0710.2581/lmg.tex
1: 
2: 
3: % Group addresses by affiliation; use superscriptaddress for long
4: % author lists, or if there are many overlapping affiliations.
5: % For Phys. Rev. appearance, change preprint to twocolumn.
6: % Choose pra, prb, prc, prd, pre, prl, prstab, or rmp for journal
7: %  Add 'draft' option to mark overfull boxes with black boxes
8: %  Add 'showpacs' option to make PACS codes appear
9: %  Add 'showkeys' option to make keywords appear
10: %\documentclass[aps,prb,preprint,groupedaddress,showpacs]{revtex4}
11: %\documentclass[aps,pre,preprint,superscriptaddress,showpacs]{revtex4}
12: \documentclass[aps,pra,twocolumn,superscriptaddress,showpacs]{revtex4}
13: \usepackage{pifont}
14: \usepackage{txfonts}
15: \usepackage{amssymb}
16: \usepackage{graphicx}% Include figure files
17: 
18: \begin{document}
19: 
20: \title{Quantum criticality of the Lipkin-Meshkov-Glick Model in terms
21: of fidelity susceptibility}
22: 
23: 
24: \author{Ho-Man Kwok}%
25: \affiliation{Department of Physics and Institute of Theoretical Physics, The
26: Chinese University of Hong Kong, Shatin, Hong Kong, China}
27: 
28: \author{Wen-Qiang Ning}%
29: \affiliation{Department of Physics and Institute of Theoretical Physics, The
30: Chinese University of Hong Kong, Shatin, Hong Kong, China}
31: \affiliation{Department of Physics, Fudan University, Shanghai 200433, China}
32: 
33: \author{Shi-Jian Gu}
34:  \email{sjgu@phy.cuhk.edu.hk}
35: \affiliation{Department of Physics and Institute of Theoretical
36: Physics, The Chinese University of Hong Kong, Shatin, Hong Kong,
37: China}
38: 
39: \author{Hai-Qing Lin}
40: \affiliation{Department of Physics and Institute of Theoretical
41: Physics, The Chinese University of Hong Kong, Shatin, Hong Kong,
42: China} \affiliation{Department of Physics, Fudan University,
43: Shanghai 200433, China}
44: 
45: \date{\today}
46: 
47: \begin{abstract}
48: We study the critical properties of the Lipkin-Meshkov-Glick Model in terms of
49: the fidelity susceptibility. By using the Holstein-Primakoff transformation, we
50: obtain explicitly the critical exponent of the fidelity susceptibility around
51: the second-order quantum phase transition point. Our results provide a rare
52: analytical case for the fidelity susceptibility in describing the universality
53: class in quantum critical behavior. The different critical exponents in two
54: phases are non-trivial results, indicating the fidelity susceptibility is not
55: always extensive.
56: \end{abstract}
57: 
58: 
59: \pacs{64.60.-i, 05.70.Fh, 75.10.-b}
60: 
61: 
62: %05.70.Fh    phase transitions: general studies
63: %03.67.Mn    entanglement production, characterization, and manipulation
64: %03.75.Gg    entanglement and decoherence Bose-Einstein condensates
65: %03.75.Hh    static properties of condensates, thermodynamical statistical
66: %            and structural properties
67: %75.10.jm    quantized spin models
68: %71.10.Fd    Lattice fermion models (Hubbard model, etc.)
69: %75.30.kz    magnetic phase boundaries (including magnetic transitions,
70: %            metamagnetism, etc)
71: %03.67.-a    quantum information
72: %64.60.-i    General studies of phase transitions (see also 63.70.+h
73: %Statistical mechanics of lattice vibrations and displacive phase transitions;
74: %for critical phenomena in solid surfaces and interfaces, and in
75: %magnetism, see 68.35.Rh, and 75.40.-s, respectively)
76: %75.10.-b General theory and models of magnetic ordering (see also
77: %05.50.+q Lattice theory and statistics)
78: 
79: 
80: %05.70.Fh, 03.67.Mn, 03.75.Gg, 03.75.Hh
81: 
82: % PACS Number
83: 
84: \maketitle
85: 
86: \section{Introduction}
87: 
88: The Lipkin-Meshkov-Glick (LMG) model \cite{LMG} was introduced in nuclear
89: physics. It describes a cluster of mutually interacting spins in a transverse
90: magnetic field. In condensed matter physics, this model is associated with a
91: system of infinite coordination number. In earlier time, scaling behaviors of
92: critical observables have been studied by mean field analysis \cite{infcoor},
93: while recently the finite-size scaling of this model was studied by the $1/N$
94: expansion in the Holstein-Primakoff single boson representation \cite{HP} and
95: by the continuous unitary transformations (CUT) \cite{CUT1,CUT2,JVidalStat}.
96: Meanwhile, a rich structure of four different regions is revealed in the
97: parameter space through a careful scrutiny on the spectrum
98: \cite{JVidalSpectra}. Besides, the quantum criticality has been investigated by
99: studying its entanglement properties \cite{QPT1st,QPT2nd,EntE1,EntE2,nilesen}.
100: Both the first- and second-order quantum phase transitions (QPTs)
101: \cite{sachdev} have been revealed, in the antiferromagnetic and the
102: ferromagnetic cases respectively \cite{QPT1st,QPT2nd}.
103: 
104: Regarding the QPT itself, the ground state of a system would
105: undergo a significant structural change at certain critical point.
106: This primary observation suggests a new description of QPTs in
107: terms of fidelity
108: \cite{HTQuan06,PZanardi06,Buonsante07,PZanardi0606130,PZanardi07,WLYou07,HQZhou07,LCVenuti07,SJGu07,SChen07,WQNing07,MFYang07,NPaunkovic07},
109: a concept introduced in quantum information theory \cite{nilesen}.
110: Mathematically it is the overlap between two ground states in
111: which their driving parameters deviate by a small amount.
112: However, the fidelity depends computationally on an arbitrarily
113: small yet finite change of the driving parameter. For this,
114: Zanardi \emph{et. al.} introduced the Riemannian metric tensor
115: \cite{PZanardi07}, while You \emph{et. al.} suggested the
116: fidelity susceptibility \cite{WLYou07}, both focus on the leading
117: term of the fidelity, in order to explain singularities in QPTs.
118: In addition, scaling analysis of these quantities has been
119: informative: it helps understanding their divergence and the
120: criticality of the system \cite{LCVenuti07}, and it also reveals
121: the intrinsic relation between the critical exponent of some
122: physical quantities and that of the fidelity susceptibility
123: \cite{SJGu07}.
124: 
125: In this paper, we explicitly compute the ground-state fidelity susceptibility
126: and its critical exponent of the LMG model. Numerical analysis
127: is also performed to check with our analytic calculations. We show that, the
128: $1/N$ expansion in the Holstein-Primakoff transformation is sufficient to determine
129: the critical exponent of the fidelity susceptibility $\chi_{_F}$. In addition, we
130: revealed two distinct critical exponents in two phases which is not a general
131: feature. Therefore, our findings not only suggest another route on understanding
132: the quantum criticality of the LMG model, but also show the fidelity susceptibility
133: is not always extensive in describing the universality class of a quantum many-body
134: system.
135: 
136: This paper consists of five sections. In Sec. \ref{sec:Ham}, we review the
137: Hamiltonian, symmetry, and conserved quantities of the LMG model. In Sec.
138: \ref{sec:criticalfs}, we diagonalize the model Hamiltonian and compute the
139: fidelity susceptibility in the anisotropic model. In Sec. \ref{sec:scalingana},
140: we perform finite size scaling analysis and discuss the
141: scaling relation between different exponents. Finally, we give a brief summary
142: in Sec. \ref{sec:sum}.
143: 
144: \section{The model Hamiltonian}
145: \label{sec:Ham}
146: 
147: The Hamiltonian of the LMG model reads
148: \begin{eqnarray}
149:  H &=&  - \frac{\lambda }{N}\sum\limits_{i < j} {\left( {\sigma _x^i \sigma _x^j
150:  + \gamma \sigma _y^i \sigma _y^j } \right)}  - h\sum\limits_i {\sigma _z^i
151:  },
152:  \\
153:  &=&  - \frac{{2\lambda }}{N}\left( {S_x^2  + \gamma S_y^2 } \right) - 2hS_z
154:  + \frac{\lambda }{2}\left( {1 + \gamma } \right),
155:  \\
156:  &=&  - \frac{\lambda }{N}\left( {1 + \gamma } \right)\left( {\textbf{S}^2  - S_z^2  - N/2} \right) - 2hS_z
157:  \nonumber \\ && - \frac{\lambda }{{2N}}\left( {1 - \gamma } \right)\left( {S_ + ^2  + S_ - ^2 } \right)
158:  , \label{eq:HLMG}
159: \end{eqnarray}
160: where $\sigma _{\kappa}\, (\kappa=x,y,z)$ are the Pauli matrices,
161: $S_\kappa = \sum _i \sigma _{\kappa}^i/2$, and $S_\pm  = S _x \pm
162: iS _y$. The prefactor $1/N$ is necessary to ensure finite energy
163: per spin in the thermodynamic limit. It is understood that the
164: total spin and the parity are the conserved quantities, i.e.,
165: \begin{eqnarray}
166: \left[ {H,S^2 } \right] = \left[ {H,\prod\limits_i {\sigma _z^i } }
167: \right] = 0  .
168: \end{eqnarray}
169: In addition, in the isotropic case $\gamma = 1$, one has $\left[ {H,S_z }
170: \right] = 0$ and simultaneous eigenstates can be found. In the main context,
171: the following parameter space is considered: $\lambda = 1, |\gamma| < 1, h
172: \geq 0$. We take $h \geq 0$ as the spectrum is invariant under the transformation
173: $h\leftrightarrow-h$. In addition, as a common practice we only consider the
174: maximum spin sector $S = N/2$ which contains the lowest energy state.
175: 
176: 
177: 
178: \section{Critical behavior of the fidelity susceptibility}
179: \label{sec:criticalfs}
180: 
181: We briefly review of the concept of the fidelity susceptibility
182: here. Suppose there is a Hamiltonian of a general form as
183: \begin{eqnarray}
184: H = H_0(\gamma) + f(h) H_I, \label{eq:generalHam}
185: \end{eqnarray}
186: for $H_I$ is defined as the driving term of the system, which
187: simply does not commute with $H_0$. The function $f(h)$ coupled
188: to $H_I$ is often considered as the linear external field
189: $f(h)=h$. Then the fidelity susceptibility is defined as
190: \cite{PZanardi07,WLYou07}
191: \begin{eqnarray}
192: \chi_{_F} =\left[\frac{d f(h)}{dh}\right]^2
193: \sum\limits_{n \ne 0} {\frac{{\left| {\left\langle {n} \right|H_I
194: \left| {0}  \right\rangle } \right|^2 }}{{\left[ {E_n  - E_0 }
195: \right]^2 }}}. \label{eq:definitionfs}
196: \end{eqnarray}
197: where $E_n$ and $|n\rangle$ stand for the $n^{\rm th}$
198: eigenenergies and eigenstates of the (whole) Hamiltonian
199: respectively.
200: 
201: The fidelity susceptibility is well-defined for a non-degenerate
202: ground state of the continuous variable $h$, but it is not suitable
203: to deal with states with good quantum numbers. The LMG model undergoes
204: ground state level crossing when $\gamma = 1$, the ground states are
205: assigned the magnetization as the quantum numbers.
206: 
207: We put our focus on the fidelity susceptibility for an arbitrary isotropy
208: $|\gamma| < 1$. One resolution is to use the Bethe-Ansatz solution
209: \cite{FPan99,JLinks03}, which is rather complicated. So we adopt the
210: $1/N$ expansion method which was used extensively by Dusuel and
211: Vidal \cite{CUT1,CUT2}, that corresponds to the large $N$ limit.
212: 
213: The $1/N$ expansion method is done under the Holstein-Primakoff
214: boson representation \cite{HP} framework. In low energy spectrum
215: the spin operators in the $S = N/2$ subspace are mapped into
216: boson operators:
217: \begin{eqnarray}
218:  S_z &=& S - a^{\dagger}a, \nonumber
219:  \\
220:  S_+ &=& (2S-a^{\dagger}a)^{1/2}a= N^{1/2}(1-a^{\dagger}a/N)^{1/2}a=S_-^{\dagger},
221: \end{eqnarray}
222: where $a$($a^{\dagger}$) is the standard bosonic annihilation
223: (creation) operator satisfying $[a,a^{\dagger}]=1$. The above
224: transformation is valid when $h\geq 1$, but when $0<h<1$ it can
225: also be used through semi-classical treatment \cite{CUT1,CUT2}.
226: This representation is also known as the spin-wave theory. It is
227: well adapted to the computation of the low-energy physics when
228: $\langle a^{\dagger}a\rangle/N \ll 1$. After inserting these
229: expressions of the spin operators in Eq. (\ref{eq:HLMG}),
230: one can approximate the square roots as one and express the
231: result in normal ordered form with respect to the boson vacuum
232: state. Keeping terms of order $(1/N)^{-1}$, $(1/N)^{-1/2}$ and
233: $(1/N)^{0}$ for $h\geq 1$ (in which the approximation is
234: justified), the Hamiltonian becomes
235: \begin{eqnarray}
236: H = -hN  + (2h-1+\gamma)a^{\dagger}a -\frac{1-\gamma}{2} \left(
237: a^{\dagger}{^2}+a^2 \right).
238: \end{eqnarray}
239: The above Hamiltonian can be diagonalized by a standard Bogoliubov
240: transformation
241: \begin{eqnarray}
242: a^{\dagger}&=&\cosh(\Theta/2)b^{\dagger}+\sinh(\Theta/2)b,\\
243: a &=& \sinh(\Theta/2)b^{\dagger}+\cosh(\Theta/2)b,
244: \end{eqnarray}
245: where $b (b^\dagger)$ is the quasi-bosonic annihilation (creation)
246: operator, and
247: \begin{eqnarray}
248: \tanh[\Theta(h\geq 1)]=\frac{1-\gamma}{2h-1+\gamma},
249: \end{eqnarray}
250: then the Hamiltonian is diagonalized as
251: \begin{eqnarray}
252: H = -h(N+1)+
253: 2\sqrt{(h-1)(h-\gamma)}\left(b^{\dagger}b+\frac{1}{2}\right).
254: \end{eqnarray}
255: Thus the low-energy spectrum of the model is mapped to the
256: spectrum of a simple harmonic oscillator. The eigenstates are just
257: $\{|n\rangle\}$, where $b^{\dagger}b|n\rangle=n|n\rangle$. We
258: consider the driving Hamiltonian $H_I$ responsible for the QPT,
259: \begin{eqnarray}
260: H_I = -\sum\limits_i{\sigma_z^i}&=&-2S_z.
261: \end{eqnarray}
262: By transforming them into combinations of $b$ and $b^\dagger$
263: operators, the fidelity susceptibility is calculated as
264: \begin{eqnarray}
265: \chi_{_F} = \frac{(1-\gamma)^2}{32(h-1)^2(h-\gamma)^2}. \label{eq:FShg1}
266: \end{eqnarray}
267: 
268: The derivation above is only valid for $h\geq1$, for $0<h<1$ the
269: calculation is actually similar to the above case of $h\geq1$,
270: provided that one first rotates the $z$ axis to bring it along
271: the classical spin direction. We do not show it explicitly here,
272: but interested readers are recommended to refer to Ref.
273: \cite{CUT1,CUT2}. We simply quote the main result, after all the
274: procedures the Hamiltonian becomes:
275: \begin{eqnarray}
276: H = -\frac{(1+h^2)}{2}N-\frac{1-\gamma}{2} +
277: 2\sqrt{(1-h^2)(1-\gamma)}\left(b^\dagger b + \frac{1}{2} \right).
278: \end{eqnarray}
279: The driving Hamiltonians also takes a different form:
280: \begin{eqnarray}
281: -\sum\limits_i{\sigma_z^i} &=& -2S_z \nonumber \\
282: &=& -2\left(-\sqrt{1-h^2}\widetilde{S_x}+h\widetilde{S_z}\right),
283: \end{eqnarray}
284: for the HP transformation is done on the $\widetilde{S}$ operators.
285: The fidelity susceptibilities are then obtained accordingly:
286: \begin{eqnarray}
287: \chi_{_F} &=&
288: \frac{N}{4\sqrt{(1-h^2)(1-\gamma)}}+\frac{h^2(h^2-\gamma)^2}{32(1-\gamma)^2(1-h^2)^2}.
289: \label{eq:FShl1}
290: \end{eqnarray}
291: 
292: Thus we obtained $\chi_{_F}$ of the anisotropic LMG model in large $N$ limit.
293: We first see the effect of isotropy to the fidelity susceptibility. It
294: dominates when $h<1$, but fades out for large $h$. Especially in the isotropic
295: limit, when $\gamma \to 1$, $\chi_{_F}$ diverges when $h<1$, but tends to zero when $h>1$.
296: This is the effect of the level-crossing points in the thermodynamic
297: limit. They together form a region of criticality, and the system undergoes
298: continuous level crossing. The fidelity susceptibility responds drastically while moving
299: along $h$. But when $h>1$, there are no further critical points,
300: $\chi_{_F}$ naturally measures zero when moving along $h$ because we have
301: $[H_0,H_I]=0$.
302: 
303: An interesting observation is $\chi_{_F}$ behaves extensively when $h<1$
304: even in the large $N$ limit. When discarding the extensive part of Eq.
305: (\ref{eq:FShl1}), we arrive a zero point at $h = \sqrt{\gamma}$, which does
306: not fit with numerical analysis [Fig. \ref{figure_fs_h05}]. This discrepancy
307: may be eliminated by adopting other transformations of the driving Hamiltonian.
308: Particularly, the flow of operators in the LMG model haven been studied by the
309: continuous unitary transformation (CUT) method \cite{CUT1,CUT2}.
310: However, such discrepancy would not hinder us from getting the correct critical exponent
311: of the fidelity susceptibility.
312: 
313: Let us emphasize the intensive property of the fidelity susceptibility, which
314: measures the average response to some driving Hamiltonians. Its divergence
315: should correspond to a critical point of a second-order QPT rather than to the
316: increasing system size. In order to predict the critical exponent correctly,
317: we should average the fidelity susceptibility whenever necessary. To the leading
318: order, Eq. (\ref{eq:FShl1}) becomes
319: \begin{eqnarray}
320: \frac{\chi_{_F}}{N} = \frac{1}{4\sqrt{(1-h^2)(1-\gamma)}}.
321: \end{eqnarray}
322: Then it comes to a key result of our paper: $\chi_{_F}$ bears
323: different critical exponents across the critical point. It diverges as
324: $(1-h)^\frac{1}{2}$ when $h<1$, $(h-1)^2$ when $h>1$. It is unlike
325: the Ising model in a transverse field \cite{PZanardi06} nor the one-dimensional
326: asymmetric Hubbard model \cite{SJGu07}, where the critical exponent is a
327: single number over the phases.
328: 
329: \begin{figure}
330: \includegraphics[width=7cm]{fs_h05}
331: \caption{(color online) The fidelity susceptibility in response to $h$ as a
332: function of $h$ at $\gamma = 0.5$. The inset denotes the difference between the
333: fidelity susceptibility and the extensive term in Eq. (\ref{eq:FShl1}).}
334: \label{figure_fs_h05}
335: \end{figure}
336: 
337: \begin{figure*}
338: \includegraphics[width=5.5cm]{fs_scaling_a05}\hspace{3mm}
339: \includegraphics[width=5.5cm]{fs_scaling_a10}\hspace{3mm}
340: \includegraphics[width=5.5cm]{fs_scaling_a15}
341: \caption{(color online) The finite size scaling analysis is
342: performed for the case of power-law divergence at $\gamma = 0.5$
343: (LEFT), $\gamma = 0$ (MIDDLE) and $\gamma = -0.5$ (RIGHT) for system
344: sizes $N=2^{n} (n=12, 13, 14, 15, 16)$. The fidelity
345: susceptibility is considered as a function of system size and
346: driving parameter is a function of $N^\nu (h-h_{\rm max})$ only,
347: with the correlation length critical exponent $\nu\simeq 0.665$.} \label{figure_3}
348: \end{figure*}
349: 
350: \begin{figure}
351: \includegraphics[width=7cm]{max_fs}
352: \caption{(color online) The finite size scaling is performed for
353: the maximum of the fidelity susceptibility.} \label{figure_4}
354: \end{figure}
355: 
356: 
357: \section{Finite size scaling analysis}
358: \label{sec:scalingana}
359: 
360: To illustrate the scaling behavior of the fidelity
361: susceptibility, we perform the exact diagonalization (ED) to
362: solve the spectrum of $H$ and then calculate the corresponding
363: fidelity susceptibility numerically.
364: 
365: Let us recall the fidelity susceptibility scaling analysis performed in
366: the asymmetric Hubbard model \cite{SJGu07}. According to the scaling ansatz
367: \cite{MAContinentinob} and the obvious power-law divergence observed in Fig.
368: \ref{figure_fs_h05}, the rescaled fidelity susceptibility
369: around its maximum point at $h_{\rm max}$ is a simple function of a scaling variable,
370: i.e.
371: \begin{eqnarray}
372: \frac{\chi_{_F{\rm max}} - \chi_{_F}}{\chi_{_F}} = f[N^\nu (h-h_{\rm max})],
373: \label{eq:scalingansatz}
374: \end{eqnarray}
375: where $f(x)$ is the scaling function and $\nu$ is the correlation
376: length critical exponent. This function is universal and does not
377: depend on the system size, as shown in Fig. \ref{figure_3} for
378: cases of $\gamma = 0.5, 0$ and $\gamma = -0.5$. Remarkably, the critical
379: exponent $\nu$ for three cases are very close. This observation
380: strongly implies that $\nu$ is a universal constant and does not
381: depend on the parameters $\gamma$ and $h$.
382: 
383: 
384: \begin{table*}
385: \caption{Scaling exponent $\mu$ at various $\gamma$, obtained by sampling
386:  system size in different range.} \label{tab:critcalexp}
387: \begin{center}
388: \begin{ruledtabular}
389: \begin{tabular}{ccccccc}
390:  $\gamma$& 0.8 & 0.5 & 0.2 & 0 & -0.2 & -0.5 \\
391: \hline $\mu (N\in[2^8, 2^{16}])$ & $1.3221\pm 0.0006$ & $1.3264\pm 0.0004$ &
392: $1.3267\pm 0.0004$ & $1.3280\pm 0.0004$
393: & $1.3283\pm 0.0003$ & $1.3285 \pm 0.0003$ \\
394: \hline $\mu (N\in[2^{12}, 2^{16}])$ & $1.3250\pm 0.0003$ & $1.3285\pm 0.0004$ &
395: $1.3295\pm 0.0002$ & $1.3299\pm 0.0002$
396: & $1.3302\pm 0.0001$ & $1.3304 \pm 0.0001$\\
397: \end{tabular}
398: \end{ruledtabular}
399: \end{center}
400: \end{table*}
401: 
402: 
403: In recent studies of the fidelity susceptibility in critical
404: phenomena, it was pointed out that the intensive fidelity
405: susceptibility scales generally like \cite{LCVenuti07,SJGu07}
406: \begin{eqnarray}
407: \chi_{_F} \propto \frac{1}{|h-h_c|^\alpha},
408: \end{eqnarray}
409: around the critical point. In the last section, we have already obtained
410: \begin{eqnarray}
411: \alpha  = \left\{ \begin{array}{ccc}
412:  2,& &h > 1 \\
413:  \frac{1}{2}, & & 0 \le h < 1 \\
414:  \end{array} \right. ,
415: \end{eqnarray}
416: which is also a universal constant. Then if the maximum point of
417: the intensive fidelity susceptibility scales like
418: \begin{eqnarray}
419: \chi_{_F{\rm max}} \propto N^\mu,
420: \end{eqnarray}
421: the scaling ansatz also implies another important relation, i.e.
422: \begin{eqnarray}
423: \alpha=\frac{\mu}{\nu} . \label{eq:scalingequality}
424: \end{eqnarray}
425: 
426: We try to confirm this equality in numerically. In Fig. \ref{figure_3}, Eq.
427: (\ref{eq:scalingansatz}) is best fitted with $\nu\simeq 0.665$. The case to
428: determine $\mu$ is more subtle. It is because Eq. (\ref{eq:scalingansatz})
429: remains the same form even for averaged $\chi_{_F}$, but the maximum of
430: $\chi_{_F}$ does not. To resolve this problem, we first determine $\mu$
431: from the ``bare" $\chi_{_F}$. By using least square fit method, we
432: evaluated ``bare" $\mu$ for at different $\gamma$. The numerical details are
433: shown in table \ref{tab:critcalexp}. However, the exponent $\mu$ does not
434: converge perfectly. We compare the $\mu$ obtained in a range of $[2^{12},
435: 2^{16}]$, and those from the range $[2^{8}, 2^{16}]$. The results converge
436: better for larger scaling regions. According to the trend of $\mu$ in larger
437: system sizes, we roughly estimate $\mu=1.33$ with three effective digits [Fig.
438: \ref{figure_4}].
439: 
440: When $h>1$, $\chi_{_F}$ is observed to be intensive [Fig.
441: \ref{figure_fs_h05}]. With the estimated $\mu$ and $\nu$, the equality
442: (\ref{eq:scalingequality}) is satisfied with $\alpha=2$. On the other hand,
443: when $h<1$, $\chi_{_F}/N$ is the intensive quantity. For $\chi_{_F}
444: \propto N^\mu$,
445: \begin{eqnarray}
446: \frac{\chi_{_F}}{N} \propto N^{(\mu-1)}.
447: \end{eqnarray}
448: Thus $\mu \simeq 0.33$, this will give the relation $\alpha =
449: 1/2$. These two values of $\alpha$ are consistent with our analytic
450: calculation in the last section.
451: 
452: The exponent $\mu$, $\nu$ can also be discussed from the scaling
453: ansatz at the critical point rather than the maximum point of a
454: finite system, as shown by Vidal, Dusuel, and Barthel
455: \cite{CUT2,JVidalStat}. Based on their approach, the critical
456: exponent $\nu$ takes the value of 1/3, and is independent of magnitude
457: of $\gamma$. Our results on the maximum simply agree with this
458: value and can be generalized to other models where the precise critical
459: point is not known.
460: 
461: Another scaling analysis is to examine how $h_{\rm max}$ tends to
462: the critical point $h_c=1$. It should scale like
463: \begin{eqnarray}
464: h_c - h_{\rm max} \propto N^{\,-\delta},
465: \end{eqnarray}
466: in the large $N$ limit. We find $\delta \simeq 0.66$ with two
467: effective digits for various $\gamma$.
468: 
469: In short, we can confirm that the exponents $\mu, \nu$, and
470: $\delta$ of the fidelity susceptibility do not depend on the
471: value of $\gamma$ and $h$. They are universal constants for the LMG
472: model and are related to the critical exponent of the fidelity susceptibility $\alpha$.
473: 
474: 
475: 
476: 
477: \section{Summary}
478: \label{sec:sum}
479: 
480: In summary, we computed explicitly the fidelity susceptibility and its
481: critical exponent of the LMG model at different isotropy.
482: We confirmed the different critical exponents in two phases numerically by
483: ED, which is a rather non-trivial result. Several scaling exponents are
484: also found in consistence with previous studies. Since the fidelity susceptibility is believed to be
485: able to characterize the universality class of quantum phenomena,
486: our results therefore provide a rare explicit case for the study of
487: fidelity susceptibility.
488: 
489: \begin{acknowledgements}
490: 
491: We are very grateful to J. Vidal for many fruitful comments. S. J. G. thanks X.
492: Wang and J. P. Cao for helpful discussions. This work was partially supported
493: RGG Grant CUHK 400906, 401504, and MOE B06011.
494: 
495: %W.Q.N. would like to thank the support of the Graduated Students' Innovation
496: %Foundation of Fudan University.
497: 
498: \end{acknowledgements}
499: 
500: \begin{thebibliography}{99}
501: \bibitem{LMG}
502: H. J. Lipkin, N. Meshkov, and A. J. Glick, Nucl. Phys. \textbf{62}, 188 (1965).
503: 
504: \bibitem{infcoor}
505: R. Botet and R. Jullien, Phys. Rev. Lett. \textbf{49}, 478 (1982); R. Botet and R. Jullien, Phys. Rev. B \textbf{28}, 3955 (1983).
506: 
507: \bibitem{HP}
508: T. Holstein and H. Primakoff, Phys. Rev. \textbf{58}, 1098 (1940).
509: 
510: \bibitem{CUT1}
511: S. Dusuel, and J. Vidal, Phys. Rev. Lett. \textbf{93}, 237204 (2004).
512: 
513: \bibitem{CUT2}
514: S. Dusuel, and J. Vidal, Phys. Rev. B \textbf{71}, 224420 (2005).
515: 
516: \bibitem{JVidalStat}
517: J. Vidal, S. Dusuel, and T. Barthel, J. Stat. Mech. P01015 (2007).
518: 
519: \bibitem{JVidalSpectra}
520: P. Ribeiro, J. Vidal, and R. Mosseri, Phys. Rev. Lett. {\bf 99}, 050402 (2007).
521: 
522: \bibitem{QPT1st}
523: J. Vidal, R. Mosseri and J. Dukelsky, Phys. Rev. A \textbf{69}, 054101 (2004).
524: 
525: \bibitem{QPT2nd}
526: J. Vidal, G. Palacios, and R. Mosseri, Phys. Rev. A \textbf{69}, 022107 (2004).
527: 
528: \bibitem{EntE1}
529: J. I. Latorre,
530: R. Or\'{u}s, E. Rico, and J. Vidal, Phys. Rev. A \textbf{71}, 064101 (2005).
531: 
532: \bibitem{EntE2}
533: T. Barthel, S. Dusuel, and J. Vidal, Phys. Rev. Lett. \textbf{97}, 220402
534: (2006).
535: 
536: \bibitem{nilesen} M. A. Nilesen and I. L. Chuang, \emph{Quantum Computation
537: and Quantum Information} (Cambridge University Press, Cambridge,
538: England, 2000)
539: 
540: \bibitem{sachdev} S. Sachdev, \emph{Quantum Phase Transitions} (Cambridge
541: University Press, Cambridge, England, 1999).
542: 
543: 
544: \bibitem{HTQuan06}
545: H. T. Quan, Z. Song, X. F. Liu, P. Zanardi, and C. P. Sun, Phys.
546: Rev. Lett. \textbf{96}, 140604 (2006).
547: 
548: \bibitem{PZanardi06}
549: P. Zanardi and N. Paunkovic, Phys. Rev. E \textbf{74}, 031123
550: (2006).
551: 
552: \bibitem{Buonsante07}
553: P. Buonsante and A. Vezzani, Phys. Rev. Lett. \textbf{98}, 110601
554: (2007).
555: 
556: \bibitem{PZanardi0606130}
557: P. Zanardi, M. Cozzini, and P. Giorda, J. Stat. Mech. \textbf{2}, L02002
558: (2007); M. Cozzini, P. Giorda, and P. Zanardi, Phys. Rev. B, \textbf{75},
559: 014439 (2007);  M. Cozzini, R. Ionicioiu, and P. Zanardi, \emph{ibid.}
560: \textbf{76}, 104420 (2007).
561: 
562: \bibitem{PZanardi07} P. Zanardi, P. Giorda, and M. Cozzini, Phys. Rev. Lett. {\bf
563: 99}, 100603 (2007).
564: 
565: \bibitem{WLYou07}
566: W. L. You, Y. W. Li, and S. J. Gu, Phys. Rev. E \textbf{76}, 022101 (2007).
567: 
568: \bibitem{HQZhou07}
569: H. Q. Zhou and J. P. Barjaktarevic, arXiv: cond-mat/0701608; H. Q. Zhou, J. H.
570: Zhao, and B. Li, arXiv:0704.2940; H. Q. Zhou, arXiv:0704.2945.
571: 
572: \bibitem{LCVenuti07}
573: L. C. Venuti and P. Zanardi, Phys. Rev. Lett. \textbf{99}, 095701
574: (2007).
575: 
576: \bibitem{SJGu07}
577: S. J. Gu, H. M. Kwok, W. Q. Ning, and H. Q. Lin, Phys. Rev. B \textbf{77}, 245109 (2008).
578: 
579: \bibitem{SChen07}
580: S. Chen, L. Wang, S. J. Gu, and Y. Wang, Phys. Rev. E {\bf 76} 061108 (2007).
581: 
582: 
583: \bibitem{WQNing07}
584: W. Q. Ning, S. J. Gu, C. Q. Wu, and H. Q. Lin, arXiv:0708.3178.
585: 
586: \bibitem{MFYang07}
587: M. F. Yang, Phys. Rev. B \textbf{76}, 180403 (R) (2007);  Y. C. Tzeng and M. F.
588: Yang, Phys. Rev. A {\bf 77}, 012311 (2008).
589: 
590: \bibitem{NPaunkovic07}
591: N. Paunkovic, P. D. Sacramento, P. Nogueira, V. R. Vieira, and V. K.
592:  Dugaev, arXiv:0708.3494.
593: 
594: 
595: \bibitem{FPan99}
596: F. Pan and J. P. Draayer, Phys. Lett. B \textbf{451}, 1 (1999).
597: 
598: \bibitem{JLinks03}
599: J. Links, H. Q. Zhou, R. H. McKenzie, and M. D. Gould, J. Phys. A \textbf{36},
600: R63 (2003).
601: 
602: \bibitem{MAContinentinob}
603: M. A. Continentino, {\it Quantum Scaling in Many-Body Systems} (World
604: Scientific Publishing, Singapore, 2001).
605: 
606: \end{thebibliography}
607: 
608: \end{document}
609: