1: %% 070110
2: %% Updated version, 070417
3:
4: \documentclass[twocolumn,prb,aps,showpacs]{revtex4}
5:
6: %\usepackage{psfrag,graphicx}
7: %\usepackage{dcolumn}
8: \usepackage{amsmath,amssymb}
9: %\usepackage{bm}
10: \usepackage{pstricks}
11: %\usepackage{hyperref}
12: \usepackage{epsfig}
13:
14: %\bibliographystyle{apsrev}
15:
16: \def\nind{\noindent}
17: \def\<{\left<}
18: \def\>{\right>}
19: \def\ket|#1>{\left|#1\right>}
20: \def\bra<#1|{\left<#1\right|}
21: \def\elem<#1|#2|#3>{\left<#1\right|#2\left|#3\right>}
22: \def\({\left(}
23: \def\){\right)}
24: \def\bul{$\bullet$ }
25:
26: \begin{document}
27:
28:
29: \title{Quantum Wavefunction Annealing of Spin Glasses on Ladders}
30:
31: \author{J. Rodr\'{\i}guez-Laguna}
32: \affiliation{International School for Advanced Studies (SISSA), Via Beirut 2-4,
33: I-34014 Trieste, Italy}
34:
35: \date{February 15, 2007}
36:
37:
38: \begin{abstract}
39: A technique inspired on quantum annealing is proposed in order to
40: obtain the classical ground state of a spin-glass by tracking the full
41: wavefunction of a given system within the subspace of matrix product
42: states (MPS), using the density matrix renormalization group
43: (DMRG). The technique is exemplified within the problem of obtention
44: of the classical ground state of an Ising spin glass on ladder
45: geometries. Its performance is evaluated and related to the
46: entanglement entropy.
47: \end{abstract}
48:
49: \pacs{
50: % 05.30.-d, % Quantum statistical mechanics
51: 73.43.Nq, % Quantum phase transitions
52: % 75.40.Cx, % Static properties (order parameter, static susceptibility,
53: % heat capacities, critical exponents, etc.)
54: 05.10.Cc % Renormalization group methods
55: 75.40.Mg, % Numerical simulation studies
56: 75.10.Nr, % Spin-glass and other random models
57: }
58:
59: \maketitle
60:
61:
62: \section{Introduction}
63:
64: Global optimization is one of the most challenging numerical
65: tasks. Simulated thermal annealing (STA) has been, for more than
66: twenty years, one of the most popular general purpose tools. Its
67: strength relies on its ability to escape metastable minima via thermal
68: fluctuations. Simulated quantum annealing (SQA)
69: \cite{Das_Chakrabarti:book} is a more recent algorithm which takes
70: profit from {\em quantum} fluctuations for the same purpose. Both
71: methods have a clear physical motivation: thermal annealing has been
72: carried out by metallurgists for millenia. On the other hand, the
73: first evidence of the superiority of quantum over thermal fluctuations
74: in order to find the global minimum of a real physical system was only
75: obtained in 1999, in the context of quantum spin glasses
76: \cite{Brooke_Sci99}.
77:
78: Numerical implementations of SQA have relied heavily on the {\em
79: quantum-classical} analogy described by the path integral Monte-Carlo
80: method (PIMC). Standard simulated thermal annealing is applied to an
81: enlarged system, which consists of several {\em replicas} of the
82: original system in interaction. Slightly different versions of the
83: method have succeeded in the obtention of the ground state of
84: classical spin glasses \cite{Kadowaki_PRE98,Santoro_SCI02}, atomic
85: clusters \cite{Lee_JPC00,Gregor_CPL05}, traveling salesman problem
86: \cite{Martonak_PRE04}, kinetically constrained systems
87: \cite{Das_PRE05} and small protein-folding problems \cite{Lee_JPC00},
88: among others. SQA has proved to work less efficiently than STA in
89: other problems, such as 3-SAT \cite{Battaglia_PRE05} and some
90: benchmark 1D potentials \cite{Stella_PRB05}.
91:
92: Despite the success of the PIMC-SQA, there are reasons to look for
93: different implementations \cite{Battaglia:chapter}. First of all, PIMC
94: simulations must be carried out at finite temperature, with lower
95: temperatures requiring a larger number or replicas and, therefore,
96: higher computational cost. Also, some quantum systems suffer from the
97: sign problem, or from problems with the Trotter break-up. Sampling
98: difficulties are always a risk for annealing schemes, and ensuring
99: ergodicity is often a highly non-trivial task.
100:
101: Advancing in this line, Green's function Monte-Carlo technique was
102: attempted by Stella and Santoro \cite{Stella_X06}. Its main
103: disadvantage is the necessity of good trial variational
104: wavefunctions. Real time evolution of the full wavefunction has been
105: implemented by Suzuki and Okada \cite{Suzuki:chapter}, making use of
106: both exact methods and the time-dependent density matrix
107: renormalization group (DMRG) algorithm \cite{White_PRL04}. For systems
108: where DMRG is efficient, loss of adiabaticity in the form of
109: Landau-Zener level crossings is the main problem of this approach.
110:
111: The proposal of this work is to perform the annealing on the full
112: wavefunction, thus overcoming most of the difficulties associated with
113: PIMC. But, as opposed to the Suzuki-Okada approach, a real time
114: evolution is not needed either. Instead, the ground state of the full
115: system is computed exactly for a high intensity of the quantum
116: fluctuations. As these fluctuations are decreased, the ground state is
117: subsequently updated, by making the necessary small modifications of
118: the previous ground state. The full wavefunction of the ground state
119: is stored in the form of a matrix product state (MPS), and is computed
120: variationally using the DMRG.
121:
122: As a benchmark case-study we have selected the obtention of the
123: classical ground state of an Ising spin-glass with couplings uniformly
124: distributed in $[-1,1]$. The quantum fluctuations are provided by a
125: transverse magnetic field. We have chosen the {\em ladders} for the
126: underlying topology, since it is the simplest case which presents
127: genuine frustration while retaining the quasi-1D character which is
128: required for the DMRG to attain its maximum efficiency. Albeit the
129: problem is in class P, we will show it to be complex enough to be
130: considered non-trivial.
131:
132: This article is organized as follows. Section \ref{model} introduces
133: our model hamiltonian, summarizing its general features as applied to
134: our featured topology. It is followed, in section \ref{qwa}, by a
135: detailed description of the proposed quantum wavefunction annealing
136: (QWA) method. In section \ref{results} we analyze the results of the
137: numerical experiments we have conducted. We conclude, in section
138: \ref{discussion}, with a discussion of the advantages and limits of
139: the method, and its possible extensions.
140:
141: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
142:
143: \section{Random Ising model in a Transverse Field on Ladders}
144: \label{model}
145:
146: Our model hamiltonian is the random Ising model in a transverse field
147: (RITF), which we define on an arbitrary graph ${\cal G}$ with $N$
148: sites \cite{Chakrabarti_Dutta_Sen:book}:
149:
150: \begin{equation}
151: H=-\sum_{\<i,j\>} J_{ij} \sigma^z_i \sigma^z_j
152: - \Gamma \sum_i \sigma^x_i
153: - \sum_i h^z_i\sigma^z_i
154: \label{ritf}
155: \end{equation}
156:
157: \nind where $\<i,j\>$ denotes nearest neighbours in ${\cal G}$, the
158: $J_{i,j}$ are independent random variables uniformly distributed in
159: $[-1,1]$, $\Gamma$ is the transverse field and $\sigma^x_i$ and
160: $\sigma^z_i$ are Pauli matrices at site $i$. We will work on the basis
161: of eigenstates of $\sigma^z$. Unless explicitly stated, we will assume
162: the longitudinal fields $h^z_i$ to be zero. The first term in
163: eq. \ref{ritf} may be considered as a potential energy, and the second
164: as a kinetic term: the magnitude of $\Gamma$ measures the intensity of
165: the quantum fluctuations. For $\Gamma=0$, the ground state is composed
166: of only two states, related by a trivial $+\Leftrightarrow -$
167: symmetry, which we term the classical ground states (CGS) of the spin
168: glass. As $\Gamma\to\infty$, on the other hand, all spins should be
169: pointing in the $X$-direction, therefore obtaining the following
170: ground state:
171:
172: \begin{equation}
173: \ket|X>=\otimes_{i=1}^N \frac{1}{\sqrt{2}} \( \ket|+>_i + \ket|->_i \)
174: \label{paramagnetic_state}
175: \end{equation}
176:
177: \nind All the wavefunction components of the ground state take the
178: same value in our basis for $\Gamma\to\infty$.
179:
180: The behaviour of this system at $T=0$ is rather well known in 1D
181: \cite{Fisher_PRB95,Fisher_PRB98} thanks to an insightful RG analysis,
182: in 2D \cite{Rieger_PRL94} and 3D \cite{Guo_PRL94} using quantum
183: Monte-Carlo, and in random graphs with fixed connectivity
184: \cite{Laguna_X06} making use of the DMRG. In all these cases, the
185: system presents a quantum spin-glass transition (QSGT) at a finite
186: value of $\Gamma=\Gamma_c$, where the energy gap vanishes. Above that
187: value, the system is said to be in a {\em quantum paramagnetic}
188: regime, whilst below it behaves as a {\em quantum spin-glass}.
189:
190: The problem of finding the global CGS for equation \ref{ritf} at
191: $\Gamma=0$ is known to be in class P for $D=1$ and for $D=2$ in the
192: absence of longitudinal magnetic field $h^z$. In the 2D case with
193: $h^z\neq 0$, the 3D case \cite{Barahona_JPA82} or in random graphs
194: with fixed connectivity \cite{Liers_PRB03}, the problem is
195: NP-complete.
196: \vspace{5mm}
197:
198: {\em Rectangular ladders} constitute the simpest topology in which the
199: system described by equation \ref{ritf} presents genuine
200: frustration. Their size is characterized by two numbers: $L\times w$,
201: where $L$ is the length and $w$ is their width, or number of
202: legs. Figure \ref{ladder:fig} shows a specimen with size $5\times
203: 2$. The classical spin-glass with $\pm J$ couplings on these ladders
204: has been analyzed by a number of authors
205: \cite{Mattis_PRL99,Uda_PRB05}. We will now discuss the complexity of
206: the energy landscape for $\Gamma=0$ and a few basic characteristics of
207: the nature of the QSGT.
208:
209: \begin{figure}
210: \psset{unit=1mm}
211: \def\la{\psline[linestyle=dashed]}
212: \def\lb{\psline}
213: \rput(-25,-10){
214: \multirput(10,0){5}{\pscircle*(0,0){1}}
215: \rput(0,-10){\multirput(10,0){5}{\pscircle*(0,0){1}}}
216: \la(0,0)(10,0)\la(0,0)(0,-10)
217: \lb(10,0)(20,0)\la(10,0)(10,-10)
218: \lb(20,0)(30,0)\lb(20,0)(20,-10)
219: \la(30,0)(40,0)\la(30,0)(30,-10)
220: \lb(40,0)(40,-10)
221: \la(0,-10)(10,-10)
222: \lb(10,-10)(20,-10)
223: \la(20,-10)(30,-10)
224: \la(30,-10)(40,-10)}
225: \vspace{25mm}
226: \caption{\label{ladder:fig}A rectangular ladder of dimension $5\times
227: 2$. Coupling constants $J_{ij}$ are associated to links of the
228: graph. In the example of the figure, dashed (continuous) lines
229: represent negative-AFM (positive-FM) links, and the system is
230: frustrated.}
231: \end{figure}
232:
233: Minimization of the classical hamiltonian ($\Gamma=0$) can be done
234: using STA or PIMC-SQA. For the system sizes under consideration, both
235: methods yield robust estimates for the CGS energy following the
236: schemes described by Santoro and coworkers\cite{Santoro_SCI02}. In
237: order to analyse the complexity of the energy landscape, we relax the
238: parameters until the system becomes non-robust, i.e.: it yields
239: different results in different runs. At this point, the annealing
240: processes provide us a series of sample configurations which
241: constitute local minima of the energy. With this purpose, we have
242: applied a STA algorithm with a multiplicative scheme for $\beta$,
243: i.e.: $\beta\to r \beta$, from $\beta_0=0.1$ to $\beta_{max}=10^6$,
244: with $r=1+10^{-5}$ and $10^4$ steps per temperature. The results for
245: three samples of a $40\times 2$ ladder are shown in figure
246: \ref{sta_ladder:fig}. Each column contains the energy values obtained
247: after the procedure was repeated 20 times, providing several different
248: local energy minima (about 10). In all the cases shown, the lowest
249: energy corresponds to the CGS. This result points to a complex energy
250: landscape for the spin-glass ladders.
251:
252: \begin{figure}
253: \epsfig{file=ladder.ps,width=6.0cm,angle=0,clip=}
254: \caption{\label{sta_ladder:fig}Some local minima of the classical
255: hamiltonian for three $40\times 2$ samples, obtained with a
256: non-robust STA in order to sample the low energy configurations.}
257: \end{figure}
258:
259: The energy per spin of the CGS seems to converge, in the case of the
260: 2-legged ladders, to $\epsilon^{(2)} \approx -0.64$. In absence of
261: frustration, this value would be $-0.75$. In the case of the ladder
262: with 4 legs, it converges to $\epsilon^{(4)} \approx -0.71$, with the
263: unfrustrated value being $-0.875$.
264: \vspace{5mm}
265:
266: At $T=0$, the system described by eq. \ref{ritf} presents a quantum
267: spin-glass transition (QSGT) at a finite value of
268: $\Gamma=\Gamma_c$. The finite-size DMRG algorithm, suitably adapted
269: for our case\cite{Laguna_X06}, can be used to characterize this
270: transition. The behavior of some relevant observables has been traced
271: in figure \ref{qsgt_ladder:fig}: (a) the energy gap $\Delta E$, (b)
272: the maximum block entropy $S_{max}$, given by
273:
274: \begin{equation}
275: S_{max}\equiv -\hbox{Tr} \rho_L \ln(\rho_L)
276: \label{entropy}
277: \end{equation}
278:
279: where $\rho_L$ is the reduced density matrix for the left half of the
280: ladder, and (c) the spin glass susceptibility $\chi_{SG}$, as defined
281: by the following formula:
282:
283: \begin{equation}
284: \chi_{SG}\equiv \frac{1}{N} \sum_{i,j} \lim_{h^z_j\to 0}
285: \( \frac{\<\sigma^z_i\>}{h^z_j} \) ^2 \;,
286: \label{xsg}
287: \end{equation}
288:
289: \noindent i.e.: a small longitudinal magnetic field $h^z_j$, applied
290: at site $j$, generates a magnetization response on each site $i$,
291: which is measured (and squared, so as to disregard its sign); the
292: results are summed over all sites $i$ and averaged over all sites $j$.
293: The value of $\Gamma_c$ changes from sample to sample. Within a single
294: sample, $\Delta E$ vanishes and $\chi_{SG}$ diverges at the same value
295: of $\Gamma_c$, as it is shown in figure \ref{qsgt_ladder:fig}. For
296: that sample, $\Gamma_c=0.6$. The entropy is seen to present a more
297: complex behavior.
298:
299: \begin{figure}
300: \epsfig{file=sgladder.ps,width=6.0cm,angle=270,clip=}
301: \caption{\label{qsgt_ladder:fig}Behavior of the energy gap $\Delta E$,
302: the maximum block entropy $S_{max}$ and the spin-glass
303: susceptibility $\chi_{SG}$ for a sample spin-glass ladder of
304: dimension $40\times 2$, as a function of $\Gamma$. The QSGT is
305: marked very clearly by both $\chi_{SG}$ and $\Delta E$. On the other
306: hand, $S_{max}$ seems to have a more erratic behavior.}
307: \end{figure}
308:
309: Some insight can be gained by tracing a few individual components of
310: the full wavefunction, which is possible within the DMRG framework
311: \cite{Laguna_X06}. Figure \ref{wfc:fig} shows that, well within the
312: paramagnetic regime, all wavefunction components take the same
313: value. As we decrease $\Gamma$, the configurations with low energy
314: start to increase its weight in the ground state wavefunction, while
315: the configurations with high energy start to decrease. At the critical
316: point, all the configurations have started their decay, except the one
317: with the lowest energy.
318:
319: \begin{figure}
320: \epsfig{file=wfc.ps,width=6.0cm,angle=270,clip=}
321: \caption{\label{wfc:fig}Wavefunction components for some selected
322: configurations, as a function of $\Gamma$, for a sample ladder with
323: size $40\times 2$. All the configurations but one are local minima
324: of the classical energy function, obtained using a non-robust
325: STA. The transition point is marked with an vertical line.}
326: \end{figure}
327:
328: More theoretical and analytical work is needed in order to fully
329: characterize this QSGT. Some comments about it are provided in section
330: \ref{discussion}. This preliminary analysis can be summarized as: (a)
331: the energy landscape of the classical problem is complex, and (b) there
332: is a quantum phase transition at a finite value of $\Gamma$.
333:
334:
335: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
336:
337: \section{Quantum Wavefunction Annealing of Spin Glass Systems}
338: \label{qwa}
339:
340: Quantum wavefunction annealing (QWA) is an implementation of simulated
341: quantum annealing (SQA) which does not require real time evolution,
342: configurations sampling or Trotter splitting. It makes extensive use
343: of matrix product states (MPS) \cite{Fannes_CMP92,Roman_JPA98} and the
344: density matrix renormalization group (DMRG) \cite{White_PRL92}. For
345: the benefit of the readers which are not acquainted with those
346: techniques, the first paragraph of this section describes a simpler
347: formulation based on the Lanczos algorithm, but which can only be used
348: for very small lattices.
349:
350: The Lanczos diagonalization method improves its performance
351: dramatically if a good seed to the real ground state of the system is
352: provided \cite{Golub:book}. The hamiltonian given by equation
353: \ref{ritf} has a simple behaviour for $\Gamma\to\infty$: its ground
354: state is known to be given by the state $\ket|X>$ defined in equation
355: \ref{paramagnetic_state}. Therefore, starting with such a state as a
356: seed, it is fairly easy for the Lanczos algorithm to obtain the exact
357: ground state for a finite but high value $\Gamma_0$. This value may be
358: decreased at small finite steps, $\Gamma\to\Gamma-\Delta\Gamma$, and
359: the ground state for the previous value of $\Gamma$ can be employed as
360: the seed for the computation of the ground state at the new value. In
361: quantum annealing, the adiabatic theorem is needed in order to prove
362: that convergence to the CGS is sure in the limit $\Delta\Gamma\to
363: 0$. In our case, a weaker result is enough: convergence is sure as
364: long as there is a finite overlap between the ground states at any two
365: consecutive values of $\Gamma$.
366:
367: The obvious drawback of the previous algorithm is the size of the
368: Hilbert space, $\dim({\cal H})=2^N$. Using the plain Lanczos
369: algorithm, all the wavefunction components should be stored and acted
370: upon. A solution is to choose a low-dimensional subspace which is
371: known to contain the ground state of the hamiltonian given by equation
372: \ref{ritf} for all values of $\Gamma$. Matrix product states (MPS) can
373: provide such a subspace in some cases. These states may be written
374: down as:
375:
376: \begin{equation}
377: \ket|\Psi>=\sum_i \hbox{tr}\( A^{s_1} A^{s_2} \cdots A^{s_N} \)
378: \ket|s_1,s_2,\cdots,s_N>
379: \label{mps}
380: \end{equation}
381:
382: \nind where $\ket|s_1,s_2,\cdots,s_N>$ are the eigenstates of
383: $\sigma^z_i$ and the $A^{s_i}$ are $2N$ matrices of dimension $m\times
384: m$, whose entries are the variational parameters of our Ansatz. The
385: dimension of the MPS subspace is, therefore, bounded by $2Nm^2$. Both
386: the computational cost and the accuracy of the method depend strongly
387: on the dimension $m$ of the matrices, whose physical meaning is the
388: following: for all possible left-right splittings of the system, the
389: ground state is approximated retaining $m$ states to represent the
390: left part and other $m$ states for the right part. For any possible
391: state of the system, its representation as an MPS becomes exact for
392: $m$ large enough. Of course, the dimensions $m$ can be made local,
393: $m_i$, if necessary.
394:
395: The DMRG is a variational method within the subspace of the MPS
396: \cite{Rommer_PRB97}. It profits from the use of a density matrix in
397: order to select the $m$ states for the left and right blocks which fit
398: best to our current approximation of the global ground state. The
399: neglected eigenvalues of the density matrix provide a way to monitor
400: the accuracy of the method. In all our applications, the tolerance to
401: the total {\em neglected probabilities} is fixed beforehand to a
402: certain value $\eta$, and the number of retained states $m(\eta)$, is
403: adapted in consonance.
404:
405: The MPS represent {\em faithfully} ground states of local 1D
406: hamiltonians \cite{Verstraete_PRB06}. In other cases, at this stage,
407: only educated experience can decide whether they are appropriate or
408: not. They are specially suited to 1D and quasi-1D problems
409: (e.g. ladders, trees), although natural extensions to higher
410: dimensions are in active development
411: \cite{Nishino_PTP01,Verstraete_X04}. A crucial tool is von Neumann's
412: {\em block entropy}, $S=-\hbox{tr}(\rho \log_2 \rho)$. In crude terms,
413: $m$ should scale as $\exp(S)$ in order to obtain an accurate DMRG
414: method. In 1D, the entropy $S$ is known to be bounded for non-critical
415: systems and to scale as $\log(N)$ for a critical one
416: \cite{Vidal_PRL02}. For higher dimensions, the {\em area law} predicts
417: the entropy to scale as $L^{d-1}$ out of criticality, where $L$ is a
418: typical dimension of the system \cite{Sredniki_PRL93}.
419:
420: As in the Lanczos case, the DMRG can benefit from a good {\em seed}
421: when searching the ground state of a hamiltonian through the use of
422: the {\em wavefunction transformations} \cite{White_PRL96}, which allow
423: to use the solution of an RG-step as an initial step for the next
424: one. Having found the ground state of a certain hamiltonian $H$, a few
425: finite-size sweeps will {\em adapt} it to become the ground state of a
426: slightly modified hamiltonian $H'$. In normal cases, this procedure is
427: not needed, since the number of sweeps required for convergence to the
428: ground state for any hamiltonian is small. It may become very useful
429: when (a) it is known that the ground state of a certain hamiltonian
430: {\em can} be represented as a MPS with low $m$ but (b) the probability
431: of the DMRG getting stuck at an excited state is very high. This is
432: the case for the ground state of \ref{ritf} in the quantum spin-glass
433: phase\cite{Laguna_X06}.
434:
435: Our proposed QWA algorithm is, therefore, the following one:
436:
437: \begin{description}
438:
439: \item[{\bf (a)}] The ground state of the system \ref{ritf} is obtained
440: for very high $\Gamma$.
441:
442: \item[{\bf (b)}] The transverse field $\Gamma$ is decreased
443: $\Gamma\to \Gamma-\Delta\Gamma$.
444:
445: \item[{\bf (c)}] A few finite-size sweeps of the DMRG are carried out,
446: which adapt the ground state to the current value of $\Gamma$.
447:
448: \item[{\bf (d)}] Go to (b) if $\Gamma$ is not yet zero.
449:
450: \item[{\bf (e)}] Measure the energy and classical ground state.
451:
452: \end{description}
453:
454: This approach is {\em deterministic}, i.e.: not limited by sampling
455: problems. It works directly with the quantum hamiltonian, therefore
456: does not require any Trotter break-up, and works directly at
457: $T=0$. Since it does not simulate real time evolution, it is not prone
458: to Landau-Zener level-crossings. Loss of adiabaticity is not,
459: therefore, a crucial issue. As it was stated, it is enough to ensure
460: that the overlap between the ground states at consecutive values of
461: $\Gamma$ is finite. A high value of $\Delta\Gamma$ is allowed far away
462: from the quantum phase transition, while a more reduced value will be
463: taken near it. Our precise adaptive reduction schedule is described in
464: the following section.
465:
466: On the negative side, it is based on a method which is specially
467: designed for 1D systems. We will say more about this in the
468: conclusions. Its main weakness stems from the inability of the MPS to
469: represent faithfully the ground state for all values of $\Gamma$ for
470: higher dimension. The efficiency and practical issues of the
471: implementation are analyzed in the next section.
472:
473:
474: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
475:
476: \section{Numerical Experiments}\label{results}
477:
478:
479: In order to study the applicability of QWA, we have generated random
480: samples of ladders with $w=2$ and $w=4$ legs, and lengths ranging from
481: $L=20$ to $L=160$. We have also considered random graphs of fixed
482: connectivity $K=3$ and, as a check, we have also analysed linear
483: chains (although the optimization problem is trivial in that
484: case). The annealing scheme has always been the same: starting with
485: $\Gamma_0=3$ and using an adaptative reduction,
486: $\Gamma\to\Gamma-\Delta\Gamma$ with
487:
488: \begin{equation}
489: \Delta\Gamma=\min(0.5,0.1/S)
490: \label{annealing_scheme}
491: \end{equation}
492:
493: \noindent where $S$ is the maximum block entropy for the previous
494: value of $\Gamma$. This way we ensure that, near the QSGT --when the
495: entropy is higher-- the steps are shorter. We insist in this fact: the
496: adiabatic theorem and Landau-Zener level crossings are not the
497: limitations of this method. Therefore, a decrease in the size of the
498: steps does not necessarily increase the probability of success.
499:
500: The last annealing step was always taken with $\Gamma_{min}=0.01$,
501: a value which is low enough for practical purposes. The obtention of
502: the minimum energy configuration was carried out measuring the
503: $z$-component of the spin polarization of all sites after convergence
504: for $\Gamma=\Gamma_{min}$. Of course, due to the obvious symmetry
505: $+\Leftrightarrow -$, the expectation value for each of these
506: components is zero. We have used a common procedure in quantum
507: spin-glass calculations: the insertion of a very small longitudinal
508: magnetic field $h^z=10^{-6}$ in a single site, selected at random, in
509: order to break the symmetry. In the QSG phase, because of the
510: divergence of the spin-glass susceptibility discussed above, an
511: infinitesimal localized magnetic field polarizes the full sample.
512:
513: Table \ref{results:table} provides the basic set of results. For each
514: geometry we have selected 20 samples and performed the QWA algorithm
515: on them. We have also chosen different values of the neglected
516: probabilities tolerance, $\eta$. For each run, QWA is said to obtain a
517: success if its minimum energy is equal to the value obtained by STA
518: and PIMC-SQA. QWA has had success, in all the attempted geometries,
519: whenever $\eta=10^{-8}$. When the tolerance was decreased, the CGS was
520: missed with higher probability, although the method can be seen to be
521: rather robust for spin-glass ladders, since the tolerance must be
522: raised up to $10^{-3}$ in order to decrease the probability of success
523: to $50\%$.
524:
525:
526: \begin{table}
527: \begin{ruledtabular}
528: \begin{tabular}{lccc}
529: Geom. & $\eta$ & Success & Time \cr
530: $20$ & $10^{-8}$ & 100\% & $1.2\pm 0.2$ \cr
531: $40$ & $10^{-8}$ & 100\% & $3.9\pm 0.8$ \cr
532: $80$ & $10^{-8}$ & 100\% & $10\pm 2$ \cr
533: $160$ & $10^{-8}$ & 100\% & $21\pm 4$ \cr
534: $320$ & $10^{-8}$ & 100\% & $56\pm 10$ \cr
535: \hline
536: \vspace{1mm}
537: $20\times 2$ & $10^{-8}$ & 100\% & $17\pm 8$ \cr
538: $40\times 2$ & $10^{-8}$ & 100\% & $60\pm 15$ \cr
539: $80\times 2$ & $10^{-8}$ & 100\% & $240\pm 30$ \cr
540: $160\times 2$ & $10^{-8}$ & 100\% & $600 \pm 110$ \cr
541: \hline
542: \vspace{1mm}
543: $40\times 2$ & $10^{-5}$ & 80\% & $26\pm 3$ \cr
544: $40\times 2$ & $10^{-3}$ & 50\% & $13\pm 1$ \cr
545: \hline
546: \vspace{1mm}
547: $20\times 4$ & $10^{-8}$ & 100\% & $1600\pm 600$ \cr
548: $40\times 4$ & $10^{-8}$ & 100\% & $6800\pm 2000$ \cr
549: $80\times 4$ & $10^{-8}$ & 100\% & $14000\pm 3000$\cr
550: $160\times 4$ & $10^{-8}$ & 100\% & $27000\pm 3000$\cr
551: \hline
552: \vspace{1mm}
553: RG-$20$ & $10^{-8}$ & 100\% & $190\pm 150$ \cr
554: RG-$100$ & $10^{-6}$ & 45\% & $15000\pm 7000$\cr
555: \end{tabular}
556: \end{ruledtabular}
557: \caption{\label{results:table}Numerical results for the QWA
558: method. The first column states the geometry of the system under
559: study: $L$ if it is a linear chain, $L\times w$ if it is a ladder, and
560: RG-$N$ if it is a random graph with connectivity $K=3$. The second
561: gives the tolerance for neglected probabilities in the DMRG,
562: $\eta$. The third provides the percentage of success. The fourth
563: column provides the average time (in seconds) for the QWA of a single
564: sample.}
565: \end{table}
566:
567: We have also tried to check the method with a different graph
568: topology: random graphs with fixed connectivity, $K=3$. For $N=20$
569: sites, employing a tolerance of $\eta=10^{-8}$, the method provides
570: again 100\% of success. If the size is increased, that tolerance can
571: not be set, since the number of states $m$ to be kept becomes
572: prohibitive. Therefore, we have carried out the experiments with
573: $\eta=10^{-6}$ for a $N=100$ sample, and the probability of success
574: reduces to $45\%$.
575:
576: The time for the QWA algorithm scales as a power law of the system
577: size: $T\approx L^{\alpha}$. For linear chains, using the results from
578: table \ref{results:table}, $\alpha\approx 1.3\pm 0.1$. For 2-legged
579: ladders, $\alpha\approx 2\pm 0.1$. In the case of the 4-legged
580: ladders, the power law fit has a higher errorbar, with a surprising
581: exponent of $\alpha\approx 1.5\pm 0.2$.
582:
583:
584:
585: \subsection{QWA and block entropy}
586:
587: This polynomial growth can be theoretically explained. For quasi-1D
588: systems at criticality, the von Neumann entropy grows as a logarithm
589: of the system size\cite{Vidal_PRL02}, $S_c(L)\approx a\log(L) +
590: b$. The value of $a$ has been related, in some random systems, to the
591: central charge of the associated conformal field
592: theory\cite{Refael_PRL04}. On the other hand, the maximum number of
593: retained states scales as the exponential of the entropy, $m\approx
594: \exp(S)$, and the time for a DMRG sweep scales as $T\approx Lm^2$. The
595: most expensive DMRG sweep for a QWA simulation takes place at the
596: critical point, and therefore we may expect $T\approx
597: L\exp(2S_c)\approx L^{2a+1}$. Thus, the theoretical prediction is
598: $\alpha\simeq 2a+1$.
599:
600: In the linear chain case, the entropy at criticality was predicted by
601: Refael and Moore \cite{Refael_PRL04} to have a coefficient
602: $a=\ln(2)/6\approx 0.11$, in agreement with our own numerical
603: measurements. Therefore, the theoretical prediction for the $\alpha$
604: exponent is $\approx 1.22$, which is not far from the $1.3\pm 0.1$
605: obtained numerically. In the case of the 2-legged ladder, there is no
606: theoretical estimate, but our own numerical simulations provide a
607: value $a\approx 0.55\pm 0.1$, thus giving an estimate for $\alpha$
608: around $2.1$, in agreement with the numerically observed value
609: $\alpha\approx 2 \pm 0.1$.
610:
611: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
612:
613: \section{Conclusions and Further Work}\label{discussion}
614:
615: In this paper we have introduced a method for the obtention of the
616: minimum energy configuration of a spin-glass, based on the annealing
617: of the full wavefunction represented as a matrix product state (MPS)
618: using the density matrix renormalization group (DMRG). The method has
619: been termed {\em quantum wavefunction annealing} (QWA), and we have
620: assessed and quantified its efficiency, by comparing its results with
621: those provided by other robust methods. For spin-glass ladders of 2
622: and 4 legs with lengths ranging from $20$ to $160$, the method has
623: always provided the optimum solution when the probability tolerance is
624: set to $\eta=10^{-8}$. The running time scales as $L^2$ for 2-legged
625: ladders, and, surprisingly, as $L^{1.5}$ for 4-legged ones. This
626: anomaly requires further clarification, perhaps in the line of thought
627: of Ferraro et al \cite{Ferraro_X07}. For random graphs with fixed
628: connectivity $K=3$, on the other hand, the tolerance has to be
629: increased in practice to $10^{-6}$, and the probability of success
630: falls to $45\%$ for $N=100$.
631:
632: The most crucial parameter which determines the success of the method
633: is the probability tolerance $\eta$. Whenever we were able to set
634: $\eta=10^{-8}$, for whichever topology, the system always attained the
635: optimum configuration. Unfortunately, for topologies which are not
636: quasi-1D, the number of retained states in DMRG, $m$, grows very fast
637: when the tolerance is decreased. Other parameters, such as the
638: annealing velocity, do not have such a direct relevance to the quality
639: of the results. Loss of adiabaticity is not the main issue for this
640: method: a finite overlap between the ground state wavefunctions at
641: consecutive annealing steps is enough to ensure convergence, as long
642: as the number of retained states $m$ is large enough for our Ansatz
643: wavefunction to represent both of them faithfully.
644:
645: Therefore, it seems reasonable to think that QWA can be converted into
646: a method to find the optimum configuration of all quasi-1D systems in
647: polynomial time. This suggests that, in fact, no quasi-1D problem can
648: be NP-complete. On the other hand, for problems which are known to be
649: NP-complete, such as the random Ising spin-glass in a 2D lattice with
650: a longitudinal field, or on random graphs of fixed connectivity, the
651: time for an algorithm which is able to obtain the optimum with
652: certainty should scale exponentially with the size of the system.
653:
654: This fact puts a limit to possible extensions of the QWA algorithm
655: described in this paper. The natural extension of the MPS are the {\em
656: tensor product states} (TPS) analyzed by Nishino and coworkers
657: \cite{Nishino_PTP01} or the pair-entangled particle states (PEPS)
658: \cite{Verstraete_X04} of Verstraete and Cirac. The computational power
659: of this type of states has been analyzed recently \cite{Schuch_X06},
660: pointing to the fact that any attempt to solve NP-complete problems
661: with an exact QWA algorithm using these states would require an
662: exponential time, which is compatible with the usually believed notion
663: that P $\neq$ NP.
664:
665: The most promising lines for future work are, therefore, in the
666: development of heuristic algorithms, running in polynomial time, which
667: may give the absolute minimum energy state with high probability for
668: {\em some} problems. A possibility is the development of a QWA
669: algorithm using TPS (or PEPS) with a fixed number of retained states
670: $m$. Another possibility is to use the DMRG approach to
671: non-equilibrium classical problems \cite{Degenhard_MMS04} to develop a
672: classical wavefunction annealing algorithm, which would solve the
673: Fokker-Planck equation associated to simulated thermal annealing.
674:
675: \begin{acknowledgments}
676: The author acknowledges G.~Santoro, A.~Trombettoni,
677: M.A.~Mart\'{\i}n-Delgado and G.~Sierra for very useful
678: discussions.
679: \end{acknowledgments}
680:
681: %\section*{References}
682:
683: % Comment if you want to insert a .bbl file directly below
684: %\bibliography{Disorder}
685:
686: \begin{thebibliography}{38}
687: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
688: \expandafter\ifx\csname bibnamefont\endcsname\relax
689: \def\bibnamefont#1{#1}\fi
690: \expandafter\ifx\csname bibfnamefont\endcsname\relax
691: \def\bibfnamefont#1{#1}\fi
692: \expandafter\ifx\csname citenamefont\endcsname\relax
693: \def\citenamefont#1{#1}\fi
694: \expandafter\ifx\csname url\endcsname\relax
695: \def\url#1{\texttt{#1}}\fi
696: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
697: \providecommand{\bibinfo}[2]{#2}
698: \providecommand{\eprint}[2][]{\url{#2}}
699:
700: \bibitem[{\citenamefont{Das and Chakrabarti}(2005)}]{Das_Chakrabarti:book}
701: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Das}} \bibnamefont{and}
702: \bibinfo{author}{\bibfnamefont{B.~K.} \bibnamefont{Chakrabarti}},
703: \emph{\bibinfo{title}{Quantum Annealing and Related Optimization Methods}},
704: Lecture Notes in Physics (\bibinfo{publisher}{Springer Verlag},
705: \bibinfo{year}{2005}).
706:
707: \bibitem[{\citenamefont{Brooke et~al.}(1999)\citenamefont{Brooke, Bitko,
708: Rosenbaum, and Aeppli}}]{Brooke_Sci99}
709: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Brooke}},
710: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Bitko}},
711: \bibinfo{author}{\bibfnamefont{T.~F.} \bibnamefont{Rosenbaum}},
712: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Aeppli}},
713: \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{284}},
714: \bibinfo{pages}{779} (\bibinfo{year}{1999}).
715:
716: \bibitem[{\citenamefont{Kadowaki and Nishimori}(1998)}]{Kadowaki_PRE98}
717: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Kadowaki}} \bibnamefont{and}
718: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Nishimori}},
719: \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{58}},
720: \bibinfo{pages}{5355} (\bibinfo{year}{1998}).
721:
722: \bibitem[{\citenamefont{Santoro et~al.}(2002)\citenamefont{Santoro,
723: {Marto\v{n}\'{a}k}, Tosatti, and Car}}]{Santoro_SCI02}
724: \bibinfo{author}{\bibfnamefont{G.~E.} \bibnamefont{Santoro}},
725: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{{Marto\v{n}\'{a}k}}},
726: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Tosatti}}, \bibnamefont{and}
727: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Car}},
728: \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{295}},
729: \bibinfo{pages}{2427} (\bibinfo{year}{2002}).
730:
731: \bibitem[{\citenamefont{Lee and Berne}(2000)}]{Lee_JPC00}
732: \bibinfo{author}{\bibfnamefont{Y.-H.} \bibnamefont{Lee}} \bibnamefont{and}
733: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Berne}}, \bibinfo{journal}{J.
734: Phys. Chem. A} \textbf{\bibinfo{volume}{104}}, \bibinfo{pages}{86}
735: (\bibinfo{year}{2000}).
736:
737: \bibitem[{\citenamefont{Gregor and Car}(2005)}]{Gregor_CPL05}
738: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Gregor}} \bibnamefont{and}
739: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Car}},
740: \bibinfo{journal}{Chem. Phys. Lett.} \textbf{\bibinfo{volume}{412}},
741: \bibinfo{pages}{125} (\bibinfo{year}{2005}).
742:
743: \bibitem[{\citenamefont{{Marto\v{n}\'{a}k}
744: et~al.}(2004)\citenamefont{{Marto\v{n}\'{a}k}, Santoro, and
745: Tosatti}}]{Martonak_PRE04}
746: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{{Marto\v{n}\'{a}k}}},
747: \bibinfo{author}{\bibfnamefont{G.~E.} \bibnamefont{Santoro}},
748: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Tosatti}},
749: \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{70}},
750: \bibinfo{pages}{057701} (\bibinfo{year}{2004}).
751:
752: \bibitem[{\citenamefont{Das et~al.}(2005)\citenamefont{Das, Chakrabarti, and
753: Stinchcombe}}]{Das_PRE05}
754: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Das}},
755: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Chakrabarti}},
756: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{R.~B.}
757: \bibnamefont{Stinchcombe}}, \bibinfo{journal}{Phys. Rev. E}
758: \textbf{\bibinfo{volume}{72}}, \bibinfo{pages}{026701}
759: (\bibinfo{year}{2005}).
760:
761: \bibitem[{\citenamefont{Battaglia
762: et~al.}(2005{\natexlab{a}})\citenamefont{Battaglia, Santoro, and
763: Tosatti}}]{Battaglia_PRE05}
764: \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Battaglia}},
765: \bibinfo{author}{\bibfnamefont{G.~E.} \bibnamefont{Santoro}},
766: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Tosatti}},
767: \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{71}},
768: \bibinfo{pages}{066707} (\bibinfo{year}{2005}{\natexlab{a}}).
769:
770: \bibitem[{\citenamefont{Stella et~al.}(2005)\citenamefont{Stella, Santoro, and
771: Tosatti}}]{Stella_PRB05}
772: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Stella}},
773: \bibinfo{author}{\bibfnamefont{G.~E.} \bibnamefont{Santoro}},
774: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Tosatti}},
775: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{72}},
776: \bibinfo{pages}{014303} (\bibinfo{year}{2005}).
777:
778: \bibitem[{\citenamefont{Battaglia
779: et~al.}(2005{\natexlab{b}})\citenamefont{Battaglia, Stella, Zagordi, Santoro,
780: and Tosatti}}]{Battaglia:chapter}
781: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Battaglia}},
782: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Stella}},
783: \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Zagordi}},
784: \bibinfo{author}{\bibfnamefont{G.~E.} \bibnamefont{Santoro}},
785: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Tosatti}},
786: in \emph{\bibinfo{booktitle}{Quantum annealing and related optimization
787: methods}}, edited by \bibinfo{editor}{\bibfnamefont{A.}~\bibnamefont{Das}}
788: \bibnamefont{and} \bibinfo{editor}{\bibfnamefont{B.~K.}
789: \bibnamefont{Chakrabarti}} (\bibinfo{publisher}{Springer Verlag},
790: \bibinfo{year}{2005}{\natexlab{b}}).
791:
792: \bibitem[{\citenamefont{Stella and Santoro}(2006)}]{Stella_X06}
793: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Stella}} \bibnamefont{and}
794: \bibinfo{author}{\bibfnamefont{G.~E.} \bibnamefont{Santoro}},
795: \bibinfo{journal}{cond-mat/0608420} (\bibinfo{year}{2006}).
796:
797: \bibitem[{\citenamefont{Suzuki and Okada}(2005)}]{Suzuki:chapter}
798: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Suzuki}} \bibnamefont{and}
799: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Okada}}, in
800: \emph{\bibinfo{booktitle}{Quantum annealing and related optimization
801: methods}}, edited by \bibinfo{editor}{\bibfnamefont{A.}~\bibnamefont{Das}}
802: \bibnamefont{and} \bibinfo{editor}{\bibfnamefont{B.~K.}
803: \bibnamefont{Chakrabarti}} (\bibinfo{publisher}{Springer Verlag},
804: \bibinfo{year}{2005}).
805:
806: \bibitem[{\citenamefont{White and Feiguin}(2004)}]{White_PRL04}
807: \bibinfo{author}{\bibfnamefont{S.~R.} \bibnamefont{White}} \bibnamefont{and}
808: \bibinfo{author}{\bibfnamefont{A.~E.} \bibnamefont{Feiguin}},
809: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{93}},
810: \bibinfo{pages}{076401} (\bibinfo{year}{2004}).
811:
812: \bibitem[{\citenamefont{Chakrabarti et~al.}(1996)\citenamefont{Chakrabarti,
813: Dutta, and Sen}}]{Chakrabarti_Dutta_Sen:book}
814: \bibinfo{author}{\bibfnamefont{B.~K.} \bibnamefont{Chakrabarti}},
815: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dutta}}, \bibnamefont{and}
816: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Sen}},
817: \emph{\bibinfo{title}{Quantum Ising phases and transitions in transverse
818: Ising models}}, Lecture Notes in Physics (\bibinfo{publisher}{Springer
819: Verlag}, \bibinfo{year}{1996}).
820:
821: \bibitem[{\citenamefont{Fisher}(1995)}]{Fisher_PRB95}
822: \bibinfo{author}{\bibfnamefont{D.~S.} \bibnamefont{Fisher}},
823: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{51}},
824: \bibinfo{pages}{6411} (\bibinfo{year}{1995}).
825:
826: \bibitem[{\citenamefont{Fisher and Young}(1998)}]{Fisher_PRB98}
827: \bibinfo{author}{\bibfnamefont{D.~S.} \bibnamefont{Fisher}} \bibnamefont{and}
828: \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Young}},
829: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{58}},
830: \bibinfo{pages}{9131} (\bibinfo{year}{1998}).
831:
832: \bibitem[{\citenamefont{Rieger and Young}(1994)}]{Rieger_PRL94}
833: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rieger}} \bibnamefont{and}
834: \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Young}},
835: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{72}},
836: \bibinfo{pages}{4141} (\bibinfo{year}{1994}).
837:
838: \bibitem[{\citenamefont{Guo et~al.}(1994)\citenamefont{Guo, Bhatt, and
839: Huse}}]{Guo_PRL94}
840: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Guo}},
841: \bibinfo{author}{\bibfnamefont{R.~N.} \bibnamefont{Bhatt}}, \bibnamefont{and}
842: \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Huse}},
843: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{72}},
844: \bibinfo{pages}{4137} (\bibinfo{year}{1994}).
845:
846: \bibitem[{\citenamefont{Rodr\'{\i}guez-Laguna and Santoro}(2006)}]{Laguna_X06}
847: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Rodr\'{\i}guez-Laguna}}
848: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.~E.}
849: \bibnamefont{Santoro}}, \bibinfo{journal}{cond-mat/0610661}
850: (\bibinfo{year}{2006}), \bibinfo{note}{submitted to PRB}.
851:
852: \bibitem[{\citenamefont{Barahona}(1982)}]{Barahona_JPA82}
853: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Barahona}},
854: \bibinfo{journal}{J. Phys. A: Math. Gen.} \textbf{\bibinfo{volume}{15}},
855: \bibinfo{pages}{3241} (\bibinfo{year}{1982}).
856:
857: \bibitem[{\citenamefont{Liers et~al.}(2003)\citenamefont{Liers, Palassini,
858: Hartmann, and J\"unger}}]{Liers_PRB03}
859: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Liers}},
860: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Palassini}},
861: \bibinfo{author}{\bibfnamefont{A.~K.} \bibnamefont{Hartmann}},
862: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{J\"unger}},
863: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{68}},
864: \bibinfo{pages}{094406} (\bibinfo{year}{2003}).
865:
866: \bibitem[{\citenamefont{Mattis and Paul}(1999)}]{Mattis_PRL99}
867: \bibinfo{author}{\bibfnamefont{D.~C.} \bibnamefont{Mattis}} \bibnamefont{and}
868: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Paul}},
869: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{83}},
870: \bibinfo{pages}{3733} (\bibinfo{year}{1999}).
871:
872: \bibitem[{\citenamefont{Uda et~al.}(2005)\citenamefont{Uda, Yoshino, and
873: Kawamura}}]{Uda_PRB05}
874: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Uda}},
875: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Yoshino}}, \bibnamefont{and}
876: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Kawamura}},
877: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{72}},
878: \bibinfo{pages}{024442} (\bibinfo{year}{2005}).
879:
880: \bibitem[{\citenamefont{Fannes et~al.}(1992)\citenamefont{Fannes, Nachtergaele,
881: and Werner}}]{Fannes_CMP92}
882: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fannes}},
883: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Nachtergaele}},
884: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Werner}},
885: \bibinfo{journal}{Commun. Math. Phys} \textbf{\bibinfo{volume}{144}},
886: \bibinfo{pages}{443} (\bibinfo{year}{1992}).
887:
888: \bibitem[{\citenamefont{Rom\'an et~al.}(1998)\citenamefont{Rom\'an, Sierra,
889: Dukelsky, and Mart\'{\i}n-Delgado}}]{Roman_JPA98}
890: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Rom\'an}},
891: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Sierra}},
892: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Dukelsky}}, \bibnamefont{and}
893: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Mart\'{\i}n-Delgado}},
894: \bibinfo{journal}{J. Phys. A: Math. Gen.} \textbf{\bibinfo{volume}{31}},
895: \bibinfo{pages}{9729} (\bibinfo{year}{1998}).
896:
897: \bibitem[{\citenamefont{White}(1992)}]{White_PRL92}
898: \bibinfo{author}{\bibfnamefont{S.~R.} \bibnamefont{White}},
899: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{69}},
900: \bibinfo{pages}{2863} (\bibinfo{year}{1992}).
901:
902: \bibitem[{\citenamefont{Golub and {van Loan}}(1996)}]{Golub:book}
903: \bibinfo{author}{\bibfnamefont{G.~H.} \bibnamefont{Golub}} \bibnamefont{and}
904: \bibinfo{author}{\bibfnamefont{C.~F.} \bibnamefont{{van Loan}}},
905: \emph{\bibinfo{title}{Matrix Computations}} (\bibinfo{publisher}{Hopkins
906: University Press}, \bibinfo{year}{1996}).
907:
908: \bibitem[{\citenamefont{Rommer and \"Ostlund}(1997)}]{Rommer_PRB97}
909: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Rommer}} \bibnamefont{and}
910: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{\"Ostlund}},
911: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{55}},
912: \bibinfo{pages}{2164} (\bibinfo{year}{1997}).
913:
914: \bibitem[{\citenamefont{Verstraete and Cirac}(2006)}]{Verstraete_PRB06}
915: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Verstraete}} \bibnamefont{and}
916: \bibinfo{author}{\bibfnamefont{J.~I.} \bibnamefont{Cirac}},
917: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{73}},
918: \bibinfo{pages}{094423} (\bibinfo{year}{2006}).
919:
920: \bibitem[{\citenamefont{Nishino et~al.}(2001)\citenamefont{Nishino, Hieida,
921: Okunishi, Maeshima, and Akutsu}}]{Nishino_PTP01}
922: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Nishino}},
923: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Hieida}},
924: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Okunishi}},
925: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Maeshima}}, \bibnamefont{and}
926: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Akutsu}},
927: \bibinfo{journal}{Prog. Theor. Phys.} \textbf{\bibinfo{volume}{105}},
928: \bibinfo{pages}{409} (\bibinfo{year}{2001}).
929:
930: \bibitem[{\citenamefont{Verstraete and Cirac}(2004)}]{Verstraete_X04}
931: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Verstraete}} \bibnamefont{and}
932: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Cirac}},
933: \bibinfo{journal}{cond-mat/0407066} (\bibinfo{year}{2004}).
934:
935: \bibitem[{\citenamefont{Vidal et~al.}(2002)\citenamefont{Vidal, Latorre, Rico,
936: and Kitaev}}]{Vidal_PRL02}
937: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Vidal}},
938: \bibinfo{author}{\bibfnamefont{J.~I.} \bibnamefont{Latorre}},
939: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Rico}}, \bibnamefont{and}
940: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kitaev}},
941: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{90}},
942: \bibinfo{pages}{227902} (\bibinfo{year}{2002}).
943:
944: \bibitem[{\citenamefont{Sredniki}(1993)}]{Sredniki_PRL93}
945: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Sredniki}},
946: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{71}},
947: \bibinfo{pages}{666} (\bibinfo{year}{1993}).
948:
949: \bibitem[{\citenamefont{White}(1996)}]{White_PRL96}
950: \bibinfo{author}{\bibfnamefont{S.~R.} \bibnamefont{White}},
951: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{77}},
952: \bibinfo{pages}{3633} (\bibinfo{year}{1996}).
953:
954: \bibitem[{\citenamefont{Refael and Moore}(2004)}]{Refael_PRL04}
955: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Refael}} \bibnamefont{and}
956: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Moore}},
957: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{93}},
958: \bibinfo{pages}{260602} (\bibinfo{year}{2004}).
959:
960: \bibitem{Ferraro_X07}A~Ferraro, A.~Garcia-Saez and A.~Acin,
961: quant-ph/0701009 (2007).
962:
963: \bibitem[{\citenamefont{Schuch et~al.}(2006)\citenamefont{Schuch, Wolf,
964: Verstraete, and Cirac}}]{Schuch_X06}
965: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Schuch}},
966: \bibinfo{author}{\bibfnamefont{M.~M.} \bibnamefont{Wolf}},
967: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Verstraete}},
968: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~I.} \bibnamefont{Cirac}},
969: \bibinfo{journal}{quant-ph/0611050} (\bibinfo{year}{2006}).
970:
971: \bibitem[{\citenamefont{Degenhard et~al.}(2004)\citenamefont{Degenhard,
972: Rodriguez-Laguna, and Santalla}}]{Degenhard_MMS04}
973: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Degenhard}},
974: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Rodriguez-Laguna}},
975: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~N.}
976: \bibnamefont{Santalla}}, \bibinfo{journal}{Mult. Model. Simul. (SIAM)}
977: \textbf{\bibinfo{volume}{3}}, \bibinfo{pages}{89} (\bibinfo{year}{2004}).
978:
979: \end{thebibliography}
980:
981:
982:
983: \end{document}
984: