cond-mat0611727/mps.tex
1: \documentclass[aps,prb,twocolumn,superscriptaddress,showpacs,amsmath]{revtex4}
2: 
3: \usepackage{epsf,latexsym}
4: \usepackage{amssymb,amsmath}
5: \usepackage{times}
6: \usepackage{bm}
7: \usepackage{verbatim}
8: %%\usepackage{amscd}
9: 
10: \newcommand{\bra}[1]{\langle #1 |}
11: \newcommand{\ket}[1]{| #1 \rangle}
12: \newcommand{\braket}[2]{\langle #1 | #2 \rangle}
13: 
14: \newcommand\R{{\mathrm {I\!R}}}
15: \newcommand\N{{\mathrm {I\!N}}}
16: \newcommand\h{{\cal H}}
17: \newcommand\f{{\mathcal{F}}}
18: \newcommand{\qed}{$\hfill \Box$}
19: 
20: \newcommand\tr{{\mbox{Tr\,}}}
21: 
22: \newcommand{\ignore}[1]{}
23: 
24: \newcommand{\ra}{{\rightarrow}}
25: \newcommand{\be}{\begin{equation}}
26: \newcommand{\ee}{\end{equation}}
27: \newcommand{\ba}{\begin{eqnarray}}
28: \newcommand{\ea}{\end{eqnarray}}
29: 
30: %%\newtheorem{proposition}{Proposition}
31: %%\newtheorem{lemma}{Lemma}
32: \newtheorem{theorem}{Theorem}
33: %%\newtheorem{corollary}{Corollary}
34: 
35: \def\CC{{\rm\kern.24em \vrule width.04em height1.46ex depth-.07ex
36:     \kern-.30em C}}
37: \def\P{{\rm I\kern-.25em P}}
38: 
39: \def\RR{{\rm
40:          \vrule width.04em height1.58ex depth-.0ex
41:          \kern-.04em R}}
42: 
43: \def\bbbone{{\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l}
44: {\rm 1\mskip-4.5mu l} {\rm 1\mskip-5mu l}}}
45: 
46: \def\bbbc{{\mathchoice {\setbox0=\hbox{$\displaystyle\rm C$}\hbox{\hbox
47: to0pt{\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
48: {\setbox0=\hbox{$\textstyle\rm C$}\hbox{\hbox
49: to0pt{\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
50: {\setbox0=\hbox{$\scriptstyle\rm C$}\hbox{\hbox
51: to0pt{\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
52: {\setbox0=\hbox{$\scriptscriptstyle\rm C$}\hbox{\hbox
53: to0pt{\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}}}
54: 
55: \def\bbbz{{\mathchoice {\hbox{$\sf\textstyle Z\kern-0.4em Z$}}
56: {\hbox{$\sf\textstyle Z\kern-0.4em Z$}}
57: {\hbox{$\sf\scriptstyle Z\kern-0.3em Z$}}
58: {\hbox{$\sf\scriptscriptstyle Z\kern-0.2em Z$}}}}
59: 
60: \newcommand{\putfig}[2]{$$\leavevmode\hbox{\epsfxsize=#2 cm
61:    \epsffile{#1.eps}}$$}
62: \newcommand{\insertfig}[2]{\leavevmode \vcenter{\hbox{\epsfxsize=#2 cm
63:    \epsffile{#1.eps}}}}
64: 
65: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
66: 
67: \begin{document}
68: 
69: \title{Quantum fidelity and quantum phase transitions in matrix product states}
70: \author{Marco Cozzini}
71: \affiliation{Dipartimento di Fisica, Politecnico di Torino, Corso Duca degli Abruzzi 24, I-10129 Torino, Italy}
72: \affiliation{Quantum Information Group, Institute for Scientific Interchange (ISI), Viale Settimio Severo 65, I-10133 Torino, Italy}
73: \author{Radu Ionicioiu}
74: \affiliation{Quantum Information Group, Institute for Scientific Interchange (ISI), Viale Settimio Severo 65, I-10133 Torino, Italy}
75: \author{Paolo Zanardi}
76: \affiliation{Quantum Information Group, Institute for Scientific Interchange
77: (ISI), Viale Settimio Severo 65, I-10133 Torino, Italy}
78: \affiliation{Department of Physics and Astronomy, University of Southern 
79: California, Los Angeles, CA 90089-0484, USA}
80: 
81: \begin{abstract}
82: Matrix product states, a key ingredient of numerical algorithms widely employed in the simulation of quantum spin chains, provide an intriguing tool for
83: quantum phase transition engineering. At critical values of the control
84: parameters on which their constituent matrices depend, singularities in the
85: expectation values of certain observables can appear, in spite of the
86: analyticity of the ground state energy.
87: For this class of generalized quantum phase transitions we test the validity
88: of the recently introduced fidelity approach, where the overlap modulus of
89: ground states corresponding to slightly different parameters is considered.
90: We discuss several examples, successfully identifying all the present
91: transitions. We also study the finite size scaling of fidelity derivatives,
92: pointing out its relevance in extracting critical exponents.
93: 
94: \end{abstract}
95: 
96: \pacs{05.30.-d,64.60.-i}
97: 
98: \maketitle
99: 
100: \section{Introduction}
101: 
102: The study of many-body quantum systems is at the same time a fascinating and
103: challenging field of research. This is due to the richness of inherently complex
104: phenomena arising in the presence of a large number of interacting particles,
105: among which quantum phase transitions (QPTs) occupy a distinguished position.
106: These transitions take place at zero temperature and are driven by an external
107: parameter, as for example the magnetic field in superconductors.
108: 
109: Recently it has been shown that the quantum fidelity -- the overlap modulus --
110: of two finite size ground states corresponding to neighboring control
111: parameters is a good indicator of QPTs.
112: Indeed, the fidelity typically drops abruptly at critical points, as a
113: consequence of the dramatic state transformation involved in a transition.
114: This approach has been tested in various contexts where the ground state can
115: be calculated exactly, either analytically or numerically. These include Dicke
116: model, one-dimensional $XY$ model in a transverse field, and general quadratic
117: fermionic Hamiltonians \cite{gs_overlap,fid_long}.
118: 
119: In the present article we further pursue this approach by analyzing
120: QPTs
121: described by matrix product states (MPSs) \cite{mps_refs}.
122: These states are at the basis of efficient numerical methods used in the
123: analysis of spin chain systems, as the density matrix renormalization group
124: (DMRG) algorithm. When considered dependent on a control parameter $g$ they can
125: give rise to \textit{generalized} QPTs, i.e., transitions where some observable
126: quantity presents a non-analytic behavior in spite of the regularity of the
127: ground state energy \cite{qpt_mps}.
128: We show that the quantum fidelity of two neighboring (in terms of $g$) MPSs is
129: an effective tool not only in detecting the critical point $g_c$ but also in
130: giving the correct scaling at $g_c$.
131: The success of the fidelity approach in analyzing MPS-QPTs further proves the
132: generality of the procedure, which, as stressed in
133: Refs.~\onlinecite{gs_overlap,fid_long},
134: does not require any a priori understanding of the structure (order parameter,
135: correlation functions, topology, etc.) of the considered system. It is also
136: worth pointing out that this method seems to have some advantage with respect
137: to other quantum information based techniques. For example, quantum phase
138: transitions in MPSs do not give rise to the logarithmic divergence of the
139: entropy of block entanglement observed in other systems \cite{block_ent},
140: thereby ruling out this quantity as a reliable transition indicator.
141: On the other hand, other entanglement measures can be used, e.g., single site
142: entanglement and concurrence, whose derivatives often provide interesting
143: information about criticality \cite{osterloh02}.
144: We discuss the results
145: obtained by these methods in one of the simple examples we analyze below.
146: As it will be shown, also these latter measures turn to be less effective than
147: the fidelity approach.
148: 
149: In the following we will first present a derivation of the overlap formula for
150: general MPSs (Sec.~\ref{sec:general}), discussing some general features of
151: these systems, and then study in detail the MPS examples introduced in
152: Ref.~\onlinecite{qpt_mps} (Sec.~\ref{sec:examples}).
153: As in Refs.~\onlinecite{gs_overlap,fid_long}, the analysis will be carried out
154: also in terms of fidelity derivatives, which are the most suitable tool to
155: observe finite size scaling properties. In particular, we will explicitly show
156: how to extract the critical exponent of the correlation length from these
157: quantities (Subsec.~\ref{subsec:IIordQPTs}), thus demonstrating the
158: independence and the completeness of our approach.
159: Finally, we will report on the entanglement analysis in
160: Subsec.~\ref{subsec:ent}.
161: 
162: \section{Fidelity for matrix product states}
163: \label{sec:general}
164: 
165: Suppose we have a closed spin chain with $N$ sites. Let $d$ be the dimension of
166: the Hilbert space at each site. The matrix product states are defined as
167: \cite{qpt_mps}
168: \be
169: \ket{g} := \ket{\psi(g)} = \frac{1}{\sqrt\mathcal{N}}\sum_{i_1,\dots,i_N=0}^{d-1}
170: \tr(A_{i_1}\dots{}A_{i_N})\ket{i_1\dots{}i_N} \ ,
171: \ee
172: where the $A_j$'s, with $j=0,\dots,d-1$, are $D\times D$ matrices, $D$ is the
173: dimension of the bonds in the so-called valence bond picture, and ${\cal
174: N}:=\sum_{i_1,\dots,i_N=0}^{d-1}|\tr(A_{i_1}\dots{}A_{i_N})|^2$ a normalization
175: factor. Since our goal is to explore quantum phase transitions in such states,
176: we assume that the matrices $A_j=A_j(g)$ depend on one or more parameters,
177: generically denoted by $g$. The overlap of two MPSs corresponding to different
178: $g$'s is given by
179: \ba
180: \lefteqn{\braket{g_1}{g_2} = [\mathcal{N}(g_1)\mathcal{N}(g_2)]^{-1/2}} \\
181: &&\times\sum_{\bm{i}\in\bbbz_d^N}\tr[A_{i_1}^*(g_1)\dots{}A_{i_N}^*(g_1)]
182: \tr[A_{i_1}(g_2)\dots{}A_{i_N}(g_2)] \ , \nonumber
183: \ea
184: where $\bm{i}=(i_1,\dots,i_N)$.
185: In order to simplify the formul\ae, we will work with the unnormalized overlap
186: \be
187: F(g_1, g_2) := \sqrt{\mathcal{N}(g_1)\mathcal{N}(g_2)}\,\braket{g_1}{g_2}
188: \label{overlap}
189: \ee
190: and we obviously have $\mathcal{N}(g_a)=F(g_a, g_a)$, $a=1,2$.
191: Using the identity $\tr(A)\tr(B)=\tr(A\otimes B)$, the overlap can be
192: computed exactly for MPSs \cite{zhou} and we obtain
193: \be\label{eq:Fsum_lam}
194: F(g_1, g_2)= \tr[E^N(g_1,g_2)]= \sum_{k=1}^{D^2} \lambda_k^N(g_1,g_2) \ ,
195: \ee
196: where $\lambda_k$ are the eigenvalues of the $D^2\times
197: D^2$ matrix
198: \be
199: E(g_1,g_2) := \sum_{i=0}^{d-1} A_i^*(g_1) \otimes A_i(g_2) \ .
200: \ee
201: The latter is the generalization of the transfer operator
202: $E_\bbbone(g):=E(g,g)$, which we assume to be diagonalizable with eigenvalues
203: $\lambda_k(g):=\lambda_k(g,g)$.
204: Also, since $A\otimes{}B$ and $B\otimes{}A$ are isospectral, one has
205: $F(g_1,g_2)=F(g_2,g_1)^*$.
206: 
207: The main object of our study will be the fidelity of two neighboring states
208: \be\label{eq:f(g;d)}
209: \f(g;\delta):= |\braket{g-\delta}{g+\delta}| =
210: \frac{|F(g-\delta,g+\delta)|}{\sqrt{\mathcal{N}(g-\delta)\mathcal{N}(g+\delta)}}
211: \ee
212: for a small variation $\delta$
213: in the parameter space spanned by $g$. Defined this way, the fidelity depends
214: not only on the parameter $g$ driving the QPT, but also on its variation
215: $\delta$.
216: Expanding the fidelity in $\delta$ we obtain
217: $\f(g;\delta)\simeq1+
218: {\partial_\delta^2\f(g;\delta)|}_{\delta=0}\delta^2/2$
219: (since $\f(g;\delta)$ reaches its maximum for $\delta=0$ the first derivative
220: vanishes).
221: Thus the rate of change of the fidelity close to a critical point is given by
222: the second derivative of the fidelity \cite{gs_overlap,fid_long}
223: %
224: \be
225: S(g):={\partial^2_\delta\f(g;\delta)|}_{\delta=0}
226: \ee
227: %
228: and this is the relevant quantity
229: we will study to determine the scaling at the critical point $g=g_c$.
230: Note that $S(g)$ is connected to the ground state variation
231: $\ket{\psi'(g)}=\partial_g\ket{\psi(g)}$.
232: For example, for real states one has the simple second order expansion
233: $\f(g;\delta)\simeq1-2\delta^2||\psi'(g)||^2$.
234: The general expression of the second derivative has the form
235: \footnote{Considering the fidelity $|\braket{g-\delta}{g+\delta}|$ corresponds
236: to taking a variation $2\delta$ between the states. This gives rise to the
237: factor $4$ in formula~(\ref{eq:S(g)general}) and differs from
238: Refs.~\onlinecite{gs_overlap,fid_long}.}
239: %
240: \be \label{eq:S(g)general}
241: S(g)= -4\,{\partial_{g_1}\partial_{g_2}\ln{F(g_1,g_2)}|}_{g_1=g_2=g} \ .
242: \ee
243: %
244: Indeed, from $\f(g;0)=1$ and ${\partial_\delta\f(g;\delta)|}_{\delta=0}=0$
245: one has
246: ${\partial_\delta^2\ln\f(g;\delta)|}_{\delta=0}={\partial_\delta^2\f(g;\delta)|}_{\delta=0}=S(g)$ and
247: then
248: %
249: \ba
250: S(g) & = & \frac{1}{2}\partial_\delta^2
251: \ln\left.\frac{F(g-\delta,g+\delta)F(g+\delta,g-\delta)}
252: {F(g-\delta,g-\delta)F(g+\delta,g+\delta)}\right|_{\delta=0} = \nonumber\\
253: & = & {(\partial_\delta^2-\partial_g^2)\ln{F(g-\delta,g+\delta)}|}_{\delta=0}
254: \ ,
255: \ea
256: %
257: from which Eq.~(\ref{eq:S(g)general}) follows immediately.
258: 
259: The general behavior of the fidelity for matrix product states in the
260: thermodynamic limit (TDL) can be inferred from Eq.~(\ref{eq:Fsum_lam}).
261: Without loss of generality we can assume the eigenvalues $\lambda_k$ of the
262: generalized transfer operator $E$ to be ordered $|\lambda_1|\geq|\lambda_2|\geq\ldots\geq|\lambda_{D^2}|$.
263: Then $F(g_1,g_2)=\sum_{k=1}^{D^2}\lambda_k^N(g_1,g_2)=
264: \lambda_1^N[1+\sum_{k=2}^{D^2}(\lambda_k/\lambda_1)^N]$ and, provided
265: $|\lambda_1|>|\lambda_k|$ for $k=2,\dots,D^2$, in the thermodynamic limit
266: $N\to\infty$ one finds $F(g_1,g_2)\sim\lambda_1^N$.
267: Since the same holds for the normalization factors $\mathcal{N}(g)=F(g,g)$,
268: whenever the maximum eigenvalues $\lambda_1$ are non-degenerate, the large $N$
269: behavior of the fidelity must be of the form
270: $\mathcal{F}(g;\delta)\sim\alpha^N$, with $0\leq\alpha\leq1$ by construction.
271: Typically, $\alpha\neq1$ and the fidelity decays exponentially in the
272: TDL \cite{zhou}.
273: If $E_\bbbone(g)$ exhibits a level crossing in the largest eigenvalue for a
274: critical coupling $g_c$, the degeneracy $|\lambda_1(g_c)|=|\lambda_2(g_c)|$
275: gives rise to a discontinuity of some expectation values in the TDL
276: (called a \textit{generalized} QPT in Ref.~\onlinecite{qpt_mps}).
277: In this case the previous discussion has to be modified in order to include all the
278: degenerate eigenvalues.
279: In general, the vanishing of the fidelity is typically
280: strongly enhanced at critical points.
281: 
282: A similar analysis can be done for the second derivative $S(g)$ of the
283: fidelity. When a single eigenvalue $\lambda_1$ dominates, in the TDL one can
284: replace $F$ with $\lambda_1^N$ in Eq.~(\ref{eq:S(g)general}), thereby
285: recovering the expected \cite{fid_long} linear scaling $\propto{N}$:
286: \be \label{eq:S(l1)}
287: S(g) \sim -4N\,
288: {\partial_{g_1}\partial_{g_2}\ln{\lambda_1 (g_1,g_2)}|}_{g_1=g_2=g} \ .
289: \ee
290: In contrast, at the critical point two or more eigenvalues of $E(g_1,g_2)$ are
291: equal in modulus.
292: For example, at $g=g_c$ one can have
293: $|\lambda_1(g_c)|=|\lambda_2(g_c)|>|\lambda_k(g_c)|$ for $k>2$. Then in the TDL
294: %
295: \ba \label{eq:S(|l1|=|l2|)}
296: S(g_c) & \sim & \lim_{g\to{}g_c}-4\,{\partial_{g_1}\partial_{g_2}
297: \ln(\lambda_1^N+\lambda_2^N)|}_{g_1=g_2=g} = \nonumber\\
298: & = & \lim_{g\to{}g_c}\left\{
299: -N^2\frac{4\lambda_1^N\lambda_2^N}{(\lambda_1^N+\lambda_2^N)^2}
300: \partial_{g_1}\!\ln\frac{\lambda_1}{\lambda_2}
301: \partial_{g_2}\!\ln\frac{\lambda_1}{\lambda_2}\right. \nonumber \\
302: && \!\!\left.\left.-4N\sum_{k=1}^2\frac{\lambda_k^N}{\lambda_1^N+\lambda_2^N}
303: \partial_{g_1}\partial_{g_2}\!\ln\lambda_k\right\}\!\right|_{g_1=g_2=g} \!\!\ ,
304: \ea
305: %
306: where $\lambda_{1,2}=\lambda_{1,2}(g_1,g_2)$.
307: Hence, at the transition one typically recovers the divergence
308: $S(g_c)\propto{}N^2$ already observed in Ref.~\onlinecite{fid_long}.
309: We also note that, by defining $\theta:=\ln(\lambda_1/\lambda_2)$, the term in
310: the second line of Eq.~(\ref{eq:S(|l1|=|l2|)}) can be rewritten in the compact
311: form $[N|\partial_{g_1}\theta|/\cosh(N\theta/2)]^2$.
312: This follows from the equality $\partial_{g_1}\ln(\lambda_1/\lambda_2)
313: \partial_{g_2}\ln(\lambda_1/\lambda_2)=
314: |\partial_{g_1}\ln(\lambda_1/\lambda_2)|^2$, due to the relation
315: $\lambda_{1,2}(g_2,g_1)=\lambda_{1,2}^*(g_1,g_2)$.
316: Then, it is clear that this term vanishes for $N\to\infty$ unless $\Re\theta=0$,
317: i.e., $|\lambda_1|=|\lambda_2|$.
318: In the latter case, instead, assuming $\partial_{g_1}\theta$ to be finite, the
319: scaling $\propto{}N^2$ is evident.
320: %
321: Finally, we point out that the limit $g\to{}g_c$ in Eq.~(\ref{eq:S(|l1|=|l2|)})
322: is necessary to deal with possible singularities in the derivatives of the
323: $\lambda_k$'s at the transition (see the examples below).
324: 
325: \section{Examples}
326: \label{sec:examples}
327: 
328: In the following we are going to study the explicit examples of MPSs described
329: in Ref.~\onlinecite{qpt_mps}. Without loss of generality we assume
330: $g_1=g-\delta<g_2=g+\delta$.
331: 
332: \subsection{Fidelity and second order MPS-QPTs}
333: \label{subsec:IIordQPTs}
334: 
335: \begin{figure}
336: \putfig{fig1}{8.6}
337: \caption{Fidelity $\f(g;\delta)=|\braket{g-\delta}{g+\delta}|$ as a function of
338: $g$ for Example 1 with $\delta=10^{-3}$ and $N=1000,2000,3000$.
339: Inset: close to the critical point, the fidelity oscillates for states
340: belonging to different phases ($\delta=10^{-3}$, $N=10^4$).}
341: \label{fig:fid2body}
342: \end{figure}
343: 
344: {\em Example 1}. Take $D=2, d=3$ and $\{ A_j \}=\{-Z,\sigma_-,g\sigma_+\}$; in
345: the following we will use the notation
346: $X,Z$ for the Pauli matrices, $\sigma_\pm$ being the corresponding raising and
347: lowering operators.
348: This one-parameter family of MPSs contains the ground state of spin-1 AKLT model
349: for $g=\pm2$ and has a critical point at $g_c=0$ (see Ref.~\onlinecite{qpt_mps}).
350: The model is also significant because it has non-local string order. The parent
351: Hamiltonian with two-body interaction is given by
352: \ba
353: H & = & \textstyle\sum_i[(2+g^2)\bm{S}_i\cdot\bm{S}_{i+1}+
354: 2(\bm{S}_i\cdot\bm{S}_{i+1})^2+ \nonumber\\
355: &&+2(4-g^2)(S_i^z)^2+(g+2)^2(S_i^zS_{i+1}^z)^2+ \nonumber\\
356: &&+g(g+2)\{S_i^zS_{i+1}^z,\bm{S}_i\cdot\bm{S}_{i+1}\}] \ ,
357: \ea
358: where $\bm{S}_i$ is the spin operator acting on the $i$-th site.
359: The eigenvalues of the transfer operator $E= Z\otimes Z+ \sigma_- \otimes
360: \sigma_- + g_1g_2 \sigma_+ \otimes \sigma_+$ are $( -1, -1, 1\pm \sqrt{g_1 g_2}
361: )$ and Eq.~(\ref{eq:Fsum_lam}) yields
362: \be
363: F(g_1,g_2)= (1+\sqrt{g_1 g_2})^N+(1-\sqrt{g_1 g_2})^N+2(-1)^N
364: \ee
365: and ${\cal N}(g)=F(g,g)= (1+g)^N+ (1-g)^N+ 2(-1)^N$.
366: 
367: Let us analyze the behavior of the fidelity $\f(g;\delta)$ defined in
368: Eq.~(\ref{eq:f(g;d)}). Since $\f(g;\delta)$ is evidently symmetric in
369: both $g$ and $\delta$, we can safely assume $g\geq0$, $\delta>0$.
370: Three cases can then be distinguished, according to the
371: behavior of the eigenvalues of the transfer operators:
372: 
373: (i) $g>\delta$, where $\f(g;\delta)\sim\alpha^N$ for $N\gg1$, with
374: $\alpha=(1+\sqrt{g^2-\delta^2})^2/\sqrt{(1+g-\delta)(1+g+\delta)}<1$;
375: 
376: (ii) $g=\delta$, where $\f(\delta;\delta)\sim{}a/(1+2\delta)^{N/2}$ for
377: $N\gg1$, with $a=2$ for $N$ even and $a=2\delta\sqrt{N(N-1)}$ for $N$ odd (this
378: is the case where the largest eigenvalue of $E(g-\delta,g+\delta)$ is
379: degenerate and the decaying behavior of the fidelity is slightly modified);
380: 
381: (iii) $g<\delta$, where two of the eigenvalues of $E(g_1,g_2)$ are complex and,
382: for $N\gg1$, the fidelity exhibits an oscillatory behavior driven by
383: $|\cos(N\varphi)|$, $\varphi=\arctan\sqrt{\delta^2-g^2}$, with a decaying
384: envelope given by $2\alpha^N$,
385: $\alpha=\sqrt{(1+\delta^2-g^2)/[(1+\delta-g)(1+\delta+g)]}<1$.
386: Explicitly, at the critical point $g=g_c=0$ the overlap becomes
387: \be
388: \braket{-\delta}{\delta} = \frac{(1+\delta^2)^{N/2} \cos(N\varphi_0)+ (-1)^N}
389: {[(1+\delta)^N+ (1-\delta)^N]/2+ (-1)^N} \ ,
390: \ee
391: where $\varphi_0=\arctan\delta$. Then
392: $\f(0;\delta)=|\braket{-\delta}{\delta}|$ (note that the overlap can be
393: negative at this point).
394: 
395: \begin{figure}
396: \putfig{fig2}{8.6}
397: \caption{Finite size scaling for $S(g)$. The second derivative $S(g)$ is
398: plotted for $N=100,200,\dots,500$ (logarithm is base 10). The peak at $g=0$ scales with $N$
399: accordingly to Eq.~(\ref{eq:S(0)2body}). The dashed line corresponds to the
400: asymptotic behavior given by Eq.~(\ref{eq:Slim2body}).
401: The inset shows the data collapsing for the same curves ($N=100,200,\dots,500$)
402: when plotted in rescaled units.}
403: \label{fig:fssS2body}
404: \end{figure}
405: 
406: The fidelity behavior in the neighborhood of the critical point is shown
407: in Fig.~\ref{fig:fid2body} for $\delta=10^{-3}$ and various values of $N$.
408: It is evident that, although for a fixed $\delta$ the fidelity vanishes in the
409: TDL for any $g$, the rate of decrease is much faster at the transition.
410: This fidelity drop provides a useful finite size precursor of the quantum
411: phase transition taking place in the infinite system.
412: In the figure inset we zoom on the oscillatory behavior in the interval
413: $g\in(-\delta,\delta)$, where states on different sides of the transition are
414: considered. Note that oscillations start to be visible only for $N\varphi>\pi$,
415: i.e., $N>N_{\mathrm{osc}}$, $N_{\mathrm{osc}}=\pi/\arctan\sqrt{\delta^2-g^2}$,
416: which for $g=0$ and $\delta\ll1$ reduces to $N_{\mathrm{osc}}\sim\pi/\delta$.
417: 
418: The second derivative $S(g)$ of the fidelity can be calculated analytically
419: using formula (\ref{eq:S(g)general}). We find
420: \ba
421: \label{eq:Slim2body}
422: S(g\neq0) & \sim & -\frac{N}{|g|(1+|g|)^2} \ , \\
423: \label{eq:S(0)2body}
424: S(g=0) & = &
425: \left\{\begin{array}{lc}
426: -N(N-1) & \mbox{($N$ even)} \\
427: -(N-2)(N-3)/3  & \mbox{($N$ odd)}
428: \end{array}\right. \ ,
429: \ea
430: where Eq.~(\ref{eq:Slim2body}) corresponds to Eq.~(\ref{eq:S(l1)}) and gives the
431: asymptotic behavior for large $N$, while Eq.~(\ref{eq:S(0)2body}) is exact.
432: As in Ref.~\onlinecite{fid_long} the critical point is identified by a
433: divergence $\propto{}N^2$ of the second derivative $S(g)$ of the fidelity.
434: 
435: In Fig.~\ref{fig:fssS2body} we plot $\log|S(g)|$ for different values of $N$. In the inset of the
436: figure the curves are plotted in rescaled units, giving rise to data collapsing
437: (lines for different values of $N$ are practically indistinguishable in the
438: inset, merging into a single thick line). This shows that $S(g)/N^2$ is
439: basically a function of $gN$ only.
440: This is reminiscent of the usual scaling behavior of diverging observables in
441: second order quantum phase transitions.
442: Indeed, if an observable $P$ of a one-dimensional system diverges algebraically
443: in the TDL for some critical coupling $g_c$, i.e.,
444: $P_\infty\sim{}c_\infty|g-g_c|^{-\rho}$, the finite size scaling Ansatz in the
445: critical region reads $P_N=N^{\rho/\nu}Q(N|g-g_c|^\nu)$, where $\nu$ is the
446: critical exponent of the bulk correlation length $\xi_c$ and $Q$ is some
447: function \cite{fss}.
448: This is due to the fact that at criticality, once properly rescaled by some
449: power of the size of the system, the observable is expected to depend only on
450: the dimensionless ratio $L/\xi_c$ between the system size $L\propto{}N$ and the
451: correlation length.
452: Since the latter diverges at the transition like $\xi_c\propto1/|g-g_c|^\nu$
453: one immediately finds $L/\xi_c\propto{}N|g-g_c|^\nu$.
454: In our case $P_N=S/N=NQ(N|g|)$ with $g_c=0$, so that we obtain $\rho=\nu=1$.
455: The result $\rho=1$ agrees with Eq.~(\ref{eq:Slim2body}), which,
456: for $g\to0$, yields $P_\infty\sim{}-|g|^{-1}$.
457: The value of $\nu$ can instead be checked by explicitly calculating the
458: correlation length
459: \footnote{The correlation length in MPSs is usually given by the formula
460: $\xi_c=1/\ln|\lambda_1/\lambda_2|$, where $\lambda_1$
461: ($\lambda_2$) is the eigenvalue of $E_\bbbone(g)$ of (second) largest modulus.
462: In this case, however, the degeneracy of $\lambda_2=\lambda_3=-1$ does not allow
463: to apply this expression.},
464: where one finds $\nu=1$, confirming the finite size scaling estimate.
465: 
466: The next examples of this subsection correspond to three-body interaction
467: Hamiltonians, again taken from Ref.~\onlinecite{qpt_mps}.
468: Qualitatively, they all feature the same behavior as Example 1:
469: in the asymptotic limit $|S(g)|/N$ diverges proportionally to $1/|g|$ at the
470: critical point $g_c=0$, while the peak height of $S(g=0)$ scales with $N^2$.
471: 
472: {\em Example 2}. Consider $D=d=2$ and
473: $A_0=(\bbbone-Z)/2+\sigma_-$, $A_0=(\bbbone+Z)/2+g\sigma_+$.
474: The three-body Hamiltonian is $\bbbz_2$ symmetric and has a critical point at
475: $g_c=0$, where the state is a Greenberger-Horne-Zeilinger (GHZ) state, and reads
476: \be
477: H = \sum_i 2(g^2-1) Z_i Z_{i+1}- (1+g)^2 X_i +
478: (g-1)^2 Z_i X_{i+1} Z_{i+2} \ .
479: \ee
480: The generalized transfer operator $E(g_1,g_2)$ has eigenvalues
481: $(0,0,1\pm\sqrt{g_1 g_2})$, very similar to Example 1. The function
482: $F(g-\delta,g+\delta)$ is then still symmetric in $g$ and $\delta$.
483: 
484: Since the largest eigenvalue $\lambda_1$ is the same as in Example 1, the TDL
485: behavior of $\f(g,\delta)$ and $S(g)$ is unchanged.
486: However, the parity dependent term $(-1)^N$ is now absent.
487: Then, the asymptotic behavior of $S(g\neq0)$ is still given by
488: Eq.~(\ref{eq:Slim2body}), while
489: the parity dependence of Eq.~(\ref{eq:S(0)2body}) is lost and one simply has
490: $S(g=0)=-2N(N-1)$.
491: The latter results follows from Eq.~(\ref{eq:S(|l1|=|l2|)}), which applies
492: exactly to this case even for finite $N$. Indeed here only two eigenvalues of
493: $E(g_1,g_2)$ are different from zero and one has the level crossing
494: $\lambda_1(g=0)=\lambda_2(g=0)$.
495: Note that, due to the divergence of the double derivative
496: $\partial_{g_1}\partial_{g_2}\ln\lambda_{1,2}$ calculated in $g_1=g_2=0$, also the
497: term in the last line of Eq.~(\ref{eq:S(|l1|=|l2|)}) gives rise to a
498: contribution $\propto{}N^2$, in spite of its apparent linearity in $N$ for
499: $|\lambda_1|=|\lambda_2|$. This explains the necessity of the limit $g\to g_c$
500: in Eq.~(\ref{eq:S(|l1|=|l2|)}).
501: 
502: {\em Example 3}. Consider $A_0=X$, $A_1=\sqrt{g}(\bbbone- Z)/2$. The
503: eigenvalues of $E(g_1,g_2)$ are
504: $(\pm1,(\sqrt{g_1g_2}\pm\sqrt{g_1g_2+4})/2)$.
505: As before, the scaling of the system can be seen from the behavior of $S(g)$
506: in the thermodynamic limit,
507: driven by $\lambda_1(g_1,g_2)=(\sqrt{g_1g_2}+\sqrt{g_1g_2+4})/2$.
508: We find
509: \ba
510: \label{eq:Slim3body}
511: S(g\neq0) & \sim & -\frac{4N}{|g| (g^2+4)^{3/2}} \ , \\
512: \label{eq:S(0)3body}
513: S(g=0) & = &
514: \left\{\begin{array}{lc}
515: -N^2/4 & \mbox{($N$ even)} \\
516: -(N^2-1)/6  & \mbox{($N$ odd)}
517: \end{array}\right. \ ,
518: \ea
519: where Eq.~(\ref{eq:Slim3body}) gives the large $N$ behavior.
520: The parity dependence of Eq.~(\ref{eq:S(0)3body}) is caused by the negative
521: eigenvalues of $E(g_1,g_2)$, giving rise to terms oscillating like $(-1)^N$.
522: 
523: {\em Example 4}. Take $A_0= \sigma_+$, $A_1=\sigma_- + \sqrt{g}(\bbbone+ Z)/2$.
524: This MPS is the ground state of the following Hamiltonian:
525: \ba
526: \nonumber
527: H&=& -\sum_i g(X_i+ X_i Z_{i+1}+ Z_i X_{i+1}+ Z_i X_{i+1} Z_{i+2}) \\
528: &+& (1+2 g^2) Z_i- 2Z_i Z_{i+1}- Z_i Z_{i+1} Z_{i+2} \ .
529: \ea
530: The eigenvalues of $E(g_1,g_2)$ are
531: $(0,0,(\sqrt{g_1g_2}\pm\sqrt{g_1 g_2+4})/2)$.
532: Hence, the asymptotic behavior of $S(g)$ is again given by
533: Eq.~(\ref{eq:Slim3body}), while $S(0)=-N^2/2$ for $N$ even and
534: $S(0)=-(N^2-1)/6$ for $N$ odd. Parity dependence arises from
535: $(\sqrt{g_1g_2}-\sqrt{g_1g_2+4})/2=-1$ for $g_1=g_2=0$.
536: 
537: \subsection{Fidelity, concurrence, and single site entanglement}
538: \label{subsec:ent}
539: 
540: While all the above examples fit in the same picture, which is typical of
541: the second order quantum phase transitions already studied in
542: Refs.~\onlinecite{gs_overlap,fid_long},
543: the next example, albeit trivial as matrix product state, features a different
544: behavior, which will serve as a basis for some additional comments on the
545: fidelity approach to QPTs.
546: 
547: {\em Example 5}. Take $D=d=2$,
548: $A_0=\begin{pmatrix}1 & 0 \cr 0 & 1+g \end{pmatrix}$,
549: $A_1=\begin{pmatrix}g^n & 0 \cr 0 & 0 \end{pmatrix}$.
550: The matrix $E(g_1,g_2)$ has the eigenvalues $1+g_1$, $1+g_2$, $(1+g_1)(1+g_2)$,
551: and $1+g_1^n g_2^n$.
552: As discussed in Sec.~\ref{sec:general}, QPTs take place when two or more
553: eigenvalues of $E_\bbbone(g)=E(g,g)$ share the largest modulus.
554: We then have to compare $1+g$, $(1+g)^2$, and $1+g^{2n}$.
555: Clearly, for $g<0$ one has $\lambda_1(g)=1+g^{2n}$, while for
556: $g\geq0$ the dominant eigenvalue is determined by the ratio
557: $r(g):=(1+g^{2n})/(1+g)^2$. The equation $r(g)=1$ can be rewritten as
558: $g(g^{2n-1}-g-2)=0$. Apart from the trivial solution $g_c=0$, for $n\geq2$ a
559: second solution $g_c'>1$ exists, so that in the region $0\leq{}g\leq{}g_c'$ one
560: has $\lambda_1(g)=(1+g)^2$, while for $g>g_c'$ one obtains again
561: $\lambda_1(g)=1+g^{2n}$. For $n=1$ the point $g_c'$ can be considered
562: at infinity and the latter level crossing never occurs.
563: 
564: In order to evaluate the fidelity we have to substitute the above eigenvalues in
565: Eq.~(\ref{eq:f(g;d)}). 
566: For simplicity, here we restrict our analysis to the fidelity of ground states
567: belonging to the same phase, avoiding to discuss the effects shown in the inset
568: of Fig.~\ref{fig:fid2body} for Example 1.
569: In the phase corresponding to the interval $g\in(0,g_c')$, where
570: $\lambda_1(g)=(1+g)^2$, one has $\lambda_1(g_1,g_2)=(1+g-\delta)(1+g+\delta)$.
571: The normalized fidelity in the TDL then results
572: $\f(g;\delta)\sim
573: [(1+g-\delta)(1+g+\delta)/\sqrt{(1+g-\delta)^2(1+g+\delta)^2}]^N=1$,
574: so that one has a constant phase in this region, due to the factorization of
575: the largest eigenvalue of $E(g-\delta,g+\delta)$.
576: Outside this region such a factorization is absent and the fidelity
577: vanishes exponentially with the size of the system.
578: 
579: The asymptotic behavior of the second derivative $S(g)$ can be calculated
580: accordingly.
581: In the interval $g\in(0,g_c')$, where the eigenvalue $(1+g)^2$ dominates, one
582: finds $S(g)=0$ in the TDL, while for $g\notin[0,g_c']$, where the leading
583: eigenvalue is given by $1+g^{2n}$, Eq.~(\ref{eq:S(l1)}) yields the formula
584: %
585: \be\label{eq:Sextriv}
586: S(g)\sim-4Nn^2g^{2n-2}/(1+g^{2n})^2 \ .
587: \ee
588: %
589: The latter function, which is evidently symmetric in $g$, has a minimum at
590: $g_{\mathrm{min}}=-[(n-1)/(n+1)]^{1/2n}$ (since $|g_{\mathrm{min}}|<g_c'$, the
591: minimum position for $g>0$ is instead given by $g_c'$ itself, as $S(g)$
592: increases monotonically for $g>g_c'$).
593: Note that, while $S(g)$ is continuous at $g=g_c=0$, where $S(0)=0$, at the second
594: critical point $g=g_c'$ one has the discontinuity
595: $\lim_{g\to(g_c')^-}S(g)=0\neq
596: \lim_{g\to(g_c')^+}S(g)=-4Nn^2(g_c'+2)/[g_c'(1+g_c')^4]$,
597: where the latter results is obtained by substituting the defining relation
598: $r(g_c')=1$ into Eq.~(\ref{eq:Sextriv}).
599: Furthermore, exactly at the critical point $g_c'$ one has the superextensive
600: scaling $S(g_c')\sim-N^2[2n-1+(n-1)g_c']^2/(1+g_c')^4$.
601: 
602: The finite size behavior of the fidelity is shown in Fig.~\ref{fig:triv_ex} for
603: $N=100$ and $\delta=10^{-3}$. Due to the approximated relation
604: $\f(g;\delta)\simeq1+S(g)\delta^2/2$ one can there recognize also the scaling of
605: $S(g)$. In particular, in the lower panel where the case $n=2$ is plotted, the
606: minimum at $g_{\mathrm{min}}$ and the superextensive scaling at
607: $g_c'(n=2)\simeq1.521$
608: are evident.
609: 
610: \begin{figure}
611: \putfig{fig3}{8.6}
612: \caption{Fidelity for the model given in Example 5, with $\delta=10^{-3}$ and
613: $N=100$; $n=1$ (top) and $n=2$ (bottom).}
614: \label{fig:triv_ex}
615: \end{figure}
616: 
617: For this simple example it is also possible to find a compact analytic expression
618: for the ground state $\ket{g}$. Due to the commutativity $[A_0, A_1]=0$, one
619: has $\sqrt\mathcal{N}\ket{g}=\sum_{\bm{i}\in\bbbz_2^N}
620: \mathrm{Tr}(A_0^{N-k}A_1^k)\ket{i_1\dots{}i_N}$ with $k=\sum_{j=1}^Ni_j$. Then,
621: by using $\mathrm{Tr}(A_0^N)=1+(1+g)^N$ and $\mathrm{Tr}(A_0^{N-k}A_1^k)=g^{nk}$
622: for $k\neq0$, one finds
623: %
624: \ba
625: \sqrt\mathcal{N}\ket{g} & \!=\! &
626: \left[1+(1+g)^N\right]\!\ket{0}^{\otimes{}N}\!+\!
627: \sum_{k=1}^Ng^{nk}{N \choose k}^{\frac{1}{2}}\ket{D_N^{(k)}} = \nonumber \\
628: & \!=\! & (1+g)^N\ket{0}^{\otimes{}N}+(1+g^{2n})^{N/2}\ket{\phi(g)}^{\otimes{}N}
629: \ , \label{eq:gs_ex5}
630: \ea
631: %
632: where $\ket{D_N^{(k)}}={N \choose k}^{-\frac{1}{2}}
633: \sum_{\bm{i}\in{}I_k}\ket{i_1\dots{}i_N}$ with
634: $I_k=\{\bm{i}\in\bbbz_2^N|\sum_{j=1}^Ni_j=k\}$ is a Dicke state \cite{dicke54}
635: and $\ket{\phi(g)}=(\ket{0}+g^n\ket{1})/\sqrt{1+g^{2n}}$ is a normalized single
636: site state.
637: Since the normalization is $\mathcal{N}=\tr{E_\bbbone^N}$,
638: depending on the dominating eigenvalue of $E_\bbbone$ three cases are
639: possible in the TDL:
640: 
641: (i) $\lambda_1(g)=(1+g)^2>\lambda_2(g)$ and
642: $\ket{g}\to\ket{0}^{\otimes{}N}$;
643: 
644: (ii) $\lambda_1(g)=1+g^{2n}>\lambda_2(g)$ and
645: $\ket{g}\to\ket{\phi(g)}^{\otimes{}N}$;
646: 
647: (iii) $\lambda_1(g)=\lambda_2(g)$ and, in
648: the non-trivial case $g=g_c'$, one has
649: \footnote{Note that $\ket{0}^{\otimes{}N}$ and $\ket{\phi(g_c')}^{\otimes{}N}$
650: become orthogonal in the TDL, in accordance with the normalization factor
651: $1/\sqrt2$.}
652: $\ket{g}\to(\ket{0}^{\otimes{}N}+\ket{\phi}^{\otimes{}N})/\sqrt2$.
653: 
654: The above analysis can be summarized in the following table, where we recall the
655: results for the largest transfer operator eigenvalue $\lambda_1(g)$ and for the
656: large $N$ behavior of the function $S(g)$ and the ground state $\ket{g}$:
657: \begin{center}
658: \begin{tabular}{|c|c|c|c|}
659: \hline
660: & $g\in[0,g_c')$ & $g=g_c'$ & $g\notin[0,g_c']$ \\
661: \hline\hline
662: $\lambda_1(g)$ & $(1+g)^2$ & $1+g^{2n}=(1+g)^2$ & $1+g^{2n}$ \\
663: \hline
664: $S(g)$ & 0 & $\propto{}N^2$ &
665: $\propto{}N$ \\
666: \hline
667: $\ket{g}$  & $\ket{0}^{\otimes{}N}$ &
668: $(\ket{0}^{\otimes{}N}+\ket{\phi(g_c')}^{\otimes{}N})/\sqrt2$ &
669: $\ket{\phi(g)}^{\otimes{}N}$ \\
670: \hline
671: \end{tabular}
672: \end{center}
673: 
674: It is worth noting that the apparent drop of the fidelity at $g_{\mathrm{min}}$
675: in the $g<0$ phase for $n\geq2$ does not correspond to a QPT. Indeed, the
676: scaling of $S(g_{\mathrm{min}})$ does not present any peculiar behavior with
677: respect to the rest of the $g<0$ phase.
678: On the contrary, the scaling of $S$ changes its nature at $g=0$ and $g=g_c'$,
679: highlighting the transitions, although in a very different way from the typical
680: behavior observed in the previous second order QPTs.
681: The transition at $g=0$ corresponds to a discontinuity in
682: $\partial_g^n\langle{X}\rangle$ (Ref.~\onlinecite{qpt_mps}), while that at
683: $g=g_c'$ to a discontinuity of $\langle{X}\rangle$ itself
684: \footnote{\protect In the limit $N\to\infty$ one has
685: $\langle{X(g)}\rangle\to0$ for $g\in[0,g_c')$,
686: $\langle{X(g_c')}\rangle\to(g_c')^n/[1+(g_c')^{2n}]$,
687: and $\langle{X(g)}\rangle\to2g^n/(1+g^{2n})$ for $g{\protect\notin}[0,g_c']$.
688: Hence, for example, $\lim_{g\to0^-}\partial_g^n\langle{X}\rangle=2n!$, while
689: $\lim_{g\to0^+}\partial_g^n\langle{X}\rangle=0$.}.
690: Due to the permutation symmetry arising from the commutativity of $A_0$, $A_1$,
691: the correlation length cannot here be defined (the usual formula
692: $\xi_c=1/\ln|\lambda_1/\lambda_2|$ does not hold).
693: In practice, the transitions of Example 5 appear to be of first order.
694: 
695: For this example we also investigate concurrence and single site
696: entanglement, mainly focusing on the critical point $g_c=0$.
697: Of course, the fact that the state is factorized in the TDL trivially implies
698: the vanishing of these quantities in this limit.
699: We are however interested in the finite size scaling of the derivatives
700: of these measures, in the spirit of Ref.~\onlinecite{osterloh02}.
701: We are motivated by various aspects.
702: First, as mentioned before, in MPS-QPTs the block entanglement entropy does not
703: exhibit the logarithmic divergence observed in other quantum phase transitions,
704: posing the question whether other entanglement properties usually signaling
705: quantum criticality behave differently in this case.
706: Second, we are interested in comparing the effectiveness of the fidelity
707: approach against other quantum information-theoretical methods.
708: The simple nature of this example allows for a fully analytical treatment.
709: 
710: {\em Concurrence.}
711: We compute the concurrence for the reduced density matrix of 2 qubits in
712: the chain (the choice of the qubits is unimportant, due to the permutation
713: symmetry mentioned above). The density matrix of two spins $\rho^{(2)}$ is
714: obtained by tracing the initial $\rho=\ket{g}\bra{g}$ over all the other spins
715: in the chain.
716: By exploiting the symmetries arising from the commutation relations $[A_i,A_j]=0$
717: and recalling that
718: $\rho^{(2)}=\sum_{i,j,k,l=0}^1\rho_{ij,kl}^{(2)}\ket{ij}\bra{kl}$ is a $4\times4$
719: real symmetric matrix, one finds
720: $\rho^{(2)}_{ij,kl}=\rho^{(2)}_{ji,kl}=\rho^{(2)}_{ij,lk}$. Together with the
721: normalization $\tr\rho^{(2)}=1$ this implies that only 5 entries of
722: $\rho^{(2)}$ are independent, e.g., $\rho_{00,00}^{(2)}$, $\rho_{00,01}^{(2)}$,
723: $\rho_{00,11}^{(2)}$, $\rho_{01,01}^{(2)}$, $\rho_{01,11}^{(2)}$,
724: significantly simplifying the calculation in the standard basis.
725: Otherwise, one can use Eq.~(\ref{eq:gs_ex5}), easily obtaining
726: %
727: \ba \label{eq:rho2}
728: \rho^{(2)} &=& [(1+g)^{2N}\ket{00}\bra{00}+
729: (1+g^{2n})^{N}\ket{\phi\phi}\bra{\phi\phi} \\
730: && +(1+g)^N(1+g^{2n})(\ket{00}\bra{\phi\phi}+\ket{\phi\phi}\bra{00})]
731: /\mathcal{N}(g) \ . \nonumber
732: \ea
733: %
734: Having $\rho^{(2)}$, the concurrence for two qubits is defined as \cite{C}
735: $C=\max\{0,\sqrt{\mu_1}-\sqrt{\mu_2}-\sqrt{\mu_3}-\sqrt{\mu_4}\}$, where
736: $\mu_i$'s are the eigenvalues, in decreasing order, of the operator
737: $\rho^{(2)}(\sigma_y \otimes \sigma_y)(\rho^{(2)})^*(\sigma_y \otimes \sigma_y)$.
738: We finally have:
739: \be
740: C(g)= \frac{2 g^{2n} |(1+g)^N| } {\mathcal{N}(g)} \ .
741: \ee
742: In the thermodynamic limit both $C(g)$ and its first derivative vanish, so that the kind of
743: transition signatures found, e.g., in Ref.~\onlinecite{osterloh02} are absent here.
744: However, for $n=1$, the fourth derivative $\partial_g^4{C|}_{g=0}$ has a
745: divergence in the TDL; for $n\geq2$ the singularity is present in even higher
746: derivatives.
747: 
748: {\em Single site entanglement.}
749: This quantity is
750: the von Neumann entropy of a single spin ${\cal
751: S}=-\mathrm{Tr}(\rho^{(1)}\log_2\rho^{(1)})$, where $\rho^{(1)}$ is the reduced
752: density matrix of a single site.
753: Again from Eq.~(\ref{eq:gs_ex5}) one finds
754: %
755: \ba
756: \rho^{(1)} &=& [(1+g)^{2N}\ket{0}\bra{0}+
757: (1+g^{2n})^{N}\ket{\phi}\bra{\phi} \\
758: && +(1+g)^N\sqrt{1+g^{2n}}(\ket{0}\bra{\phi}+\ket{\phi}\bra{0})]
759: /\mathcal{N}(g) \ , \nonumber
760: \ea
761: %
762: whose eigenvalues are $\lambda_\pm=(1\pm\sqrt{1-4\det\rho^{(1)}})/2$, with
763: $\lambda_+=1-\lambda_-$ and
764: %
765: \be\label{eq:det(rho1)}
766: \det\rho^{(1)}=\frac{g^{2n}(1+g)^{2N}[(1+g^{2n})^{N-1}-1]}{\mathcal{N}^2(g)} \ .
767: \ee
768: %
769: The single site entanglement entropy
770: $\mathcal{S}=-\lambda_+\log_2\lambda_+-\lambda_-\log_2\lambda_-$ vanishes in the
771: thermodynamic limit at $g=0$, together with its first derivative.
772: For $n=1$, one finds a divergence in $\partial_g^4{\mathcal{S}|}_{g=0}$, while the
773: divergence is shifted to higher derivatives for $n\geq2$, similarly to the
774: concurrence. The considered divergence is however due to the functional form of
775: the von Neumann entropy. Indeed, $\lambda_-\to0$ for $g\to0$, giving rise to a
776: singularity in the logarithm.
777: 
778: We conclude this analysis by briefly discussing the TDL of these entanglement
779: measures at the critical point $g=g_c'$, where the state is not factorized
780: (see the table above).
781: The single site reduced density matrix in the TDL is simply
782: $\rho^{(1)}(g_c')=(\ket{0}\bra{0}+\ket{\phi}\bra{\phi})/2$ and
783: $\det\rho^{(1)}(g_c')=[1-1/(1+g_c')^2]/4$.
784: Then $\lambda_\pm(g_c')=[1\pm1/(1+g_c')]/2$ and $\mathcal{S}(g_c')\neq0$.
785: Note however that $C(g_c')=0$ in the TDL, reminiscent of the behavior of the
786: GHZ state.
787: This can be seen from the simple TDL of Eq.~(\ref{eq:rho2}), namely,
788: $\rho^{(2)}(g_c')=(\ket{00}\bra{00}+\ket{\phi\phi}\bra{\phi\phi})/2$.
789: 
790: \section{Conclusions}
791: 
792: In conclusion we have investigated the quantum fidelity of slightly different
793: matrix product states dependent on a parameter $g$ in the context of quantum
794: phase transitions. 
795: For generic MPSs, the overlap can be related analytically to the eigenvalues of a
796: generalized transfer operator, among which the largest in modulus plays a crucial
797: role. If the latter is non-degenerate (in modulus), the fidelity typically
798: exhibits an exponential decay in the thermodynamic limit and its second
799: derivative is proportional to the size $N$ of the system.
800: If instead more eigenvalues share the largest modulus, a quantum phase transition
801: can take place and the fidelity second derivative $S(g)$ generally scales with
802: $N^2$.
803: We have demonstrated this behavior in the exhaustive analysis of some simple
804: examples, taken from Ref.~\onlinecite{qpt_mps}, where possible exceptions have
805: also been pointed out.
806: Moreover, we have shown that the second derivative of the fidelity is a useful
807: quantity for the quantitative analysis of the scaling properties of the system at
808: the transition point. From the finite size scaling behavior of $S(g)$ we have
809: indeed estimated the critical exponent for the correlation length, finding
810: agreement with the explicit calculation.
811: Finally, for one of the considered examples, we have analyzed both concurrence and 
812: single site entanglement.
813: Although the latter quantities did provide signatures of the undergoing quantum
814: phase transitions, this information was hidden in high order derivatives, making
815: it much more difficult to extract than the fidelity (or its second derivative).
816: 
817: The fidelity analysis has the advantage of providing a unified framework to
818: detect very different types of phase transitions.
819: For the concrete examples analyzed here, singularities in the fidelity are
820: related to level crossings in the transfer operator, recovering the known
821: transition mechanism for matrix product states.
822: The results presented here further contribute to demonstrate the generality of
823: the fidelity approach to quantum phase transitions \cite{FiniteT}, supporting
824: the findings of Refs.~\onlinecite{gs_overlap,fid_long}.
825: 
826: \acknowledgments
827: 
828: We thank L.~Amico and P.~Giorda for valuable discussions and comments.
829: 
830: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
831: 
832: \begin{thebibliography}{}
833: 
834: \bibitem{gs_overlap}
835: P.~Zanardi and N.~Paunkovi\'c,
836: \pre {\bf 74}, 031123 (2006);
837: P.~Zanardi, M.~Cozzini, and P.~Giorda,
838: quant-ph/0606130.
839: 
840: \bibitem{fid_long}
841: M.~Cozzini, P.~Giorda, and P.~Zanardi,
842: quant-ph/0608059.
843: 
844: \bibitem{mps_refs}
845: I.~Affleck, T.~Kennedy, E.H.~Lieb and H.~Tasaki,
846: 	\prl {\bf 59}, 799 (1987);
847: 	Commun. Math. Phys. {\bf 115}, 477 (1988);
848: M.~Fannes, B.~Nachtergaele and R.F.~Werner,
849: 	Europhys. Lett. {\bf 10}, 633 (1989);
850: 	Commun. Math. Phys. {\bf 144}, 443 (1992);
851: S.~\"Ostlund and S.~Rommer,
852: 	\prl {\bf 75}, 3537 (1995);
853: S.~Rommer and S.~\"Ostlund,
854: 	\prb {\bf 55}, 2164 (1997);
855: D.~P\'erez-Garc\'ia, F.~Verstraete, M.M.~Wolf, and J.I.~Cirac,
856: 	quant-ph/0608197.
857: 
858: \bibitem{qpt_mps} M.M.~Wolf, G.~Ortiz, F.~Verstraete, J.I.~Cirac,
859: \prl {\bf 97}, 110403 (2006).
860: 
861: \bibitem{block_ent}
862: G.~Vidal, J.I.~Latorre, E.~Rico, and A.~Kitaev,
863: 	\prl {\bf 90}, 227902 (2003);
864: G.~Refael and J.E.~Moore,
865: 	\prl {\bf 93}, 260602 (2004);
866: V.~Korepin,
867: 	\prl {\bf 92}, 096402 (2004).
868: 
869: \bibitem{osterloh02} A.~Osterloh, L.~Amico, G.~Falci, and R.~Fazio,
870: Nature {\bf 416}, 608 (2002).
871: 
872: \bibitem{zhou} H.-Q.~Zhou and J.P.~Barjaktarevic, private communication.
873: 
874: \bibitem{fss}
875: M.E.~Fisher and M.N.~Barber,
876: \prl {\bf 28}, 1516 (1972);
877: M.N.~Barber,
878: in \textit{Phase Transitions and Critical Phenomena},
879: edited by C.~Domb and J.L.~Lebowitz
880: (Academic Press, London, 1983),
881: Vol. 8, p. 145.
882: 
883: \bibitem{C} W.K.~Wootters, \prl {\bf 80}, 2245 (1998).
884: 
885: \bibitem{dicke54}
886: R.H.~Dicke,
887: Phys. Rev. {\bf 93}, 99 (1954).
888: 
889: \bibitem{FiniteT} By using Uhlmann's fidelity for mixed states, one can extend
890: this kind of analysis to finite temperature systems and therefore even to
891: classical phase transitions. See P.~Giorda and P.~Zanardi, in preparation.
892: 
893: \end{thebibliography}
894: 
895: \end{document}
896: