possibility (4), and they can be
759: consistently extrapolated to a non-zero value. If we would insist on
760: power law fits we would find a value of $\alpha_{A}=0.15$ for $A(L)$
761: and a ridiculous value of $\alpha_{B}=0.03$ for $B(L)$.
762:
763: It is remarkable that $A(L)$, that should tend to the link overlap in
764: the $L\to\infty$ limit, tends to $0.242$, to be compared with the
765: value $0.245\pm 0.015$ (see equation (\ref{E-TWOFOURFIVE})) that we
766: have derived from a completely different analysis: the picture that
767: emerges from these data give consistent support to a RSB physical
768: scenario. Also it is worth to comment more on the power fit of figure
769: (\ref{AB}): even if they are compatible with the data it looks clear
770: they are only minimally consistent with the physical behavior shown by
771: the measured points. Typically these fits stay basically constant for
772: all measured $L$ values, and than they forecast a sharp descend at
773: very high $L$ values: the polynomial fits show on the contrary a
774: consistent behavior in the measured $L$ range and in the $L\to\infty$
775: limit.
776:
777: The second scenario we have discussed is in complete variance with
778: the behavior of the correlation function data. The exponent $\lambda$
779: should be $.3$, not $1.5$, as indicated by the data. This conclusion
780: is in perfect agreement with the results of the hole probability
781: distribution.
782:
783: We are now ready to analyze the possible correctness of the fourth
784: scenario we have discussed (the interface is a dense set in the
785: continuum limit). The results of figure (\ref{DeltaG}) show that we
786: should have $D_{s}=D-\lambda \approx 1.5$. The fact that the
787: interface would have a fractal dimension less than 2 is strange, but
788: not impossible: the interface could look like the surface of
789: T-lymphocytes, which is full of microvilli, i.e. quasi one dimensional
790: objects.
791:
792: However in this case the correlation function (and therefore the
793: correlation functions at fixed distance) should go to zero as
794: $L^{-\alpha}$: the correlation functions at distances of order $L$
795: should go to zero as $L^{-2\alpha}$. The scaling observed in figure
796: (\ref{DeltaG}) should not be there. The two exponents $\alpha_{A}$ and
797: $\alpha_{B}$ should be both equal to $0.3$, which we have already seen
798: it is not true.
799:
800: It is evident that also the exotic possibility (4) is not compatible
801: with the observed form of the link overlap correlations (which are
802: perfectly compatible with simple scaling laws without anomalies), and
803: the only possibilities is given by the behavior implied by the RSB
804: scenario.
805:
806: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
807: \subsection {The Overlap Correlation Functions}
808:
809: Here we consider the overlap correlation function. In this case we can
810: define two interesting correlation functions: the transverse
811: correlation function
812:
813: \begin{equation}
814: C_{T}(d,L)={1 \over 2 L^{3}}\sum_{x,y,z}
815: \overline{q(x,y,z)\ q(x,y,z+d)+q(x,y,z)\ q(x,y+d,z)}\ ,
816: \end{equation}
817: and the perpendicular correlation function
818:
819: \begin{equation}
820: C_{P}(d,L)={1 \over L^{3}}\sum_{x,y,z}
821: \overline{q(x,y,z)\ q(x+d,y,z)\ S(x,x+d)}\ ,
822: \end{equation}
823: where $S(x,x+d)=\pm 1 $ is factor which implements gauge invariance:
824: it is equal to $-1$ if the line which connects the points $x$ and
825: $x+d$ crosses the plane where boundary conditions are changed.
826:
827: The two correlations functions are respectively periodic and
828: anti-periodic functions of $D$ with period $L$ ($C_{P}(L/2,L)=0$).
829:
830: \begin{figure}
831: \centering
832: \includegraphics[width=0.5\textwidth,angle=270]{fig6.ps}
833: \caption[a]{The overlap-overlap correlation functions divided times
834: the factor $\left(\frac{1}{d}+\frac{1}{L-d}\right)^{0.32}$ as a
835: function of $\frac{d}{L}$. The upper dots are for the perpendicular
836: correlation function divided an additional factor
837: $(1-\frac{2d}{L})$. Lines are best fits (see the text).
838: \protect\label{overcf}}
839: \end{figure}
840:
841: We notice that
842:
843: \begin{equation}
844: q_{P}(L)=C_{P}(1,L),\ \ \ \ q_{T}(L)=C_{T}(1,L).
845: \end{equation}
846: Consequently an analysis of the combined $d$ and $L$ dependence of the
847: overlap-overlap correlation functions can give information on the
848: origin of the $L$ dependence of the link overlap, which coincide with
849: the overlap-overlap correlation function at distance $1$.
850:
851: According to the previous analysis of the window overlap we expect
852: that the two correlations functions to scale as
853:
854: \begin{equation}
855: C(d,L)=f(dL^{-1}) (d^{-1}+(L-d)^{-1})^{-\delta} \ .
856: \label{fitC}
857: \end{equation}
858: The plot of figure (\ref{overcf}) shows a very good scaling, and two
859: parameter polynomial fits work very well. Transversal and
860: perpendicular correlation functions go to the same limit when $x\to
861: 0$. Since the perpendicular correlation function is identically zero
862: in $d=\frac{L}2$, we also plot the perpendicular $C$ divided times
863: $(1-\frac{2d}{L})$, removing in this way the main $\frac{d}{L}$
864: dependence, which comes from a kinematical effect: it goes very
865: smoothly to the correct limit when $x\to 0$.
866:
867: The extrapolated values are $0.732 \pm 0.008$ for the transverse
868: correlation function and $0.722 \pm 0.005$ for the perpendicular one.
869: This is well consistent with the estimate quoted before (and obtained
870: by measuring a completely different quantity) for the link overlap,
871: $q_{l} = 0.755 \pm 0.015$.
872:
873: As a consistency check we plot in figure (\ref{meta}) the correlation
874: function $C_{T}(L/2,L)$ as function of $L$. The data can be fitted
875: very well by simple power fit $C\ L^{-\delta}$, with $\delta \simeq
876: 0.32$.
877:
878: It is clear that we have identified a physical mechanism which
879: naturally generates an $\frac{1}{L}$ dependence in $q_{l}$
880: (correlation functions do usually feel the size of the system): simple
881: scaling laws are valid and the ground state structure is non
882: trivial. On the contrary a trivial structure (in the sense of the
883: droplet model) would imply that the quantities $f(s)$ extrapolate at
884: $1$ when $s$ goes to zero (which is evidently not true) or they do not
885: satisfy a simple scaling and there is a crossover region of length
886: $\xi(L)$ which goes to infinity with $L$. Both these possibilities are
887: not compatible with the data which strongly suggested the RSB
888: scenario.
889:
890: \begin{figure}
891: \centering
892: \includegraphics[width=0.5\textwidth,angle=270]{fig_cll2.ps}
893: \caption[a]{The correlation $C_{T}(L/2,L)$ as function
894: $L$ $=$ $4$, $6$, $8$, $10$, $12$, $14$.
895: \protect\label{meta} }
896: \end{figure}
897:
898: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
899: \section{Conclusions}
900:
901: We have studied the overlap among the two ground states obtained under
902: periodic and anti-periodic boundary conditions for $3D$ lattice of
903: linear sizes $L$ up to $L=14$. We have analyzed many different
904: observables: the average of the link overlap, the probability that the
905: a cube of size $M$ or a plane hits the interface, the window overlap
906: probability distribution and the correlation functions of the link
907: overlap and the overlap.
908:
909: The droplet model in the usual form assumes that the interface density
910: (which is one minus the link overlap) goes to zero for large $L$ as
911: $L^{-\alpha}$, and that the interface is a fractal with dimension
912: $D^{s}=D-\alpha=3-\alpha$. A small subset of our numerical data is
913: consistent with the possibility the interface density could go to zero
914: as $L^{-\alpha}$, but other data cannot be fitted as a simple power
915: and strongly suggest that the interface density remains finite in the
916: limit $L \to \infty$. The possibility that $D_{s}=3-\alpha$ is
917: completely excluded by the data. Some of the data are compatible with
918: $\alpha=0.3$ and $D_{s}=1.5$, but we have shown that this exotic
919: possibility is not compatible with the rest of data.
920:
921: The RSB scenario is perfectly compatible with the whole set of data.
922: The appropriate scaling laws for all the variables are satisfied with
923: amazing accuracy. The observed $L$ dependence of the interface
924: density can be easily explained if we consider the scaling law
925: appropriate for the overlap overlap correlation function. We also
926: want to notice the coherence of these results and the recent detection
927: of large scale excitations in spin glass ground states
928: \cite{KRZMAR,PALYOU2}.
929:
930: The behavior we find connects smoothly with the results obtained by
931: simulations at finite temperature \cite{MAPARIRUZU}. For example we
932: find that the overlap-overlap correlation decays as the distance to a
933: power $\delta=.3$, which is not very far from the value obtained from
934: simulation at finite temperature. Other results in this direction can
935: be found in \cite{MAPARU}.
936:
937: We think that we have conclusively shown that the ground state
938: structure of $3D$ Ising spin glasses with Gaussian quenched random
939: couplings is not trivial.
940:
941: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
942: \section*{Acknowledgments}
943: We are grateful to M. Palassini and P. Young for a very useful
944: correspondence.
945:
946: \begin{references}
947:
948: \bibitem{RSB}
949: G. Parisi,
950: Phys. Rev. Lett. {\bf 43}, 1754 (1979);
951: J. Phys. A {\bf 13}, 1101, 1887, L115 (1980);
952: Phys. Rev. Lett. {\bf 50}, 1946 (1983);
953: M. M\'ezard, G. Parisi and M. A. Virasoro,
954: {\em Spin Glass Theory and Beyond}
955: (World Scientific, Singapore 1987).
956:
957: \bibitem{MAPARIRUZU}
958: E. Marinari, G. Parisi, F. Ricci Tersenghi,
959: J.J. Ruiz Lorenzo and F. Zuliani,
960: J. Stat. Phys. {\bf 98}, 973 (2000),
961: cond-mat/9906076.
962:
963: %first work on ground states
964:
965: \bibitem{RIEGER}
966: H. Rieger,
967: {\em Frustrated Systems: Ground State Properties via
968: Combinatorial Optimization}, in
969: {\em Lecture Notes in Physics} {\bf 501} (Springer-Verlag,
970: Heidelberg 1998).
971:
972: \bibitem{MARTIN}
973: J. Houdayer and O. Martin,
974: Phys. Rev. Lett. {\bf 83}, 1030 (1999),
975: cond-mat/9901276.
976:
977: \bibitem{YOUNG}
978: M. Palassini and A. P. Young,
979: Phys. Rev. B {\bf 60}, R9919 (1999),
980: cond-mat/9904206.
981:
982: \bibitem{MIDDLETON}
983: A. Middleton,
984: Phys. Rev. Lett. {\bf 83}, 1672 (1999),
985: cond-mat/9904285.
986:
987: \bibitem{PALYOU}
988: M. Palassini and A. P. Young,
989: Phys. Rev. Lett. {\bf 83}, 5126 (1999),
990: cond-mat/9906323.
991:
992: \bibitem{MAPA}
993: E. Marinari and G. Parisi,
994: to be published.
995:
996: \bibitem{DROPLET}
997: W. L. McMillan,
998: J. Phys. C {\bf 17}, 3179 (1984);
999: A. J. Bray and M. A. Moore,
1000: in {\em Heidelberg Colloquium on Glassy Dynamics and Optimization},
1001: L. Van Hemmen and I. Morgenstern eds.
1002: (Springer-Verlag, Heidelberg 1986);
1003: D. S. Fisher and D. A. Huse,
1004: Phys. Rev. B {\bf 38}, 386 (1988).
1005:
1006: \bibitem{WINDOW}
1007: E. Marinari, G. Parisi, F. Ricci-Tersenghi and J. J. Ruiz-Lorenzo,
1008: J. Phys. A {\bf 31}, L481 (1998).
1009:
1010: \bibitem{KRZMAR}
1011: F. Krzakala and O. C. Martin,
1012: cond-mat/0002055.
1013:
1014: \bibitem{PALYOU2}
1015: M. Palassini and A. P. Young,
1016: cond-mat/0002134.
1017:
1018: \bibitem{MAPARU}
1019: E. Marinari, G. Parisi and J. J. Ruiz-Lorenzo,
1020: to be published.
1021:
1022: \end{references}
1023:
1024: \end{document}
1025: