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


Parameterized Bipartite Entanglement Measure

Zhi-Wei Wei [email protected] School of Mathematical Sciences, Capital Normal University, 100048 Beijing, China    Shao-Ming Fei [email protected] School of Mathematical Sciences, Capital Normal University, 100048 Beijing, China Max-Planck-Institute for Mathematics in the Sciences, 04103 Leipzig, Germany
Abstract

We propose a novel parameterized entanglement measure α\alpha-concurrence for bipartite systems. By employing positive partial transposition and realignment criteria, we derive analytical lower bounds for the α\alpha-concurrence. Moreover, we calculate explicitly the analytic expressions of the α\alpha-concurrence for isotropic states and Werner states.

I introduction

As a distinguishing feature of quantum mechanics, quantum entanglement is an important resource schrodinger1935gegenwartige ; PhysRev.47.777 ; PhysPhys195 in quantum computation and quantum information processing nielsen2002quantum such as quantum dense coding PhysRevLett.69.2881 , clock synchronization PRL852010 , quantum teleportation PhysRevLett.70.1895 , quantum secret sharing PhysRevA.59.1829 and quantum cryptography PRL67661 . One of the problems in quantum entanglement theory is to quantify the entanglement of a bipartite system. A reasonable entanglement measure EE should fulfill PhysRevLett.78.2275 ; PhysRevA.57.1619 ; PhysRevA.56.R3319 ; PhysRevA.59.141 ; PRL842014 : (E1) E(ρ)0E\left(\rho\right)\geqslant 0, with the equality holding iff ρ\rho is separable; (E2) EE is convex, i.e., E(tρ+(1t)σ)tE(ρ)+(1t)E(σ)E\left(t\rho+\left(1-t\right)\sigma\right)\leqslant tE\left(\rho\right)+\left(1-t\right)E\left(\sigma\right) for any states ρ\rho and σ\sigma, where t[0,1]t\in[0,1]; (E3) EE does not increase under local operation and classical communication (LOCC), i.e., E(ρ)E(Λ(ρ))E\left(\rho\right)\geqslant E\left(\Lambda\left(\rho\right)\right) for any completely positive and trace preserving (CPTP) LOCC Λ\Lambda; (E4) EE does not increase on average under stochastic LOCC, i.e., E(ρ)kpkE(ρk)E\left(\rho\right)\geqslant\sum_{k}p_{k}E\left(\rho_{k}\right), where pk=TrAkρAkp_{k}=\mathrm{Tr}A_{k}\rho A_{k}^{\dagger}, ρk=AkρAk/pk\rho_{k}=A_{k}\rho A_{k}^{\dagger}/p_{k} and AkA_{k} are the Kraus operators such that kAkAk=I\sum_{k}A_{k}^{\dagger}A_{k}=I with II the identity.

Several reasonable entanglement measures have been presented in the past few years, such as concurrence PhysRevLett.78.5022 ; PhysRevA.64.042315 , entanglement of formation PhysRevA.54.3824 ; PhysRevLett.80.2245 ; horodecki2001entanglement , negativity PhysRevA.65.032314 ; PhysRevLett.95.090503 and Rényi-α\alpha entropy of entanglement gour2007dual ; Kim_2010 . Generally it is formidably difficult to calculate analytically the degree of entanglement for arbitrary given states. Analytical results are often available for two-qubit states PhysRevLett.80.2245 or special higher-dimensional mixed states PhysRevLett.85.2625 ; PhysRevA.64.062307 ; PhysRevA.67.012307 for several entanglement measures PhysRevA.68.062304 ; buchholz2016evaluating . Efforts have been also made towards the analytical lower bounds of entanglement measures like the concurrence PhysRevLett.95.210501 ; Li_Guo_2009 . In PhysRevLett.95.040504 analytical lower bound of the concurrence has been derived by employing the positive partial transpose (PPT) and realignment criteria PhysRevLett.77.1413 ; PhysRevA.59.4206 ; rudolph2004computable ; rudolph2005further ; chen2003matrix . In this paper, inspired by Tsallis-qq entropy of entanglement raggio1995properties ; PhysRevA.81.062328 and parameterized entanglement monotone qq-concurrence PhysRevA.103.052423 , we propose a novel parameterized entanglement measure α\alpha-concurrence for any 0α1/20\leqslant\alpha\leqslant 1/2. Based on PPT and realignment criteria, we also obtain an analytic lower bound of the α\alpha-concurrence for general bipartite systems. Finally, we calculate the α\alpha-concurrence of isotropic states and Werner states analytically.

II α\alpha-concurrence

Let AB\mathcal{H}_{A}\otimes\mathcal{H}_{B} be an arbitrary d×dd\times d dimensional bipartite Hilbert space associated with subsystems AA and BB. Any pure state on the AB\mathcal{H}_{A}\otimes\mathcal{H}_{B} can be written as in the Schmidt form,

|ψ=i=1rλi|aibi,\ket{\psi}=\sum_{i=1}^{r}\sqrt{\lambda_{i}}\ket{a_{i}b_{i}}, (1)

where i=1rλi=1\sum_{i=1}^{r}\lambda_{i}=1 with λi>0\lambda_{i}>0, rr is the Schmidt rank, 1rd1\leqslant r\leqslant d. {|ai}\left\{\ket{a_{i}}\right\} and {|bi}\left\{\ket{b_{i}}\right\} are the local bases associated with the subsystems AA and BB, respectively nielsen2002quantum .

𝐷𝑒𝑓𝑖𝑛𝑖𝑡𝑖𝑜𝑛\mathit{Definition}. For any pure state |ψ\ket{\psi} given in (1), the α\alpha-concurrence is defined by

Cα(|ψ)=TrρAα1C_{\alpha}\left(\ket{\psi}\right)=\mathrm{Tr}\rho_{A}^{\alpha}-1 (2)

for any 0α1/20\leqslant\alpha\leqslant 1/2, where ρA=TrB|ψψ|\rho_{A}=\mathrm{Tr}_{B}|\psi\rangle\langle\psi|.

From (2), for a pure state |ψ\ket{\psi} given by (1) one has

Cα(|ψ)=i=1rλiα1,C_{\alpha}\left(\ket{\psi}\right)=\sum_{i=1}^{r}\lambda_{i}^{\alpha}-1, (3)

where Cα(|ψ)C_{\alpha}\left(\ket{\psi}\right) satisfies 0Cα(|ψ)d1α10\leqslant C_{\alpha}\left(\ket{\psi}\right)\leqslant d^{1-\alpha}-1. It is obvious that the lower bound is attained if and only if |ψ\ket{\psi} is a separable state, that is, |ψ=|aibi\ket{\psi}=\ket{a_{i}b_{i}} for some |ai\ket{a_{i}} and |bi\ket{b_{i}}. While the upper bound is achieved for the maximally entangled pure states |Ψ+=1di=1d|aibi\ket{\Psi^{+}}=\frac{1}{\sqrt{d}}\sum_{i=1}^{d}\ket{a_{i}b_{i}}.

For a general mixed state ρ\rho on the Hilbert space AB\mathcal{H}_{A}\otimes\mathcal{H}_{B}, the α\alpha-concurrence is given by the convex-roof extension,

Cα(ρ)=min{pi,|ψi}ipiCα(|ψi),C_{\alpha}\left(\rho\right)=\min_{\left\{p_{i},\ket{\psi_{i}}\right\}}\sum_{i}p_{i}C_{\alpha}\left(\ket{\psi_{i}}\right), (4)

where the infimum is taken over all possible pure-state decompositions of ρ=ipi|ψiψi|\rho=\sum_{i}p_{i}|\psi_{i}\rangle\langle\psi_{i}|, with ipi=1\sum_{i}p_{i}=1 and pi>0p_{i}>0. Before showing that Cα(ρ)C_{\alpha}\left(\rho\right) defined in (4) is indeed a bona fide entanglement measure, we first present the following lemma, see proof in Appendix A.

𝐿𝑒𝑚𝑚𝑎 1\mathit{Lemma\ 1}. The function Fα(ρ)=Trρα1F_{\alpha}\left(\rho\right)=\mathrm{Tr}\rho^{\alpha}-1 is concave, that is,

Fα(ipiρi)ipiFα(ρi)F_{\alpha}\left(\sum_{i}p_{i}\rho_{i}\right)\geqslant\sum_{i}p_{i}F_{\alpha}\left(\rho_{i}\right) (5)

for any 0α1/20\leqslant\alpha\leqslant 1/2, where {pi}\set{p_{i}} is a probability distribution and ρi\rho_{i} are density matrices. The equality holds if and only if all ρi\rho_{i} are the same for all pi>0p_{i}>0.

By using the above Lemma 1, we have the following theorem.

𝑇ℎ𝑒𝑜𝑟𝑒𝑚 1\mathit{Theorem\,1}. The α\alpha-concurrence Cα(ρ)C_{\alpha}\left(\rho\right) given in (4) is a well defined parameterized entanglement measure.

𝑃𝑟𝑜𝑜𝑓\mathit{Proof}. We need to verify that Cα(ρ)C_{\alpha}\left(\rho\right) fulfills the following four requirements.

(E1) If ρ\rho is an entangled state, then there is at least one entangled pure state |ψ\ket{\psi} in any pure state decomposition of ρ\rho. Thus Cα(ρ)>0C_{\alpha}\left(\rho\right)>0. Otherwise, Cα(ρ)=0C_{\alpha}\left(\rho\right)=0 for separable states.

(E2) Consider ρ=tρ1+(1t)ρ2\rho=t\rho_{1}+\left(1-t\right)\rho_{2}. Let ρ1=ipi|ψiψi|\rho_{1}=\sum_{i}p_{i}|\psi_{i}\rangle\langle\psi_{i}| (ρ2=jqj|ϕjϕj|\rho_{2}=\sum_{j}q_{j}|\phi_{j}\rangle\langle\phi_{j}|) be the optimal pure state decomposition of Cα(ρ1)C_{\alpha}\left(\rho_{1}\right) (Cα(ρ2)C_{\alpha}\left(\rho_{2}\right)) with ipi=1\sum_{i}p_{i}=1 (jqj=1\sum_{j}q_{j}=1) and pi>0p_{i}>0 (qj>0q_{j}>0). We have

Cα(ρ)\displaystyle C_{\alpha}\left(\rho\right) i=1ktpiCα(|ψi)+j=1l(1t)qjCα(|ϕj)\displaystyle\leqslant\sum_{i=1}^{k}tp_{i}C_{\alpha}\left(\ket{\psi_{i}}\right)+\sum_{j=1}^{l}\left(1-t\right)q_{j}C_{\alpha}\left(\ket{\phi_{j}}\right)
=tCα(ρ1)+(1t)Cα(ρ2),\displaystyle=tC_{\alpha}\left(\rho_{1}\right)+\left(1-t\right)C_{\alpha}\left(\rho_{2}\right), (6)

where the first inequality is due to that i=1ktpi|ψiψi|+j=1l(1t)qj|ϕjϕj|\sum_{i=1}^{k}tp_{i}|\psi_{i}\rangle\langle\psi_{i}|+\sum_{j=1}^{l}\left(1-t\right)q_{j}|\phi_{j}\rangle\langle\phi_{j}| is also a pure state decomposition of ρ\rho.

(E3) We adopt the approach given in MINTERT2005207 to show that our entanglement measure does not increase under LOCC. Denote λψ\vec{\lambda}_{\psi} (λϕ\vec{\lambda}_{\phi}) the Schmidt vector given by the squared Schmidt coefficients of the state |ψ\ket{\psi} (|ϕ\ket{\phi}) in the decreasing order. It has been shown that the state |ϕ\ket{\phi} can be prepared starting from the state |ψ\ket{\psi} under LOCC if and only if λψ\vec{\lambda}_{\psi} is majorized by λϕ\vec{\lambda}_{\phi} PhysRevLett.83.436 , λψλϕ\vec{\lambda}_{\psi}\prec\vec{\lambda}_{\phi}, where the majorization means that the components [λψ]i\left[\lambda_{\psi}\right]_{i} ([λϕ]i\left[\lambda_{\phi}\right]_{i}) of λψ\vec{\lambda}_{\psi} (λϕ\vec{\lambda}_{\phi}), listed in nonincreasing order, satisfy i=1j[λψ]ii=1j[λϕ]i\sum_{i=1}^{j}\left[\lambda_{\psi}\right]_{i}\leqslant\sum_{i=1}^{j}\left[\lambda_{\phi}\right]_{i} for 1<jd1<j\leqslant d, with equality for j=dj=d.

Since the entanglement cannot increase under LOCC, any entanglement measure EE has to satisfy that E(ψ)E(ϕ)E\left(\psi\right)\geqslant E\left(\phi\right) whenever λψλϕ\vec{\lambda}_{\psi}\prec\vec{\lambda}_{\phi}. This condition, known as the 𝑆𝑐ℎ𝑢𝑟𝑐𝑜𝑛𝑐𝑎𝑣𝑖𝑡𝑦\mathit{Schur\,concavity}, is satisfied if and only if EE, given as a function of the squared Schmidt coefficients λi\lambda_{i}’s ANDO1989163 , is invariant under the permutations of any two arguments and satisfies

(λiλj)(EλiEλj)0\displaystyle\left(\lambda_{i}-\lambda_{j}\right)\left(\frac{\partial E}{\partial\lambda_{i}}-\frac{\partial E}{\partial\lambda_{j}}\right)\leqslant 0 (7)

for any two components λi\lambda_{i} and λj\lambda_{j} of λ\vec{\lambda}.

For any pure state |ψ\ket{\psi} given by (1), the α\alpha-concurrence is obviously invariant under the permutations of the Schmidt coefficients for any 0α1/20\leqslant\alpha\leqslant 1/2. Since

(λiλj)(CαλiCαλj)\displaystyle\left(\lambda_{i}-\lambda_{j}\right)\left(\frac{\partial C_{\alpha}}{\partial\lambda_{i}}-\frac{\partial C_{\alpha}}{\partial\lambda_{j}}\right)
=α(λiλj)(λiα1λjα1)0\displaystyle=\alpha\left(\lambda_{i}-\lambda_{j}\right)\left(\lambda_{i}^{\alpha-1}-\lambda_{j}^{\alpha-1}\right)\leqslant 0

for any two components λi\lambda_{i} and λj\lambda_{j} of the squared Schmidt coefficients of |ψ\ket{\psi}, we have Cα(|ψ)Cα(Λ|ψ)C_{\alpha}\left(\ket{\psi}\right)\geqslant C_{\alpha}\left(\Lambda\ket{\psi}\right) for any LOCC Λ\Lambda and 0α1/20\leqslant\alpha\leqslant 1/2.

Next let ρ=ipi|ψiψi|\rho=\sum_{i}p_{i}|\psi_{i}\rangle\langle\psi_{i}| be the optimal pure state decomposition of Cα(ρ)C_{\alpha}\left(\rho\right) with ipi=1\sum_{i}p_{i}=1 and pi>0p_{i}>0. We obtain

Cα(ρ)\displaystyle C_{\alpha}\left(\rho\right) =ipiCα(|ψi)\displaystyle=\sum_{i}p_{i}C_{\alpha}\left(\ket{\psi_{i}}\right)
ipiCα(Λ|ψi)\displaystyle\geqslant\sum_{i}p_{i}C_{\alpha}\left(\Lambda\ket{\psi_{i}}\right)
Cα(Λ(ρ))\displaystyle\geqslant C_{\alpha}\left(\Lambda\left(\rho\right)\right)

for any 0α1/20\leqslant\alpha\leqslant 1/2, where the last inequality is from the definition (4).

(E4) Let ρ=ipi|ψiψi|\rho=\sum_{i}p_{i}|\psi_{i}\rangle\langle\psi_{i}| be the optimal pure state decomposition of Cα(ρ)C_{\alpha}\left(\rho\right) with ipi=1\sum_{i}p_{i}=1 and pi>0p_{i}>0. Consider stochastic LOCC protocol given by Kraus operators AkA_{k} with kAkAk=I\sum_{k}A_{k}^{\dagger}A_{k}=I. We have

Cα(ρ)\displaystyle C_{\alpha}\left(\rho\right) =ipiCα(|ψi)\displaystyle=\sum_{i}p_{i}C_{\alpha}\left(\ket{\psi_{i}}\right)
i,kpipk|iCα(|ψik)\displaystyle\geqslant\sum_{i,k}p_{i}p_{k|i}C_{\alpha}\left(\ket{\psi_{i}^{k}}\right)
=i,kpkpi|kCα(|ψik)\displaystyle=\sum_{i,k}p_{k}p_{i|k}C_{\alpha}\left(\ket{\psi_{i}^{k}}\right)
=kpk(ipi|kCα(|ψik))\displaystyle=\sum_{k}p_{k}\left(\sum_{i}p_{i|k}C_{\alpha}\left(\ket{\psi_{i}^{k}}\right)\right)
kpkCα(ρk),\displaystyle\geqslant\sum_{k}p_{k}C_{\alpha}\left(\rho^{k}\right), (8)

where pk|i=TrAk|ψiψi|Akp_{k|i}=\mathrm{Tr}A_{k}|\psi_{i}\rangle\langle\psi_{i}|A_{k}^{\dagger} is the probability of obtaining the outcome kk with |ψik=Ak|ψi/pk|i\ket{\psi_{i}^{k}}=A_{k}\ket{\psi_{i}}/\sqrt{p_{k|i}}, and pk=TrAkρAkp_{k}=\mathrm{Tr}A_{k}\rho A_{k}^{\dagger} is the probability of obtaining the outcome kk with ρk=AkρAk/pk\rho^{k}=A_{k}\rho A_{k}^{\dagger}/p_{k}. The first inequality is due to the concavity of the Lemma 1, since TrB|ψiψi|=kpk|iTrB|ψikψik|\mathrm{Tr}_{B}|\psi_{i}\rangle\langle\psi_{i}|=\sum_{k}p_{k|i}\mathrm{Tr}_{B}|\psi_{i}^{k}\rangle\langle\psi_{i}^{k}|. The last inequality is from the definition of (4), since ipi|k|ψikψik|=ρk\sum_{i}p_{i|k}|\psi_{i}^{k}\rangle\langle\psi_{i}^{k}|=\rho^{k}. \hfill\qed

In PhysRevA.103.052423 the parameterized entanglement monotone qq-concurrence Cq(|ψ)C_{q}\left(\ket{\psi}\right) for any pure state |ψ\ket{\psi} defined in (1) has been introduced, Cq(|ψ)=1TrρAqC_{q}\left(\ket{\psi}\right)=1-\mathrm{Tr}\rho_{A}^{q}, where q2q\geqslant 2. It seems that our α\alpha-concurrence defined by (2), Cα(|ψ)=TrρAα1C_{\alpha}\left(\ket{\psi}\right)=\mathrm{Tr}\rho_{A}^{\alpha}-1, is in some sense dual to the qq-concurrence as the parameter α[0,1/2]\alpha\in[0,1/2], while the parameter q2q\geqslant 2. Nevertheless, these two concurrences characterize the quantum entanglement in different aspects, even though they are both derived from the Tsallis-qq of entanglement PhysRevA.81.062328 . For large enough qq, the qq-concurrence Cq(ρ)C_{q}(\rho) converges to the constant 11 for any entangled state ρ\rho, while the α\alpha-concurrence Cα(ρ)C_{\alpha}(\rho) not for any α[0,1/2]\alpha\in[0,1/2]. Particularly, for α=0\alpha=0, the measure C0(|ψ)=r1C_{0}\left(\ket{\psi}\right)=r-1 for any pure state |ψ\ket{\psi} given in (1), where rr is the Schmidt rank of the state |ψ\ket{\psi}, which is solely determined by the Schmidt rank of the bipartite pure state |ψ\ket{\psi}. Therefore, the α\alpha-concurrences with different α\alpha provide different characterizations of the feature of entanglement.

III bounds on α\alpha-concurrence

Owing to the optimization in the calculation of the entanglement measures, it is generally difficult to obtain analytical expressions of the entanglement measures for general mixed states. In this section, we derive analytical lower bounds for the α\alpha-concurrence based on PPT and realignment criteria PhysRevLett.77.1413 ; HORODECKI19961 ; PhysRevA.59.4206 ; rudolph2004computable ; rudolph2005further ; chen2003matrix .

A bipartite state can be written as ρ=ijklρij,kl|ijkl|\rho=\sum_{ijkl}\rho_{ij,kl}|ij\rangle\langle kl|, where the subscripts ii and kk are the row and column indices for the subsystem AA, respectively, while jj and ll are such indices for the subsystem BB. The PPT criterion says that if the state ρ\rho is separable, then the partial transposed matrix ρΓ=ijklρij,kl|ilkj|\rho^{\Gamma}=\sum_{ijkl}\rho_{ij,kl}|il\rangle\langle kj| with respect to the subsystem BB is non-negative, ρΓ0\rho^{\Gamma}\geqslant 0. While the realignment criterion says that the realigned matrix of ρ\rho, (ρ)=ijklρij,kl|ikjl|\mathcal{R}\left(\rho\right)=\sum_{ijkl}\rho_{ij,kl}|ik\rangle\langle jl|, satisfies that (ρ)11\|\mathcal{R}\left(\rho\right)\|_{1}\leqslant 1 if ρ\rho is separable, where X1\|X\|_{1} denotes the trace norm of matrix XX, X1=TrXX\|X\|_{1}=\mathrm{Tr}\sqrt{XX^{\dagger}}.

For a pure state |ψ\ket{\psi} given by (1), it is straightforward to obtain that PhysRevLett.95.040504

1ρΓ1=(ρ)1=(i=1rλi)2r,1\leqslant\left\|\rho^{\Gamma}\right\|_{1}=\left\|\mathcal{R}\left(\rho\right)\right\|_{1}=\left(\sum_{i=1}^{r}\sqrt{\lambda_{i}}\right)^{2}\leqslant r, (9)

where ρ=|ψψ|\rho=|\psi\rangle\langle\psi|. In particular, for α=1/2\alpha=1/2, the 1/21/2-concurrence becomes C12(|ψ)=i=1rλi1C_{\frac{1}{2}}\left(\ket{\psi}\right)=\sum_{i=1}^{r}\sqrt{\lambda_{i}}-1. One has then

C12(|ψ)ρΓ11r+1\displaystyle C_{\frac{1}{2}}\left(\ket{\psi}\right)\geqslant\frac{\left\|\rho^{\Gamma}\right\|_{1}-1}{\sqrt{r}+1} (10)

for any pure state |ψ\ket{\psi} on the AB\mathcal{H}_{A}\otimes\mathcal{H}_{B}.

𝑇ℎ𝑒𝑜𝑟𝑒𝑚 2\mathit{Theorem\ 2}. For any mixed state ρ\rho on the AB\mathcal{H}_{A}\otimes\mathcal{H}_{B}, the α\alpha-concurrence Cα(ρ)C_{\alpha}\left(\rho\right) satisfies

Cα(ρ)d1α1d1[max(ρΓ1,(ρ)1)1].C_{\alpha}\left(\rho\right)\geqslant\frac{d^{1-\alpha}-1}{d-1}\left[\max\left(\|\rho^{\Gamma}\|_{1},\|\mathcal{R}\left(\rho\right)\|_{1}\right)-1\right]. (11)

𝑃𝑟𝑜𝑜𝑓\mathit{Proof}. For a pure state |ψ\ket{\psi} given in (1), let us analyze the monotonicity of the following function,

g(α)=i=1rλiα1r1α1g\left(\alpha\right)=\frac{\sum_{i=1}^{r}\lambda_{i}^{\alpha}-1}{r^{1-\alpha}-1} (12)

for any 0α1/20\leqslant\alpha\leqslant 1/2, where r>1r>1. The first derivative of g(α)g\left(\alpha\right) with respect to α\alpha is given by

gα\displaystyle\frac{\partial g}{\partial\alpha} =Grα(r1α1)2,\displaystyle=\frac{G_{r\alpha}}{\left(r^{1-\alpha}-1\right)^{2}}, (13)

where

Grα=iλiαlnλi(r1α1)+(iλiα1)r1αlnr.G_{r\alpha}=\sum_{i}\lambda_{i}^{\alpha}\ln\lambda_{i}\left(r^{1-\alpha}-1\right)+\left(\sum_{i}\lambda_{i}^{\alpha}-1\right)r^{1-\alpha}\ln r. (14)

Employing the Lagrange multiplies PhysRevA.67.012307 under constraints i=1rλi=1\sum_{i=1}^{r}\lambda_{i}=1 and λi>0\lambda_{i}>0, one has that there is only one stable point λi=1/r\lambda_{i}=1/r for every i=1,,ri=1,\cdots,r, for which Grα=0G_{r\alpha}=0 for any 0α1/20\leqslant\alpha\leqslant 1/2. Since the second derivative at this point,

2Grαλi2|λi=1r\displaystyle\frac{\partial^{2}G_{r\alpha}}{\partial\lambda_{i}^{2}}\Big{|}_{\lambda_{i}=\frac{1}{r}}
=r2α{α(α1)lnr+(2α1)(r1α1)}\displaystyle=r^{2-\alpha}\left\{\alpha\left(\alpha-1\right)\ln r+\left(2\alpha-1\right)\left(r^{1-\alpha}-1\right)\right\}
<0\displaystyle<0 (15)

for any 0α1/20\leqslant\alpha\leqslant 1/2, the maximum extreme value point is just the maximum value point. g(α)g\left(\alpha\right) is a decreasing function for any 0α1/20\leqslant\alpha\leqslant 1/2, since g/α0\partial g/\partial\alpha\leqslant 0.

We have

Cα(|ψ)\displaystyle C_{\alpha}\left(\ket{\psi}\right) r1α1r1C12(|ψ)\displaystyle\geqslant\frac{r^{1-\alpha}-1}{\sqrt{r}-1}C_{\frac{1}{2}}\left(\ket{\psi}\right)
r1α1r1(σΓ11)\displaystyle\geqslant\frac{r^{1-\alpha}-1}{r-1}\left(\|\sigma^{\Gamma}\|_{1}-1\right)
d1α1d1(σΓ11),\displaystyle\geqslant\frac{d^{1-\alpha}-1}{d-1}\left(\|\sigma^{\Gamma}\|_{1}-1\right), (16)

where σ=|ψψ|\sigma=|\psi\rangle\langle\psi|, the second inequality is due to (10), the last inequality is due to that r1α1r1\frac{r^{1-\alpha}-1}{r-1} is a decreasing function with respect to rr. Assume ρ=ipi|φiφi|\rho=\sum_{i}p_{i}|\varphi_{i}\rangle\langle\varphi_{i}| is the optimal pure state decomposition for Cα(ρ)C_{\alpha}\left(\rho\right). Then

Cα(ρ)\displaystyle C_{\alpha}\left(\rho\right) =ipiCα(|φi)\displaystyle=\sum_{i}p_{i}C_{\alpha}\left(\ket{\varphi_{i}}\right)
d1α1d1ipi(σiΓ11)\displaystyle\geqslant\frac{d^{1-\alpha}-1}{d-1}\sum_{i}p_{i}\left(\|\sigma_{i}^{\Gamma}\|_{1}-1\right)
d1α1d1(ρΓ11),\displaystyle\geqslant\frac{d^{1-\alpha}-1}{d-1}\left(\|\rho^{\Gamma}\|_{1}-1\right), (17)

where σi=|φiφi|\sigma_{i}=|\varphi_{i}\rangle\langle\varphi_{i}|, the last inequality is due to the convex property of the trace norm and ρΓ1\|\rho^{\Gamma}\|\geq 1 in (9).

Similar to (III) and (III), we obtain from (9) that

Cα(ρ)d1α1d1((ρ)11)C_{\alpha}\left(\rho\right)\geqslant\frac{d^{1-\alpha}-1}{d-1}\left(\|\mathcal{R}\left(\rho\right)\|_{1}-1\right) (18)

for any 0α1/20\leqslant\alpha\leqslant 1/2.

Combining (III) and (18), we complete the proof. \hfill\qed

IV α\alpha-concurrence for Isotropic and Werner states

In this section, we compute the α\alpha-concurrence for isotropic states and Werner states. Let EE be a convex-roof extended quantum entanglement measure. Denote SS the set of states and PP the set all pure states in SS. Let GG be a compact group acting on SS by (U,ρ)UρU\left(U,\rho\right)\mapsto U\rho U^{\dagger}. Assume that the measure EE defined on PP is invariant under the operations of GG. One can define the projection 𝐏\mathbf{P}:SSS\rightarrow S by 𝐏ρ=𝑑UUρU\mathbf{P}\rho=\int dUU\rho U^{\dagger} with the standard (normalized) Haar measure dUdU on GG, and the function η\eta on 𝐏S\mathbf{P}S by

η(ρ)=min{E(|Ψ):|ψP,𝐏|ψψ|=ρ}.\eta\left(\rho\right)=\min\left\{E\left(\ket{\Psi}\right):\ket{\psi}\in P,\,\mathbf{P}|\psi\rangle\langle\psi|=\rho\right\}. (19)

Then for ρ𝐏S\rho\in\mathbf{P}S, we have

E(ρ)=co(η(ρ)),E\left(\rho\right)=co\left(\eta\left(\rho\right)\right), (20)

where co(f)co\left(f\right) is the convex-roof extension of a function ff. In other words, it is the convex hull of ff.

IV.1 Isotropic states

The isotropic states ρF\rho_{F} are given by PhysRevA.59.4206 ,

ρF=1Fd21(I|ΨΨ|)+F|ΨΨ|,\rho_{F}=\frac{1-F}{d^{2}-1}\left(I-|\Psi\rangle\langle\Psi|\right)+F|\Psi\rangle\langle\Psi|, (21)

where |Ψ=1di=1d|ii\ket{\Psi}=\frac{1}{\sqrt{d}}\sum_{i=1}^{d}\ket{ii} and F=Ψ|ρF|ΨF=\braket{\Psi}{\rho_{F}}{\Psi}. ρF\rho_{F} is separable if and only if 0F1/d0\leqslant F\leqslant 1/d PhysRevLett.85.2625 . Inspired by the techniques adopted in PhysRevLett.85.2625 ; PhysRevA.67.012307 ; PhysRevA.64.062307 ; PhysRevA.68.062304 , we have, see Appendix B,

ηα(ρF)=(Fd)1α1\eta_{\alpha}\left(\rho_{F}\right)=\left(Fd\right)^{1-\alpha}-1 (22)

for any 0α1/20\leqslant\alpha\leqslant 1/2, where F>1/dF>1/d.

Obviously, the second derivative of (22) with respect to FF is non-positive for any 0α1/20\leqslant\alpha\leqslant 1/2. Hence, ηα(ρF)\eta_{\alpha}\left(\rho_{F}\right) is concave in the whole regime F(1/d,1]F\in\left(1/d,1\right]. The Cα(ρF)C_{\alpha}\left(\rho_{F}\right) is the largest convex function that is upper bounded by ηα(ρF)\eta_{\alpha}\left(\rho_{F}\right), which is constructed in the following way. Find the line that passes through the points (F=1/d,ηα=0)\left(F=1/d,\eta_{\alpha}=0\right) and (F=1,ηα=d1α1)\left(F=1,\eta_{\alpha}=d^{1-\alpha}-1\right) of ηα(ρF)\eta_{\alpha}\left(\rho_{F}\right) for any 0α1/20\leqslant\alpha\leqslant 1/2. Thus, we have the following analytical formula of the α\alpha-concurrence for isotropic states.

𝐿𝑒𝑚𝑚𝑎 2\mathit{Lemma\ 2}. The α\alpha-concurrence for isotropic states ρFdd(d2)\rho_{F}\in\mathbb{C}^{d}\otimes\mathbb{C}^{d}\ \left(d\geqslant 2\right) is given by

Cα(ρF)={0,F1/d,d1α1d1(dF1),F>1/d,C_{\alpha}\left(\rho_{F}\right)=\begin{cases}0,&F\leqslant 1/d,\\[5.69054pt] \displaystyle\frac{d^{1-\alpha}-1}{d-1}\left(dF-1\right),&F>1/d,\end{cases} (23)

where 0α1/20\leqslant\alpha\leqslant 1/2 and d2d\geqslant 2.

Since ρFΓ1=(ρF)1=Fd\|\rho_{F}^{\Gamma}\|_{1}=\|\mathcal{R}\left(\rho_{F}\right)\|_{1}=Fd for F>1/dF>1/d rudolph2005further ; PhysRevA.65.032314 , surprisingly the lower bound of (11) is just exactly the (23) for every 0α1/20\leqslant\alpha\leqslant 1/2 with d2d\geqslant 2.

The concurrence C(ρF)C\left(\rho_{F}\right) of isotropic states has been derived in PhysRevA.67.012307 , C(ρF)=2d(d1)(dF1)C\left(\rho_{F}\right)=\sqrt{\frac{2}{d\left(d-1\right)}}\left(dF-1\right) for any F>1/dF>1/d. Fig. 1 exhibits the relations between the concurrence and the α\alpha-concurrence of isotropic states for α=0\alpha=0 and 1/21/2.

Refer to caption
Figure 1: Entanglement of isotropic states. The green (blue) surface stands for C0(ρF)C_{0}\left(\rho_{F}\right) (C1/2(ρF)C_{1/2}\left(\rho_{F}\right)), the red surface is for the concurrence C(ρF)C\left(\rho_{F}\right).

Especially, it shows that the C0(ρF)=C(ρF)C_{0}\left(\rho_{F}\right)=C\left(\rho_{F}\right) with d=2d=2, and the concurrence of isotropic states is less than the 0-concurrence with d>2d>2. Moreover, we notice that the 1/21/2-concurrence is bigger than the concurrence with d5.1508d\geq 5.1508.

IV.2 Werner states

The Werner states are of the form,

ρW=\displaystyle\rho_{W}= 2(1W)d(d+1)(k=1d|kkkk|+i<j|Ψij+Ψij+|)\displaystyle\frac{2\left(1-W\right)}{d\left(d+1\right)}\left(\sum_{k=1}^{d}|kk\rangle\langle kk|+\sum_{i<j}|\Psi_{ij}^{+}\rangle\langle\Psi_{ij}^{+}|\right)
+2Wd(d1)i<j|ΨijΨij|,\displaystyle+\frac{2W}{d\left(d-1\right)}\sum_{i<j}|\Psi_{ij}^{-}\rangle\langle\Psi_{ij}^{-}|, (24)

where |Ψij±=(|ij±|ji)/2\ket{\Psi_{ij}^{\pm}}=(\ket{ij}\pm\ket{ji})/\sqrt{2} and W=Tr(ρWi<j|ΨijΨij|)W=\mathrm{Tr}(\rho_{W}\sum_{i<j}|\Psi_{ij}^{-}\rangle\langle\Psi_{ij}^{-}|) PhysRevA.68.062304 . ρW\rho_{W} is separable if and only if 0W1/20\leqslant W\leqslant 1/2 PhysRevA.64.062307 ; PhysRevA.40.4277 . For W>1/2W>1/2, we have, see Appendix C,

ηα(ρW)=(2W)1α1\eta_{\alpha}\left(\rho_{W}\right)=\left(2W\right)^{1-\alpha}-1 (25)

for any 0α1/20\leqslant\alpha\leqslant 1/2.

It is direct to verify that the second derivative of (25) with respect to WW is non-positive, namely, ηα(ρW)\eta_{\alpha}\left(\rho_{W}\right) is concave. Similar to (23), we have

𝐿𝑒𝑚𝑚𝑎 3\mathit{Lemma\ 3}. The α\alpha-concurrence for Werner states ρWdd(d2)\rho_{W}\in\mathbb{C}^{d}\otimes\mathbb{C}^{d}\ \left(d\geqslant 2\right) is given by

Cα(ρW)={0,W1/2,(21α1)(2W1),W>1/2,C_{\alpha}\left(\rho_{W}\right)=\begin{cases}0,&W\leqslant 1/2,\\[2.84526pt] \left(2^{1-\alpha}-1\right)\left(2W-1\right),&W>1/2,\end{cases} (26)

where 0α1/20\leqslant\alpha\leqslant 1/2.

We remark that for W>1/2W>1/2, the lower bound of (11) for Werner states is given by

Cα(ρW)2(d1α1)d(d1)(2W1).\displaystyle C_{\alpha}\left(\rho_{W}\right)\geqslant\frac{2\left(d^{1-\alpha}-1\right)}{d\left(d-1\right)}\left(2W-1\right). (27)

Accounting to (26), we obtain

(21α1)(2W1)2(d1α1)d(d1)(2W1),\displaystyle\left(2^{1-\alpha}-1\right)\left(2W-1\right)\geqslant\frac{2\left(d^{1-\alpha}-1\right)}{d\left(d-1\right)}\left(2W-1\right), (28)

where equality holds if d=2d=2, and the inequality holds strictly for higher dimensional quantum systems.

The concurrence of Werner states has been obtained in PhysRevA62044302 , C(ρW)=2W1C\left(\rho_{W}\right)=2W-1 for W>1/2W>1/2. It is direct to find that Cα(ρW)=(21α1)C(ρW)C_{\alpha}\left(\rho_{W}\right)=\left(2^{1-\alpha}-1\right)C\left(\rho_{W}\right) for any 0α1/20\leq\alpha\leq 1/2. Moreover, the entanglement of formation for Werner states is given by PhysRevA.64.062307 , EF(ρW)=H2[12(12W(1W))]E_{F}\left(\rho_{W}\right)=H_{2}[\frac{1}{2}(1-2\sqrt{W\left(1-W\right)})]. In Fig. 2 we illustrate the relations among the concurrence, entanglement of formation and α\alpha-concurrence of Werner states for α=0\alpha=0 and 1/21/2. The entanglement of formation for Werner states EF(ρW)E_{F}\left(\rho_{W}\right) is always upper bounded by the C0(ρW)=C(ρW)C_{0}\left(\rho_{W}\right)=C\left(\rho_{W}\right), and larger than the 1/21/2-concurrence C1/2(ρW)C_{1/2}\left(\rho_{W}\right) for W0.6W\geq 0.6.

Refer to caption
Figure 2: Entanglement for Werner states. The dashed (green) line stands for C0(ρW)=C(ρW)C_{0}\left(\rho_{W}\right)=C\left(\rho_{W}\right). The solid (red) line is for the entanglement of formation EF(ρW)E_{F}\left(\rho_{W}\right), the dot dashed (blue) line is for the 1/21/2-concurrence C1/2(ρW)C_{1/2}\left(\rho_{W}\right).

V summary

We have introduced the concept of α\alpha-concurrence and shown that the α\alpha-concurrence is a well defined entanglement measure. Analytical lower bounds of the α\alpha-concurrence for general mixed states have been derived based on PPT and realignment criterion. Specifically, we have derived explicit formulae for the α\alpha-concurrence of isotropic states and Werner states. Interestingly our lower bounds are exact for isotropic states and Werner states with d=2d=2. Our parameterized entanglement measure α\alpha-concurrence gives a family of entanglement measures and enriches the theory of quantum entanglement, which may highlight further researches on the study of quantifying quantum entanglement and the related investigations like monogamy and polygamy relations in entanglement distribution, as well as the physical understanding of quantum correlations.


Acknowledgments

This work is supported by the National Natural Science Foundation of China (NSFC) under Grant Nos. 12075159 and 12171044; Beijing Natural Science Foundation (Grant No. Z190005); Academy for Multidisciplinary Studies, Capital Normal University; Shenzhen Institute for Quantum Science and Engineering, Southern University of Science and Technology (No. SIQSE202001), the Academician Innovation Platform of Hainan Province.

Appendix A Proof of Lemma 1

We only need to prove that Fα(tρ+(1t)σ)tFα(ρ)+(1t)Fα(σ)F_{\alpha}\left(t\rho+\left(1-t\right)\sigma\right)\geqslant tF_{\alpha}\left(\rho\right)+\left(1-t\right)F_{\alpha}\left(\sigma\right) for any t[0,1]t\in\left[0,1\right]. Since for any concave function ff, the Tr[f(ρ)]\mathrm{Tr}\left[f\left(\rho\right)\right] is also concave RevModPhys.50.221 , f(x)=xαf\left(x\right)=x^{\alpha} with x[0,1]x\in\left[0,1\right] is concave for any 0α1/20\leqslant\alpha\leqslant 1/2. Let ρ=tρ+(1t)σ=jqj|ϕjϕj|{\rho}^{\prime}=t\rho+\left(1-t\right)\sigma=\sum_{j}q_{j}|\phi_{j}\rangle\langle\phi_{j}|, ρ=ipi|ψiψi|\rho=\sum_{i}p_{i}|\psi_{i}\rangle\langle\psi_{i}| and σ=krk|ξkξk|\sigma=\sum_{k}r_{k}|\xi_{k}\rangle\langle\xi_{k}| be corresponding eigendecompositions. We have

Fα(ρ)\displaystyle F_{\alpha}\left({\rho}^{\prime}\right) =j[tϕj|ρ|ϕj+(1t)ϕj|σ|ϕj]α1\displaystyle=\sum_{j}\left[t\braket{\phi_{j}}{\rho}{\phi_{j}}+\left(1-t\right)\braket{\phi_{j}}{\sigma}{\phi_{j}}\right]^{\alpha}-1
j{tϕj|ρ|ϕjα+(1t)ϕj|σ|ϕjα}1\displaystyle\geqslant\sum_{j}\left\{t\braket{\phi_{j}}{\rho}{\phi_{j}}^{\alpha}+\left(1-t\right)\braket{\phi_{j}}{\sigma}{\phi_{j}}^{\alpha}\right\}-1
=tj(ipi|ϕj|ψi|2)α\displaystyle=t\sum_{j}\left(\sum_{i}p_{i}|\braket{\phi_{j}}{\psi_{i}}|^{2}\right)^{\alpha}
+(1t)j(krk|ϕj|ξk|2)α1\displaystyle\ \ \ +\left(1-t\right)\sum_{j}\left(\sum_{k}r_{k}|\braket{\phi_{j}}{\xi_{k}}|^{2}\right)^{\alpha}-1
ti,j|ϕj|ψi|2piα+(1t)j,k|ϕj|ξk|2rkα1\displaystyle\geqslant t\sum_{i,j}|\braket{\phi_{j}}{\psi_{i}}|^{2}p_{i}^{\alpha}+\left(1-t\right)\sum_{j,k}|\braket{\phi_{j}}{\xi_{k}}|^{2}r_{k}^{\alpha}-1
=tipiα+(1t)krkα1\displaystyle=t\sum_{i}p_{i}^{\alpha}+\left(1-t\right)\sum_{k}r_{k}^{\alpha}-1
=tFα(ρ)+(1t)Fα(σ),\displaystyle=tF_{\alpha}\left(\rho\right)+\left(1-t\right)F_{\alpha}\left(\sigma\right), (29)

where the two inequalities follow from the concavity of ff. The equality holds if ρ\rho and σ\sigma are identical. \hfill\qed

Appendix B ηα\eta_{\alpha} for Isotropic states

Let 𝒯iso\mathcal{T}_{iso} be the (UU)\left(U\otimes U^{*}\right)-twirling operator defined by 𝒯iso(ρ)=𝑑U(UU)ρ(UU)\mathcal{T}_{iso}\left(\rho\right)=\int dU\left(U\otimes U^{\ast}\right)\rho\left(U\otimes U^{\ast}\right)^{\dagger}, where dUdU denotes the standard Haar measure on the group of all d×dd\times d unitary operations. Then the operator satisfies that 𝒯iso(ρ)=ρF(ρ)\mathcal{T}_{iso}\left(\rho\right)=\rho_{F\left(\rho\right)} with F(ρ)=Ψ|ρ|ΨF\left(\rho\right)=\braket{\Psi}{\rho}{\Psi}. One has 𝒯iso(ρF)=ρF\mathcal{T}_{iso}\left(\rho_{F}\right)=\rho_{F} PhysRevA.64.062307 ; PhysRevA.59.4206 ; PhysRevA.68.062304 . Applying 𝒯iso\mathcal{T}_{iso} to the pure state |ψ\ket{\psi} given in (1), |ψ=i=1rλiUAUB|ii\ket{\psi}=\sum_{i=1}^{r}\sqrt{\lambda_{i}}U_{A}\otimes U_{B}\ket{ii} with |ai=UA|i\ket{a_{i}}=U_{A}\ket{i} and |bi=UB|i\ket{b_{i}}=U_{B}\ket{i}, we have

𝒯iso(|ψψ|)=ρF(|ψψ|)=ρF(λ,V),\mathcal{T}_{iso}\left(|\psi\rangle\langle\psi|\right)=\rho_{F\left(|\psi\rangle\langle\psi|\right)}=\rho_{F\left(\vec{\lambda},V\right)}, (30)

where V=UATUBV=U_{A}^{T}U_{B} and

F(λ,V)=|Ψ|ψ|2=1d|i=1rλiVii|2F\left(\vec{\lambda},V\right)=|\braket{\Psi}{\psi}|^{2}=\frac{1}{d}\Big{|}\sum_{i=1}^{r}\sqrt{\lambda_{i}}V_{ii}\Big{|}^{2} (31)

with Vij=i|V|jV_{ij}=\braket{i}{V}{j}, and λ\vec{\lambda} is the Schmidt vector of (1). Then the function η\eta defined in (19) becomes

ηα(ρF)=min{λ,V}{Cα(λ):1d|i=1rλiVii|2=F}.\eta_{\alpha}\left(\rho_{F}\right)=\min_{\set{\vec{\lambda},V}}\left\{C_{\alpha}\left(\vec{\lambda}\right):\frac{1}{d}\Big{|}\sum_{i=1}^{r}\sqrt{\lambda_{i}}V_{ii}\Big{|}^{2}=F\right\}. (32)

It has been proved that the minimum above is attained for V=IV=I PhysRevA.68.062304 . Therefore, we have

ηα(ρF)=minλ{Cα(λ):1d|i=1rλi|2=F}.\displaystyle\eta_{\alpha}\left(\rho_{F}\right)=\min_{\vec{\lambda}}\left\{C_{\alpha}\left(\vec{\lambda}\right):\frac{1}{d}\Big{|}\sum_{i=1}^{r}\sqrt{\lambda_{i}}\Big{|}^{2}=F\right\}. (33)

For F(0,1d]F\in(0,\frac{1}{d}], one can always chose suitable UAU_{A} and UBU_{B} such that λ1=1\lambda_{1}=1, and hence ηα(ρF)=0\eta_{\alpha}\left(\rho_{F}\right)=0. For F(1d,1]F\in(\frac{1}{d},1], similar to PhysRevA.103.052423 ; PhysRevLett.85.2625 , by using the Lagrange multipliers PhysRevA.67.012307 one can minimize (33) subject to the constraints

i=1rλi=1,i=1rλi=Fd\displaystyle\sum_{i=1}^{r}\lambda_{i}=1,~{}~{}~{}\sum_{i=1}^{r}\sqrt{\lambda_{i}}=\sqrt{Fd} (34)

with Fd1Fd\geqslant 1. An extremum is attained when

(λi)2α1+Λ1λi+Λ2=0,\left(\sqrt{\lambda_{i}}\right)^{2\alpha-1}+\Lambda_{1}\sqrt{\lambda_{i}}+\Lambda_{2}=0, (35)

where Λ1\Lambda_{1} and Λ2\Lambda_{2} denote the Lagrange multipliers.

It is evident that f(λi)=(λi)2α1f\left(\sqrt{\lambda_{i}}\right)=\left(\sqrt{\lambda_{i}}\right)^{2\alpha-1} is a convex function of λi\sqrt{\lambda_{i}} for any 0α1/20\leqslant\alpha\leqslant 1/2. Since a convex function and a linear function cross at most two points, equation (35) has at most two possible nonzero solutions for λi\sqrt{\lambda_{i}}. Let γ\gamma and δ\delta be these two positive solutions with γ>δ\gamma>\delta. The Schmidt vector λ={λ1,λ2,,λr,0,,0}\vec{\lambda}=\left\{\lambda_{1},\lambda_{2},\dots,\lambda_{r},0,\cdots,0\right\} has the form,

λj={γ2,j=1,2,,n,δ2,j=n+1,,n+m,0,j=n+m+1,,d,\lambda_{j}=\begin{cases}\gamma^{2},&j=1,2,...,n,\\ \delta^{2},&j=n+1,...,n+m,\\ 0,&j=n+m+1,...,d,\end{cases} (36)

where r=n+mdr=n+m\leqslant d and n1n\geqslant 1. The minimization problem of (33) has been reduced to the following minimum problem,

ηα(ρF)=minn,mCαnm(F)\eta_{\alpha}\left(\rho_{F}\right)=\min_{n,m}C_{\alpha}^{nm}\left(F\right) (37)

with

Cαnm(F)=nγ2α+mδ2α1,C_{\alpha}^{nm}\left(F\right)=n\gamma^{2\alpha}+m\delta^{2\alpha}-1, (38)

subject to the constraints

nγ2+mδ2=1,nγ+mδ=Fd.\displaystyle n\gamma^{2}+m\delta^{2}=1,~{}~{}~{}n\gamma+m\delta=\sqrt{Fd}. (39)

By solving Eq. (39), we obtain

γnm±(F)=nFd±nm(n+mFd)n(n+m),\gamma_{nm}^{\pm}\left(F\right)=\frac{n\sqrt{Fd}\pm\sqrt{nm\left(n+m-Fd\right)}}{n\left(n+m\right)}, (40)
δnm±(F)=mFdnm(n+mFd)m(n+m).\delta_{nm}^{\pm}\left(F\right)=\frac{m\sqrt{Fd}\mp\sqrt{nm\left(n+m-Fd\right)}}{m\left(n+m\right)}. (41)

The relation γnm±(F)=δmn(F)\gamma_{nm}^{\pm}\left(F\right)=\delta_{mn}^{\mp}\left(F\right) suggests that we only need to consider the cases γnm:=γnm+(F)\gamma_{nm}:=\gamma_{nm}^{+}\left(F\right) and δnm=δnm+(F)\delta_{nm}=\delta_{nm}^{+}\left(F\right), which are real for Fdn+mFd\leqslant n+m. On the other hand, since δnm\delta_{nm} should be non-negative, we must have FdnFd\geqslant n. Therefore, we see that δnmFd/(n+m)γnm\delta_{nm}\leqslant\sqrt{Fd}/\left(n+m\right)\leqslant\gamma_{nm}, in consistent with the assumption γ>δ\gamma>\delta. Here, n1n\geqslant 1 as n=0n=0 is ill defined.

We seek is the minimum of Cαnm(F)C_{\alpha}^{nm}\left(F\right) over all possible nn and mm, by minimizing CαnmC_{\alpha}^{nm} on the parallelogram defined by 1nFd1\leqslant n\leqslant Fd and Fdn+mdFd\leqslant n+m\leqslant d. Note that the parallelogram collapses to a line when Fd=1Fd=1, i.e., the separable boundary. We have γnmδnm0\gamma_{nm}\geqslant\delta_{nm}\geqslant 0 in the parallelogram. Moreover, γnm=δnm\gamma_{nm}=\delta_{nm} if and only if n+m=Fdn+m=Fd; while δnm=0\delta_{nm}=0 if and only if n=Fdn=Fd.

When α=1/2\alpha=1/2, we see from Eqs. (38) and (39) that Eq. (22) holds without any optimization. When α=0\alpha=0, C0nm(F)=n+m1C_{0}^{nm}\left(F\right)=n+m-1 and Eq. (22) satisfied with the constraint conditions. From Eq.(39) the derivatives of γnm\gamma_{nm} and δnm\delta_{nm} with respect to nn and mm are given by,

γn\displaystyle\frac{\partial\gamma}{\partial n} =12n2γδγ2γδ,\displaystyle=\frac{1}{2n}\frac{2\gamma\delta-\gamma^{2}}{\gamma-\delta},
δn\displaystyle\frac{\partial\delta}{\partial n} =12mγ2γδ,\displaystyle=-\frac{1}{2m}\frac{\gamma^{2}}{\gamma-\delta},
δm\displaystyle\frac{\partial\delta}{\partial m} =12m2γδδ2γδ,\displaystyle=-\frac{1}{2m}\frac{2\gamma\delta-\delta^{2}}{\gamma-\delta},
γm\displaystyle\frac{\partial\gamma}{\partial m} =12nδ2γδ.\displaystyle=\frac{1}{2n}\frac{\delta^{2}}{\gamma-\delta}. (42)

Hence, using Eq. (38) we have the partial derivatives of Cαnm(F)C_{\alpha}^{nm}\left(F\right) with respect to nn and mm,

Cαnmn=(1α)γ2α+αγ2δγ2α2δ2α2γδ,\frac{\partial C_{\alpha}^{nm}}{\partial n}=\left(1-\alpha\right)\gamma^{2\alpha}+\alpha\gamma^{2}\delta\frac{\gamma^{2\alpha-2}-\delta^{2\alpha-2}}{\gamma-\delta}, (43)
Cαnmm=(1α)δ2α+αδ2γγ2α2δ2α2γδ.\frac{\partial C_{\alpha}^{nm}}{\partial m}=\left(1-\alpha\right)\delta^{2\alpha}+\alpha\delta^{2}\gamma\frac{\gamma^{2\alpha-2}-\delta^{2\alpha-2}}{\gamma-\delta}. (44)

By lengthy calculations, we have

Cαnmm=δ2α+1γδ{(1α)(γδ1)+α(γδ)2α1αγδ}.\frac{\partial C_{\alpha}^{nm}}{\partial m}=\frac{\delta^{2\alpha+1}}{\gamma-\delta}\left\{\left(1-\alpha\right)\left(\frac{\gamma}{\delta}-1\right)+\alpha\left(\frac{\gamma}{\delta}\right)^{2\alpha-1}-\alpha\frac{\gamma}{\delta}\right\}. (45)

Denote t=γδt=\frac{\gamma}{\delta}. One has t1t\geqslant 1. Let

g(t)=(1α)(t1)+αt2α1αt.g\left(t\right)=\left(1-\alpha\right)\left(t-1\right)+\alpha t^{2\alpha-1}-\alpha t. (46)

We have g(1)=0g\left(1\right)=0 and

gt=(12α)(1αt2α2).\frac{\partial g}{\partial t}=\left(1-2\alpha\right)\left(1-\alpha t^{2\alpha-2}\right). (47)

Set h(t)=1αt2α2h\left(t\right)=1-\alpha t^{2\alpha-2} with h(1)=1α>0h\left(1\right)=1-\alpha>0. We obtain

ht=α(2α2)t2α30.\frac{\partial h}{\partial t}=-\alpha\left(2\alpha-2\right)t^{2\alpha-3}\geqslant 0. (48)

From Eqs. (48) and (47), combining with Eq.(46) we have

Cαnmm0.\frac{\partial C_{\alpha}^{nm}}{\partial m}\geqslant 0. (49)

Now corresponding to moving perpendicularly to and parallel to the n+m=n+m=constant boundaries of the parallelogram, we make a parameter transformation, u=nmu=n-m and v=n+mv=n+m. The derivative of CαnmC_{\alpha}^{nm} with respect to uu is given by

Cαnmu=12(CαnCαm)\displaystyle\frac{\partial C_{\alpha}^{nm}}{\partial u}=\frac{1}{2}\left(\frac{\partial C_{\alpha}}{\partial n}-\frac{\partial C_{\alpha}}{\partial m}\right)
=12{γ2α1[(1α)γ+αδ]δ2α1[(1α)δ+αγ]}\displaystyle=\frac{1}{2}\left\{\gamma^{2\alpha-1}\left[\left(1-\alpha\right)\gamma+\alpha\delta\right]-\delta^{2\alpha-1}\left[\left(1-\alpha\right)\delta+\alpha\gamma\right]\right\}
=γ2α2{1α+αδγ(δγ)2α(1α+αγδ)}.\displaystyle=\frac{\gamma^{2\alpha}}{2}\left\{1-\alpha+\alpha\frac{\delta}{\gamma}-\left(\frac{\delta}{\gamma}\right)^{2\alpha}\left(1-\alpha+\alpha\frac{\gamma}{\delta}\right)\right\}.

Set x=δγx=\frac{\delta}{\gamma} with 0x10\leqslant x\leqslant 1. Let

f(x)=1α+αxx2α(1α+αx1)f\left(x\right)=1-\alpha+\alpha x-x^{2\alpha}\left(1-\alpha+\alpha x^{-1}\right) (50)

with f(1)=0f\left(1\right)=0. We have the derivative of f(x)f\left(x\right) respect to xx,

fx=α{1[2x2α1(1α+αx1)x2α2]}.\frac{\partial f}{\partial x}=\alpha\left\{1-\left[2x^{2\alpha-1}\left(1-\alpha+\alpha x^{-1}\right)-x^{2\alpha-2}\right]\right\}. (51)

Again let

k(x)=2x2α1(1α+αx1)x2α2k\left(x\right)=2x^{2\alpha-1}\left(1-\alpha+\alpha x^{-1}\right)-x^{2\alpha-2} (52)

with k(1)=1k\left(1\right)=1. Then

kx=x2α3l(x),\frac{\partial k}{\partial x}=x^{2\alpha-3}l\left(x\right), (53)

where

l(x)=2(2α1)(xαx)(1α)(4α2).l\left(x\right)=2\left(2\alpha-1\right)\left(x-\alpha x\right)-\left(1-\alpha\right)\left(4\alpha-2\right). (54)

We have l(1)=0l\left(1\right)=0 and

lx=2(2α1)(1α)0.\frac{\partial l}{\partial x}=2\left(2\alpha-1\right)\left(1-\alpha\right)\leqslant 0. (55)

From Eqs. (55) and (53), combining Eq. (51) we obtain f/x0\partial f/\partial x\geqslant 0. Then for any x[0,1]x\in[0,1] we have f(x)f(1)=0f\left(x\right)\leqslant f\left(1\right)=0. Therefore,

Cαnmu0.\frac{\partial C_{\alpha}^{nm}}{\partial u}\leqslant 0. (56)

From Eqs. (49) and (56), the minimum of Cαnm(F)C_{\alpha}^{nm}\left(F\right) is obtained when mm is the minimum and uu is the maximum. These results imply that the minimum of Cαnm(F)C_{\alpha}^{nm}\left(F\right) occurs at the vertex of n=Fdn=Fd and m=0m=0. Specifically, since γnm=δnm\gamma_{nm}=\delta_{nm} on the boundary n+m=Fdn+m=Fd where Eqs. (49) and (56) are both hold, we have n=n+m=Fd{n}^{\prime}=n+m=Fd and m=0{m}^{\prime}=0. In this way, we derive an analytical expression of the function ηα(ρF)\eta_{\alpha}\left(\rho_{F}\right) as follows,

ηα(ρF)=(Fd)1α1.\eta_{\alpha}\left(\rho_{F}\right)=\left(Fd\right)^{1-\alpha}-1. (57)

Appendix C ηα\eta_{\alpha} for Werner states

Let 𝒯wer(ρ)=𝑑U(UU)ρ(UU)\mathcal{T}_{wer}\left(\rho\right)=\int dU\left(U\otimes U\right)\rho\left(U^{\dagger}\otimes U^{\dagger}\right) be the (UU)\left(U\otimes U\right)-twirling transformations PhysRevA.64.062307 . Then the Werner states defined in (IV.2) satisfy that, in analogous to the isotropic states, 𝒯wer(ρ)=ρW(ρ)\mathcal{T}_{wer}\left(\rho\right)=\rho_{W\left(\rho\right)}, where W(ρ)=Tr(ρi<j|ΨijΨij|)W\left(\rho\right)=\mathrm{Tr}\left(\rho\sum_{i<j}|\Psi_{ij}^{-}\rangle\langle\Psi_{ij}^{-}|\right) and 𝒯wer(ρW)=ρW\mathcal{T}_{wer}\left(\rho_{W}\right)=\rho_{W} PhysRevA.59.4206 ; PhysRevA.68.062304 . Applying 𝒯wer\mathcal{T}_{wer} to the pure state |ψ\ket{\psi} defined in (1), |ψ=i=1rλiUAUB|ii\ket{\psi}=\sum_{i=1}^{r}\sqrt{\lambda_{i}}U_{A}\otimes U_{B}\ket{ii}, we have

𝒯wer(|ψψ|)=ρW(|ψψ|)=ρW(λ,Λ),\mathcal{T}_{wer}\left(|\psi\rangle\langle\psi|\right)=\rho_{W\left(|\psi\rangle\langle\psi|\right)}=\rho_{W\left(\vec{\lambda},\Lambda\right)}, (58)

where Λ=UAUB\Lambda=U_{A}^{\dagger}U_{B} and

W(λ,Λ)\displaystyle W\left(\vec{\lambda},\Lambda\right) =i<j|Ψij|ψ|2\displaystyle=\sum_{i<j}|\braket{\Psi_{ij}^{-}}{\psi}|^{2}
=12i<j|λiΛjiλjΛij|2\displaystyle=\frac{1}{2}\sum_{i<j}|\sqrt{\lambda_{i}}\Lambda_{ji}-\sqrt{\lambda_{j}}\Lambda_{ij}|^{2} (59)

with Λij=i|Λ|j\Lambda_{ij}=\braket{i}{\Lambda}{j}. Then the function η\eta defined in (19) becomes

ηα(ρW)=min{λ,Λ}{Cα(λ):W(λ,Λ)=W}.\eta_{\alpha}\left(\rho_{W}\right)=\min_{\set{\vec{\lambda},\Lambda}}\left\{C_{\alpha}\left(\vec{\lambda}\right):W\left(\vec{\lambda},\Lambda\right)=W\right\}. (60)

By W(λ,Λ)=WW\left(\vec{\lambda},\Lambda\right)=W we have

2W\displaystyle 2W =1i=1rλi|Λii|22i<jλiλjRe(ΛijΛji)\displaystyle=1-\sum_{i=1}^{r}\lambda_{i}|\Lambda_{ii}|^{2}-2\sum_{i<j}\sqrt{\lambda_{i}\lambda_{j}}\mathrm{Re}\left(\Lambda_{ij}\Lambda_{ji}^{\ast}\right)
1+2i<jλiλj|Re(ΛijΛji)|\displaystyle\leqslant 1+2\sum_{i<j}\sqrt{\lambda_{i}\lambda_{j}}|\mathrm{Re}\left(\Lambda_{ij}\Lambda_{ji}^{\ast}\right)|
1+2i<jλiλj\displaystyle\leqslant 1+2\sum_{i<j}\sqrt{\lambda_{i}\lambda_{j}}
=|i=1rλi|2,\displaystyle=|\sum_{i=1}^{r}\sqrt{\lambda_{i}}|^{2}, (61)

where Re(z)\mathrm{Re}\left(z\right) is the real part of zz.

Note that the equalities in (C) hold if only the two nozero components Λ01=1\Lambda_{01}=1 and Λ10=1\Lambda_{10}=-1, and λ=(λ1,λ2,0,,0)\vec{\lambda}=\left(\lambda_{1},\lambda_{2},0,\cdots,0\right), which give rise to the optimal minimum of (60) PhysRevA.64.062307 . Therefore, (60) becomes

ηα(ρW)=minλ{Cα(λ):|i=12λi|2=2W}.\eta_{\alpha}\left(\rho_{W}\right)=\min_{\vec{\lambda}}\left\{C_{\alpha}\left(\vec{\lambda}\right):|\sum_{i=1}^{2}\sqrt{\lambda_{i}}|^{2}=2W\right\}. (62)

For W(0,12]W\in(0,\frac{1}{2}], we can always chose suitable UAU_{A} and UBU_{B} to have that λ1=1\lambda_{1}=1, which results in ηα(ρW)=0\eta_{\alpha}\left(\rho_{W}\right)=0. For W>1/2W>1/2, one minimizes (62) subject to the constraints

i=12λi=1,i=12λi=2W\displaystyle\sum_{i=1}^{2}\lambda_{i}=1,~{}~{}~{}\sum_{i=1}^{2}\sqrt{\lambda_{i}}=\sqrt{2W} (63)

with W>1/2W>1/2. The rest of the calculation is the similar to the one in Appendix B. We only need to set d=2d=2 and F=WF=W. In this way, we can obtain

ηα(ρW)=(2W)1α1.\eta_{\alpha}\left(\rho_{W}\right)=\left(2W\right)^{1-\alpha}-1. (64)

References