Local quantum Fisher information and quantum correlation in the mixed-spin Heisenberg XXZ chain

Peng-Fei Wei, Qi Luo, Huang-Qiu-Chen Wang, Shao-Jie Xiong, Bo Liu, Zhe Sun

Front. Phys. ›› 2024, Vol. 19 ›› Issue (2) : 21201.

PDF(6821 KB)
Front. Phys. All Journals
PDF(6821 KB)
Front. Phys. ›› 2024, Vol. 19 ›› Issue (2) : 21201. DOI: 10.1007/s11467-023-1336-9
RESEARCH ARTICLE
RESEARCH ARTICLE

Local quantum Fisher information and quantum correlation in the mixed-spin Heisenberg XXZ chain

Author information +
History +

Abstract

We study the local quantum Fisher information (LQFI) in the mixed-spin Heisenberg XXZ chain. Both the maximal and minimal LQFI are studied and the former is essential to determine the accuracy of the quantum parameter estimation, the latter can be well used to characterize the discord-type quantum correlations. We investigate the effects of the temperature and the anisotropy parameter on the maximal LQFI and thus on the accuracy of the parameter estimation. Then we make use of the minimal LQFI to study the discord-type correlations of different site pairs. Different dimensions of the subsystems cause different values of the minimal LQFI which reflects the asymmetry of the discord-type correlation. In addition, the site pairs at different positions of the spin chains have different minimal LQFI, which reveals the influence of the surrounding spins on the bipartite quantum correlation. Our results show that the LQFI obtained through a simple calculation process provides a convenient way to investigate the discord-type correlation in high-dimensional systems.

Graphical abstract

Keywords

local quantum Fisher information / quantum correlation / mixed-spin Heisenberg XXZ chain

Cite this article

Download citation ▾
Peng-Fei Wei, Qi Luo, Huang-Qiu-Chen Wang, Shao-Jie Xiong, Bo Liu, Zhe Sun. Local quantum Fisher information and quantum correlation in the mixed-spin Heisenberg XXZ chain. Front. Phys., 2024, 19(2): 21201 https://doi.org/10.1007/s11467-023-1336-9

1 Introduction

Quantum metrology, as a rapidly developing field of quantum technology, can achieve higher precision and sensitivity measurements than classical proposals. According to the quantum Cramér−Rao theorem, the quantum Fisher information (QFI) lies at the heart of quantum metrology, which provides the lower bound on the variance of an unbiased estimator. In recent years, various proposals of quantum metrology and quantum parameter estimation have been successfully implemented in different physical systems [1-12]. In addition to quantum metrology, the QFI has been well applied in other important issues, such as entanglement witness [13-15], quantum phase transition detection [16,17], and non-Markovianity measure [18].
Recently, the concept of local quantum Fisher information (LQFI) is proposed in Ref. [19], which defines the QFI of a bipartite system while the parameter to be estimated is introduced via a local unitary evolution. The minimum LQFI is found to be able to naturally quantify the guaranteed sensitivity that a bipartite probe state allows in an interferometric configuration. It can also be used to quantify the discord-type quantum correlation and provide an experimental demonstration of the usefulness of discord in sensing applications. In Ref. [20], the minimum LQFI is used to characterize the non-Markovianity of open systems in local decoherence channels. In Ref. [21], LQFI is used to describe the quantum-correlation dynamics of two non-interacting qubits driven by a single classical field pattern with random phases. In Ref. [22], LQFI is used to analyze the influence of cavity dissipation and spontaneous emission on two-magnon dynamics for different magnon−magnon and photon−magnon couplings. Then, the LQFI was investigated in the anisotropic XY Heisenberg spin chain [23], where the LQFI was found to depend on the temperature and the coupling parameter. The LQFI was also studied in the Heisenberg XYZ chain with Dzyaloshinskii−Moriya interaction [24-28], where the authors found the Dzyaloshinskii−Moriya interaction can enhance the value of LQFI. However, most of the aforementioned works mainly focused on the two-site spin-1/2 Heisenberg chains. There is currently a lack of research on the situations of high-dimensional systems, e.g., the spin length is larger than 1/2 and the total site number is larger than two. In this work we for the first time consider the LQFI in the (1/2,S) mixed-spin chain, then the LQFI corresponding to the spin-S subsystem will present different behaviors from the spin-1/2 subsystem, and which can reveal the asymmetry of the discord-type correlation. Moreover, by increasing the total site number of the spin chain, the LQFI of different site pairs can be considered, which will exhibit the influence of the surrounding sites on the LQFI.
This paper is organized as follows. In Section 2, we introduce the definition of the LQFI and the calculation method of the maximal and minimal LQFI. In Section 3, we investigate the LQFI in the two-site (1/2,S) mixed-spin XXZ Heisenberg chain and analytically calculate both of the maximal and minimal LQFI. In Section 4, we numerically study the relationship between the maximal LQFI of the subsystem and the temperature T as well as the anisotropy parameter Jz. In Section 5, we compare the quantum discord with the minimal and maximal LQFI and further consider the minimal LQFI of different site pairs. The (1/2,S) mixed-spin cases with S=1,3/2,2 are considered and the total size of the spin chain is enlarged to 6 sites. In Section 6, we add the second and third nearest-neighbor coupling and discuss the influence on the minimum LQFI. Conclusions are given in Section 7.

2 Local quantum Fisher information

The parameterized state ρθ can be expressed by ρθ=UθρUθ, where the initial state ρ=i=1Dλi|ψiψi| does not hold the parameter θ, and Uθ is the unitary operator to lead into the parameter θ. Here, D is the dimension of the support set of ρ, λi and |ψi are the i-th eigenvalue of ρ and the corresponding eigenstate, respectively. Then QFI is expressed as [29,30]
F=i=1D4λiψi|H2|ψii,j=1D8λiλjλi+λj|ψi|H|ψj|2.
Here, H:=i(θUθ)Uθ is a Hermitian operator.
In this paper, we consider a bipartite state ρAB in the Hilbert space HANAHBNB, where NA and NB denote the dimensions of the subsystems A and B, respectively. Assuming the dynamic evolution of the subsystem A is UA=eiθH~A with the local Hamiltonian H~A=HAIB, the LQFI of the subsystem A is written as
F(ρAB,HA)=4Tr(ρH~A2)i,j=1D8λiλjλi+λj|ψi|H~A|ψj|2.
For a 2×NB-dimensional bipartite state, the local Hamiltonian is chosen as HA=xs, where s=(σx,σy,σz)/2 with the Pauli matrices {σx,y,z}, and x=(x1,x2,x3) with |x|=1. Then, LQFI of subsystem A can be rewritten as
F(ρAB,HA)=1m,n=13i,j=1D2λiλjλi+λjψi|σmI|ψj×ψj|σnI|ψixmxn.
Here, λi(j) are the eigenvalues of density matrix ρAB with the corresponding eigenvectors |ψi(j).
The LQFI of the subsystem A is rewritten as
F(ρAB,HA)=xTWAx.
Here, the matrix WA is a real symmetric matrix, whose elements are
WmnA=δmni,j=1D2λiλjλi+λjψi|σmI|ψjψj|σnI|ψi,
where δmn=1 (for m=n) and δmn=0 (for mn). We use μmax and μmin to represent the maximum and minimal eigenvalues of WA, respectively. Thus, we can define the maximal and the minimal LQFI of the subsystem A as
QAmax=maxHAF=μmax,QAmin=minHAF=μmin.
Let us consider the subsystem B. The unitary evolution is UB=eiθH~B with the local Hamiltonian H~B=IAHB. The local Hamiltonian is HB=yS, where S is the spin operator with the spin length S>1/2 and also let |y|=1. Similarly, the LQFI of the subsystem B is
F(ρAB,HB)=yTWBy.
Here, WB is also a real symmetric matrix with the elements as
WmnB=i=1D2λiψi|I(SmSn+SnSm)|ψii,j=1D8λiλjλi+λjψi|ISm|ψjψj|ISn|ψi.
Similarly, the maximal and the minimal LQFI of the subsystem B (i.e., QBmax and QBmin) can be obtained by the maximal and minimal eigenvalues of the matrix WB.

3 LQFI in Heisenberg model

The LQFI in the Heisenberg spin model has been widely studied [23-28]. In this paper, we consider the (1/2, S) mixed-spin Heisenberg XXZ chain with only the nearest-neighbor interaction under the open boundary condition. The system consists of two kinds of spins with different spin lengths alternating on a chain. Its Hamiltonian components are
Hα=i=1n/2(siαSiα+Siαsi+1α),(α=x,y,z)H=Hx+Hy+JzHz,
where the spin length of the spins s=1/2 and S>1/2. Jz is the anisotropy parameter and n is the number of sites (n is chosen as even numbers in this work). The open boundary condition is considered, i.e., we set sn/2+1α=0.
Let us first focus on the two-site case, i.e., n=2, and the spin-1/2 and spin-S are denoted by the subsystems A and B, respectively. Then the density matrix of this system in thermal equilibrium can be described by the Gibbs state at the temperature T, i.e., ρ(T)=exp(βH)/Z with the partition function ZTr[exp(βH)], and β=1/(kBT) (here the Boltzmann constant is set as kB=1). When we choose the eigenstates of the operator sz+Sz to be the basis, i.e., {|1/2,1, |1/2,0, |1/2,1, |1/2,1, |1/2,0, |1/2,1}, the eigenvalues of the Hamiltonian (9) can be easily obtained as E1=E2=Jz/2, E3=E4=(Jz+Jz2+8)/4, and E5=E6=(JzJz2+8)/4. Based on the definition in Eqs. (5) and (8), one can find that the off-diagonal matrix elements of WA and WB for the subsystems A and B are zero and the diagonal matrix elements are obtained as
W11(22)A=18Z[ab(a+b)(2E52+1)+ac(a+c)(2E32+1)+2bcE32(b+c)(2E32+1)2+bE52+cE322(2E32+1)(2E52+1)],W11(22)B=4Z[a+4b(ba)E52(a+b)(2E52+1)+4c(ca)E32(a+c)(2E32+1)4bc(2E321)2(b+c)(2E32+1)2+2(E32b+E52c)3(b+c)(2E32+1)(2E52+1)],W33A(B)=8(bc)2Z(b+c)(2E32+1)(2E52+1).
Here, a=exp(βE1), b=exp(βE3), and c=exp(βE5). Thus, the maximal and the minimal LQFI for the subsystems A and B can be obtained from the diagonal elements of WA and WB. Then one can find that the diagonal elements are W33A=W33B.
Limiting temperature cases. Now, let us discuss the limiting temperature cases:
(i) When the temperature T0, we should take different regions of Jz into account. For Jz>1, the diagonal elements are
W11(22)A=4E32(E32+1)(2E32+1)2,W11(22)B=8E344E32+4(2E32+1)2,W33A(B)=8E32(2E32+1)2.
Thus, in this case, we have QAmax=W33A and QAmin=W11A for Jz(1,1], while QAmax=W11A and QAmin=W33A for Jz(1,). For the subsystem B, we have QBmax=W11B and QBmin=W33B for Jz(1,0][1,), while QBmax=W33B and QBmin=W11B for Jz(0,1). Since the diagonal elements W33A=W33B, the behavior of QAmax (QAmin) can be consistent with that of QBmax (QBmin) in some regions of Jz, e.g., we have QAmax=QBmax for Jz[0,1] and QAmin=QBmin for Jz[1,).
For Jz<1, we have
W11(22)A=1,W11(22)B=2,W33A(B)=0,
thus, QAmax=1, QAmin=0, QBmax=2, and QBmin=0.
When we consider the special point Jz=1, all the diagonal elements of WA and WB equal to 4/9, which means QAmax=QBmax=QAmin=QBmin=4/9. At another special point Jz=1, there is QAmax=QBmax=QAmin=QBmin=8/9. The special results can be attributed to the high degeneracy of the eigenstates at Jz=±1 where the anisotropy disappears.
(ii) For the limit temperature T, whatever the value of Jz is, all the diagonal elements of WA and WB equal to zero, and thus QAmax=QAmin=QBmax=QBmin=0. It implies that the thermal fluctuation will eventually destroy all the quantum Fisher information and quantum correlations.

4 Relationship between the maximum LQFI and the parameters

Based on the method of Eqs. (5)−(8), we can obtain the maximal LQFI of the subsystems A and B (QA,Bmax) even for high-dimensional operators. In this part, we will discuss the relationship between QA,Bmax and the temperature T as well as the anisotropy parameter Jz. The maximal QFI is essential to determine the precision of quantum parameter estimations.

4.1 Maximal LQFI versus different temperatures

We numerically study the effect of the temperature on the maximal LQFI. In Fig.1, the two-site case (n=2) in Eq. (9) is considered and the anisotropy parameter is set as Jz=3,2,1/2. The spin length is s=1/2 for the subsystem A and S=1, 3/2, and 2 for the subsystem B, respectively. The maximum LQFI for the subsystems A and B (QAmax and QBmax) are defined by Eq. (6). As shown in Fig.1, with the increase of temperature T, both QAmax (the red line with circles) and QBmax (the blue line with stars) gradually decrease to zero.
Fig.1 For the two-site case of the Hamiltonian in Eq. (9), QAmax and QBmax vary with temperature T. We choose the spin length s=1/2 for the subsystem A and S=1 (a, d, g), S=3/2 (b, e, h), and S=2 (c, f, k) for the subsystem B. The anisotropy parameters is Jz=3 (a)−(c), Jz=2 (d)−(f), and Jz=1/2 (g)−(k).

Full size|PPT slide

In addition, our results clearly show that the subsystems A and B with different dimensions hold different LQFI. We find that at the low temperatures, when |Jz|>1, the value of QBmax is larger than that of QAmax and the difference between them increases as the spin length S increases. However, when |Jz|1, the rule may be broken, e.g., when Jz=1/2 the difference between QBmax and QAmax in the case of S=2 [see Fig.1(k)] is smaller than that of S=3/2 [see Fig.1(h)] at the low temperatures. When the temperature rises high enough, the difference between the QAmax and QBmax will disappears gradually for all the values of Jz. While, larger Jz and larger spin length S may lead to a higher threshold temperature at which QAmax and QBmax become consistent with each other.

4.2 Maximal LQFI versus different values of the parameter Jz

We take into account the effect of the anisotropy parameters Jz on the maximal LQFI for the temperature T=0.5. In Fig.2(a)−(c), the spin length s=1/2 for the subsystem A and S=1,3/2,2 for the subsystem B. Due to the different dimensions of the subsystems A and B, the values of QAmax and QBmax are different. When |Jz|>1, QBmax (the blue line with stars) is bigger than QAmax (the red line with circles). Moreover, both of the QAmax and QBmax increase with the increase of the absolute value |Jz|. That means the maximal quantum Fisher information, i.e., the accuracy of the parameter estimation, can be improved by increasing the absolute value of Jz. In the region |Jz|1, the dependence of QA,Bmax on the absolute value |Jz| may be different, e.g., larger |Jz| can lead to smaller values of QA,Bmax.
Fig.2 For the two-site case of the Hamiltonian in Eq. (9) with T=0.5, in subfigures (a−c) QA,Bmax and in subfigures (d−f) the corresponding eigen energy levels vary with Jz. The spin length is s=1/2 for the subsystem A and S=1 (a, d), S=3/2 (b, e), and S=2 (c, f) for subsystem B.

Full size|PPT slide

In the range of Jz[1,1], QAmax equals to QBmax. This may be due to the special energy level structure in this region. We show the energy levels versus parameter Jz in Fig.2(d)−(f), where one can find the energy levels intersect at the point of Jz=±1, i.e., the high degeneracy of the eigenstates occurs at the points. In the region Jz[1,1], all the energy levels are close to each other. At finite temperatures, the eigenstates become to be mixed together with the weights corresponding to their eigenenergies, and the degenerate states hold equal weights. The special structure of the mixed state for Jz[1,1] may be one of the reasons why the difference between QAmax and QBmax disappears. We find that it requires a threshold temperature. For the case of two-site (1/2,1) mixed spin chain, we numerically find that when the temperature reaches about Tth0.178, QAmax and QBmax become consistent in the region Jz[1,1]. When the temperature is lower than that, the behaviors of the QA and QB tend to follow the analytical results shown in the case of T0 in Eqs. (11) and (12). For the two sites (1/2,3/2), the threshold temperature is about Tth0.317, and for the two sites (1/2,2), the threshold temperature becomes Tth0.429. We show more numerical results of different mixed-spin cases for different temperatures in Appendix A. In addition, we also increase the spin length of the subsystem A to s=1, and investigate the change rule of Jz. Compared with the result of Fig.2, it can be found that the increase of spin length of the subsystem A will also bring some different behavior rules, see Appendix A for details.

5 LQFI and quantum correlation

5.1 LQFI and quantum discord

For a bipartite state, quantum discord is defined as [31]
DQAB(ρAB)minΠiA[I(ρAB)J(ρAB)ΠiA].
Here, I(ρAB)=S(ρA)+S(ρB)S(ρAB) represents the mutual information between the subsystems A and B [32]. S(ρ)Tr(ρlog2ρ) is the von Neumann entropy of the density matrix ρ, and ρA(B) is the reduced density matrix of the subsystem. J(ρAB)ΠiA=S(ρB)S(ρB|ΠiA) is the amount of information gained about the subsystem B by measuring the subsystem A. {ΠiA} are the measurement operators corresponding to the von Neumann measurement on the subsystem A. S(ρB|ΠiA) is the conditional entropy after the measurement performed on the subsystem A. Generally, DQ is asymmetric for the different subsystems A and B, which is therefore defined by DQAB or DQBA for distinction [32,33]. Quantum discord can be used to describe the quantum correlation even when the quantum entanglement disappears. Especially, in dissipative systems, the quantum discord is more robust than quantum entanglement [34,35]. The quantum discord has been well applied in many systems to describe the quantum correlations [36-42]. For the general states, the analytical and numerical calculation of quantum discord is complicated because of the optimization process over all the local generalized measurements. Therefore, it is only possible to obtain the analytical expressions of quantum discord for certain classes of states.
It has been proved that the minimal LQFI satisfies all the established criteria to be a measure of discord-type quantum correlations [19]. Thus, in this work, we employ the minimal LQFI to characterize the bipartite correlation in the mixed-spin chain. We first compare the behaviors of the quantum discord DQAB, the minimal LQFI (QAmin), and the maximal LQFI (QAmax) versus different parameters Jz. Different temperatures are also considered.
In Fig.3, for the two-site (1/2,1) mixed-spin chain with spin-1/2 for the subsystem A and spin-1 for the subsystem B, we numerically calculate the quantum discord DQAB versus different Jz, and the temperature is set to T=0.05, 0.2, 0.4, and 0.6 [see Fig.3(a)]. Then, the results of QAmin are shown in Fig.3(b). One can find that DQAB and QAmin follow a basically consistent behavior. They both reach the peak value at Jz=1 and reach the secondary peak at Jz=1.
Fig.3 For the two-site (1/2,1) mixed-spin chain, DQAB in (a), QAmin in (b), and QAmax in (c) vary with the anisotropy parameter Jz at different temperature T=0.05,0.2,0.4,0.6.

Full size|PPT slide

For comparison, we numerically calculate the maximal LQFI (QAmax) in Fig.3(c). Different from QAmin (DQAB), QAmax reaches the minimum values at Jz=±1. In the regions of |Jz|>1, one can find that the increasing (decreasing) QAmax corresponds to the decreasing (increasing) QAmin and DQAB. However, the situation is different for the region of |Jz|1, where the behaviors of QAmin and DQAB become nonmonotonic, and the three quantities QAmin, DQAB, and QAmax can increase together for some values of Jz.
We can understand the phenomena that when Jz<1, the system tends to the ferromagnetic structure, i.e., the ground state can be treated as a direct product state of states with consistent spin orientations at all sites. Obviously, there is no quantum correlation, thus DQAB and QAmin have zero values. When the temperature increases, different spin orientations will occur. This leads to a certain quantum correlation in the system state. As the temperature continues to rise, more excited states are mixed in. This makes the system tend towards a completely mixed state, thereby disrupting the quantum correlation. On the other hand, in the case of antiferromagnetism such as Jz>1, the ground state tends to be the superposition state or mixed state consisting of the components with the opposite-oriented spins alternately on the chain. Therefore, the ground state initially has quantum correlations, and the increasing temperature will weaken the quantum correlation.
The maximal LQFI (QAmax) is different from the minimal LQFI QAmin, which is related to the accuracy of parameter estimation. As the temperature increases, the thermal fluctuation will destroy the quantum resource and thus reduce the accuracy of the quantum parameter estimation. Therefore, in all the considered regions of Jz, increasing temperature always has a negative effect on QAmin.
In order to show the effect of the temperature more clearly, we calculate DQAB, QAmin, and QAmax versus temperature in Fig.4 and find that in the region of Jz<1, the dependence of QAmin and DQAB on temperature is nonmonotonic. QAmin and DQAB first increase with the temperature from zero values to their maximum values, then decrease gradually. However, the influence of the temperature T on QAmax is obviously different from that on QAmin and DQAB. Raising the temperature always lowers the value of QAmax regardless of the value of Jz.
Fig.4 For the two-site (1/2,1) mixed-spin chain, DQAB in (a), QAmin in (b), and QAmax in (c) vary with different temperatures for different parameters Jz.

Full size|PPT slide

5.2 The minimal LQFI of different site pairs

Since the minimal QFI (QA,Bmin) can be a measure of discord-type quantum correlations, we make use of QA,Bmin to study the bipartite correlations of a pair of sites on the spin chain. Several effects will be considered, such as the subsystem dimension, the total number of the spin sites, the position of the site pairs, and the distance between the two sites.
First, we take into account the impact of the total number of the spin sites on the minimal LQFI. We choose the site number as n=2,4,6, and two types of spin with s=1/2 (corresponding to the subsystem A) and S=1 (corresponding to the subsystem B) alternately distribute on the spin chain. The site pair considered by us consists of the 1-st and 2-nd sites (denoted by “sites 1&2”). Then, we show QA,Bmin of the sites 1&2 varying with Jz at temperature T=0.5 in Fig.5(a) and (b). In most cases within the range of Jz, the QAmin of n=4 (the blue line with stars) and n=6 (the red line with circles) are below the case of n=2 (the green line with rhombus). There is only a slight difference between the two curves of n=4 and n=6. We also check the 8-site chain and find the similar phenomena with those of n=4 and n=6. The spin chain in this work is considered under the open boundary condition, and the sites far from the chosen sites have a weak correlation with them. Thus, increasing the total site number to 6 and 8 only introduces a slight influence on the minimal LQFI of the sites 1&2.
Fig.5 The subgraphs (a, b) show the QA,Bmin of the site 1&2 varies with Jz for different total site number n=2,4,6. In (c, d), the 6-site spin chain is considered and the results show the QA,Bmin of the sites 1&2, 3&4, and 5&6 varying with Jz. In (e, f) the 6-site spin chain is considered and the results show the QA,Bmin of the sites 1&2, 1&4, and 1&6 varies with Jz. The subgraphs (g−i) show the results of the negativity which is used to describe the entanglement of the site pairs. Here, we choose spin-1/2 for the subsystem A and spin-1 for the subsystem B at temperature T=0.5.

Full size|PPT slide

Comparing QAmin and QBmin in subfigures (a) and (b), one may find some different behaviors between them. For example, in the region of Jz[1,0], the values of QBmin of n=4 and n=6 can be slightly larger than those of n=2. The difference between QAmin and QBmin also reflects the asymmetry of the discord-type correlations. We show the results of the two-site (1/2,3/2) and (1/2,2) mixed-spin chain in Appendix B, where more obvious differences between QAmin and QBmin can be found.
Second, we focus on the different positions of the spin pairs. Three nearest-neighbour pairs, i.e., 1&2, 3&4, and 5&6, are considered. Then, we calculate QA,Bmin versus Jz at temperature T=0.5 and show the results in Fig.5(c) and (d). We find that different positions of the site pairs provide different values of QA,Bmin, which reflects the different influences of the surrounding sites when the spin pair under consideration is placed at the edges or in the center of the spin chain. Different values and behaviors between QAmin and QBmin can also be found.
In addition, we consider the effect of the distance between the two sites. In Fig.5(e) and (f), the QA,Bmin of the spin pairs consist of the sites 1&2 (green line with rhombus), the sites 1&4 (blue line with stars), and the sites 1&6 (red line with circles). Obviously, QA,Bmin decrease with the increase of the distance between the sites of the subsystems A and B. In this type of spin chain, the bipartite correlations cannot exist between two sites at a long distance.
For comparison, the entanglement (measured by the negativity [43,44]) is discussed and the results are shown in Fig.5(g)−(i). One can find that in the region of Jz[3,1], QA,Bmin becomes evidently larger than zero, while the negativity keeps a zero value. In addition, entanglement can be easily suppressed or destroyed. When we increase the distance between the two sites under consideration, the entanglement (measured by the negativity) disappears for the site pair 1&4 in the whole region of Jz [see Fig.5(i)]. By contrast, the discord-type correlation (measured by QA,Bmin) presents nonzero values in an approximate region of Jz(2,2). In the case of two-site (1/2,3/2) and (1/2,2) mixed-spin chain shown in Appendix B, one can see the phenomena more clearly.

6 LQFI in multiple nearest-neighbor coupling Heisenberg model

In this section, we have discussions on the impact of the second and third nearest-neighbor coupling. Then, the corresponding Hamiltonian can be expressed as
Hα=i=1n/2[J1(siαSiα+Siαsi+1α)+J2(siαsi+1α+SiαSi+1α)+J3(siαSi+1α+Siαsi+2α)],(α=x,y,z)
where J1, J2, and J3 are the first, second, and third nearest-neighbor interaction intensities, respectively. For simplicity, we set Jk=1,(k=1,2,3). The open boundary condition is considered, i.e., we set sn/2+1α=0, Sn/2+1α=0, sn/2+2α=0.
Then, for the Hamiltonian with the second and third nearest-neighbor coupling, we calculated the QA,Bmin of the nearest-neighbor sites, i.e., 1&2, 3&4, and 5&6, varying with Jz [see Fig.6(a) and (b)]. Compared to Fig.5(c,d), quite different results are shown in Fig.6(a) and (b). Especially, the curve of the QA,Bmin between 1&2 sites (green line with rhombus) becomes the lowest one, while the QA,Bmin between 5&6 sites become the highest one.
Fig.6 The subgraphs (a, b) show the QA,Bmin of the sites 1&2, 3&4, and 5&6 varying with Jz. The subgraphs (c, d) show the QA,Bmin of the sites 1&2, 1&4, and 1&6 varying with Jz. Here, we choose spin-1/2 for the subsystem A and spin-1 for the subsystem B at temperature T=0.5.

Full size|PPT slide

We also considered the QA,Bmin of the 1&2, 1&4, and 1&6 sites varying with Jz [see Fig.6(c) and (d)]. Compared to Fig.5(e) and (f), the QA,Bmin of the 1&2 sites (green line with rhombus) in Fig.6(c) and (d) significantly decreases, while the QA,Bmin of the 1&6 sites (red line with circles) and the QA,Bmin of the 1&4 sites (blue line with stars) increase. Thus, the second and third nearest-neighbor coupling present remarkable effect on the minimal LQFI (i.e., the discord-type correlation), i.e., they enhance the QA,Bmin of the high-order-neighbor sites and suppress the QA,Bmin of the nearest-neighbor sites.

7 Conclusion

We investigate the LQFI and bipartite quantum correlations in the mixed-spin Heisenberg XXZ chain. Both the maximal LQFI (QA,Bmax) and the minimal LQFI (QA,Bmin) are considered in this work. The former determines the accuracy of the parameter estimation and the latter can measure the discord-type quantum correlation.
We find that the increasing temperature will suppress the maximal LQFI (QA,Bmax) to zero values. The dependence of QA,Bmax on the anisotropy parameter Jz is complicated. In the region |Jz|>1, the larger |Jz| leads to the larger QA,Bmax. While, in the region |Jz|1, the dependence of QA,Bmax on the absolute value |Jz| may be different, e.g., larger |Jz| can lead to smaller values of QA,Bmax. Different dimensions of the subsystems can lead to different values of QA,Bmax.
Then we make use of the minimal LQFI (QA,Bmin) to characterize the discord-type correlation. Different dimensional subsystems can provide different values of QA,Bmin, which reflects the asymmetry of the discord-type correlation. In most regions of the anisotropy parameter Jz, the higher dimensional subsystem can provide larger QA,Bmin. We find that the temperature can affect QA,Bmin either positively or negatively, which depends on the value of the Jz. When the spin chain with more sites (e.g., n=6) is considered, we find that the site pair at different positions of the chain holds different values of QA,Bmin. The QA,Bmin of the nearest-neighbour sites is larger than that of the site pairs consisting next-nearest-neighbour or farther apart sites. Our results show that the minimal LQFI can well describe the discord-type correlation, especially in the case of the high dimensions. Compared with the original definition of quantum discord, the calculation of the minimal LQFI is much simpler since there is no need for the optimization process over all the local generalized measurements.

References

[1]
V. Giovannetti, S. Lloyd, L. Maccone. Quantum-enhanced measurements: Beating the standard quantum limit. Science, 2004, 306(5700): 1330
CrossRef ADS Google scholar
[2]
D. Braun, G. Adesso, F. Benatti, R. Floreanini, U. Marzolino, M. W. Mitchell, S. Pirandola. Quantum-enhanced measurements without entanglement. Rev. Mod. Phys., 2018, 90(3): 035006
CrossRef ADS Google scholar
[3]
L. Pezzè, A. Smerzi, M. K. Oberthaler, R. Schmied, P. Treutlein. Quantum metrology with nonclassical states of atomic ensembles. Rev. Mod. Phys., 2018, 90(3): 035005
CrossRef ADS Google scholar
[4]
J. P. Dowling, K. P. Seshadreesan. Quantum optical technologies for metrology, sensing, and imaging. J. Lightwave Technol., 2015, 33(12): 2359
CrossRef ADS Google scholar
[5]
J. Liu, M. Zhang, H. Chen, L. Wang, H. Yuan. Optimal scheme for quantum metrology. Adv. Quantum Technol., 2022, 5(1): 2100080
CrossRef ADS Google scholar
[6]
R. Demkowicz-Dobrzański, L. Maccone. Using entanglement against noise in quantum metrology. Phys. Rev. Lett., 2014, 113(25): 250801
CrossRef ADS Google scholar
[7]
L. Pezzé, A. Smerzi. Ultrasensitive two-mode interferometry with single-mode number squeezing. Phys. Rev. Lett., 2013, 110(16): 163604
CrossRef ADS Google scholar
[8]
R. Schnabel, N. Mavalvala, D. E. McClelland, P. K. Lam. Quantum metrology for gravitational wave astronomy. Nat. Commun., 2010, 1(1): 121
CrossRef ADS Google scholar
[9]
S. D. Huver, C. F. Wildfeuer, J. P. Dowling. Entangled Fock states for robust quantum optical metrology, imaging, and sensing. Phys. Rev. A, 2008, 78(6): 063828
CrossRef ADS Google scholar
[10]
M. Ahmadi, D. E. Bruschi, C. Sab’ın, G. Adesso, I. Fuentes. Relativistic quantum metrology: Exploiting relativity to improve quantum measurement technologies. Sci. Rep., 2014, 4(1): 4996
CrossRef ADS Google scholar
[11]
Z. Sun, J. Ma, X. M. Lu, X. G. Wang. Fisher information in a quantum-critical environment. Phys. Rev. A, 2010, 82(2): 022306
CrossRef ADS Google scholar
[12]
M. Zhang, H. M. Yu, H. D. Yuan, X. G. Wang, R. Demkowicz-Dobrzański, J. Liu. QuanEstimation: An open-source toolkit for quantum parameter estimation. Phys. Rev. Res., 2022, 4(4): 043057
CrossRef ADS Google scholar
[13]
S. J. Gu. Fidelity approach to quantum phase transitions. Int. J. Mod. Phys. B, 2010, 24(23): 4371
CrossRef ADS Google scholar
[14]
T. L. Wang, L. N. Wu, W. Yang, G. R. Jin, N. Lambert, F. Nori. Quantum Fisher information as a signature of the superradiant quantum phase transition. New J. Phys., 2014, 16(6): 063039
CrossRef ADS Google scholar
[15]
U. Marzolino, T. Prosen. Fisher information approach to non-equilibrium phase transitions in a quantum XXZ spin chain with boundary noise. Phys. Rev. B, 2017, 96(10): 104402
CrossRef ADS Google scholar
[16]
P. Hyllus, W. Laskowski, R. Krischek, C. Schwemmer, W. Wieczorek, H. Weinfurter, L. Pezzé, A. Smerzi. Fisher information and multiparticle entanglement. Phys. Rev. A, 2012, 85(2): 022321
CrossRef ADS Google scholar
[17]
G. Tóth. Multipartite entanglement and high-precision metrology. Phys. Rev. A, 2012, 85(2): 022322
CrossRef ADS Google scholar
[18]
X. M. Lu, X. Wang, C. P. Sun. Quantum Fisher information flow and non-Markovian processes of open systems. Phys. Rev. A, 2010, 82(4): 042103
CrossRef ADS Google scholar
[19]
D. Girolami, A. M. Souza, V. Giovannetti, T. Tufarelli, J. G. Filgueiras, R. S. Sarthour, D. O. Soares-Pinto, I. S. Oliveira, G. Adesso. Quantum discord determines the interferometric power of quantum states. Phys. Rev. Lett., 2014, 112(21): 210401
CrossRef ADS Google scholar
[20]
H. S. Dhar, M. N. Bera, G. Adesso. Characterizing non-Markovianity via quantum interferometric power. Phys. Rev. A, 2015, 91(3): 032115
CrossRef ADS Google scholar
[21]
L. P. Chen, Y. N. Guo. Dynamics of local quantum uncertainty and local quantum fisher information for a two-qubit system driven by classical phase noisy laser. J. Mod. Opt., 2021, 68(4): 217
CrossRef ADS Google scholar
[22]
A. B. A. Mohamed, H. Eleuch. Dynamics of two magnons coupled to an open microwave cavity: Local quantum Fisher- and local skew-information coherence. Eur. Phys. J. Plus, 2022, 137(7): 853
CrossRef ADS Google scholar
[23]
A.SlaouiL. BakmouM.DaoudR.Ahl Laamara, A comparative study of local quantum Fisher information and local quantum uncertainty in Heisenberg XY model, Phys. Lett. A 383(19), 2241 (2019)
[24]
N. Habiballah, Y. Khedif, M. Daoud. Local quantum uncertainty in XYZ Heisenberg spin models with Dzyaloshinski‒Moriya interaction. Eur. Phys. J. D, 2018, 72(9): 154
CrossRef ADS Google scholar
[25]
F. Ozaydin, A. A. Altintas. Parameter estimation with Dzyaloshinski‒Moriya interaction under external magnetic fields. Opt. Quantum Electron., 2020, 52(2): 70
CrossRef ADS Google scholar
[26]
F. Ozaydin, A. A. Altintas. Quantum metrology: Surpassing the shot-noise limit with Dzyaloshinskii‒Moriya interaction. Sci. Rep., 2015, 5(1): 16360
CrossRef ADS Google scholar
[27]
S. Haseli. Local quantum Fisher information and local quantum uncertainty in two-qubit Heisenberg XYZ chain with Dzyaloshinskii‒Moriya interactions. Laser Phys., 2020, 30(10): 105203
CrossRef ADS Google scholar
[28]
A. V. Fedorova, M. A. Yurischev. Behavior of quantum discord, local quantum uncertainty, and local quantum Fisher information in two-spin-1/2 Heisenberg chain with DM and KSEA interactions. Quantum Inform. Process., 2022, 21(3): 92
CrossRef ADS Google scholar
[29]
J. Liu, X. X. Jing, W. Zhong, X. G. Wang. Quantum Fisher information for density matrices with arbitrary ranks. Commum. Theor. Phys., 2014, 61(1): 45
CrossRef ADS Google scholar
[30]
M.A. NielsenI.L. Chuang, Quantum Computation and Quantum Information, Cambridge: Cambridge University Press, 2000
[31]
H. Ollivier, W. H. Zurek. Quantum discord: A measure of the quantumness of correlations. Phys. Rev. Lett., 2001, 88(1): 017901
CrossRef ADS Google scholar
[32]
B. Groisman, S. Popescu, A. Winter. Quantum, classical, and total amount of correlations in a quantum state. Phys. Rev. A, 2005, 72(3): 032317
CrossRef ADS Google scholar
[33]
B. Dakić, V. Vedral, Č. Brukner. Necessary and sufficient condition for nonzero quantum discord. Phys. Rev. Lett., 2010, 105(19): 190502
CrossRef ADS Google scholar
[34]
X.M. LuZ. J. XiZ.SunX.G. Wang, Geometric measure of quantum discord under decoherence, Quantum Inf. Comput. 10(11–12), 0994 (2010)
[35]
T. Werlang, S. Souza, F. F. Fanchini, C. J. Villas Boas. Robustness of quantum discord to sudden death. Phys. Rev. A, 2009, 80(2): 024103
CrossRef ADS Google scholar
[36]
A. Datta, A. Shaji, C. M. Caves. Quantum discord and the power of one qubit. Phys. Rev. Lett., 2008, 100(5): 050502
CrossRef ADS Google scholar
[37]
S. L. Luo, S. S. Fu. Geometric measure of quantum discord. Phys. Rev. A, 2010, 82(3): 034302
CrossRef ADS Google scholar
[38]
L. Henderson, V. Vedral. Classical, quantum and total correlations. J. Phys. Math. Gen., 2001, 34(35): 6899
CrossRef ADS Google scholar
[39]
Y. X. Chen, Z. Yin. Thermal quantum discord in anisotropic Heisenberg XXZ model with Dzyaloshinskii‒Moriya interaction. Commum. Theor. Phys., 2010, 54(1): 60
CrossRef ADS Google scholar
[40]
Q. Chen, C. Zhang, S. Yu, X. X. Yi, C. H. Oh. Quantum discord of two-qubit X states. Phys. Rev. A, 2011, 84(4): 042313
CrossRef ADS Google scholar
[41]
M. Ali, A. R. P. Rau, G. Alber. Quantum discord for two-qubit X states. Phys. Rev. A, 2010, 81(4): 042105
CrossRef ADS Google scholar
[42]
S. L. Braunstein, C. M. Caves. Statistical distance and the geometry of quantum states. Phys. Rev. Lett., 1994, 72(22): 3439
CrossRef ADS Google scholar
[43]
G. Vidal, R. F. Werner. Computable measure of entanglement. Phys. Rev. A, 2002, 65(3): 032314
CrossRef ADS Google scholar
[44]
Z. Sun, X. G. Wang, A. Z. Hu, Y. Q. Li. Entanglement properties in mixed-spin Heisenberg systems. Physica A, 2006, 370(2): 483
CrossRef ADS Google scholar

Declarations

The authors declare that they have no competing interests and there are no conflicts.

Acknowledgements

The work was supported by the National Natural Science Foundation of China (NSFC) under Grant No. 12175052 and the Postdoctoral Science Foundation of China (No. 2022M722794).

RIGHTS & PERMISSIONS

2023 Higher Education Press
AI Summary AI Mindmap
PDF(6821 KB)

853

Accesses

1

Citations

Detail

Sections
Recommended

/