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

Nonlinear steering criteria for arbitrary two-qubit quantum systems

Guo-Zhu Pan111[email protected] School of Electrical and Photoelectric Engineering, West Anhui University, Lu’an, 237012, China    Ming Yang222[email protected] School of Physics & Materials Science, Anhui University, Hefei 230601, China    Hao Yuan School of Electrical and Photoelectric Engineering, West Anhui University, Lu’an, 237012, China School of Physics & Materials Science, Anhui University, Hefei 230601, China    Gang Zhang School of Electrical and Photoelectric Engineering, West Anhui University, Lu’an, 237012, China    Jun-Long Zhao School of Physics & Materials Science, Anhui University, Hefei 230601, China
Abstract

Abstract: By employing Pauli measurements, we present some nonlinear steering criteria applicable for arbitrary two-qubit quantum systems and optimized ones for symmetric quantum states. These criteria provide sufficient conditions to witness steering, which can recover the previous elegant results for some well-known states. Compared with the existing linear steering criterion and entropic criterion, ours can certify more steerable states without selecting measurement settings or correlation weights, which can also be used to verify entanglement as all steerable quantum states are entangled.

Quantum steering, Nonlocality, Entanglement, Covariance matrices
pacs:
03.65.Ud, 03.67.Mn, 42.50.Dv

I Introduction

Quantum steering describes the ability of one observer to nonlocally affect the other observer’s state through local measurements, which was first noted by Einstein, Podolsky and Rosen (EPR) for arguing the completeness of quantum mechanics in 1935 ein , and later introduced by Schrödinger in response to the well-known EPR paradoxsch . After being formalized by Wiseman et al. with a local hidden variable (LHV)-local hidden state model in 2007 wis , quantum steering has attracted increasing attention and been explored widely. Steerable states were shown to be advantageous for tasks involving secure quantum teleportation rei ; ros , quantum secret sharing walk ; kog , one-sided device-independent quantum key distribution bra and channel discrimination pia .

Quantum steering is one form of quantum correlations intermediate between quantum entanglement horo and Bell nonlocality bell . It has been demonstrated that a quantum state which is Bell nonlocal must be steerable, and a quantum state which is steerable must be entangled jone ; brun . One distinct feature of quantum steering which differs from entanglement and Bell nonlocality is asymmetry. That is, there exists the case when Alice can steer Bob’s state but Bob cannot steer Alice’s state, which is referred to as one-way steerable and has been demonstrated in theory bow and experiment han ; wol .

Quantum steering is the failure description of the local hidden variable-local hidden state models to reproduce the correlation between two subsystems, which can be witnessed by quantum steering criteria. Recently, a lot of steering criteria have been developed to distinguish steerable quantum states from unsteerable ones. In Ref. sau , the linear steering criteria was introduced for qubit states. In Ref. sch2 , the steering criteria from entropic uncertainty relations were derived, which can be applicable for both discrete and continuous variable systems. Subsequently, the steering criteria via covariance matrices of local observables ji and local uncertainty relations zhen in arbitrary-dimensional quantum systems were presented. Recently, Refs. zhe1 ; zhe2 generalized the linear steering criteria to high-dimensional systems. Although these criteria work well for a number of quantum states, most of them require constructing appropriate measurement settings or correlation weights in practice, which increases the complexities of the detecting inevitably. The development of the universal criterion to detect steering is still one vexed question.

In this paper, we first present some steering criteria applicable for arbitrary two-qubit quantum systems, then optimize them for symmetric quantum states, and finally we provide a broad class of explicit examples including two-qubit Werner states, Bell diagonal states, and Gisin states. Compared with the existing linear steering criterion and entropic criterion, ours can certify more steerable states without selecting measurement settings or correlation weights, which can also be used to verify entanglement as all steerable quantum states are entangled.

II Nonlinear steering criteria for arbitrary two-qubit quantum systems

Suppose two separate parties, Alice and Bob, share a two-qubit quantum state on a composite Hilbert space =AB\mathcal{H}=\mathcal{H}_{A}\otimes\mathcal{H}_{B}. The steering is defined by the failure description of all possible local hidden variable-local hidden state models in the form wis ; jone

P(a,b|A,B;W)=λP(a|A;λ)P(b|B;ρλ)pλ,P(a,b|A,B;W)=\sum_{\lambda}P(a|A;\lambda)P(b|B;\rho_{\lambda})p_{\lambda}, (1)

where P(a,b|A,B;W)P(a,b|A,B;W) are joint probabilities for Alice and Bob’s measurements AA and BB, with the results aa and bb, respectively; pλp_{\lambda} and P(a|A;λ)P(a|A;\lambda) denote some probability distributions involving the LHV λ\lambda, and P(b|B;ρλ)P(b|B;\rho_{\lambda}) denotes the quantum probability of outcome bb given measurement BB on state ρλ\rho_{\lambda}. WW represents the bipartite state under consideration. In other words, a quantum state will be steerable if it does not satisfy Eq.(1). Within the formulation, we propose a nonlinear steering criterion that can be used to certify a wide range of steerable quantum states for two-qubit quantum systems.

Theorem 1. If a given two-qubit quantum state is unsteerable from Alice to Bob (or Bob to Alice), the following inequality holds:

i=13j=13σiσj21,\sum\limits_{i=1}\limits^{3}\sum\limits_{j=1}\limits^{3}\langle\sigma_{i}\otimes\sigma_{j}\rangle^{2}\leq 1, (2)

where σi,j\sigma_{i,j} (i,j=1,2,3i,j=1,2,3) are Pauli operators.

Proof. Suppose Alice and Bob share a two-qubit quantum state ρAB\rho_{AB} on a composite Hilbert space, both of them perform NN measurements on their own states, which are denoted by AkA_{k} and BlB_{l}, respectively. Here BlB_{l} is a quantum observable while AkA_{k} have no such constraint, k(l)k(l) (k(l)=1,2,,Nk(l)=1,2,\cdots,N) labels the kth (lth) measurement setting for Alice (Bob). If the state is unsteerable from Alice to Bob, we have the following inequality

k=1Nl=1NAkBl2\displaystyle\sum\limits_{k=1}\limits^{N}\sum\limits_{l=1}\limits^{N}\langle A_{k}\otimes B_{l}\rangle^{2} (3)
=\displaystyle= k=1Nl=1N(ak,blakblP(ak,bl|Ak,Bl;ρAB))2\displaystyle\sum\limits_{k=1}\limits^{N}\sum\limits_{l=1}\limits^{N}\left(\sum\limits_{a_{k},b_{l}}a_{k}b_{l}P(a_{k},b_{l}|A_{k},B_{l};\rho_{AB})\right)^{2}
\displaystyle\leq λ(pλk=1N[akakP(ak|Ak,λ)]2l=1N[blblP(bl|Bl,ρλ)]2)\displaystyle\sum\limits_{\lambda}\left(p_{\lambda}\sum\limits_{k=1}\limits^{N}\left[\sum\limits_{a_{k}}a_{k}P(a_{k}|A_{k},\lambda)\right]^{2}\sum\limits_{l=1}\limits^{N}\left[\sum\limits_{b_{l}}b_{l}P(b_{l}|B_{l},\rho_{\lambda})\right]^{2}\right)
=\displaystyle= λpλ(k=1NAkλ2l=1NBlρλ2)\displaystyle\sum\limits_{\lambda}p_{\lambda}\left(\sum\limits_{k=1}\limits^{N}\langle A_{k}\rangle^{2}_{\lambda}\sum\limits_{l=1}\limits^{N}\langle B_{l}\rangle^{2}_{\rho_{\lambda}}\right)
\displaystyle\leq ηλpλ(k=1NAk2λ)max{ρλ}(l=1NBlρλ2)\displaystyle\eta\sum\limits_{\lambda}p_{\lambda}\left(\sum\limits_{k=1}\limits^{N}\langle A_{k}^{2}\rangle_{\lambda}\right)\max\limits_{\{\rho_{\lambda}\}}\left(\sum\limits_{l=1}\limits^{N}\langle B_{l}\rangle^{2}_{\rho_{\lambda}}\right)
=\displaystyle= ηk=1NAk2CB=ηCACB,\displaystyle\eta\sum\limits_{k=1}\limits^{N}\langle A_{k}^{2}\rangle C_{B}=\eta C_{A}C_{B},

where CA=k=1NAk2,CB=max{ρλ}(l=1NBlρλ2)C_{A}=\sum\limits_{k=1}\limits^{N}\langle A_{k}^{2}\rangle,C_{B}=\max\limits_{\{\rho_{\lambda}\}}\left(\sum\limits_{l=1}\limits^{N}\langle B_{l}\rangle^{2}_{\rho_{\lambda}}\right). The parameter η\eta (0η10\leq\eta\leq 1) is a constant, which is used to adjust the value to the appropriate bound. The first inequality follows from the fact k=1Nl=1N(αkβl)2k=1Nαk2l=1Nβl2\sum_{k=1}^{N}\sum_{l=1}^{N}(\alpha_{k}\beta_{l})^{2}\leq\sum_{k=1}^{N}\alpha_{k}^{2}\sum_{l=1}^{N}\beta_{l}^{2}. The second inequality follows from the definition of CBC_{B} and the fact Ak2λAkλ2\langle A_{k}^{2}\rangle_{\lambda}\geq\langle A_{k}\rangle_{\lambda}^{2}. If the observables AkA_{k} and BlB_{l} are restricted to Pauli matrices, i.e., Ak(Bl)={σ1,σ2,σ3}A_{k}(B_{l})=\{\sigma_{1},\sigma_{2},\sigma_{3}\}, one has straightforwardly CA=3C_{A}=3 and CB=1C_{B}=1, so Eq.(3) reduces to

i=13j=13σiσj2η.\sum\limits_{i=1}\limits^{3}\sum\limits_{j=1}\limits^{3}\langle\sigma_{i}\otimes\sigma_{j}\rangle^{2}\leq\eta^{\prime}. (4)

where η=3η\eta^{\prime}=3\eta.

As we know, quantum entanglement, quantum steering, and Bell nonlocality are equivalent in the case of pure states wis ; jone ; ysx . For an arbitrary quantum steering criterion, it is preferable to be a sufficient and necessary condition to detect pure states zhe1 ; zhe2 ; zhen . In order to obtain the optimal value of the parameter η\eta^{\prime}, we introduce the pure states as reference states. For any two-qubit state, it can be expressed as

ρAB=14(𝕀+i=13ci0σi𝕀+j=13c0j𝕀σj+i=13j=13cijσiσj),\rho_{AB}=\frac{1}{4}(\mathbb{I}+\sum_{i=1}^{3}c_{i0}\sigma_{i}\otimes\mathbb{I}+\sum_{j=1}^{3}c_{0j}\mathbb{I}\otimes\sigma_{j}+\sum_{i=1}^{3}\sum_{j=1}^{3}c_{ij}\sigma_{i}\otimes\sigma_{j}), (5)

where |cij|1|c_{ij}|\leq 1 for i,j=0,1,2,3i,j=0,1,2,3. For arbitrary pure states ρAB\rho_{AB}, one has straightforwardly i=13ci02+j=13c0j2+i=13j=13cij2=3\sum_{i=1}^{3}c_{i0}^{2}+\sum_{j=1}^{3}c_{0j}^{2}+\sum_{i=1}^{3}\sum_{j=1}^{3}c_{ij}^{2}=3 due to the fact tr(ρAB2)=1tr(\rho_{AB}^{2})=1. Next we consider two cases, one is that ρAB\rho_{AB} be pure separable states, then one achieves i=13σi𝕀2=i=13ci02=1,i=13𝕀σj2=i=13c0j2=1\sum_{i=1}^{3}\langle\sigma_{i}\otimes\mathbb{I}\rangle^{2}=\sum_{i=1}^{3}c_{i0}^{2}=1,\sum_{i=1}^{3}\langle\mathbb{I}\otimes\sigma_{j}\rangle^{2}=\sum_{i=1}^{3}c_{0j}^{2}=1, and i=13j=13σiσj2=i=13j=13cij2=1\sum_{i=1}^{3}\sum_{j=1}^{3}\langle\sigma_{i}\otimes\sigma_{j}\rangle^{2}=\sum_{i=1}^{3}\sum_{j=1}^{3}c_{ij}^{2}=1, which result in η1\eta^{\prime}\geq 1 due to the fact that all pure separable states are unsteerable. The other is that ρAB\rho_{AB} be pure entangled states, then one attains i=13σi𝕀2=i=13ci02<1,i=13𝕀σj2=i=13c0j2<1\sum_{i=1}^{3}\langle\sigma_{i}\otimes\mathbb{I}\rangle^{2}=\sum_{i=1}^{3}c_{i0}^{2}<1,\sum_{i=1}^{3}\langle\mathbb{I}\otimes\sigma_{j}\rangle^{2}=\sum_{i=1}^{3}c_{0j}^{2}<1, and i=13j=13σiσj2=i=13j=13cij2>1\sum_{i=1}^{3}\sum_{j=1}^{3}\langle\sigma_{i}\otimes\sigma_{j}\rangle^{2}=\sum_{i=1}^{3}\sum_{j=1}^{3}c_{ij}^{2}>1, which result in η1\eta^{\prime}\leq 1 due to the fact that all pure entangled states are steerable zhe1 ; zhe2 ; zhen . So the optimal value of the parameter η=1\eta^{\prime}=1. This gives the proof of Theorem 1.

In this way, we derive the steering criterion for arbitrary two-qubit quantum systems. Whatever strategies Alice and Bob choose, a violation of inequality (2) would imply steering.

In the following we further develop steering criterion by introducing quantum correlation matrix of local observables. Given a quantum state ρ\rho and observables {Ok}(k=1,2,,n)\{O_{k}\}(k=1,2,...,n), an n×nn\times n symmetric covariance matrix γ\gamma is defined as ji

γkk(ρ)=(OkOk+OkOk)/2OkOk.\gamma_{kk^{\prime}}(\rho)=(\langle O_{k}O_{k^{\prime}}\rangle+\langle O_{k^{\prime}}O_{k}\rangle)/2-\langle O_{k}\rangle\langle O_{k^{\prime}}\rangle. (6)

Now, let us consider a composite system ρAB\rho_{AB} and a set observables {Om}={σiσj}(i,j=1,2,3,m=3(i1)+j)\{O_{m}\}=\{\sigma_{i}\otimes\sigma_{j}\}(i,j=1,2,3,m=3(i-1)+j). Similarly, the covariance matrix can be constructed as

γmm(ρAB)=(OmOm+OmOm)/2OmOm.\gamma_{mm^{\prime}}(\rho_{AB})=(\langle O_{m}O_{m^{\prime}}\rangle+\langle O_{m^{\prime}}O_{m}\rangle)/2-\langle O_{m}\rangle\langle O_{m^{\prime}}\rangle. (7)

Obviously, the diagonal elements of the covariance matrix stand for the variance of the observables {Om}\{O_{m}\}.

Corollary 1. If a given quantum state ρAB\rho_{AB} is unsteerable, the sum of the eigenvalues of the covariance matrix γmm(ρAB)\gamma_{mm^{\prime}}(\rho_{AB}) of the observables {Om}={σiσj}(i,j=1,2,3,m=3(i1)+j)\{O_{m}\}=\{\sigma_{i}\otimes\sigma_{j}\}(i,j=1,2,3,m=3(i-1)+j) must satisfied

k=19λk8,\sum\limits_{k=1}\limits^{9}\lambda_{k}\geq 8, (8)

where λk\lambda_{k} is the eigenvalue of the covariance matrix γmm(ρAB)\gamma_{mm^{\prime}}(\rho_{AB}).

Proof. For an unsteerable state ρAB\rho_{AB}, one has i=13j=13σiσj21\sum_{i=1}^{3}\sum_{j=1}^{3}\langle\sigma_{i}\otimes\sigma_{j}\rangle^{2}\leq 1 according to Theorem 1, which results in i=13j=13δ2(σiσj)8\sum_{i=1}^{3}\sum_{j=1}^{3}\delta^{2}(\sigma_{i}\otimes\sigma_{j})\geq 8, where δ2(σiσj)=(σiσj)2σiσj2\delta^{2}(\sigma_{i}\otimes\sigma_{j})=\langle(\sigma_{i}\otimes\sigma_{j})^{2}\rangle-\langle\sigma_{i}\otimes\sigma_{j}\rangle^{2} is the variance of the observable σiσj\sigma_{i}\otimes\sigma_{j}. To prove the corollary 1, we introduce the principal components analysis (PCA) pea ; hot ; jol , which is a mathematical procedure that transforms a number of possibly correlated variables into a number of uncorrelated variables called principal components. The first principal component accounts for as much of the variability in the data as possible, and each succeeding component accounts for as much of the remaining variability as possible. Similar to classical PCA, for the quantum covariance matrix γmm(ρAB)\gamma_{mm^{\prime}}(\rho_{AB}), the variances of principal components correspond to the eigenvalues of the covariance matrix, i.e., k=19λk=k=19δ2Pk\sum_{k=1}^{9}\lambda_{k}=\sum_{k=1}^{9}\delta^{2}P_{k}, where PkP_{k} is the principal component of the covariance matrix γmm(ρAB)\gamma_{mm^{\prime}}(\rho_{AB}), and k=19δ2Pk=i=13j=13δ2(σiσj)\sum_{k=1}^{9}\delta^{2}P_{k}=\sum_{i=1}^{3}\sum_{j=1}^{3}\delta^{2}(\sigma_{i}\otimes\sigma_{j}), one has k=19λk=i=13j=13δ2(σiσj)\sum_{k=1}^{9}\lambda_{k}=\sum_{i=1}^{3}\sum_{j=1}^{3}\delta^{2}(\sigma_{i}\otimes\sigma_{j}). So one attains k=19λk8\sum_{k=1}^{9}\lambda_{k}\geq 8 for an unsteerable state. A detailed proof is provided in the Appendix A.

III Optimized steering criteria for symmetric two-qubit quantum systems

Symmetry is another central concept in quantum theory gro , which can be used to simplify the study of the entanglement sometimes voll ; stoc ; toth . A bipartite quantum state ρ\rho is called symmetric if it is permutationally invariant, i.e., FρF=ρF\rho F=\rho, here F=ij|ijji|F=\sum_{ij}|ij\rangle\langle ji| is the flip operator. In the following we optimize the steering criterion for symmetric two-qubit quantum states.

Theorem 2. If a given symmetric two-qubit quantum state is unsteerable from Alice to Bob (or Bob to Alice), the following inequality holds:

i=13σiσi21,\sum\limits_{i=1}\limits^{3}\langle\sigma_{i}\otimes\sigma_{i}\rangle^{2}\leq 1, (9)

where σi\sigma_{i} (i=1,2,3i=1,2,3) are Pauli operators.

proof. For arbitrary symmetric two-qubit quantum state, one has σiσj=0\langle\sigma_{i}\otimes\sigma_{j}\rangle=0, where i,j=1,2,3,iji,j=1,2,3,i\neq j. So Theorem 1 reduces to Theorem 2.

Corollary 2. If a given symmetric two-qubit quantum state ρAB\rho_{AB} is unsteerable, the sum of the eigenvalues of the covariance matrix γmm(ρAB)\gamma_{mm^{\prime}}(\rho_{AB}) of the observables {Oi}={σiσi}(i=1,2,3)\{O_{i}\}=\{\sigma_{i}\otimes\sigma_{i}\}(i=1,2,3) must satisfy

k=13λk2,\sum\limits_{k=1}\limits^{3}\lambda_{k}\geq 2, (10)

where λk\lambda_{k} is the eigenvalue of the covariance matrix γmm(ρAB)\gamma_{mm^{\prime}}(\rho_{AB}). A brief proof of our theorem is specified below.

proof. For a symmetric unsteerable state ρAB\rho_{AB}, one has i=13δ2σiσi21\sum_{i=1}^{3}\delta^{2}\langle\sigma_{i}\otimes\sigma_{i}\rangle^{2}\leq 1 from Eq.(9), which results in i=13δ2(σiσi)2\sum_{i=1}^{3}\delta^{2}(\sigma_{i}\otimes\sigma_{i})\geq 2. For the quantum covariance matrix γmm(ρAB)\gamma_{mm^{\prime}}(\rho_{AB}), one has k=13λk=i=13δ2(σiσi)\sum_{k=1}^{3}\lambda_{k}=\sum_{i=1}^{3}\delta^{2}(\sigma_{i}\otimes\sigma_{i}) according to PCA. So one get k=13λk2\sum_{k=1}^{3}\lambda_{k}\geq 2 for a symmetric unsteerable state .

IV Illustrations of generic examples

(i) Werner state. Consider two-qubit Werner states wern , which can be written as

ρW=p|ψ+ψ+|+(1p)𝕀/4,\rho_{W}=p|\psi^{+}\rangle\langle\psi^{+}|+(1-p)\mathbb{I}/4, (11)

where |ψ+=(1/2)(|00+|11)|\psi^{+}\rangle=(1/\sqrt{2})(|00\rangle+|11\rangle) is Bell state and 𝕀\mathbb{I} is the identity, 0p10\leq p\leq 1. The Werner states are entangled iff p>1/3p>1/3, steerable iff p>1/2p>1/2 wis , and Bell nonlocal if p>1/2p>1/\sqrt{2}. According to symmetry of the Werner state and our Theorem 2, we achieve p>3/3p>\sqrt{3}/3 for successful steering under the Pauli measurements {σ1,σ2,σ3}\{\sigma_{1},\sigma_{2},\sigma_{3}\}. Our results are in agreement with the results of Ref. zhe1 ; zhe2 ; zhen , which implies that the nonlinear steering criterion is qualified for witnessing steering .

(ii) Bell diagonal states. Suppose now that Alice and Bob share a Bell diagonal state as follows:

ρbd=14(𝕀+i=13ciσiσi),\rho_{bd}=\frac{1}{4}(\mathbb{I}+\sum_{i=1}^{3}c_{i}\sigma_{i}\otimes\sigma_{i}), (12)

where σi\sigma_{i} (i=1,2,3)(i=1,2,3) are Pauli operators and |ci|1|c_{i}|\leq 1 for i=1,2,3i=1,2,3. According to Theorem 2, we find that ρbd\rho_{bd} are steerable if ici2>1\sum_{i}c_{i}^{2}>1. In this case, the local uncertainty relations steering criterion can be written as iδ2(σiB)C2(σiA,σiB)/δ2(σiA)>2\sum_{i}\delta^{2}(\sigma_{i}^{B})-C^{2}(\sigma_{i}^{A},\sigma_{i}^{B})/\delta^{2}(\sigma_{i}^{A})>2 zhen , where δ2(A)=A2A2\delta^{2}(A)=\langle A^{2}\rangle-\langle A\rangle^{2} is the variance and C(A,B)=ABABC(A,B)=\langle AB\rangle-\langle A\rangle\langle B\rangle is the covariance. The violation is ici2>1\sum_{i}c_{i}^{2}>1 and the corresponding states are steerable. Likely for the linear criterion we have |iωiσiAσiB|3|\sum_{i}\omega_{i}\langle\sigma_{i}^{A}\otimes\sigma_{i}^{B}\rangle|\geq\sqrt{3} with ωi{±1}\omega_{i}\in\{\pm 1\} sau , and the violation implies |c1±c2±c3|>3|c_{1}\pm c_{2}\pm c_{3}|>\sqrt{3}. For entropic criterion we have iH(σiB|σiA)>2\sum_{i}H(\sigma_{i}^{B}|\sigma_{i}^{A})>2 sch2 , where H(B|A)=ap(a|A)H(B|A=a)H(B|A)=\sum_{a}p(a|A)H(B|A=a) and H()H(\cdot) denotes von Neumann entropy. The violation is i(1+ci)log(1+ci)+(1ci)log(1ci)>2\sum_{i}(1+c_{i})log(1+c_{i})+(1-c_{i})log(1-c_{i})>2. It can be checked that our criterion performs equivalently well as the local uncertainty relations steering criterion, which certifies more steerable states than the linear criterion and the entropic criterion (Fig.1).

Refer to caption
Figure 1: The performances of different quantum steering criteria for the Bell diagonal states under the conditions c1=c3c_{1}=c_{3}. The area inside the brown solid lines denotes Bell diagonal states (BDS). The red solid line, blue circled line, green dashed line, cyan dotted line are given by the nonlinear steering criterion (NLC), local uncertainty relations criterion (LUR), linear criterion (LC), entropic criterion (EC), respectively. States in the left side of these lines are steerable. It is clear that the NLC performs equivalently well as the LUR criterion, which certifies more steerable states than the LC and EC.

(iii) Asymmetric entangled states. Let us consider Gisin states gis , which can be expressed as

ρG=p|ψθψθ|+(1p)ρs,\rho_{G}=p|\psi_{\theta}\rangle\langle\psi_{\theta}|+(1-p)\rho_{s}, (13)

where ψθ=sinθ|01+cosθ|10\psi_{\theta}=sin\theta|01\rangle+cos\theta|10\rangle, ρs=12|0000|+12|1111|\rho_{s}=\frac{1}{2}|00\rangle\langle 00|+\frac{1}{2}|11\rangle\langle 11|. In Fig.2, we show the performances of the nonlinear steering criterion (Theorem 1), the local uncertainty relations steering criterion zhen , the linear criterion sau and the entropic criterion sch2 for the Gisin states. It follows from straightforward calculation that the nonlinear steering criterion certifies more steerable states than the linear criterion and entropic criterion.

Refer to caption
Figure 2: The performances of different quantum steering criteria for the Gisin states. The cyan dotted line, green dashed line, red solid line, blue dashed line are given by the EC, LC, NLC, LUR criterion, respectively. States above these lines are steerable. It is clear that the NLC certifies more steerable states than the LC and EC.

V Conclusion

In summary, we have proposed some nonlinear steering criteria applicable for arbitrary two-qubit quantum systems and optimized ones for symmetric quantum states. These criteria can be used to detect a wide range of steerable quantum states under Pauli measurements. Compared with the existing linear steering criterion and the entropic criterion, ours can certify more steerable states without selecting measurement settings or correlation weights, which can also be used to verify entanglement as all steerable quantum states are entangled.

Acknowledgments

This work is supported by the National Natural Science Foundation of China (NSFC) under Grant No. 11947102, the Natural Science Foundation of Anhui Province under Grant Nos. 2008085MA16 and 2008085QA26, the Key Program of West Anhui University under Grant No.WXZR201819, the Research Fund for high-level talents of West Anhui University under Grant No.WGKQ202001004.

Appendix A Proof of the equation k=19λk=i=13j=13δ2(σiσj)\sum_{k=1}^{9}\lambda_{k}=\sum_{i=1}^{3}\sum_{j=1}^{3}\delta^{2}(\sigma_{i}\otimes\sigma_{j})

In order to prove the Eq. k=19λk=i=13j=13δ2(σiσj)\sum_{k=1}^{9}\lambda_{k}=\sum_{i=1}^{3}\sum_{j=1}^{3}\delta^{2}(\sigma_{i}\otimes\sigma_{j}), we extend principal components analysis to quantum correlation matrix γmm(ρAB)\gamma_{mm^{{}^{\prime}}}(\rho_{AB}) of local observables {Om}={σiσj}(i,j=1,2,3,m=3(i1)+j)\{O_{m}\}=\{\sigma_{i}\otimes\sigma_{j}\}(i,j=1,2,3,m=3(i-1)+j). As in classical correlation analysis, the principal components on a matrix space can be expressed as

Pj=a1jO1+a2jO2++a9jO9,P_{j}=a_{1j}O_{1}+a_{2j}O_{2}+...+a_{9j}O_{9}, (14)

where j=1,2,,9j=1,2,...,9. iaijaij=1\sum_{i}a_{ij}^{*}a_{ij}=1, and iaijaik=0\sum_{i}a_{ij}^{*}a_{ik}=0 for jkj\neq k.

To achieve the first principal component, we use the Lagrange multiplier technique to find the maximum of a function. The Lagrangean function is defined as

L(a)=tr[ρ(a11O1+a21O2++a91O9)2]{tr[ρ(a11O1+a21O2++a91O9)]}2\displaystyle L(a)=tr[\rho(a_{11}O_{1}+a_{21}O_{2}+...+a_{91}O_{9})^{2}]-\{tr[\rho(a_{11}O_{1}+a_{21}O_{2}+...+a_{91}O_{9})]\}^{2}
+λ1(1a112a212a912),\displaystyle+\lambda_{1}(1-a_{11}^{2}-a_{21}^{2}-...-a_{91}^{2}),\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (15)

where λ1\lambda_{1} are the Lagrange multipliers. The necessary conditions for the maximum are

La11=0;La21=0;;La91=0.\frac{\partial L}{\partial a_{11}}=0;\frac{\partial L}{\partial a_{21}}=0;...;\frac{\partial L}{\partial a_{91}}=0. (16)

By using the properties of the trace, we obtain

Lai1=2ai1tr(ρOi2)2ai1[tr(ρOi)]2+k=1,,9,kiak1[tr(ρOiOk)+tr(ρOkOi)]\displaystyle\frac{\partial L}{\partial a_{i1}}=2a_{i1}tr(\rho O_{i}^{2})-2a_{i1}[tr(\rho O_{i})]^{2}+\sum\limits_{k=1,...,9,k\neq i}a_{k1}[tr(\rho O_{i}O_{k})+tr(\rho O_{k}O_{i})]
k=1,,9,kiak1[tr(ρOi)tr(ρOk)+tr(ρOk)tr(ρOi)]2λ1ai1=0.\displaystyle-\sum\limits_{k=1,...,9,k\neq i}a_{k1}[tr(\rho O_{i})tr(\rho O_{k})+tr(\rho O_{k})tr(\rho O_{i})]-2\lambda_{1}a_{i1}=0.\ \ \ \ \ \ \ \ \ \ \ \ (17)

By rearranging the above expression, we get

ai1[tr(ρOi2)(tr(ρOi))2]+{k=1,,9,kiak1[tr(ρOiOk)+tr(ρOkOi)]}/2\displaystyle a_{i1}[tr(\rho O_{i}^{2})-(tr(\rho O_{i}))^{2}]+\{\sum\limits_{k=1,...,9,k\neq i}a_{k1}[tr(\rho O_{i}O_{k})+tr(\rho O_{k}O_{i})]\}/2
k=1,,9,kiak1tr(ρOi)tr(ρOk)=λ1ai1.\displaystyle-\sum\limits_{k=1,...,9,k\neq i}a_{k1}tr(\rho O_{i})tr(\rho O_{k})=\lambda_{1}a_{i1}.\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (18)

For i=1,,9i=1,...,9, the following eigenvalue problem is obtained in compact form:

γa1=λ1a1,\gamma\textbf{\emph{a}}_{1}=\lambda_{1}\textbf{\emph{a}}_{1}, (19)

where a1=(a11,a21,,a91)\textbf{\emph{a}}_{1}=(a_{11},a_{21},...,a_{91})^{\prime}, γij=(OiOj+OjOi)/2OiOj\gamma_{ij}=(\langle O_{i}O_{j}\rangle+\langle O_{j}O_{i}\rangle)/2-\langle O_{i}\rangle\langle O_{j}\rangle, which is exactly the quantum covariance matrix as defined in Eq.(6). It shows that a1\textbf{\emph{a}}_{1} should be chosen to be an eigenvector of the covariance matrix γ\gamma, with eigenvalue λ1\lambda_{1}. The variance of the first principal component is

V(P1)=tr(a1γa1)=λ1.V(P_{1})=tr(\textbf{\emph{a}}_{1}^{\dagger}\gamma\textbf{\emph{a}}_{1})=\lambda_{1}. (20)

Therefore, in order to obtain the maximum of the variance, a1\textbf{\emph{a}}_{1} should be chosen as the eigenvector corresponding to the largest eigenvalue λ1\lambda_{1} of the covariance matrix. Similarly, for the second principal component, in order to obtain the second maximum of the variance, a2\textbf{\emph{a}}_{2} should be chosen as the eigenvector corresponding to the second largest eigenvalue λ2\lambda_{2} of the covariance matrix. This is fully consistent with the classical principal components analysis since the variances correspond to the eigenvalues of the covariance matrix.

For a arbitrary covariance matrix γij(ρAB)\gamma_{ij}(\rho_{AB}) of local observables {Om}={σiσj}(i,j=1,2,3,m=3(i1)+j)\{O_{m}\}=\{\sigma_{i}\otimes\sigma_{j}\}(i,j=1,2,3,m=3(i-1)+j), the variance of the observables OmO_{m} can be analytically given as m=19δ2(Om)=i=1Nδ2Pi\sum_{m=1}^{9}\delta^{2}(O_{m})=\sum_{i=1}^{N}\delta^{2}P_{i} due to the fact jaijaij=1\sum_{j}a_{ij}^{*}a_{ij}=1. As i=1Nδ2Pi=i=1Nλi\sum_{i=1}^{N}\delta^{2}P_{i}=\sum_{i=1}^{N}\lambda_{i}, one achieves m=19δ2(Om)=i=19λi\sum_{m=1}^{9}\delta^{2}(O_{m})=\sum_{i=1}^{9}\lambda_{i}.

References

  • (1) A. Einstein, B. Podolsky, and N. Rosen, Phys. Rev. 1935, 47, 777.
  • (2) E. Schrödinger, Proc. Cambridge Philos. Soc. 1936, 32, 446.
  • (3) H. M. Wiseman, S. J. Jones, and A. C. Doherty, Phys. Rev. Lett. 2007, 98, 140402.
  • (4) M. D. Reid, Phys. Rev. A 2013, 88, 062338.
  • (5) Q. He, L. Rosales-Za´\acute{a}rate, G. Adesso, and M. D. Reid, Phys. Rev. Lett. 2015, 115, 180502.
  • (6) N. Walk, S. Hosseini, J. Geng, O. Thearle, J. Y. Haw, S. Armstrong, S. M. Assad, J. Janousek, T. C. Ralph, T. Symul, H. M. Wiseman, and P. K. Lam, Optica 2016, 3,634.
  • (7) I. Kogias, Y. Xiang, Q. He, and G. Adesso, Phys. Rev. A 2017, 95, 012315.
  • (8) C. Branciard, E. G. Cavalcanti, S. P. Walborn, V. Scarani, and H. M. Wiseman, Phys. Rev. A 2012, 85, 010301(R).
  • (9) M. Piani and J. Watrous, Phys. Rev. Lett. 2015, 114, 060404.
  • (10) R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki, Rev. Mod. Phys. 2009, 81, 865.
  • (11) J. S. Bell, Physics 1964, 1, 195.
  • (12) S. J. Jones, H. M. Wiseman, and A. C. Doherty, Phys. Rev. A 2007, 76, 052116.
  • (13) N. Brunner, D. Cavalcanti, S. Pironio, V. Scarant, and S. Wehner, Rev. Mod. Phys. 2014, 86, 419.
  • (14) J. Bowles, T. Vértesi, M. T. Quintino, and N. Brunner, Phys. Rev. Lett. 2014, 112, 200402.
  • (15) V. Händchen, T. Eberle, S. Steinlechner, A. Samblowski, T. Franz, R. F. Werner, and R. Schnabel, Nature Photonics 2012, 6, 596.
  • (16) S. Wollmann, N. Walk, A. J. Bennet, H. M. Wiseman, and G. J. Pryde, Phys. Rev. Lett. 2016, 116, 160403.
  • (17) D. J. Saunders, S. J. Jones, H. M. Wiseman, and G. J. Pryde, Nat. Phys. 2010, 6, 845.
  • (18) J. Schneeloch, C. J. Broadbent, S. P. Walborn, E. G. Cavalcanti, and J. C. Howell, Phys. Rev. A 2013, 87, 062103.
  • (19) S. W. Ji, J. Lee, J. Park, and H. Nha, Phys. Rev. A 2015, 92, 062130.
  • (20) Y. Z. Zhen, Y. L. Zheng, W. F. Cao, L. Li, Z. B. Chen, N. L. Liu, and K. Chen, Phys. Rev. A 2016, 93, 012108.
  • (21) Y. L. Zheng, Y. Z. Zhen, Z. B. Chen, N. L. Liu, K. Chen, and J. W. Pan, Phys. Rev. A 2017, 95, 012142.
  • (22) Y. L. Zheng, Y. Z. Zhen, W. F. Cao, L. Li, Z. B. Chen, N. L. Liu, and K. Chen, Phys. Rev. A 2017, 95, 032128.
  • (23) S. X. Yu, Q. Chen, C. J. Zhang, C. H. Lai, and C. H. Oh, Phys. Rev. Lett. 109, 120402 (2012).
  • (24) K. Pearson, The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 1901, 2, 559.
  • (25) H. Hotelling, Journal of Educational Psychology 1933, 24, 417.
  • (26) I. T. Jolliffe, Principal Component Analysis, Series: Springer Series in Statistics, 2002, 2nd ed., Springer.
  • (27) D. J. Gross, Phys. Today 1995, 48(12), 46 (1995).
  • (28) K. G. H. Vollbrecht and R. F. Werner, Phys. Rev. A 2001, 64, 062307.
  • (29) J. K. Stockton, J. M. Geremia, A. C. Doherty, and H. Mabuchi, Phys. Rev. A 2003, 67, 022112.
  • (30) G. Táth and O. Gühne, Phys. Rev. Lett. 2009, 102, 170503.
  • (31) R. F. Werner, Phys. Rev. A 1989, 40, 4277.
  • (32) N. Gisin, Phys. Lett. A 1996, 210, 151.