This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Bell-type correlation at quantum phase transitions in spin-1 chain

Dongkeun Lee Department of Physics, Sogang University, 35, Baekbeom-ro, Mapo-gu, Seoul, 04107, Republic of Korea Research Institute for Basic Science, Sogang University, 35, Baekbeom-ro, Mapo-gu, Seoul, 04107, Republic of Korea    Wonmin Son [email protected] Department of Physics, Sogang University, 35, Baekbeom-ro, Mapo-gu, Seoul, 04107, Republic of Korea
Abstract

For the identification of non-trivial quantum phase, we exploit a Bell-type correlation that is applied to the one-dimensional spin-1 XXZ chain. It is found that our generalization of bipartite Bell correlation can take a decomposed form of transverse spin correlation together with high-order terms. The formulation of density-matrix renormalisation group is utilized to obtain the ground state of a given Hamiltonian with non-trivial phase. Subsequently Bell-SLK-type generalized correlation is evaluated through the analysis of the matrix product state. Diverse classes of quantum phase transitions in the spin-1 model are identified precisely through the evaluation of the first and the second moments of the generalized Bell correlations. The role of high-order terms in the criticality has been identified and their physical implications for the quantum phase has been revealed.

preprint: APS/123-QED

I Introduction

Quantum correlation, that has no classical counterparts, plays a pivotal role in studying the many-body systems. Investigation of such a complex massive system through the correlation is a major research topic in the statistical and the condensed matter physics as they provide useful view for the quantification of physical properties in a large scaled systems De Chiara and Sanpera (2018). Especially, the trait of quantum phase and its phase transition driven by the quantum fluctuation in correlation have been attracted much interests throughout the many years as they are closely related to macroscopic quantum phenomena such as superconductivities and topological state of a matter. They are analyzed through the various measures that has been rooted at the quantum information theory.

Quantum entanglement is one of the most crucial concepts that has been extensively studied in the last two decades to comprehend the quantum phase transitions and the exotic quantum properties in condensed matter systems Amico et al. (2008). Concurrence, one of the most used measures for two-qubit entanglement, is exploited to estimate the quantum criticality and, further, its scaling behaviour is also analyzed near the critical point of the one-dimensional XY model Osterloh et al. (2002); Osborne and Nielsen (2002). Entanglement entropy of a block in various many-body systems also has been discussed during the analysis using the conformal field theory Eisert et al. (2010). The entanglement entropy demonstrates logarithmic divergence at critical points, whereas it saturates for noncritical systems associated with the area law Eisert et al. (2010); Vidal et al. (2003). Subsequently, entanglement instabilities demonstrates the origin of the quantum criticality precisely and their analytic dynamics as the interaction parameters are modified Son et al. (2009).

It has been well known that all pure entangled states for two-qubit systems violate CHSH-Bell inequality Capasso et al. (1973). On the other hands, the non-local character of mixed entangled states are not easily quantifiable since it does not necessarily violate Bell inequalities Werner (1989). It means that quantum entanglement is not necessarily equivalent to the concept of Bell non-locality Werner (1989); Augusiak et al. (2015). Thus, it can be said that the concept of entanglement is closely related to the non-locality while they addresses the different physical quantities in general.

Unlikely to the many-body entanglement, that has been subjected to extensive investigation, the Bell non-locality in many-body systems has been attracted less attention relatively so far. It is because generalization of non-locality is needed to consider all the possible local measurements and locate their non-trivial combinations of the joint probabilities. It is only recent that the non-localities in many-body systems get to make an active progress. At first, Bell correlations for many-body systems are applied to two-site reduced density states which are generated from the ground state in one-dimensional (1D) spin-1/2 XY model Batle and Casas (2010) and spin-1/2 XXZ model Justino and de Oliveira (2012). As a result, it has been known that the Bell correlation demonstrates the non-analyticity at the critical point while, due to the monogamy characteristics of the correlation, any bipartite Bell inequality is not violated by the translational invariant many-body systems de Oliveira et al. (2013).

Differently from the bi-partite quantum non-locality, when multipartite correlation is considered, the violation of local realism is possible to be observed, especially in the 1D spin-1/2 model of the nearest-neighbor interactions Sun et al. (2014a, b). It is also quite recent that non-locality in the Lipkin-Meshkov-Glick (LMG) model can be found through multipartite Mermin inequalities Bao et al. (2020). The violation is manifested when the external magnetic field is adjusted. Furthermore, multipartite non-locality, that takes into account one- and two-body correlations imposing on the permutational symmetry, has been found in the Dicke state as well as the ground state of the isotopic LMG model Tura et al. (2014, 2015). The violation of the permutational invariant Bell inequality has been experimentally tested by the spin squeezed state that is realized by the large number of atoms in the Bose-Einstein condensate state Schmied et al. (2016). It is also found that the maximal violation is observed at the quantum criticality in the Ising model of infinite-range interactions Piga et al. (2019).

It is in two-dimensional local Hilbert space that the most Bell correlations are elucidated in aforementioned studies. In other words, Bell correlations in d-dimensional systems of many particles have been rarely investigated, whereas the other measures of quantum information theory, such as fidelity and quantum coherence, are used to characterize the quantum phase transitions (QPTs) many times in spin-1 chains or more Tzeng and Yang (2008); Goli et al. (2013); Malvezzi et al. (2016). In this work, we use a generalized high-order Bell correlation obtained from high-dimensional spin measurements which is proposed by D. Collins et al Collins et al. (2002) and W. Son et al Son et al. (2006); Bae and Son (2018). Especially, we analyze its Bell correlation in the spin-1 chain to identify its exotic quantum phase near the point of QPTs.

In particular, we choose 1D spin-1 XXZ model with the onsite anisotropy, that is one of the most studied spin-1 chain model. Higher-spin XXZ system have richer quantum phases in various different parameter regimes and certain parts of the system are need to be depicted with more complicated physics like symmetry-protected topological phases Kjäll et al. (2013). Since the ground state of spin-1 chains are not possible to be obtained analytically, we utilize numerical techniques, the density-matrix renormalization group (DMRG) method, for the ground state of spin-1 system which are described in matrix product state (MPS) representations. As a result, we found various QPTs that is accurately indicated by analyzing the derivatives of the CGLMP-SLK correlation : FM-to-XY, Haldane-to-AFM, Haldane-to-large-D, and XY1-to-XY2 quantum phase transitions. The first-order QPTs (FM-to-XY transitions) shows discontinuity of the CGLMP-SLK correlation. The Ising-type (Haldane-to-AFM and XY1-to-XY2) QPTs occur where the second derivative of the correlation vanishes. The Haldane-to-large-D transition, also referred to as the Gaussian phase transition, emerges when ^r/D=0\partial\langle\hat{\mathcal{B}}_{r}\rangle/\partial D=0. The criticality of the correlation is caused by the fact that both the derivatives of the spin-spin correlations Cr(1)C^{(1)}_{r} and Cr(2)C^{(2)}_{r} are simultaneously zero in the criticality. Moreover, we discuss whether violation of the CGLMP-SLK inequality can be detected even in three-dimensional systems.

The outline of this paper is given as following: In Sec.II, we starts with the system hamiltonian of the 1D spin-1 XXZ model with the onsite anisotropy. There, we briefly explains its phases at the different parameter regime and their criticalities at the QPTs. In sec.III, we presents the Bell correlation in the CGLMP-Bell inequality from the generalized SLK-Mermin-Bell formalism described in an operator form of nonlocal order parameter. In Sec.IV, we introduce the matrix product state representations for the ground state of the infinite-size system along with the iDMRG method. Then, using the state obtained, we present our results for the description of Bell correlations in overall parameter space and analysis in the vicinity of the criticality in Sec.V, and we conclude this paper in Sec.VI.

II The 1D spin-1 XXZ model with the on-site anisotropy

We now start with the description of the system to be analyzed. The Hamiltonian for the one-dimensional spin-1 XXZ model with the on-site anisotropy is of the form

H^=jJ(S^jxS^j+1x+S^jyS^j+1y+JzS^jzS^j+1z)+Dj(S^jz)2,\displaystyle\hat{H}=\sum_{j}J(\hat{S}^{x}_{j}\hat{S}^{x}_{j+1}+\hat{S}^{y}_{j}\hat{S}^{y}_{j+1}+J_{z}\hat{S}^{z}_{j}\hat{S}^{z}_{j+1})+D\sum_{j}(\hat{S}^{z}_{j})^{2}, (1)

where S^ja\hat{S}^{a}_{j} for a=x,y,za=x,y,z denotes the spin-1 operators at site jj. In the spin-z basis, the spin-1 operators are given by

S^x=12\displaystyle\hat{S}^{x}=\frac{1}{\sqrt{2}} (010101010),S^y=12i(010101010),\displaystyle\begin{pmatrix}0&1&0\\ 1&0&1\\ 0&1&0\end{pmatrix},\;\;\hat{S}^{y}=\frac{1}{\sqrt{2}i}\begin{pmatrix}0&1&0\\ -1&0&1\\ 0&-1&0\end{pmatrix},
S^z=(100000001).\displaystyle\;\;\;\;\;\;\;\hat{S}^{z}=\begin{pmatrix}1&0&0\\ 0&0&0\\ 0&0&-1\end{pmatrix}. (2)

The summation in Eq.(1) illustrates the one-dimensional spin-1 XXZ chain with the exchange anisotropy JzJ_{z} and we set J=1J=1 for the unit energy scale.

The second sum in Eq.(1) describes an on-site anisotropy where the constant D means the uniaxial single-site anisotropy. It has been known that this model can be in the various phases including the antiferromagnetic (AFM) phase, the ferromagnetic (FM) phase, XY phase, the Haldane phase, and the large-D phase. Additionally, the ground state of the system undergoes four types of the QPTs, as varying the parameters JzJ_{z} and DD. The ground-state phase diagram of the spin-1 chain Eq.(1) is investigated in the earlier study Chen et al. (2003). Here we now briefly demonstrate the characteristics of the different phases and QPTs in the given parameter space.

The AFM and FM phases :

The antiferromagnetic (AFM) and ferromagnetic (FM) phases describes the system with the magnetic order in it. The fact that all spins in the FM phase are aligned in the same direction gives rise to S^jz=S^j+1z\langle\hat{S}^{z}_{j}\rangle=\langle\hat{S}^{z}_{j+1}\rangle for all jj. The spins in FM phase can be quantified by the order parameter S^jz\langle\hat{S}^{z}_{j}\rangle. In the case of AFM phase, spins at the nearest neighbor site jj and j+1j+1 are aligned in the opposite directions and thereby they satisfies S^jz=S^j+1z\langle\hat{S}^{z}_{j}\rangle=-\langle\hat{S}^{z}_{j+1}\rangle.

The XY phase and BKT-type phase transition :

The XY phase is that all spins which interact to the nearest neighbors lie in the xyxy plane and are not allowed to point in the spin-zz direction. At a finite temperature TcT_{c}, Berezinskii, Kosterlitz, and Thouless (BKT) firstly depicted the phase transition without any symmetry breaking in the classical 2D XY model Berezinsky (1970); Kosterlitz and Thouless (1973); Kosterlitz (1974). While decaying exponentially above the temperature TcT_{c}, the spin-spin correlation at TTcT\leq T_{c} is expected to decay in a power law, S+Srη\langle S^{+}S^{-}\rangle\sim r^{-\eta} where 0<η1/40<\eta\leq 1/4. These behaviors for the classical XY model still hold for the quantum spin chains Schulz (1986). The XY phase in the 1D spin-SS XXZ model (Eq.(1) for any spin SS and D=0D=0) is prevalent for 1Jz0-1\leq J_{z}\leq 0 Kjäll et al. (2013); Alcaraz and Moreo (1992): the XY phase of the spin-1/21/2 XXZ chain covers the wider range 1Jz1-1\leq J_{z}\leq 1 while that of the spin-1 system belongs to the range 1Jz0-1\leq J_{z}\leq 0. This feature has an influence on the BKT-type quantum phase transition which it takes place at Jz=0J_{z}=0 for S=1S=1 Kitazawa et al. (1996) unlike at Jz=1J_{z}=1 for S=1/2S=1/2. In Eq.(1), two types of the XY phase are manifested Schulz (1986). In the XY1 phase, the spin-spin correlation S^+S^\langle\hat{S}^{+}\hat{S}^{-}\rangle displays a power-law decay as we mentioned above. For large negative DD, the XY2 phase appears where (S^+)2(S^)2\langle(\hat{S}^{+})^{2}(\hat{S}^{-})^{2}\rangle decays following the power law behavior.

The Haldane phase and the large-D phase :

A variety of phases in many-body systems can be discerned by local parameters that demonstrates the spontaneous symmetry breaking. However, it is the Haldane’s conjecture Haldane (1983a, b) that 1D chain of arbitrary integer spin has the gapped phases where any spontaneous symmetry breaking is not manifested. Since then, surveying the characteristics of the Haldane phases becomes significant topic to be investigated.

Later it has been known that the most renown order parameters to capture the Haldane phase is the string order parameter

𝒪strS^iαeiπl=i+1j1S^lαS^jα,\mathcal{O}_{str}\equiv\langle\hat{S}^{\alpha}_{i}e^{i\pi\sum_{l=i+1}^{j-1}\hat{S}^{\alpha}_{l}}\hat{S}^{\alpha}_{j}\rangle, (3)

for α{x,y,z}\alpha\in\{x,y,z\} in the presence of the Z2×Z2Z_{2}\times Z_{2} symmetry Nijs and Rommelse (1989); Kennedy and Tasaki (1992). It is remarkable that this quantity, an expectation value of the nonlocal operator has a nonzero value even in the thermodynamic limit 111Here, a word, nonlocal, is not the same as the Bell nonlocality, a violation of Bell inequality in Sec.III. It is also mentioned in a paper, N. Brunner et al., Rev. Mod. Phys. 86, 419 (2014).

Also, Affleck et al. suggested the exactly solvable spin-1 model, called the AKLT state, that exhibits the Haldane phase Affleck et al. (1987, 1988). The model is described by the Hamiltonian

H=jSjSj+1+13(SjSj+1)2H=\sum_{j}\vec{S}_{j}\cdot\vec{S}_{j+1}+\frac{1}{3}\left(\vec{S}_{j}\cdot\vec{S}_{j+1}\right)^{2} (4)

whose ground state tend to be valanced bonds state with four-fold edge degeneracies. The unexpected properties like edge states and a string order are investigated by using its exact ground state that has non-trivial phases originated from topological order in the system. On the course of the discussion, it is notable that, for the non-trivial phases, bi-quadratic terms are appeared in the interaction Hamiltonian as it is needed to identify the high order moments in the higher spin system.

At the D>>1D>>1 limit in Eq.(1), the ground state becomes large-D phase whose state takes a rather trivial product form |000|00\cdots 0\rangle. As it is argued in Gu and Wen (2009); Chen et al. (2011a, b), the state can be identified by the trivial phase factor from the projective representations of symmetry groups in the Schmidt-state basis. Unlikely to the large-D phase, Pollmann et al. demonstrate that the entanglement spectrum in the Haldane phase of spin-1 chains is doubly degenerated by using the projective representations in MPS language Pollmann et al. (2010, 2012).

III Bell-type correlation from the generalized nonlocality criteria

Bell-type inequalities examine the correlation between two parties of a composite quantum state and determine whether a given quantum state is nonlocal through the violation of local realistic bounds. Clauser-Horne-Shimony-Holt (CHSH) version of Bell inequality is given for the simplest symmetric states where two dichotomic measurements are performed for each of the parties Clauser et al. (1969). If the CHSH correlation admits the local hidden variable model, its absolute value does not exceed the local realistic bound 2. Taking pure entangled states into account, one can find violation of the CHSH inequality and its maximal violation 222\sqrt{2} derived by Tsirelson Cirel’son (1980). For the case of two-qubit mixed states, entangled ones do not always violates this inequality Werner (1989).

When it comes to the systems of arbitrary dimension dd, the inequality derived by Collinset. al. provides a generalized version of CHSH inequality Collins et al. (2002). Further generalization of CHSH for the multiparty dd-dimesional systems has been made by Son et. al. which is also consistent to the CGLMP Son et al. (2006); Bae and Son (2018). In these scenarios, all the parties are permitted to perform two different measurements at each site with dd-outcome measurements.

From here, the correlation on the left hand side of the inequality for d×dd\times d system is said to be CGLMP correlation as it reduced into the one found by CGMLP Collins et al. (2002). Our description is made not by set of probabilities but by the form of correlation functions. It can be proved that the representations are equivalent as long as an appropriate collection of the correlations is made through the right choices of weighting coefficients Bae and Son (2018).

In the scenario of the CGLMP inequality, there are two observers (j=1,2j=1,2) and each of them has two measurement choices Vj{Aj,Bj}V_{j}\in\{A_{j},B_{j}\}. The measurement operators for the jj-th party are given as

V^jα=0d1ωα|αVjα|,\hat{V}_{j}\equiv\sum^{d-1}_{\alpha=0}\omega^{\alpha}|\alpha\rangle_{V_{j}}\langle\alpha|, (5)

where dd is the index of the number of possible outcomes, ωα\omega^{\alpha} is the eigenvalue, and |αVj|\alpha\rangle_{V_{j}} is the corresponding eigenvector. One observer chooses the Fourier transformed basis and the other observer takes the inverse Fourier basis. The Fourier transformed basis is given as

|αVj=1dβ=0d1ω(α+ϕVj)β|β,\displaystyle|\alpha\rangle_{V_{j}}=\frac{1}{\sqrt{d}}\sum^{d-1}_{\beta=0}\omega^{-(\alpha+\phi_{V_{j}})\beta}|\beta\rangle, (6)

where ω=exp(2πi/d)\omega=\exp(2\pi i/d) and ϕVj\phi_{V_{j}} denotes a phase shift for a measurement V^j\hat{V}_{j}. Here, the basis |β|\beta\rangle is nothing but the computational basis like the spin-zz basis for d=2d=2.

As to be maximal tests, four different measurement choices can be made through four different phase shifts that are ϕA1=0\phi_{A_{1}}=0, ϕB1=1/2\phi_{B_{1}}=1/2,ϕA2=1/4\phi_{A_{2}}=-1/4, and ϕB2=1/4\phi_{B_{2}}=1/4. Based upon this expression, CGLMP correlation can be described by the expectation value of the following composite operator,

^=<n=1d1fn(A^1n+ωn/2B^1n)(A^2n\displaystyle\langle\hat{\mathcal{B}}\rangle=\Bigg{<}\sum^{d-1}_{n=1}f_{n}\left(\hat{A}^{n}_{1}+\omega^{n/2}\hat{B}^{n}_{1}\right)\otimes\left(\hat{A}^{n}_{2}\right. +ωn/2B^2n)>+c.c.\displaystyle+\left.\omega^{n/2}\hat{B}^{n}_{2}\right)^{\dagger}\Bigg{>}+c.c.

when the weight function becomes fn1/2d1ωn4sec(nπ/2d)f_{n}\equiv\frac{1/2}{d-1}\omega^{\frac{n}{4}}\sec(n\pi/2d). As it is addressed in Bae and Son (2018), the weight fnf_{n} can be specified for the convex sum of the probability distributions suggested by Collins et. al. Collins et al. (2002). Bell correlation Eq.(III) for the case of d=2d=2 is obviously equivalent to one in the CHSH inequality. Local realistic bounds of the CGLMP correlation for d=3d=3 are given by 4^2-4\leq\langle\hat{\mathcal{B}}\rangle\leq 2, which can be derived from fnf_{n} Bae and Son (2018).

It is convenient to investigate the structure of the operator ^\langle\hat{\mathcal{B}}\rangle in Eq.(III) by introducing the nn-level lowering operator J^jnβ=nd1|βjβn|\hat{J}^{n}_{j}\equiv\sum^{d-1}_{\beta=n}|\beta\rangle_{j}\langle\beta-n| for jj-th party. From Eq.(5) and Eq.(6), one can derive the identities such that

12(A^1n+ωn/2B^1n)\displaystyle\frac{1}{2}(\hat{A}^{n}_{1}+\omega^{n/2}\hat{B}^{n}_{1}) =J^1n,\displaystyle=\hat{J}^{n}_{1}, (7)
12(A^2n+ωn/2B^2n)\displaystyle\frac{1}{2}(\hat{A}^{n}_{2}+\omega^{n/2}\hat{B}^{n}_{2}) =ωn/4J^.n\displaystyle=\omega^{n/4}\hat{J}^{n}_{.} (8)

Thus, the operator ^\hat{\mathcal{B}} can be simplified as

^=2d1n=1d1sec[nπ2d](J^1nJ^2n+J^1nJ^2n).\displaystyle\hat{\mathcal{B}}=\frac{2}{d-1}\sum^{d-1}_{n=1}\sec\left[\frac{n\pi}{2d}\right]\left(\hat{J}^{n}_{1}\otimes\hat{J}^{n\dagger}_{2}+\hat{J}^{n\dagger}_{1}\otimes\hat{J}^{n}_{2}\right). (9)

This operator holds for the local basis of dimension dd. In this work, we analyze the spin-1 chain where the dimension of local basis is set to be three, i.e., d=3d=3. When the set of states {|β|β=0,1,2}\{|\beta\rangle\;|\;\beta=0,1,2\} is regarded as the computational basis, the matrix representation of J^\hat{J} and J^2\hat{J}^{2} is given by

J^=(000100010)andJ^2=(000000100)\displaystyle\hat{J}=\begin{pmatrix}0&0&0\\ 1&0&0\\ 0&1&0\end{pmatrix}\;\;\;\text{and}\;\;\;\hat{J}^{2}=\begin{pmatrix}0&0&0\\ 0&0&0\\ 1&0&0\end{pmatrix} (10)

where J^S^/2=(S^xiS^y)/2\hat{J}\equiv\hat{S}^{-}/\sqrt{2}=(\hat{S}^{x}-i\hat{S}^{y})/\sqrt{2} and J^2J^J^\hat{J}^{2}\equiv\hat{J}\cdot\hat{J}. The operators shift the higher spin state into the lower spin state in one step and two steps respectively. The choice of measurements AiA_{i} and BiB_{i} in Eq.(III) using the corresponding ϕAi\phi_{A_{i}} and ϕBi\phi_{B_{i}} lead to the annihilation operators of higher order and they identifies all the matrix elements of the given state. It can be found that the convex sum of the correlation function provides the symmetric selection of relevant probabilities for the correlations in the parties as it is addressed in Bae and Son (2018).

IV MPS Representation and iDMRG method

In the next step, it is required to investigate the correlation of ground state for the physical properties of many-body systems in Eq. (1). We utilize a matrix product state (MPS) representation, which helps to identify the Schmidt coefficients of the many-body pure state as they are in the product forms of the matrices Pérez-García et al. (2007); Schollwöck (2011); Orús (2014). A quantum state |Ψ|\Psi\rangle on the 1D systems of length NN can be written in canonical MPS form Vidal (2003):

|Ψ=s1,,sNΓs1[1]Λ[1]Γs2[2]Λ[N1]ΓsN[N]|s1sN,\displaystyle|\Psi\rangle=\sum_{s_{1},\cdots,s_{N}}\Gamma^{[1]}_{s_{1}}\Lambda^{[1]}\Gamma^{[2]}_{s_{2}}\cdots\Lambda^{[N-1]}\Gamma^{[N]}_{s_{N}}|s_{1}\cdots s_{N}\rangle, (11)

where Γsj[j]\Gamma^{[j]}_{s_{j}} is χj1×χj\chi_{j-1}\times\chi_{j} matrices and |sj|s_{j}\rangle is the jjth-site local basis of dimension dd with the physical indices sj=0,,d1s_{j}=0,\ldots,d-1. When the elements of matrices Γsj[j]\Gamma^{[j]}_{s_{j}} for any sjs_{j} are expressed as Γsj;μj1μj[j]\Gamma^{[j]}_{s_{j};\mu_{j-1}\mu_{j}}, we call the subscript μ\mu the bond index and its dimension χ\chi the bond dimension. The matrices Γ[1]\Gamma^{[1]} and Γ[N]\Gamma^{[N]} at the ends (j=1,Nj=1,N) become vectors whose dimensions are 1×χ11\times\chi_{1} and χN1×1\chi_{N-1}\times 1, respectively. In the case of the periodic boundary conditions, these matrices are of dimensions χN×χ1\chi_{N}\times\chi_{1} and χN1×χN\chi_{N-1}\times\chi_{N} and there is an additional Λ[N]\Lambda^{[N]} connected to them by summing over χN\chi_{N} Verstraete et al. (2004). The matrices Λ[j]\Lambda^{[j]} are real, diagonal, and of dimension χj×χj\chi_{j}\times\chi_{j}. These matrices should satisfy the relation such that

sjΓsj[j](Λ[j])2Γsj[j]=sjΓsj[j](Λ[j1])2Γsj[j]=𝟙,\displaystyle\sum_{s_{j}}\Gamma^{[j]}_{s_{j}}\left(\Lambda^{[j]}\right)^{2}\Gamma^{[j]\dagger}_{s_{j}}=\sum_{s_{j}}\Gamma^{[j]\dagger}_{s_{j}}\left(\Lambda^{[j-1]}\right)^{2}\Gamma^{[j]}_{s_{j}}=\mathds{1}, (12)

which stems from the normalization condition ψ|ψ=1\langle\psi|\psi\rangle=1. In this paper, we are interested in the 1D infinite chains. The infinite matrix product states (iMPS) can be represented by

|Ψ={𝐬}Tr[ΓsjΛΓsj+1Λ]|sjsj+1,\displaystyle|\Psi\rangle=\sum_{\{\mathbf{s}\}}\text{Tr}\left[\cdots\Gamma_{s_{j}}\Lambda\Gamma_{s_{j+1}}\Lambda\cdots\right]|\cdots s_{j}s_{j+1}\cdots\rangle, (13)

where the superscript [j][j] on each matrices is omitted since the boundary effect in the thermodynamic limit is insignificant and it is invariant under the translational symmetry.

The diagonal elements λμj\lambda_{\mu_{j}} in Λ[j]\Lambda^{[j]} are nothing but the Schmidt coefficients and signify the entanglement spectrum of two partitions where the one consist of spins from 11 to jj and the other ranges from j+1j+1 to NN. The Schmidt state of Eq.(11) can be written as

|Ψ=μj=0χj1λμj|μjL|μjR,\displaystyle|\Psi\rangle=\sum^{\chi_{j}-1}_{\mu_{j}=0}\lambda_{\mu_{j}}|\mu_{j}\rangle_{L}|\mu_{j}\rangle_{R}, (14)

where the nonnegative λμj\lambda_{\mu_{j}} satisfy the property μjλμj2=1\sum_{\mu_{j}}\lambda^{2}_{\mu_{j}}=1. The Schmidt bases of the left partition and the right partition are given as

|μjL\displaystyle|\mu_{j}\rangle_{L} =s1,,sjΓs1[1]Λ[1]Γsj[j]|s1sj,\displaystyle=\sum_{s_{1},\cdots,s_{j}}\Gamma^{[1]}_{s_{1}}\Lambda^{[1]}\cdots\Gamma^{[j]}_{s_{j}}|s_{1}\cdots s_{j}\rangle, (15)
|μjR\displaystyle|\mu_{j}\rangle_{R} =sj+1,,sNΓsj+1[j+1]Λ[j+1]ΓsN[N]|sj+1sN.\displaystyle=\sum_{s_{j+1},\cdots,s_{N}}\Gamma^{[j+1]}_{s_{j+1}}\Lambda^{[j+1]}\cdots\Gamma^{[N]}_{s_{N}}|s_{j+1}\cdots s_{N}\rangle. (16)

It is notable that the matrix Λ[j]\Lambda^{[j]} signifies the bipartite entanglement of two subsystems LL and RR. This is because the entanglement entropy is expressed as S(|Ψ)=μjλμj2logλμj2S(|\Psi\rangle)=-\sum_{\mu_{j}}\lambda^{2}_{\mu_{j}}\log\lambda^{2}_{\mu_{j}} from its definition.

For the 1D gapped systems, the ground state can be efficiently approximated by this MPS representation Verstraete and Cirac (2006) and the entanglement entropy of the ground state of a bipartite system LL and RR is finite according to the area law Hastings (2007). When the system lies in the vicinity of the quantum criticality, the entanglement entropy for the 1D infinite systems diverges logarithmically Vidal et al. (2003). That is, a lack amount of entanglement that comes from a small bond dimension χ\chi causes the inefficiency in the MPS representation to describe the ground state at the criticality. Nevertheless, the scaling properties of finite entanglement at the quantum criticality is suggested based on the conformal field theory Pollmann et al. (2009).

Refer to caption

(a)

Refer to caption

(b)

Figure 1: Graphical descriptions for the expectation value of (a) the operator ^i,i+r\hat{\mathcal{B}}_{i,i+r} Eq.(17) and (b) the spin-spin correlation (S^iα)n(S^i+rβ)n\langle(\hat{S}^{\alpha}_{i})^{n}(\hat{S}^{\beta}_{i+r})^{n}\rangle

In order to calculate the CGLMP correlation in spin chains which describes the two-body correlation among the many particles, the matrix product states (MPS) representation is applied for the depiction of the ground state and the matrix product operator (MPO) representation for the operator ^\hat{\mathcal{B}}. Together with the MPS representation for the ground state, the operator ^i,i+r\hat{\mathcal{B}}_{i,i+r} between parties ii and i+ri+r can be described by the MPO representation that reads

^i,i+r=𝒥^iL𝟙r1𝒥^i+rR,\displaystyle\hat{\mathcal{B}}_{i,i+r}=\hat{\mathcal{J}}^{L}_{i}\mathds{1}^{r-1}\hat{\mathcal{J}}^{R}_{i+r}, (17)

where

𝒥^iL=(23J^i23J^i2J^i22J^i2),\displaystyle\hat{\mathcal{J}}^{L}_{i}=\begin{pmatrix}\sqrt{\frac{2}{\sqrt{3}}}~{}~{}\hat{J}_{i}&\sqrt{\frac{2}{\sqrt{3}}}~{}~{}\hat{J}^{\dagger}_{i}&\sqrt{2}~{}~{}\hat{J}^{2}_{i}&\sqrt{2}~{}~{}\hat{J}^{\dagger 2}_{i}\end{pmatrix},
𝒥^i+rR=(23J^i+r23J^i+r2J^i+r22J^i+r2),\displaystyle\hat{\mathcal{J}}^{R}_{i+r}=\begin{pmatrix}\sqrt{\frac{2}{\sqrt{3}}}~{}~{}\hat{J}^{\dagger}_{i+r}\\ \sqrt{\frac{2}{\sqrt{3}}}~{}~{}\hat{J}_{i+r}\\ \sqrt{2}~{}~{}\hat{J}^{\dagger 2}_{i+r}\\ \sqrt{2}~{}~{}\hat{J}^{2}_{i+r}\end{pmatrix},

and 𝟙\mathds{1} denotes the identity matrix of dimension 4×44\times 4 as shown in Fig. 1(a). Here, 𝒥^L\hat{\mathcal{J}}^{L} and 𝒥^R\hat{\mathcal{J}}^{R} can be obtained by using the method according to Fröwis et al. Fröwis et al. (2010) and schematic representation of the operator expectation in MPS form is depicted in Fig.1.

iDMRG method

Density-matrix renormalization group (DMRG) method is one of the most widely used numerical methods in one-dimensional systems for infinite size as well as finite size systemsWhite (1992, 1993). Later, McCulloch showed its infinite version based on the MPO form of the Hamiltonian invariant under the translational symmetry McCulloch . Thus, it is sufficient to just consider an unit cell of two sites in the lattice to find the ground state. The main idea of infinite-size DMRG (iDMRG) method is that (i) the ground state for the two-site Hamiltonian is obtained, (ii) it is applied to update the environment and reconstruct the two-site Hamiltonian together with the updated environment, and (iii) this procedure is repeated until the singular-value matrix of the two-site ground state converges. During this process, the Lanczos method is applied to solve the eigenvalue problem for the two-site Hamiltonian and the singular value decomposition is used to obtain the two-site ground state in MPS representation. The specific procedure of the iDMRG algorithm is well-described in various previous works Schollwöck (2011); Hauschild and Pollmann (2018). Here, we simulate the iDMRG method and calculate some expectation values by using TeNPy Library (version 0.4.1) Hauschild and Pollmann (2018). We choose the bond dimension χ=200\chi=200 and specially confine a parameter for convergence such that the relative difference of entanglement entropy in each sweep is not over 10710^{-7}.

V Quantum criticalities through the correlation in non-locality tests

At first, let us present the CGLMP correlation ^i,i+r\langle\hat{\mathcal{B}}_{i,i+r}\rangle that is obtained by the expectation value of the operator ^\hat{\mathcal{B}} for the ground state of Eq.(1) in the thermodynamic limit. The function is plotted as varying the parameters JzJ_{z} and DD for the spins at different sites distanced rr. Here, the class of matrices Γsj\Gamma_{s_{j}} and Λ\Lambda for the ground state can be obtained from the iDMRG method mentioned in Sec.IV. The CGLMP correlation ^i,i+r\langle\hat{\mathcal{B}}_{i,i+r}\rangle is calculated through the summation of all tensors Γ\Gamma and Λ\Lambda with 𝒥^iL\hat{\mathcal{J}}^{L}_{i} and 𝒥^i+rR\hat{\mathcal{J}}^{R}_{i+r}. A graphical description for the evaluation of ^i,i+r\langle\hat{\mathcal{B}}_{i,i+r}\rangle is given in Fig.1. In the figures, the identity matrix 𝟙L\mathds{1}^{L} comes from the overlap αi1|αi1=δα,αL\langle\alpha^{\prime}_{i-1}|\alpha_{i-1}\rangle=\delta^{L}_{\alpha^{\prime},\alpha} in Eq.(12) and Eq.(15) and the same thing is obtained for 𝟙R\mathds{1}^{R}.

As a result, we here have obtained the characteristics of the CGLMP correlation ^\langle\hat{\mathcal{B}}\rangle for the particular set of fixed parameters D=0D=0, Jz=1J_{z}=1 and Jz=1J_{z}=-1. In the parameter region, various phases and QPTs have been identified. The criticalities can be clearly found through the non-analyticity of the correlation function in the first and the second order. Moreover the contribution of higher moment correlation for the criticalities is possible to be discriminated. However, we found that it is not possible to observe the violation of the CGLMP-type inequality in the ground state of Hamiltonian Eq.(1). They are presented in Fig.2 (a), Fig.3 (a) and Fig. 4 (a). Such non-violation of the CGMLP criteria is consistent to the cases of CHSH in the earlier studies of 1D spin-1/2 XY chain Batle and Casas (2010) and XXZ chain Justino and de Oliveira (2012).

The reduced density matrices for two sites in the model of translational symmetry are not possible to violate bipartite Bell inequality. Nevertheless, the CGLMP correlation ^r\langle\hat{\mathcal{B}}_{r}\rangle can be applied to elucidate the physical properties for the various quantum phases in the system of Eq.(1). One of the rationales for the versatile quantification is that Eq.(9) for d=3d=3 is nothing but the linear combination of spin-spin correlation functions in the higher moments where the nn-level lowering operator J^jn\hat{J}^{n}_{j} is regarded as the spin-1 lowering operator (S^)jn(\hat{S}^{-})^{n}_{j} in this model. It is, thus, given by

^r=2(13Cr(1)+12Cr(2)),\displaystyle\langle\hat{\mathcal{B}}_{r}\rangle=2\left(\frac{1}{\sqrt{3}}\,C^{(1)}_{r}+\frac{1}{2}C^{(2)}_{r}\right), (18)

where the transverse spin-spin correlations Cr(1)C^{(1)}_{r} and Cr(2)C^{(2)}_{r} are given by

Cr(1)S^i+S^i+randCr(2)(S^i+)2(S^i+r)2.\displaystyle C^{(1)}_{r}\equiv\langle\hat{S}^{+}_{i}\hat{S}^{-}_{i+r}\rangle~{}~{}~{}\mbox{and}~{}~{}~{}C^{(2)}_{r}\equiv\langle(\hat{S}^{+}_{i})^{2}(\hat{S}^{-}_{i+r})^{2}\rangle. (19)

In Eq.(18), we drop the index ii because these correlation functions depend not on the site ii but on the relative distance rr between two spins due to the translational symmetry in the thermodynamic limit. Based upon the MPS representation, we discuss below how the CGLMP correlation can be used as an indicator of the QPTs by identifying the extremum and inflection points Wolf et al. (2006). It is also discussed how the Bell-type correlations varies as the distance rr changed, especially at the point of QPT, in the following subsections.

Refer to caption
Refer to caption
Figure 2: (Color online) (a) the CGLMP correlation ^r\langle\hat{\mathcal{B}}_{r}\rangle and (b) the spin-spin correlations Cr(2)C^{(2)}_{r} and (c) Cr(2)C^{(2)}_{r} as varying the parameter JzJ_{z} at D=0D=0 for odd distances.

V.1 The CGLMP correlation at D=0D=0

Let us analyze our evaluation of the CGLMP correlation at D=0D=0 as varying the anisotropic exchange interaction JzJ_{z}. The CGLMP correlation is zero in the FM phase (Jz<1.0J_{z}<-1.0) which is trivial to be illustrated. A discontinuous change of ^r\langle\hat{\mathcal{B}}_{r}\rangle can be detected at Jz=1J_{z}=-1 where the first-order QPT occurs at the point. The spins in the AFM phase (Jz1.18J_{z}\gtrsim 1.18) are inclined to locate the spins alternatively in either |+|+\rangle or ||-\rangle site by site. A very large JzJ_{z} gives rise to nearly vanishing value of the transverse spin-spin correlations and the CGLMP correlation becomes consequently close to zero. The inflection point of the CGLMP correlation can be found at Jz=1.18J_{z}=1.18, which turns the AFM phase into different phases. This type of transition is similar to the criticality in the model of two-dimensional Ising interaction. Moreover, the critical point Jz=1.18J_{z}=1.18 agrees with known results from the previous analysis using the DMRG techniques Malvezzi et al. (2016); Heng Su et al. (2012); Ueda et al. (2008).

Refer to caption
Refer to caption
Refer to caption
Figure 3: (a) the CGLMP correlation ^r\langle\hat{\mathcal{B}}_{r}\rangle and the first derivative of ^r\langle\hat{\mathcal{B}}_{r}\rangle with respect to DD near (b) D=0.31D=-0.31 and (c) D=0.97D=0.97 as varying the parameter DD at Jz=1J_{z}=1 for odd distances.

The crucial region for the non-trivial phase is lied at the interval between the FM and the AFM phases. For spin-1/2 system, only XY phase exists in this interval, while there are the XY phase and the Haldane phase for the spin-1 systems. Therefore, more phase transition point exists between the XY phase and the Haldane phase whereas the transition type is going to be the BKT type QPT Schulz (1986). However, both Cr(1)C^{(1)}_{r} and Cr(2)C^{(2)}_{r} does not show the critical behaviour solely in the region because minimum or the inflection points are not appeared as it can be seen in Fig.2(b)-(c). Therefore, the CGLMP correlation does not identify the BKT-type QPTs near the point Jz=0J_{z}=0 at which the QPT in the spin-1 XXZ chain is expected to occur Chen et al. (2003). Instead of identifying the transition point from the CGLMP correlation directly, the BKT-type QPT can be indicated by an indirect manner such as vanishing string order parameter or a local maximum of the entanglement entropy Heng Su et al. (2012); Ueda et al. (2008). In this case, it is demonstrated that the correlation function Cr(1)C_{r}^{(1)} evaluated from iDMRG method fits a function arη+bar^{\eta}+b near Jz=0J_{z}=0. Through the extrapolation, the critical exponents ηc=1/4\eta_{c}=1/4 of the BKT-type QPT can be found at Jz=0.02J_{z}=0.02.

In comparison to the spin-1 case, it is intriguing that BKT-type QPT can be identified by the CHSH correlation in the spin-1/2 XXZ chain Justino and de Oliveira (2012). At Jz=1J_{z}=1, the spin-1/2 XXZ chain lies at BKT-type QPT and has the SU(2) symmetry. It is possible that the SU(2) symmetry, rather than criticality, is involved in the minimum of the CHSH correlation for Jz=1J_{z}=1. To be more specific, while S^ixS^i+rx<S^izS^i+rz\langle\hat{S}^{x}_{i}\hat{S}^{x}_{i+r}\rangle<\langle\hat{S}^{z}_{i}\hat{S}^{z}_{i+r}\rangle in the AFM phase (Jz>1J_{z}>1), S^ixS^i+rx>S^izS^i+rz\langle\hat{S}^{x}_{i}\hat{S}^{x}_{i+r}\rangle>\langle\hat{S}^{z}_{i}\hat{S}^{z}_{i+r}\rangle in the XY phase (Jz<1J_{z}<1). Due to the SU(2) symmetry, S^ixS^i+rx=S^izS^i+rz\langle\hat{S}^{x}_{i}\hat{S}^{x}_{i+r}\rangle=\langle\hat{S}^{z}_{i}\hat{S}^{z}_{i+r}\rangle for Jz=1J_{z}=1, which affects the maximization of the CHSH correlation with respect to the measurement directions in each phase.

Let us discuss the CGLMP correlation ^r\langle\hat{\mathcal{B}}_{r}\rangle as a function of the distance rr between two sites. The decaying behavior of the CGLMP correlations are affected by that of the spin-spin correlations Cr(1)C^{(1)}_{r} and Cr(2)C^{(2)}_{r}. The spin-spin correlation Cr(1)C^{(1)}_{r}, rather than Cr(2)C^{(2)}_{r} in Fig.2(b), dominates ^r\langle\hat{\mathcal{B}}_{r}\rangle without the onsite anisotropy. Then, the CGLMP correlation as a function of rr exhibits the even-odd oscillation. Notable feature of the CGLMP correlation is shown in the XY1 phase: ^r\langle\hat{\mathcal{B}}_{r}\rangle still has a finite value due to a power-law decay of Cr(1)C^{(1)}_{r}, whereas in other phases it drastically approaches to zero at a large distance rr (see Fig.2). For the Haldane phase and the AFM phase, the CGLMP correlations ^r\langle\hat{\mathcal{B}}_{r}\rangle are also characterized by an exponential decay, which results from the behavior of the spin-spin correlations Cr(1)C^{(1)}_{r} and Cr(2)C^{(2)}_{r}. In Fig.2(a), CGLMP correlation in the Haldane phase decays less slowly than the AFM phase. These features corresponds to the fact that the spin-spin correlation function decays exponentially in the AFM phase and one also has an exponential decaying with an additional factor r\sqrt{r} in the Haldane phase, i.e. Cr(1)(1)rr1/2exp(r/ξ)C^{(1)}_{r}\sim(-1)^{r}r^{-1/2}\exp(-r/\xi) at the Heisenberg point (Jz=1J_{z}=1 and D=0D=0) Haldane (1983b); Nomura (1989). The qualitative forms of the correlation functions Cr(1)C^{(1)}_{r} and Cr(2)C^{(2)}_{r} including various exponents can be expected by using the bosonization technique, an analytical approach based on the effective field theory with defects Schulz (1986) and its exponent η\eta and correlation lengths ξ\xi can have precise values by using the DMRG technique Heng Su et al. (2012); White and Huse (1993).

Refer to caption
Refer to caption
Refer to caption
Figure 4: (a) the CGLMP correlation ^r\langle\hat{\mathcal{B}}_{r}\rangle, (b) the spin-spin correlations Cr(1)C^{(1)}_{r} and Cr(2)C^{(2)}_{r} as varying the parameter DD, and (c) the first derivative of ^r\langle\hat{\mathcal{B}}_{r}\rangle with respect to DD at Jz=0.1J_{z}=-0.1 for odd distances rr.

V.2 The CGLMP correlation at Jz=1J_{z}=1

The Hamiltonian in Eq.(1) goes through three different phases by tuning the single-ion anisotropy DD at Jz=1J_{z}=1: the AFM, the Haldane, and the large-D phase. In addition, there are two types of QPTs that corresponds to an extremum point (either minimum for an odd rr or maximum for an even rr) and an inflection point of ^r\langle\hat{\mathcal{B}}_{r}\rangle for each rr in Fig.3(a). These critical points are influenced by ones for both correlations Cr(1)C^{(1)}_{r} and Cr(2)C^{(2)}_{r}. In the previous section, the characteristics of the Haldane and the AFM phase have been discussed and for Jz=1J_{z}=1 we can also detect the Ising-type QPT at Dc10.31D_{c_{1}}\simeq-0.31 between them, a peak of ^r/D\partial\langle\hat{\mathcal{B}}_{r}\rangle/\partial D for small rr compared with the fidelity Tzeng and Yang (2008). So, we are going to concentrate on the characteristics of the Haldane phase, and its relationship with the large-D phase. It is noteworthy in Fig.3(a) that the phase transition between the Haldane phase and the large-D phase, often called the Gaussian phase transition.

The string order parameter, consisting of spin operators, has a nonzero value only in the Haldane phase and thereby is widely used to discriminate the Haldane and large-D phase in this model. Later, the Haldane phase and the large-D phase are featured by a nontrivial topological phase and a trivial one under the symmetry protection, respectively Gu and Wen (2009); Pollmann et al. (2012). It is suggested that the order parameter which reflects the symmetry-protected topological order in the presence of Z2×Z2Z_{2}\times Z_{2} symmetry becomes -1 for the Haldane and 1 for the large-D phase Pollmann et al. (2012). Furthermore, other various studies have revealed the boundary of the Haldane and large-D phases. The structure for the energy levels has been mainly analyzed to capture the accurate transition point Dc2=0.975D_{c_{2}}=0.975 Ueda et al. (2008). and also, the behaviors of the ground-energy state have been treated by investigating the fidelity (Dc2=0.97D_{c_{2}}=0.97) Tzeng and Yang (2008), the entanglement entropy (Dc2=0.96845(8)D_{c_{2}}=0.96845(8)) Hu et al. (2011), which they all use the DMRG method.

In our simulation (see Fig. 3(a)), Dc2D_{c_{2}} the minimum points of ^r\langle\hat{\mathcal{B}}_{r}\rangle which come from those of spin-spin correlations Cr(1)C^{(1)}_{r} and Cr(2)C^{(2)}_{r} are captured around the phase transition between the Haldane and the large-D phase. For r=1r=1, we can find the critical point Dc2=0.914D_{c_{2}}=0.914 by evaluating the minimum of ^r=1\langle\hat{\mathcal{B}}_{r=1}\rangle. This critical point moves to Dc2=0.968D_{c_{2}}=0.968 at r=81r=81 by using the linear interpolation near the criticality in Fig.3(c). This critical point converges to Dc2=0.969D_{c_{2}}=0.969 until the CGLMP correlation decays to zero as the distance rr goes to infinity in Fig.3(c). Thus, ^r\langle\hat{\mathcal{B}}_{r}\rangle deserves as a indicator of the QPT for sufficiently distant parties. It is also remarkable that peak shape in Fig.3(a) remains on the criticality even at a large distance rr. Specifically, the CGLMP correlation ^r\langle\hat{\mathcal{B}}_{r}\rangle as a function of rr decays more slowly in the vicinity of the critical point than in the noncritical region. It is because the correlation functions Cr(1)C^{(1)}_{r} and Cr(2)C^{(2)}_{r} decays as a power law near the criticality, whereas it decays exponentially in the Haldane and the large-D phase Schulz (1986); Nijs and Rommelse (1989) and the exponents are well-approximated by using DMRG method Ueda et al. (2008); Hu et al. (2011).

V.3 The CGLMP correlation at Jz=0.1J_{z}=-0.1

The spin-1 Hamiltonian Eq.(1) involves two different types of XY phases, the XY1 phase and XY2 phases Schulz (1986). The former which we have been discuss above is located at small DD and the latter lies along the large negative DD.

In order to investigate the CGLMP correlation in two types of XY phases and their transition, let us take the anisotropy JzJ_{z} to be 0.1-0.1 and change the on-site anisotropy DD. It is known that for large negative DD, decay of the correlation function Cr(2)C^{(2)}_{r} obeys power laws while the correlation Cr(1)C^{(1)}_{r} decays exponentially, which coincides with the XY2 phase Schulz (1986). In Fig.4(c), we can detect the Ising-type criticality between the XY1 and XY2 phases Dc2.10D_{c}\simeq-2.10 by solving 2^r/D2=0\partial^{2}\langle\hat{\mathcal{B}}_{r}\rangle/\partial D^{2}=0. The farther the distance rr is, the deeper the peak of ^r/D\partial\langle\hat{\mathcal{B}}_{r}\rangle/\partial D is in the vicinity of this criticality. This detection quite exactly corresponds to the level crossing point of two energy gaps, one of which consists of the excitation (or the quantum number) Mz=±2M^{z}=\pm 2 in the XY2 phase and the other has Mz=±1M^{z}=\pm 1 in the XY1 phase Chen et al. (2003).

VI Conclusion

We have investigated the CGLMP correlations in the 1D XXZ model with onsite anisotropy especially near the quantum phase transitions. In the case of spin-1, the CGLMP correlation obtained from local measurements in the Fourier bases can be interestingly interpreted as a linear combination of the first-order and second-order transverse spin-spin correlations. By analyzing the derivatives of the CGLMP correlation, one can accurately indicate various QPTs: FM-to-XY, Haldane-to-AFM, Haldane-to-large-D, and XY1-to-XY2 quantum phase transitions. The first-order QPTs (FM-to-XY transitions) shows discontinuity of the CGLMP correlation. The Ising-type (Haldane-to-AFM and XY1-to-XY2) QPTs occur where the second derivative of the CGLMP correlation vanishes. The Haldane-to-large-D transition, also referred to as the Gaussian phase transition, emerges when ^r/D=0\partial\langle\hat{\mathcal{B}}_{r}\rangle/\partial D=0. The criticality of the CGLMP correlation is caused by the fact that both the derivatives of the spin-spin correlations Cr(1)C^{(1)}_{r} and Cr(2)C^{(2)}_{r} are simultaneously zero in the criticality. Moreover, decaying behaviors of the CGLMP correlation with respect to the distance rr in each phase can be predicted from the field-theoretical approach Schulz (1986). However, the BKT-type QPT at Jz=0J_{z}=0 cannot be indicated by finding critical points of the derivatives of the CGLMP correlation. This is contrast to the CHSH correlation in spin-1/2 XXZ chain Justino and de Oliveira (2012).

Another crucial result is that there is unfortunately nonviolation of CGLMP inequality in all phases of spin-1 chain, that is, the CGLMP correlations for the ground states of spin-1 chain do not surpass the local realistic bounds. Even though we do not optimize the measurements for the ground states in each phase, this nonviolation coincides with the one in the 1D spin-1/2 XY chain Batle and Casas (2010) and XXZ chain Justino and de Oliveira (2012) invariant under translational symmetry. The reduced density matrices on two sites which display mixed entangled states are not sufficient to violate CGLMP inequality in the spin-1 chain, although many-body ground states are pure and highly entangled in this model. As the CHSH correlation has monogamy trade-off relation de Oliveira et al. (2013), monogamy of CGLMP correlation can be a candidate to explain the non-violation of CGLMP inequality in the spin-1 chain. Therefore, in order to detect nonlocality in this model, one should consider multipartite Bell correlations for arbitrary dimensions. This gets a clue from the result that multipartite Bell correlations for spin-1/2 invariant under the permutational symmetry Tura et al. (2014) do not satisfy the local hidden variable model in the Ising model with long-range interactions Schmied et al. (2016); Piga et al. (2019). We leave that part as a future investigation that is quite consistent to our current approach demonstrated in this work.

Acknowledgements.
We thank Y. Jo and K. Bae for useful discussions. This work was supported by the National Research Council of Science and Technology (NST) (Grant No. CAP-15-08-KRISS), NRF and Samsung.

References

  • De Chiara and Sanpera (2018) G. De Chiara and A. Sanpera, Rep. Prog. Phys. 81, 074002 (2018).
  • Amico et al. (2008) L. Amico, R. Fazio, A. Osterloh,  and V. Vedral, Rev. Mod. Phys. 80, 517 (2008).
  • Osterloh et al. (2002) A. Osterloh, L. Amico, G. Falci,  and R. Fazio, Nature 416, 608 (2002).
  • Osborne and Nielsen (2002) T. J. Osborne and M. A. Nielsen, Phys. Rev. A 66, 032110 (2002).
  • Eisert et al. (2010) J. Eisert, M. Cramer,  and M. B. Plenio, Rev. Mod. Phys 82, 277 (2010).
  • Vidal et al. (2003) G. Vidal, J. I. Latorre, E. Rico,  and A. Kitaev, Phys. Rev. Lett. 90, 227902 (2003).
  • Son et al. (2009) W. Son, L. Amico, F. Plastina,  and V. Vedral, Phys. Rev. A 79, 022302 (2009)0807.1602 .
  • Capasso et al. (1973) V. Capasso, D. Fortunato,  and F. Selleri, Int. J. Theor. Phys. 7, 319 (1973).
  • Werner (1989) R. F. Werner, Phys. Rev. A 40, 4277 (1989).
  • Augusiak et al. (2015) R. Augusiak, M. Demianowicz, J. Tura,  and A. Acín, Phys. Rev. Lett. 115, 030404 (2015).
  • Batle and Casas (2010) J. Batle and M. Casas, Phys. Rev. A 82, 062101 (2010).
  • Justino and de Oliveira (2012) L. Justino and T. R. de Oliveira, Phys. Rev. A 85, 052128 (2012).
  • de Oliveira et al. (2013) T. R. de Oliveira, A. Saguia,  and M. S. Sarandy, Europhys. Lett. 100, 60004 (2013).
  • Sun et al. (2014a) Z.-Y. Sun, Y.-Y. Wu, J. Xu, H.-L. Huang, B.-F. Zhan, B. Wang,  and C.-B. Duan, Phys. Rev. A 89, 022101 (2014a).
  • Sun et al. (2014b) Z.-Y. Sun, S. Liu, H.-L. Huang, D. Zhang, Y.-Y. Wu, J. Xu, B.-F. Zhan, H.-G. Cheng, C.-B. Duan,  and B. Wang, Phys. Rev. A 90, 062129 (2014b).
  • Bao et al. (2020) J. Bao, B. Guo, H.-G. Cheng, M. Zhou, J. Fu, Y.-C. Deng,  and Z.-Y. Sun, Phys. Rev. A 101, 012110 (2020).
  • Tura et al. (2014) J. Tura, R. Augusiak, A. B. Sainz, T. Vértesi, M. Lewenstein,  and A. Acín, Science 344, 1256 (2014).
  • Tura et al. (2015) J. Tura, R. Augusiak, A. B. Sainz, B. Lücke, C. Klempt, M. Lewenstein,  and A. Acín, Ann. Phys. 362, 370 (2015).
  • Schmied et al. (2016) R. Schmied, J.-D. Bancal, B. Allard, M. Fadel, V. Scarani, P. Treutlein,  and N. Sangouard, Science 352, 441 (2016).
  • Piga et al. (2019) A. Piga, A. Aloy, M. Lewenstein,  and I. Frérot, Phys. Rev. Lett. 123, 170604 (2019).
  • Tzeng and Yang (2008) Y.-C. Tzeng and M.-F. Yang, Phys. Rev. A 77, 012311 (2008).
  • Goli et al. (2013) V. M. L. D. P. Goli, S. Sahoo, S. Ramasesha,  and D. Sen, J. Phys: Condens. Matter 25, 125603 (2013)1208.0706 .
  • Malvezzi et al. (2016) A. L. Malvezzi, G. Karpat, B. Çakmak, F. F. Fanchini, T. Debarba,  and R. O. Vianna, Phys. Rev. B 93, 184428 (2016).
  • Collins et al. (2002) D. Collins, N. Gisin, N. Linden, S. Massar,  and S. Popescu, Phys. Rev. Lett. 88, 040404 (2002).
  • Son et al. (2006) W. Son, J. Lee,  and M. S. Kim, Phys. Rev. Lett. 96, 060406 (2006)quant-ph/0503164 .
  • Bae and Son (2018) K. Bae and W. Son, Phys. Rev. A 98, 022116 (2018).
  • Kjäll et al. (2013) J. A. Kjäll, M. P. Zaletel, R. S. K. Mong, J. H. Bardarson,  and F. Pollmann, Phys. Rev. B 87, 235106 (2013).
  • Chen et al. (2003) W. Chen, K. Hida,  and B. C. Sanctuary, Phys. Rev. B 67, 104401 (2003).
  • Berezinsky (1970) V. Berezinsky, Zh. Eksp. Teor. Fiz. 32, 493 (1970).
  • Kosterlitz and Thouless (1973) J. M. Kosterlitz and D. J. Thouless, J. Phys. C 6, 1181 (1973).
  • Kosterlitz (1974) J. M. Kosterlitz, J. Phys. C 7, 1046 (1974).
  • Schulz (1986) H. J. Schulz, Phys. Rev. B 34, 6372 (1986).
  • Alcaraz and Moreo (1992) F. C. Alcaraz and A. Moreo, Phys. Rev. B 46, 2896 (1992).
  • Kitazawa et al. (1996) A. Kitazawa, K. Nomura,  and K. Okamoto, Phys. Rev. Lett. 76, 4038 (1996).
  • Haldane (1983a) F. D. M. Haldane, Phys. Rev. Lett. 50, 1153 (1983a).
  • Haldane (1983b) F. D. M. Haldane, Phys. Lett. A 93, 464 (1983b).
  • Nijs and Rommelse (1989) M. d. Nijs and K. Rommelse, Phys. Rev. B 40, 4709 (1989).
  • Kennedy and Tasaki (1992) T. Kennedy and H. Tasaki, Phys. Rev. B 45, 304 (1992).
  • Note (1) Here, a word, nonlocal, is not the same as the Bell nonlocality, a violation of Bell inequality in Sec.III. It is also mentioned in a paper, N. Brunner et al., Rev. Mod. Phys. 86, 419 (2014).
  • Affleck et al. (1987) I. Affleck, T. Kennedy, E. H. Lieb,  and H. Tasaki, Phys. Rev. Lett. 59, 799 (1987).
  • Affleck et al. (1988) I. Affleck, T. Kennedy, E. H. Lieb,  and H. Tasaki, Commun. Math. Phys. 115, 477 (1988).
  • Gu and Wen (2009) Z.-C. Gu and X.-G. Wen, Phys. Rev. B 80, 155131 (2009).
  • Chen et al. (2011a) X. Chen, Z.-C. Gu,  and X.-G. Wen, Phys. Rev. B 83, 035107 (2011a).
  • Chen et al. (2011b) X. Chen, Z.-C. Gu,  and X.-G. Wen, Phys. Rev. B 84, 235128 (2011b).
  • Pollmann et al. (2010) F. Pollmann, A. M. Turner, E. Berg,  and M. Oshikawa, Phys. Rev. B 81, 064439 (2010).
  • Pollmann et al. (2012) F. Pollmann, E. Berg, A. M. Turner,  and M. Oshikawa, Phys. Rev. B 85, 075125 (2012).
  • Clauser et al. (1969) J. F. Clauser, M. A. Horne, A. Shimony,  and R. A. Holt, Phys. Rev. Lett. 23, 880 (1969).
  • Cirel’son (1980) B. S. Cirel’son, Lett. Math. Phys. 4, 93 (1980).
  • Pérez-García et al. (2007) D. Pérez-García, F. Verstraete, M. M. Wolf,  and J. I. Cirac, Quantum Info. Comput. 7, 401 (2007).
  • Schollwöck (2011) U. Schollwöck, Ann. Phys. 326, 96 (2011).
  • Orús (2014) R. Orús, Ann. Phys. 349, 117 (2014).
  • Vidal (2003) G. Vidal, Phys. Rev. Lett. 91, 147902 (2003).
  • Verstraete et al. (2004) F. Verstraete, D. Porras,  and J. I. Cirac, Phys. Rev. Lett. 93, 227205 (2004).
  • Verstraete and Cirac (2006) F. Verstraete and J. I. Cirac, Phys. Rev. B 73, 094423 (2006).
  • Hastings (2007) M. B. Hastings, J. Stat. Mech.: Theory Exp. 2007, P08024 (2007).
  • Pollmann et al. (2009) F. Pollmann, S. Mukerjee, A. M. Turner,  and J. E. Moore, Phys. Rev. Lett. 102, 255701 (2009).
  • Fröwis et al. (2010) F. Fröwis, V. Nebendahl,  and W. Dür, Phys. Rev. A 81, 062337 (2010).
  • White (1992) S. R. White, Phys. Rev. Lett. 69, 2863 (1992).
  • White (1993) S. R. White, Phys. Rev. B 48, 10345 (1993).
  • (60) I. P. McCulloch, arXiv:0804.2509 .
  • Hauschild and Pollmann (2018) J. Hauschild and F. Pollmann, SciPost Phys. Lect. Notes , 5 (2018), code available from https://github.com/tenpy/tenpyarXiv:1805.00055 .
  • Wolf et al. (2006) M. M. Wolf, G. Ortiz, F. Verstraete,  and J. I. Cirac, Phys. Rev. Lett. 97, 110403 (2006).
  • Heng Su et al. (2012) Y. Heng Su, S. Young Cho, B. Li, H.-L. Wang,  and H.-Q. Zhou, J. Phys. Soc. Jpn 81, 074003 (2012).
  • Ueda et al. (2008) H. Ueda, H. Nakano,  and K. Kusakabe, Phys. Rev. B 78, 224202 (2008).
  • Nomura (1989) K. Nomura, Phys. Rev. B 40, 2421 (1989).
  • White and Huse (1993) S. R. White and D. A. Huse, Phys. Rev. B 48, 3844 (1993).
  • Hu et al. (2011) S. Hu, B. Normand, X. Wang,  and L. Yu, Phys. Rev. B 84, 220402 (2011).