0711.4651/rg.tex
1: \documentclass[twocolumn,showpacs,floatfix,superscriptaddress,amsmath,amssymb,prl]{revtex4}
2: \usepackage{txfonts}
3: \usepackage{amssymb}
4: \usepackage{graphicx}
5: \usepackage{hyperref}
6: \usepackage{ulem}
7:  \usepackage{overpic}
8:  \usepackage{psfrag}
9: \newcommand{\be}{\begin{equation}}
10: \newcommand{\ee}{\end{equation}}
11: 
12: \newcommand{\added}[1]{\textsf{#1}}
13: 
14: \begin{document}
15: 
16: 
17: \title{Singularities in fidelity surfaces for quantum phase transitions: a geometric perspective}
18: 
19: \author{Huan-Qiang Zhou}
20: \affiliation{Centre for Modern Physics and Department of Physics,
21: Chongqing University, Chongqing 400044, The People's Republic of
22: China}
23: \author{Jian-Hui Zhao}
24: \affiliation{Centre for Modern Physics and Department of Physics,
25: Chongqing University, Chongqing 400044, The People's Republic of
26: China}
27: \author{Hong-Lei Wang}
28: \affiliation{Centre for Modern Physics and Department of Physics,
29: Chongqing University, Chongqing 400044, The People's Republic of
30: China}
31: \author{Bo Li}
32: \affiliation{Centre for Modern Physics
33: and Department of Physics, Chongqing University, Chongqing 400044,
34: The People's Republic of China}
35: \begin{abstract}
36: The fidelity per site between two ground states of a quantum lattice
37: system corresponding to different values of the control parameter
38: defines a surface embedded in a Euclidean space. The Gaussian
39: curvature naturally quantifies quantum fluctuations that destroy
40: orders at transition points. It turns out that quantum fluctuations
41: wildly distort the fidelity surface near the transition points, at
42: which the Gaussian curvature is singular in the thermodynamic limit.
43: As a concrete example, the one-dimensional quantum Ising model in a
44: transverse field is analyzed.  We also perform a finite size scaling
45: analysis for the transverse Ising model of finite sizes. The scaling
46: behavior for the Gaussian curvature is numerically checked and the
47: correlation length critical exponent is extracted, which is
48: consistent with the conformal invariance at the critical point.
49: 
50: \end{abstract}
51: \pacs{03.67.-a, 05.70.Fh, 64.60.Ak}
52: 
53: \date{\today}
54: \maketitle
55: 
56: Quantum phase transitions (QPTs) have been a research topic subject
57: to intense study,  since their significant role was realized in
58: accounting for high-$T_c$ superconductors, fractional quantum Hall
59: liquids, and quantum magnets~\cite{sachdev,wen}. Recently,
60: significant advances have been made in attempt to clarify the
61: connection between quantum many-body physics and quantum information
62: science. This provides a new perspective to investigate QPTs from
63: \textit{entanglement}~\cite{osborne,vidal,levin,entanglement1,entanglement2}
64: and \textit{fidelity}~\cite{zanardi,zjp,zhou,zov,more}, basic
65: notions of quantum information science~\cite{nielsen} and turns out
66: to be very insightful in our understanding of QPTs in a variety of
67: quantum lattice systems in condensed matter.
68: 
69: Conventionally, orders and fluctuations provide a proper language to
70: study QPTs, with order parameters being the key to quantify quantum
71: fluctuations. Instead, the fidelity approach is based on state
72: distinguishability arising from the orthogonality of different
73: ground states in the thermodynamic limit. In fact, the ground state
74: fidelity for a quantum system may be mapped onto the partition
75: function of the equivalent classical statistical lattice model with
76: the same geometry~\cite{zov}. Thus, the fidelity per site is
77: well-defined in the thermodynamic limit, and its singularities
78: unveil transition points, at which the system under consideration
79: undergoes QPTs. Therefore, a practical means is now available to map
80: out the ground state phase diagram for a quantum lattice system
81: without prior knowledge of order parameters. An intriguing question
82: is how to characterize singularities in the fidelity per site.
83: Indeed, a proper answer to this question will shed new light on our
84: understanding of QPTs.
85: 
86: 
87: In this paper, we present an \textit{intrinsic} characterization of
88: singularities in the fidelity per site in terms of Riemannian
89: geometry. For this purpose, we first \textit{define} a fidelity
90: surface as a surface embedded in a Euclidean space,  which in turn
91: is determined by the average fidelity per lattice site between two
92: ground states of a quantum lattice system as a function of the
93: control parameters. This makes the whole machinery developed in
94: differential geometry of curves and surfaces available to study
95: QPTs. As it is well known, the Gaussian curvature, or equivalently,
96: the Ricci scalar curvature for the surfaces embedded in Euclidean
97: spaces, is a fundamental concept used to measure how curved a
98: surface is. Therefore, the Gaussian curvature is expected to
99: naturally quantifies quantum fluctuations that destroy orders at
100: transition points.  We discuss the global behaviors of the Gaussian
101: curvature. It turns out that quantum fluctuations wildly distort the
102: fidelity surfaces near the transition points. Generically,
103: precursors of QPTs occur in the Gaussian curvature for finite-size
104: systems. In the thermodynamic limit, the Gaussian curvature becomes
105: singular at transition points. The one-dimensional quantum Ising
106: model in a transverse field is exploited to explicitly illustrate
107: the theory. We also perform a finite size scaling analysis for the
108: Gaussian curvature with different lattice sizes to extract the
109: correlation length critical exponent.
110: 
111: {\it Fidelity surfaces.} For a quantum lattice system described by a
112: Hamiltonian $H(\lambda)$, with $\lambda$ a control parameter. Here
113: we restrict ourselves to discuss the simplest case with one single
114: control parameter, although the extension to multiple control
115: parameters is straightforward. For two ground states
116: $|\psi(\lambda_1)\rangle$ and $|\psi(\lambda_2)\rangle$
117: corresponding to different values of the control parameter
118: $\lambda$, the fidelity is defined as $F(\lambda_1, \lambda_2)\equiv
119: |\langle\psi(\lambda_2)|\psi(\lambda_1)\rangle|$. For a large but
120: finite $L$, the fidelity $F$ asymptotically scales as $F(\lambda_1,
121: \lambda_2) \sim d^L(\lambda_1, \lambda_2)$, where the scaling
122: parameter $d(\lambda_1, \lambda_2)$ characterizes how fast the
123: fidelity changes when the thermodynamic limit is
124: approached~\cite{zhou}. Physically, it is the fidelity per site.
125: Here note that the contribution from each site to $F(\lambda_1,
126: \lambda_2)$ is multiplicative. Following~\cite{zov}, the ground
127: state fidelity for a quantum system is nothing but the partition
128: function of the equivalent classical statistical lattice model with
129: the same geometry, if one utilizes the tensor network
130: representations of ground state many-body wave functions. Therefore,
131: $d(\lambda_1,\lambda_2)$ may be interpreted as the partition
132: function per site~\cite{baxter}, which is well-defined in the
133: thermodynamic limit:
134: \begin{equation} \ln d(\lambda_1,\lambda_2) = \lim_{L
135: \rightarrow \infty} \ln F(\lambda_1,\lambda_2) /L.
136: \end{equation}
137: The fidelity per site $d(\lambda_1,\lambda_2)$ satisfies the
138: properties: (1) symmetry under interchange $\lambda_1
139: \longleftrightarrow \lambda_2$; (2) $d(\lambda_1,\lambda_1)=1$; and
140: (3) $0\leq d(\lambda_1,\lambda_2) \leq 1$.
141: 
142: For simplicity, let us assume that the system undergoes a QPT at
143: $\lambda_c$. If $|\psi(\lambda_1)\rangle$ and
144: $|\psi(\lambda_2)\rangle$ are in the same phase, then they flow to
145: the same stable fixed point in the sense of renormalization group
146: theory, and so their difference arises from quantum fluctuations
147: depending on the details of the system. On the other hand, if
148: $|\psi(\lambda_1)\rangle$ and $|\psi(\lambda_2)\rangle$ are in
149: different phases, then they flow to two different stable fixed
150: points. Therefore, they possess different orders, although quantum
151: fluctuations originate from the same unstable fixed point
152: $\lambda_c$~\cite{relevant}. Imagine that if there were no quantum
153: fluctuations, then $d(\lambda_1,\lambda_2)$ would be simply 1 when
154: $|\psi(\lambda_1)\rangle$ and $|\psi(\lambda_2)\rangle$ are in the
155: same phase; otherwise, when $|\psi(\lambda_1)\rangle$ and
156: $|\psi(\lambda_2)\rangle$ are in different phases,
157: $d(\lambda_1,\lambda_2)$ would take the minimum value corresponding
158: to the two stable fixed points. For continuous QPTs, quantum
159: fluctuations are strong enough such that no orders survive at the
160: transition point, so $d(\lambda_1,\lambda_2)$ is continuous, but
161: displays singularities, whereas for the first order QPTs,
162: $d(\lambda_1,\lambda_2)$ remains to be discontinuous at the
163: transition point.  An interesting observation is to regard the
164: fidelity per site, $d(\lambda_1,\lambda_2)$, as a two-dimensional
165: surface embedded in the three-dimensional Euclidean space, with a
166: Riemannian metric induced from the Euclidean metric. Our aim is to
167: give an \textit{intrinsic} characterization of singularities in such
168: a fidelity surface in terms of Riemannian geometry.
169: 
170: {\it Differential geometry of the two-dimensional surfaces embedded
171: in the three-dimensional Euclidean space.} Let us briefly recall the
172: fundamentals of differential geometry of surfaces embedded in
173: Euclidean spaces~\cite{novikov}. For a two-dimensional surface
174: embedded in a three-dimensional Euclidean space:
175: $z=f(\lambda_1,\lambda_2)$, the first fundamental form on the
176: surface is
177: \begin{equation}
178: dl^{2}=g_{ij}d\lambda^{i}d\lambda^{j}=E(du)^{2}+2F(du
179: dv)+G(dv)^{2}\label{frist},\end{equation} where $g_{ij}$ is the
180: Riemannian metric on the surface: $g_{11}=1+f_{\lambda_1}^{2}$,
181: $g_{12}=g_{21}=f_{\lambda_1}f_{\lambda_2}$, and
182: $g_{22}=1+f_{\lambda_2}^{2}$. Here the subscripts $\lambda_1$ and
183: $\lambda_2$ denote partial differentiations with respect to
184: $\lambda_1$ and $\lambda_2$, respectively. In terms of the
185: co-ordinates $u=\lambda_1$ and $v=\lambda_2$, we have $E=g_{11}$,
186: $F=g_{12}=g_{21}$ and $G=g_{22}$. Suppose the surface is given in
187: parametric form: $r=r(u, v)$. Then, the vector product $r_{u} \times
188: r_{v}$ is a non-zero vector perpendicular to the surface at each
189: non-singular point; define $m$ to be a unit vector in the normal
190: direction, then one has $r_{u} \times r_{v} =|r_{u} \times r_{v}|m$.
191: For a curve $r=r(u(l),v(l))$ on the surface, the projection of the
192: second order derivative $\ddot{r}$ of $r$ with respect to the arc
193: length $l$ on the normal to the surface leads to the second
194: fundamental form as follows
195: \begin{equation}
196: \langle \ddot{r},m \rangle
197: (dl)^2=b_{ij}d\lambda^{i}d\lambda^{j}=X(du)^{2}+2Y du dv+Z (dv)^{2},
198: \label{second}
199: \end{equation}
200: if a surface is given in the form $z=f(\lambda_1,\lambda_2)$ with
201: $\lambda_1=u$, $\lambda_2=v$, and $r(u,v)=(u,v,f(u,v))$. Therefore,
202: we have $X=b_{11}=f_{\lambda_1\lambda_1}/ \sqrt
203: {1+f_{\lambda_1}^{2}+f_{\lambda_2}^{2}}$,
204: $Y=b_{12}=b_{21}=f_{\lambda_1\lambda_2}/ \sqrt
205: {1+f_{\lambda_1}^{2}+f_{\lambda_2}^{2}}$ and
206: $Z=b_{22}=f_{\lambda_2\lambda_2}/ \sqrt
207: {1+f_{\lambda_1}^{2}+f_{\lambda_2}^{2}}$.
208: 
209: 
210: The eigenvalues of the pair of quadratic forms (\ref{frist}) and
211: (\ref{second}) are the {\it principal curvatures} of the surface at
212: the point under investigation. The product of the principal
213: curvatures is the {\it Gaussian curvature} $K$ of the surface at the
214: point, and their sum the {\it mean curvature}. The principal
215: curvatures $k_{1}$ and $k_{2}$ are the solutions of equation:
216: \begin{equation}
217: {\rm det}(Q-k G)=0 \label{det}, \end{equation} where $Q=(b_{ij})$ is
218: the matrix of the second fundamental form, and $G=(g_{ij})$. Since
219: the first fundamental form is positive definite, its matrix $G$ is
220: non-singular. Hence ${\rm det}(Q-k G)={\rm det} G \; {\rm
221: det}(G^{-1}Q - k \cdot I)$, we deduce that the Gaussian curvature
222: $K=k_{1} k_{2}={\rm det}(G^{-1}Q)={\rm det}Q / {\rm det}G$ and the
223: mean curvature $M=k_{1}+k_{2}={\rm tr}(G^{-1} Q)$. Therefore, the
224: Gaussian curvature $K$ and the mean curvature $M$ take the form:
225: \begin{equation} K=\frac {f_{\lambda_1\lambda_1} f_{\lambda_2\lambda_2}-
226: f_ {\lambda_1 \lambda_2}^{2}}{\left(1 + f_{\lambda_1}^{2} +
227: f_{\lambda_2}^{2}\right)^{2}}, \label{gaussian curvature}
228: \end{equation}
229: and
230: \begin{equation} M=\frac {(1+f_{\lambda_2}^{2})f_{\lambda_1\lambda_1}+ (1+f_{\lambda_1}^{2})f_{\lambda_2\lambda_2}-
231: 2f_{\lambda_1}f_{\lambda_2}f_ {\lambda_1 \lambda_2}^{2}}{\left(1 +
232: f_{\lambda_1}^{2} + f_{\lambda_2}^{2}\right)^{\frac{3}{2}}},
233: \label{mean curvature}
234: \end{equation}
235: respectively. We notice that the sign of the {\it Gaussian
236: curvature} $K$ is the same as the sign of the determinant:
237: $f_{\lambda_1 \lambda_1} f_{\lambda_2\lambda_2}- f_
238: {\lambda_1\lambda_2}^{2}$, i.e., the {\it Hessian} of $z=
239: f(\lambda_1,\lambda_2)$
240: 
241: It follows that, in contrast with the mean curvature $M$, the
242: Gaussian curvature $K$ of a surface may be expressed in terms of the
243: induced metric on the surface alone, and is therefore an intrinsic
244: invariant of the surface~\cite{novikov}. In addition, a
245: two-dimensional surface in a three-dimensional space may also be
246: regarded as a differentiable manifold endowed with a Riemannian
247: metric induced from the Euclidean metric. The Ricci scalar curvature
248: $R$ is twice the Gaussian curvature $K$: $R=2K$.
249: 
250: {\it Global behaviors of the Gaussian curvature $K$ for a fidelity
251: surface.} Now we consider the (logarithmic function of) fidelity per
252: site, $\ln d(\lambda_1,\lambda_2)$, as a two-dimensional surface
253: embedded in the three-dimensional Euclidean space:
254: $z=f(\lambda_1,\lambda_2) \equiv \ln d(\lambda_1,\lambda_2)$.  The
255: Gaussian curvature $K(\lambda_1,\lambda_2)$ for such a fidelity
256: surface may be used to quantify how strong quantum fluctuations are
257: in given quantum many-body ground states, thus providing an
258: intrinsic characterization of singularities in the fidelity surface.
259: Indeed, as justified in Refs.~\cite{zjp,zhou,zov}, the fidelity per
260: site $d(\lambda_1,\lambda_2)$  is singular when $\lambda_1
261: (\lambda_2)$ crosses $\lambda_c$ for a fixed $\lambda_2 (\lambda_1)$
262: in the thermodynamic limit. Therefore the Gaussian curvature
263: $K(\lambda_1,\lambda_2)$ for the fidelity surface is singular at
264: $\lambda_1=\lambda_c$ and/or $\lambda_2=\lambda_c$ in the
265: thermodynamic limit. Generically, we have: $(1)$
266: $K(\lambda_1,\lambda_2)>0$, there is a neighborhood of the point
267: throughout which the surface lies on one sides of the tangent plane
268: at the points; $(2)$ $K(\lambda_1,\lambda_2)<0$, then the surface
269: intersects the tangent plane at the point arbitrarily close to the
270: point. If the surface is strictly convex, then we say that the
271: Gaussian curvature $K(\lambda_1,\lambda_2)$ is positive at every
272: point of the surface. That is what happens if $\lambda_1$ and
273: $\lambda_2$ are away from the transition point.  However, if
274: $\lambda_1$ and $\lambda_2$ are close to the transition point, then
275: the Gaussian curvature $K(\lambda_1,\lambda_2)$ can be negative.
276: 
277: For finite-size systems, the Gaussian curvature
278: $K(\lambda_1,\lambda_2)$ remains to be smooth, although the
279: precursors of QPTs occur as anomalies in the Gaussian curvature
280: $K(\lambda_1,\lambda_2)$. The anomalies get more pronounced when the
281: thermodynamic limit is approached. We may take advantage of this
282: fact to perform finite size scaling to extract the correlation
283: length critical exponent.
284: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
285: %%1
286: \begin{figure}[ht]
287: \begin{overpic}[width=72mm,totalheight=42mm]{fig1a.eps}
288: \put(30,12){$(a)$}
289: \end{overpic}
290: \hspace{0in}
291: \begin{overpic}[width=72mm,totalheight=42mm]{fig1b.eps}
292: \put(30,12){$(b)$}
293: \end{overpic}
294:   \caption{(color online) The behavior near the critical
295:   point $\lambda_c=1$ is analyzed for the Gaussian curvature $K(\lambda_1,\lambda_2)$ of the quantum transverse Ising model
296:   for various lattice sizes.
297:   The curves shown correspond to different lattice sizes $L=201,401,1201,2001$, and $\infty$.
298:   The peaks (dips) get more pronounced in the left (right) side with increasing system size.
299:   The Gaussian curvature
300: $K(\lambda_1,\lambda_2)$ diverges at the critical point $\lambda_1 =
301: \lambda_c$ for the infinite-size system ($L= \infty$). Upper panel:
302: Here  $K(\lambda_1,\lambda_2)$ is regarded as a function of
303: $\lambda_1$ for $\lambda_2 =0.6$ and $\gamma =1$. Lower panel: Here
304: $K(\lambda_1,\lambda_2)$ is regarded as a function of $\lambda_1$
305: for $\lambda_2 =0.6$ and $\gamma =1/2$.} \label{fig1}
306: \end{figure}
307: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
308: 
309: {\it Quantum $XY$ spin 1/2 model.} The quantum $XY$ spin model is
310: described by the Hamiltonian
311: \begin{equation}
312: H= -\sum_{j=-M}^M \left( \frac {1+\gamma}{2} \sigma^x_j
313: \sigma^x_{j+1} + \frac {1-\gamma}{2} \sigma^y_j \sigma^y_{j+1}  +
314: \lambda \sigma^z_j \right). \label{HXY}
315: \end{equation}
316: Here $\sigma_j^x, \sigma_j^y$ and $\sigma_j^z$ are the Pauli
317: matrices at the $j$-th lattice site. The parameter $\gamma$ denotes
318: an anisotropy in the nearest-neighbor spin-spin interaction, whereas
319: $\lambda$ is an external magnetic field. The Hamiltonian (\ref{HXY})
320: may be exactly diagonalized~\cite{lieb,pfeuty} for any finite size
321: $L$ with $L=2M+1$. In the thermodynamic limit $L \rightarrow
322: \infty$, $\ln d(\lambda_1,\lambda_2)$ takes the form~\cite{zjp}:
323: \begin{equation}
324: \ln d(\lambda_1,\lambda_2) = \frac {1}{2\pi} \int ^\pi_0 d\alpha \ln
325: {\cal F} (\lambda_1,\lambda_2;\alpha),\label{dforising}
326: \end{equation}
327: where ${\cal F}(\lambda_1,\lambda_2;\alpha)=\cos [\vartheta
328: (\lambda_1;\alpha)-\vartheta(\lambda_2;\alpha)]/2,$ with $\cos
329: \vartheta(\lambda;\alpha)=(\cos \alpha - \lambda)/\sqrt {(\cos
330: \alpha -\lambda)^2+\gamma^2 \sin^2 \alpha}$~\cite{finitefidelity}.
331: 
332: 
333: Now it is straightforward to calculate the Gaussian curvature
334: $K(\lambda_1,\lambda_2)$ for the fidelity surface of the quantum
335: $XY$ spin chain. In Fig.~\ref{fig1}, we plot
336: $K(\lambda_1,\lambda_2=0.6)$ for the fidelity surface of the quantum
337: $XY$ model ($\gamma=1$ for the upper panel and $\gamma=1/2$ for the
338: lower panel). One observes that $K(\lambda_1,\lambda_2=0.6)$ is
339: divergent as a function of $\lambda_1$ at the critical point
340: $\lambda_c=1$ for the infinite-size system $L=\infty$, indicating
341: that the fidelity surface is wildly distorted, due to strong quantum
342: fluctuations near the critical point. This is true for any nonzero
343: $\gamma$, consistent with the fact that the quantum $XY$ model for
344: any nonzero $\gamma$ belongs to the same universality class as the
345: quantum transverse Ising model. That is, there is a critical line
346: $\gamma\neq 0$ and $\lambda_c=1$; only one (second-order) critical
347: point $\lambda_c=1$ separates two gapful phases: (spin reversal)
348: $Z_2$ symmetry-breaking and symmetric phases.
349: 
350: 
351: 
352: 
353: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
354: %%2
355: \begin{figure}[ht]
356: \begin{overpic}[width=42mm,totalheight=42mm]{fig2a.eps}
357: \put(55,25){$(a)$}
358: \end{overpic}
359: \hspace{0in}
360: \begin{overpic}[width=40mm,totalheight=40mm]{fig2b.eps}
361: \put(55,25){$(b)$}
362: \end{overpic}
363: \caption{(color online) (a) The peaks values of the Gaussian
364: curvature $K(\lambda_1,\lambda_2)$ of the quantum transverse Ising
365: model for large lattice sizes scale as $L/(\ln L)^4$. (b) The dips
366: values of the Gaussian curvature $K(\lambda_1,\lambda_2)$ of the
367: quantum transverse Ising model for large lattice sizes scale as
368: $L/(\ln L)^4$. In both cases, $\lambda_2 =0.6$ and $\gamma =1$.}
369: \label{fig2}
370: \end{figure}
371: 
372: 
373: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
374: 
375: 
376: {\it Finite size scaling analysis for the Gaussian curvature $K$.}
377: We focus on the quantum Ising universality class. The order
378: parameter, i.e., magnetization $\langle \sigma ^x \rangle$ is
379: non-zero for $\lambda < 1$, and otherwise zero. At the critical
380: point, the correlation length $\xi \sim |\lambda -\lambda_c
381: |^{-\nu}$ with $\nu = 1$~\cite{pfeuty}. In order to analyze how the
382: Gaussian curvature $K(\lambda_1,\lambda_2)$ behaves near the
383: critical point $\lambda_c =1$, we perform a finite size scaling
384: analysis for the quantum transverse Ising model.
385: 
386: As already observed, the drastic change of the ground state wave
387: functions  makes the Gaussian curvature $K(\lambda_1,\lambda_2)$
388: divergent when the system undergoes the second order QPT at the
389: critical point $\lambda_c =1$ in the thermodynamic limit. However,
390: for finite-size systems, $K(\lambda_1,\lambda_2)$ remains to be
391: smooth for the quantum $XY$ model. In Fig.~{\ref{fig1}}, the
392: numerical results are also plotted for the Gaussian curvature
393: $K(\lambda_1,\lambda_2)$ with different system sizes, where
394: $\lambda_2=0.6$ and $\gamma=1$ (upper panel) and $\lambda_2=0.6$ and
395: $\gamma=1/2$ (lower panel). More precisely, in the thermodynamic
396: limit, $K(\lambda_1,\lambda_2)$ (as a function of $\lambda_1$ for a
397: fixed $\lambda_2$) diverges at the critical point
398: $\lambda_1=\lambda_c$:
399: \begin{equation}
400: K(\lambda_1,\lambda_2) \sim \frac{1}{|\lambda_1 -\lambda_c| (\ln
401: |\lambda_1 -\lambda_c|)^{4}} \label{infinite}.
402: \end{equation}
403: However, there is no divergence for finite-size systems, but there
404: are clear anomalies, featuring two quasi-critical values $\lambda_p$
405: and $\lambda_d$, one at each side of the critical point.   On the
406: left (right) side, the so-called quasi-critical points $\lambda_p$
407: ($\lambda_d$) approach the critical value as $\lambda_p \approx
408: 1-1.6149 L^{-1.03531}$($\lambda_d \approx 1+9.69198L^{-0.974152}$),
409: with the values at peaks (dips) diverging with increasing system
410: size $L$:
411: \begin{equation}
412: K(\lambda_1,\lambda_2)\big|_{\lambda_1=\lambda_{p(d)}} =
413: k_{p(d)}\frac{L}{(\ln L)^{4}} + {\rm constant}. \label{finite}
414: \end{equation}
415: Here the prefactor $k_{p(d)}$ is non-universal in the sense that it
416: depends on $\lambda_2$ and $\gamma$.  We emphasize that
417: Eq.~(\ref{finite}) follows from  Eq.~(\ref{infinite}), if we take
418: into account the fact that the model is conformally invariant at the
419: critical point. Indeed, on the one hand,  from Eq.~(\ref{infinite})
420: and the correlation length $\xi \sim |\lambda -\lambda_c |^{-\nu}$
421: with $\nu = 1$, we have $K(\lambda_1,\lambda_2) \sim \xi/(\ln
422: \xi)^4$.  On the other hand, the conformal invariance requires the
423: scale invariance: $\xi / L = \xi' / L'$.  The numerical results are,
424: respectively, plotted for
425: $K(\lambda_1,\lambda_2)|_{\lambda_1=\lambda_{p(d)}}$ in
426: Fig.~\ref{fig2} and  for $\lambda_{p(d)}$ in Fig.~\ref{fig3} with
427: $\lambda_2 = 0.6$ and $\gamma = 1$.  The same is also true for any
428: nonzero $\gamma$.  This shows that, consistent with the exact
429: result, the correlation length critical exponent $\nu$ equals 1, as
430: long as $\gamma$ is nonzero.
431: 
432: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
433: %%3
434: \begin{figure}[ht]
435: \begin{overpic}[width=42mm,totalheight=42mm]{fig3a.eps}
436: \put(55,25){$(a)$}
437: \end{overpic}
438: \hspace{0in}
439: \begin{overpic}[width=40mm,totalheight=40mm]{fig3b.eps}
440: \put(55,25){$(b)$}
441: \end{overpic}
442: \caption{(color online) (a): The positions of the peaks approach the
443: critical
444:   point $\lambda_c =1$ with increasing system size $L$ as $\lambda_p \approx 1-1.61490 L^{-1.03531}$.  (b) The positions of the dips approach the critical
445:   point $\lambda_c =1$  with increasing system size $L$ as $\lambda_d \approx 1+9.69198 L^{-0.974152}$. In both cases,  $\lambda_2 =0.6$ and $\gamma =1$.} \label{fig3}
446: \end{figure}
447: 
448: 
449: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
450: 
451: 
452: 
453: 
454: {\it Conclusions.} We have shown that singularities in fidelity
455: surfaces may be \textit{intrinsically} characterized in terms of
456: Riemannian geometry, based on the fidelity description of QPTs.
457: Generically, the Ricci curvature tensor for finite-size systems is
458: analytic and it exhibits singularities at transition points in the
459: thermodynamic limit, as reflected in the Ricci scalar curvature that
460: blows up when the system size tends to $\infty$. This opens up the
461: possibility to exploit the theory of Ricci flows~\cite{perelman} to
462: characterize QPTs in condensed matter theory. The one-dimensional
463: quantum Ising model in a transverse field is exploited as an example
464: to explicitly illustrate the theory~\cite{explanation}, and a finite
465: size scaling analysis has been performed for the Ricci scalar
466: curvature with different lattice sizes, and the correlation length
467: critical exponent has been extracted, consistent with the known
468: exact value.
469: 
470: 
471: We thank John Paul Barjaktarevi\v{c}, Sam Young Cho and John
472: Fjaerestad for helpful discussions and comments. The support from
473: the National Natural Science Foundation of China is acknowledged.
474: 
475: 
476: 
477: 
478: \begin{thebibliography}{99}
479: 
480: \bibitem{sachdev} S. Sachdev, \textit{Quantum Phase Transitions}
481: (Cambridge University Press, 1999, Cambridge).
482: \bibitem{wen} X.-G.
483: Wen, \textit{Quantum Field Theory of Many-Body Systems}, (Oxford
484: University Press, 2004, Oxford).
485: 
486: \bibitem{osborne} T.J. Osborne and M.A. Nielsen, Phys. Rev. A
487: \textbf{66}, 032110 (2002); A. Osterloh, L. Amico, G. Falci, and R.
488: Fazio, Nature \textbf{416}, 608 (2002).
489: \bibitem{vidal} G. Vidal, J. I. Latorre, E. Rico, and A. Kitaev, Phys. Rev. Lett.
490: \textbf{90}, 227902 (2003).
491: \bibitem{levin} A. Kitaev and J.
492: Preskill, Phys. Rev. Lett. \textbf{96}, 110404 (2006); M. Levin and
493: X.-G. Wen, Phys. Rev. Lett. \textbf{96}, 110405 (2006).
494: %\bibitem{preskill} J. Preskill, J. Mod. Opt. \textbf{47}, 127
495: %(2000).
496:  \bibitem{entanglement1} K. E. Korepin, Phys. Rev. Lett. \textbf{92},
497:  096402 (2004); G. C. Levine, Phys. Rev. Lett. \textbf{93},
498:  266402 (2004); G. Raefael and J. E. Moore, Phys. Rev. Lett. \textbf{93},
499:  260602 (2004); P. Calabrese and J. Cardy, J. Stat. Mech. P06002
500:  (2004);  H.-Q. Zhou, T. Barthel, J.O. Fjaerestad, and U. Schollw\"{o}ck, Phys. Rev. A  \textbf{74}, 050305(R) (2006)..
501: \bibitem{entanglement2} F. Verstraete, M.A. Martin-Delgado, and J. I. Cirac, Phys. Rev. Lett.
502: \textbf{92}, 087201 (2004); H. Barnum, E. Knill, G. Ortiz, R. Somma,
503: and L. Viola, Phys. Rev. Lett. \textbf{92}, 107902 (2004); W. D\"ur,
504: L. Hartmann, M. Hein, M. Lewenstein, and H.J. Briegel, Phys. Rev.
505: Lett. \textbf{94}, 097203 (2005).
506: 
507: \bibitem{zanardi} P. Zanardi and N. Paunkovi\'{c}, Phys. Rev. E
508: \textbf{74}, 031123 (2006).
509: \bibitem{zjp} H.-Q. Zhou and J.P. Barjaktarevi\v{c}, Fidelity
510: and quantum phase transitions, cond-mat/0701608.
511: 
512: \bibitem{zhou} H.-Q. Zhou, J.-H. Zhao, and B. Li, arXiv:0704.2940[condmat.
513: stat-mech]; H.-Q. Zhou, arXiv:0704.2945[condmat. stat-mech].
514: \bibitem{zov} H.-Q. Zhou, R. Or\'us, and G. Vidal,  arXiv:0709.4596[condmat. stat-mech].
515: \bibitem{more} P. Zanardi, M. Cozzini, and P. Giorda, cond-mat/ 0606130; N. Oelkers
516: and J. Links, Phys. Rev. B 75, 115119 (2007); M. Cozzini, R.
517: Ionicioiu, and P. Zanardi, cond-mat/0611727; L. Campos Venuti and P.
518: Zanardi, Phys. Rev. Lett. 99, 095701 (2007); P. Buonsante and A.
519: Vezzani, Phys. Rev. Lett. 98, 110601 (2007); S.J. Gu, H.M. Kwok,
520: W.Q. Ning, and H.Q. Lin, arXiv:0706.2495[condmat. stat-mech]; S.
521: Chen, L. Wang, S.-J. Gu, and Y.P. Wang, arXiv:0706.0072[condmat.
522: stat-mech]; N. Paunkovi\'{c} and V.R. Vieira
523: arXiv:0707.4667[condmat. stat-mech]; M.F. Yang,
524: arXiv:0707.4574[condmat. stat-mech]; Y.C. Tzeng and M.F. Yang,
525: arXiv:0709.1518[condmat. stat-mech].
526: \bibitem{nielsen} M.A. Nielsen
527: and I.L. Chuang, \textit{Quantum Computation and Quantum
528: Information} (Cambridge University Press, 2000, Cambrige).
529: \bibitem{baxter} R.J. Baxter, cond-mat/0611167.
530: \bibitem{relevant} Actually, in the fidelity approach~\cite{zjp}, we
531: prefer to introduce \textit{relevant information} and
532: \textit{irrelevant information} as the counterparts of
533: \textit{orders} and \textit{fluctuations} in the conventional
534: theory, since the fidelity, unlike order parameters, is \textit{not}
535: physical observable, although both may be used to quantify quantum
536: fluctuations.
537: 
538: \bibitem{novikov} B.A. Dubrovin, A.T. Fomenko, and S.P. Novikov, \textit{Modern
539: Geometry--Methods and Applications}, Part 1, Second edition,
540: (Springer-Verlag, 1985, New York).
541: \bibitem{lieb} E. Lieb, T. Schultz, and D. Mattis, Ann. Phys.
542: \textbf{60}, 407 (1961).
543: \bibitem{pfeuty} P. Pfeuty, Ann. Phys.
544: \textbf{57}, 79 (1970).
545: \bibitem{finitefidelity} The explicit
546: expression for $\ln d(\lambda_1,\lambda_2)$ was also given in
547: Refs.~\cite{zjp,zhou} for the quantum $XY$ model of finite sizes,
548: which is needed for the evaluation of the Gaussian curvature
549: $K(\lambda_1,\lambda_2)$ for the corresponding fidelity surfaces.
550: \bibitem{perelman} G. Perelman, arXiv:math/0307245;
551: arXiv:math/0303109; arXiv:math/0211159.
552: 
553: \bibitem{explanation} Although we restrict ourselves to discuss an
554: exactly solvable model in one spatial dimension, the theory in
555: principle applies to \textit{any} quantum lattice systems in
556: \textit{any} spatial dimensions~\cite{zov} by using the
557: newly-developed tensor network algorithms~\cite{tensornetwork}.
558: 
559: 
560: \bibitem{tensornetwork} G. Vidal, Phys. Rev. Lett. 98, 070201 (2007);  S. Singh, H.-Q. Zhou, and G. Vidal,
561: cond-mat/0701427; J. Jordan, R. Or¡äus, G. Vidal, F. Verstraete, and
562: J.I. Cirac, arXiv:cond-mat/0703788;  F. Verstraete and G. Vidal,
563: private communication.
564: 
565: 
566: 
567: 
568: 
569: 
570: 
571: 
572: 
573: 
574: 
575: \end{thebibliography}
576: \end{document}
577: