Abstract

Quantum entanglement plays crucial roles in quantum information processing. Quantum entangled states have become the key ingredient in the rapidly expanding field of quantum information science. Although the nonclassical nature of entanglement has been recognized for many years, considerable efforts have been taken to understand and characterize its properties recently. In this review, we introduce some recent results in the theory of quantum entanglement. In particular separability criteria based on the Bloch representation, covariance matrix, normal form and entanglement witness, lower bounds, subadditivity property of concurrence and tangle, fully entangled fraction related to the optimal fidelity of quantum teleportation, and entanglement distillation will be discussed in detail.

1. Introduction

Entanglement is the characteristic trait of quantum mechanics, and it reflects the property that a quantum system can simultaneously appear in two or more different states [1]. This feature implies the existence of global states of composite system which cannot be written as a product of the states of individual subsystems. This phenomenon [2], now known as “quantum entanglement,” plays crucial roles in quantum information processing [3]. Quantum entangled states have become the key ingredient in the rapidly expanding field of quantum information science, with remarkable prospective applications such as quantum computation [3, 4], quantum teleportation [59], dense coding [10], quantum cryptographic schemes [1113], entanglement swapping [1418], and remote states preparation (RSP) [1924]. All such effects are based on entanglement and have been demonstrated in pioneering experiments.

It has become clear that entanglement is not only the subject of philosophical debates, but also a new quantum resource for tasks which cannot be performed by means of classical resources. Although considerable efforts have been taken to understand and characterize the properties of quantum entanglement recently, the physical character and mathematical structure of entangled states have not been satisfactorily understood yet [25, 26]. In this review we mainly introduce some recent results related to our researches on several basic questions in this subject.

(1) Separability of Quantum States
We first discuss the separability of a quantum states; namely, for a given quantum state, how we can know whether or not it is entangled.

For pure quantum states, there are many ways to verify the separability. For instance, for a bipartite pure quantum state the separability is easily determined in terms of its Schmidt numbers. For multipartite pure states, the generalized concurrence given in [27] can be used to judge if the state is separable or not. In addition separable states must satisfy all possible Bell inequalities [28].

For mixed states we still have no general criterion. The well-known PPT (partial positive transposition) criterion was proposed by Peres in 1996 [29]. It says that for any bipartite separable quantum state the density matrix must be positive under partial transposition. By using the method of positive maps Horodecki et al. [30] showed that the Peres' criterion is also sufficient for and bipartite systems. And for higher dimensional states, the PPT criterion is only necessary. Horodecki [31] has constructed some classes entangled states with positive partial transposes for and systems. States of this kind are said to be bound entangled (BE). Another powerful operational criterion is the realignment criterion [32, 33]. It demonstrates a remarkable ability to detect many bound entangled states and even genuinely tripartite entanglement [34]. Considerable efforts have been made in finding stronger variants and multipartite generalizations for this criterion [3539]. It was shown that PPT criterion and realignment criterion are equivalent to the permutations of the density matrix's indices [34]. Another important criterion for separability is the reduction criterion [40, 41]. This criterion is equivalent to the PPT criterion for composite systems. Although it is generally weaker than the PPT, the reduction criteria have tight relation to the distillation of quantum states.

There are also some other necessary criteria for separability. Nielsen and Kempe [42] presented a necessary criterion called majorization: the decreasing ordered vector of the eigenvalues for is majorized by that of or alone for a separable state. That is, if a state is separable, then , . Here denotes the decreasing ordered vector of the eigenvalues of . A -dimensional vector is majorized by , , if for and the equality holds for . Zeros are appended to the vectors such that their dimensions are equal to the one of .

In [31], another necessary criterion called range criterion was given. If a bipartite state acting on the space is separable, then there exists a family of product vectors such that (i) they span the range of ; (ii) the vector spans the range of , where denotes complex conjugation in the basis in which partial transposition was performed and is the partially transposed matrix of with respect to the subspace . In particular, any of the vectors belongs to the range of .

Recently, some elegant results for the separability problem have been derived. In [4345], a separability criteria based on the local uncertainty relations (LURs) was obtained. The authors show that, for any separable state , where or are arbitrary local orthogonal and normalized operators (LOOs) in . This criterion is strictly stronger than the realignment criterion. Thus more bound entangled quantum states can be recognized by the LUR criterion. The criterion is optimized in [46] by choosing the optimal LOOs. In [47] a criterion based on the correlation matrix of a state has been presented. The correlation matrix criterion is shown to be independent of PPT and realignment criterion [48], that is, there exist quantum states that can be recognized by correlation criterion while the PPT and realignment criterion fail. The covariance matrix of a quantum state is also used to study separability in [49]. It has been shown that the LUR criterion, including the optimized one, can be derived from the covariance matrix criterion [50].

(2) Measure of Quantum Entanglement
One of the most difficult and fundamental problems in entanglement theory is to quantify entanglement. The initial idea to quantify entanglement was connected with its usefulness in terms of communication [51]. A good entanglement measure has to fulfill some conditions [52]. For bipartite quantum systems, we have several good entanglement measures such as Entanglement of Formation (EOF), Concurrence, and Tangle ctc. For two-qubit systems it has been proved that EOF is a monotonically increasing function of the concurrence and an elegant formula for the concurrence was derived analytically by Wootters [53]. However with the increasing dimensions of the subsystems the computation of EOF and concurrence become formidably difficult. A few explicit analytic formulae for EOF and concurrence have been found only for some special symmetric states [5458].

The first analytic lower bound of concurrence for arbitrary dimensional bipartite quantum states was derived by Mintert et al. in [59]. By using the positive partial transposition (PPT) and realignment separability criterion, analytic lower bounds on EOF and concurrence for any dimensional mixed bipartite quantum states have been derived in [60, 61]. These bounds are exact for some special classes of states and can be used to detect many bound entangled states. In [62] another lower bound on EOF for bipartite states has been presented from a new separability criterion [63]. A lower bound of concurrence based on local uncertainty relations (LURs) criterion is derived in [64]. This bound is further optimized in [46]. The lower bound of concurrence for tripartite systems has been studied in [65]. In [66, 67] the authors presented lower bounds of concurrence for bipartite systems by considering the “two-qubit" entanglement of bipartite quantum states with arbitrary dimensions. It has been shown that this lower bound has a tight relationship with the distillability of bipartite quantum states. Tangle is also a good entanglement measure that has a close relation with concurrence, as it is defined by the square of the concurrence for a pure state. It is also meaningful to derive tight lower and upper bounds for tangle [68].

In [69] Mintert et al. proposed an experimental method to measure the concurrence directly by using joint measurements on two copies of a pure state. Then Walborn et al. presented an experimental determination of concurrence for two-qubit states [70, 71], where only one-setting measurement is needed, but two copies of the state have to be prepared in every measurement. In [72] another way of experimental determination of concurrence for two-qubit and multiqubit states has been presented, in which only one copy of the state is needed in every measurement. To determine the concurrence of the two-qubit state used in [70, 71], also one-setting measurement is needed, which avoids the preparation of the twin states or the imperfect copy of the unknown state, and the experimental difficulty is dramatically reduced.

(3) Fidelity of Quantum Teleportation and Distillation
Quantum teleportation, or entanglement-assisted teleportation, is a technique used to transfer information on a quantum level, usually from one particle (or series of particles) to another particle (or series of particles) in another location via quantum entanglement. It does not transport energy or matter, nor does it allow communication of information at super luminal (faster than light) speed.

In [57], Bennett et al. first presented a protocol to teleport an unknown qubit state by using a pair of maximally entangled pure qubit state. The protocol is generalized to transmit high-dimensional quantum states [8, 9]. The optimal fidelity of teleportation is shown to be determined by the fully entangled fraction of the entangled resource which is generally a mixed state. Nevertheless similar to the estimation of concurrence, the computation of the fully entangled fraction for a given mixed state is also very difficult.

The distillation protocol has been presented to get maximally entangled pure states from many entangled mixed states by means of local quantum operations and classical communication (LQCC) between the parties sharing the pairs of particles in this mixed state [7376]. Bennett et al. first derived a protocol to distill one maximally entangled pure Bell state from many copies of not maximally entangled quantum mixed states in [73] in 1996. The protocol is then generalized to distill any bipartite quantum state with higher dimension by M. Horodecki and P. Horodecki in 1999 [77]. It is proven that a quantum state can be always distilled if it violates the reduced matrix separability criterion [77].

This review mainly contains three parts. In Section 2 we investigate the separability of quantum states. We first introduce several important separability criteria. Then we discuss the criteria by using the Bloch representation of the density matrix of a quantum state. We also study the covariance matrix of a quantum density matrix and derive separability criterion for multipartite systems. We investigate the normal forms for multipartite quantum states at the end of this section and show that the normal form can be used to improve the power of these criteria. In Section 3 we mainly consider the entanglement measure concurrence. We investigate the lower and upper bounds of concurrence for both bipartite and multipartite systems. We also show that the concurrence and tangle of two entangled quantum states will be always larger than that of one, even if both of the two states are bound entangled (not distillable). In Section 4 we study the fully entangled fraction of an arbitrary bipartite quantum state. We derive precise formula of fully entangled fraction for two-qubit system. For bipartite system with higher dimension we obtain tight upper bounds which can not only be used to estimate the optimal teleportation fidelity but also help to improve the distillation protocol. We further investigate the evolution of the fully entangled fraction when one of the bipartite system undergoes a noisy channel. We give a summary and conclusion in the last section.

2. Separability Criteria and Normal Form

A multipartite pure quantum state is said to be fully separable if it can be written as

where and are reduced density matrices defined as , . This is equivalent to the condition where .

A multipartite quantum mixed state is said to be fully separable if it can be written as

where are the reduced density matrices with respect to the systems , respectively, , and . This is equivalent to the condition where are normalized pure states of systems , respectively, , and .

For pure states, the definition (2.1) itself is an operational separability criterion. In particular, for bipartite case, there are Schmidt decompositions.

Theorem 2.1 (see Schmidt decomposition in [78]). Suppose that is a pure state of a composite system, , then there exist orthonormal states for system and orthonormal states for system such that where are nonnegative real numbers satisfying , known as Schmidt coefficients.

and are called Schmidt bases with respect to and . The number of nonzero values is called Schmidt number, also known as Schmidt rank, which is invariant under unitary transformations on system or system . For a bipartite pure state , is separable if and only if the Schmidt number of is one.

For multipartite pure states, one has no such Schmidt decomposition. In [79] it has been verified that any pure three-qubit state can be uniquely written as

with normalization condition , where , . Equation (2.6) is called generalized Schmidt decomposition.

For mixed states it is generally very hard to verify whether a decomposition like (2.3) exists. For a given generic separable density matrix, it is also not easy to find the decomposition (2.3) in detail.

2.1. Separability Criteria for Mixed States

In this section we introduce several separability criteria and the relations among themselves. These criteria have also tight relations with lower bounds of entanglement measures and distillation that will be discussed in the next section.

2.1.1. Partial Positive Transpose Criterion

The positive partial transpose (PPT) criterion provided by Peres [29] says that if a bipartite state is separable, then the new matrix with matrix elements defined in some fixed product basis as is also a density matrix (i.e., it has nonnegative spectrum). The operation , called partial transpose, just corresponds to the transposition of the indices with respect to the second subsystem . It has an interpretation as a partial time reversal [80].

Afterwards Horodecki et al. showed that Peres' criterion is also sufficient for and bipartite systems [30]. This criterion is now called PPT or Peres-Horodecki (P-H) criterion. For high-dimensional states, the P-H criterion is only necessary. Horodecki has constructed some classes of families of entangled states with positive partial transposes for and systems [31]. States of this kind are said to be bound entangled (BE).

2.1.2. Reduced Density Matrix Criterion

Cerf et al. [81] and M. Horodecki and P. Horodecki [82], independently, introduced a map ), which gives rise to a simple necessary condition for separability in arbitrary dimensions, called the reduction criterion. If is separable, then where , . This criterion is simply equivalent to the P-H criterion for composite systems. It is also sufficient for and systems. In higher dimensions the reduction criterion is weaker than the P-H criterion.

2.1.3. Realignment Criterion

There is yet another class of criteria based on linear contractions on product states. They stem from the new criterion discovered in [33, 83] called computable cross-norm (CCN) criterion or matrix realignment criterion which is operational and independent on PPT test [29]. If a state is separable, then the realigned matrix with elements has trace norm not greater than one: Quite remarkably, the realignment criterion can detect some PPT entangled (bound entangled) states [33, 83] and can be used for construction of some nondecomposable maps. It also provides nice lower bound for concurrence [61].

2.1.4. Criteria Based on Bloch Representations

Any Hermitian operator on an -dimensional Hilbert space can be expressed according to the generators of the special unitary group [84]. The generators of can be introduced according to the transition-projection operators , where , , are the orthonormal eigenstates of a linear Hermitian operator on . Set where and . We get a set of operators which satisfies the relations and thus generate the [85].

Any Hermitian operator in can be represented in terms of these generators of as

where is a unit matrix and and is called Bloch vector. The set of all the Bloch vectors that constitute a density operator is known as the Bloch vector space .

A matrix of the form (2.13) is of unit trace and Hermitian, but it might not be positive. To guarantee the positivity restrictions must be imposed on the Bloch vector. It is shown that is a subset of the ball of radius , which is the minimum ball containing it, and that the ball of radius is included in [86], that is,

Let the dimensions of systems , and be , and , respectively. Any tripartite quantum states can be written as

where , are the generators of and ; , and are operators of .

Theorem 2.2. Let , and , . For a tripartite quantum state with representation (2.15), one has [87]

Proof. Since , and , , we have that and are positive Hermitian operators. Let . Then and . The partial trace of over (and ) should be also positive. Hence

Formula (2.16) is valid for any tripartite state. By setting in (2.16), one can get a result for bipartite systems.

Corollary 2.3. Let , which can be generally written as , then, for any with , .

A separable tripartite state can be written as From (2.13) it can also be represented as

where and are real vectors on the Bloch sphere satisfying and .

Comparing (2.15) with (2.19), we have

For any real matrix and real matrix satisfying and , we define a new matrix

where is a transformation acting on an matrix by

Using , we define a new operator :

where , and .

Theorem 2.4. If is separable, then [87] .

Proof. From (2.20) and (2.23) we get A straightforward calculation gives rise to As and , we get Therefore is still a density operator, that is, .

Theorem 2.4 gives a necessary separability criterion for general tripartite systems. The result can be also applied to bipartite systems. Let , . For any real matrix satisfying and any state , we define where .

Corollary 2.5. For , if there exists an with such that , then must be entangled.

For systems, the above corollary is reduced to the results in [88]. As an example we consider the istropic states If we choose to be , we get that is entangled for .

For tripartite case, we take the following mixed state as an example: where . Taking , we have that is entangled for .

In fact the criterion for systems [88] is equivalent to the PPT criterion [89]. Similarly Theorem 2.4 is also equivalent to the PPT criterion for systems.

2.1.5. Covariance Matrix Criterion

In this subsection we study the separability problem by using the covariance matrix approach. We first give a brief review of covariance matrix criterion proposed in [49]. Let and be -dimensional complex vector spaces and a bipartite quantum state in . Let (resp., ) be observables on (resp., ) such that they form an orthonormal normalized basis of the observable space, satisfying (resp., ). Consider the total set . It can be proven that [44]

The covariance matrix is defined with entries

which has a block structure [49]

where , and . Such covariance matrix has a concavity property: for a mixed density matrix with and , one has .

For a bipartite product state , in (2.32) is zero. Generally if is separable, then there exist states on on and such that

where , .

For a separable bipartite state, it has been shown that [49]

Criterion (2.34) depends on the choice of the orthonormal normalized basis of the observables. In fact the term has an upper bound which is invariant under unitary transformation and can be attained by choosing proper local orthonormal observable basis, where stands for the Ky Fan norm of , , with denoting the transpose and conjugation. It has been shown in [46] that if is separable, then

From the covariance matrix approach, we can also get an alternative criterion. From (2.32) and (2.33) we have that if is separable, then Hence all the minor submatrices of must be positive. Namely, one has that is, . Summing over all , and using (2.30), we get

That is,

where stands for the Euclid norm of , that is, .

Formulae (2.35) and (2.39) are independent and could be complement. When (2.39) can recognize the entanglement but (2.35) cannot. When (2.35) can recognize the entanglement while (2.39) cannot.

The separability criteria based on covariance matrix approach can be generalized to multipartite systems. We first consider the tripartite case . Take observables on , respectively, on , respectively, on . Set . The covariance matrix defined by (2.31) has then the following block structure:

where , , , , , and .

Theorem 2.6. If is fully separable, then [90]

Proof. For a tripartite product state , and in (2.42) are zero. If is fully separable, then there exist states in , in , and in , and such that , where , , and , that is, Thus all the minor submatrices of must be positive. Selecting one with two rows and columns from the first two block rows and columns of , we have that is, . Summing over all , and using (2.30), we get which proves (2.43). Equations (2.44) and (2.45) can be similarly proved.
From (2.50) we also have . Therefore Note that . By using that for any matrix and any unitary [91], we have .
Let be the singular value decomposition of . Make a transformation of the orthonormal normalized basis of the local orthonormal observable space and . In the new basis we have Then (2.52) becomes which proves (2.46). Equations (2.47) and (2.48) can be similarly treated.

We consider now the case that is bipartite separable.

Theorem 2.7. If is a bipartite separable state with respect to the bipartite partition of the sub-systems and (resp., and ; resp., and ), then (2.43), (2.44) and (2.46), (2.47) (resp., (2.44), (2.45) and (2.47), (2.48); resp., (2.43), (2.45) and (2.46), (2.48)) must hold [90].

Proof. We prove the case that is bipartite separable with respect to the system and systems partition. The other cases can be similarly treated. In this case the matrices and in the covariance matrix (2.42) are zero. takes the form . Define , . has a form where and , . By using the concavity of covariance matrix we have Accounting to the method used in proving Theorem , we get (2.43), (2.44), and (2.46), (2.47).

From Theorems 2.6 and 2.7 we have the following corollary.

Corollary 2.8. If two of the inequalities (2.43), (2.44), and (2.45) (or (2.46), (2.47), and (2.48)) are violated, then the state must be fully entangled.

The result of Theorem 2.6 can be generalized to general multipartite case . Define , where , () are the normalized generators of satisfying and acting on the th system , . Denote as the set of all . Then the covariance matrix of can be written as

where and for .

For a product state , , , in (2.57) are zero matrices. Define Then for a fully separable multipartite state , one has from which we have the following separability criterion for multipartite systems.

Theorem 2.9. If a state is fully separable, then the following inequalities must be fulfilled for any [90].

2.2. Normal Form of Quantum States

In this subsection we show that the correlation matrix (CM) criterion can be improved from the normal form obtained under filtering transformations. Based on CM criterion entanglement witness in terms of local orthogonal observables (LOOs) [92] for both bipartite and multipartite systems can be also constructed.

For bipartite case, with , , and is mapped to the following form under local filtering transformations [93]:

where are arbitrary invertible matrices. This transformation is also known as stochastic local operations assisted by classical communication (SLOCC). By the definition it is obvious that filtering transformation will preserve the separability of a quantum state.

It has been shown that under local filtering operations one can transform a strictly positive into a normal form [94]:

where and and are some traceless orthogonal observables. The matrices and can be obtained by minimizing the function where and . In fact, one can choose , and , where . Then by iterations one can get the optimal and . In particular, there is a matlab code available in [95].

For bipartite separable states , the CM separability criterion [96] says that

where is an matrix with , stands for the trace norm of , s are the generators of and have been chosen to be normalized, and .

As the filtering transformation does not change the separability of a state, one can study the separability of instead of . Under the normal form (2.62) the criterion (2.64) becomes

In [44] a separability criterion based on local uncertainty relation (LUR) has been obtained. It says that, for any separable state ,

where s are LOOs such as the normalized generators of and for . The criterion is shown to be strictly stronger than the realignment criterion [61]. Under the normal form (2.62) criterion (2.66) becomes that is,

As holds for any and , from (2.65) and (2.68) it is obvious that the CM criterion recognizes entanglement better when the normal form is taken into account.

We now consider multipartite systems. Let be a strictly positive density matrix in and . can be generally expressed in terms of the generators [97] as

where with appears at the th position and

The generalized CM criterion says that if in (2.69) is fully separable, then

for . The KF norm is defined by where is a kind of matrix unfolding of .

The criterion (2.71) can be improved by investigating the normal form of (2.69).

Theorem 2.10. By filtering transformations of the form where , followed by normalization, any strictly positive state can be transformed into a normal form [98]:

Proof. Let be the sets of density matrices of the subsystems. The cartesian product consisting of all product density matrices with normalization , , is a compact set of matrices on the full Hilbert space . For the given density matrix we define the following function of : The function is well defined on the interior of where . As is assumed to be strictly positive, we have . Since is compact, we have with a lower bound depending on .
It follows that on the boundary of where at least one of the s satisfies . It follows further that has a positive minimum on the interior of with the minimum value attained for at least one product density matrix with , . Any positive density matrix with can be factorized in terms of Hermitian matrices as where . Denote that , so that . Set and define
We see that, when , has a minimum and
Since is stationary under infinitesimal variations about the minimum, it follows that for all infinitesimal variations subjected to the constraint , which is equivalent to , , using for a given matrix . Thus, can be represented by the generators, . It follows that for any and . Hence the terms proportional to in (2.69) disappear.

Corollary 2.11. The normal form of a product state in must be proportional to the identity.

Proof. Let be such a state. From (2.74), we get that Therefore for a product state we have

As an example for separability of multipartite states in terms of their normal forms (2.74), we consider the PPT entangled edge state [79] mixed with noises Select , and . Using the criterion in [97] we get that is entangled for . But after transforming to its normal form (2.74), the criterion can detect entanglement for .

Here we indicate that the filtering transformation does not change the PPT property. Let be PPT, that is, and . Let be the normal form of . From (2.61) we have For any vector , we have where .   can be proved similarly. This property is also valid for multipartite case. Hence a bound entangled state will be bound entangled under filtering transformations.

2.3. Entanglement Witness Based on Correlation Matrix Criterion

Entanglement witness (EW) is another way to describe separability. Based on CM criterion we can further construct entanglement witness in terms of LOOs. EW [92] is an observable of the composite system such that (i) nonnegative expectation values in all separable states and (ii) at least one negative eigenvalue (can recognize at least one entangled state). Consider bipartite systems in with .

Theorem 2.12. For any properly selected LOOs and , is an EW [98], where and

Proof. Let be a separable state, where are normalized generators of with , . Any other LOOs fulfilling (2.88) can be obtained from these s through orthogonal transformations , where and are orthogonal matrices. We have where we have used for any unitary in the first inequality and the CM criterion in the second inequality.
Now let be a state in which violates the CM criterion. Denote as the singular values of . By singular value decomposition, one has , where is a diagonal matrix with . Now choose LOOs to be , for and . We obtain where the CM criterion has been used in the last step.

As the CM criterion can be generalized to multipartite form [97], we can also define entanglement witness for multipartite system in . Set . Choose LOOs for with and define

where . One can prove that (2.91) is an EW candidate for multipartite states. First we assume that . Note that, for any , there must exist an elementary transformation such that . Then for an N-partite separable state we have for any , where we have taken into account that is orthogonal and for any unitary at the first inequality. The second inequality is due to the generalized CM criterion.

By choosing proper LOOs, it is also easy to show that has negative eigenvalues. For example, for three-qubit case, taking the normalized Pauli matrices as LOOs, one finds a negative eigenvalue of , .

3. Concurrence and Tangle

In this section, we focus on two important measures: concurrence and tangle (see [99, 100]). An elegant formula for concurrence of two-qubit states is derived analytically by Wootters [53, 101]. This quantity has recently been shown to play an essential role in describing quantum phase transition in various interacting quantum many-body systems [102, 103] and may affect macroscopic properties of solids significantly [104, 105]. Furthermore, concurrence also provides an estimation [106, 107] for the entanglement of formation (EOF) [76], which quantifies the required minimally physical resources to prepare a quantum state.

Let (resp., ) be an -(resp., -) dimensional complex vector space with , (resp., ), as an orthonormal basis. A general pure state on is of the form

where satisfy the normalization .

The concurrence of (3.1) is defined by [27, 108, 109] where . The definition is extended to general mixed states by the convex roof

For two-qubits systems, the concurrence of is given by

where , is the complex conjugate of , and is the Pauli matrix,

For a mixed two-qubit quantum state , the entanglement of formation has a simple relation with the concurrence [53, 101] where , where the s are the eigenvalues, in decreasing order, of the Hermitian matrix and .

Another entanglement measure called tangle is defined by

for a pure state . For mixed state , the definition is given by

For multipartite state , , the concurrence of is defined by [110, 111]

where labels all different reduced density matrices.

Up to constant factor (3.9) can be also expressed in another way. Let denote a -dimensional vector space with basis , . An -partite pure state in is generally of the form

Let and (resp., and ) be subsets of the subindices of , associated to the same sub Hilbert spaces but with different summing indices. (or ) and (or ) span the whole space of the given subindix of . The generalized concurrence of is then given by [27]

where and stands for the summation over all possible combinations of the indices of and .

For a mixed multipartite quantum state, in , the corresponding concurrence is given by the convex roof:

3.1. Lower and Upper Bounds of Concurrence

Calculations of the concurrence for general mixed states are extremely difficult. However, one can try to find the lower and the upper bounds to estimate the exact values of the concurrence [46, 61, 64, 65].

3.1.1. Lower Bound of Concurrence from Covariance Matrix Criterion

In [61] a lower bound of has been obtained as

where and stand for partial transpose with respect to subsystem and the realignment, respectively. This bound is further improved based on local uncertainty relations [64]

where and are any set of local orthonormal observables, .

Bound (3.14) again depends on the choice of the local orthonormal observables. This bound can be optimized, in the sense that a local orthonormal observable-independent up bound of the right-hand side of (3.14) can be obtained.

Theorem 3.1. Let be a bipartite state in . Then satisfies [90]

Proof. The other orthonormal normalized basis of the local orthonormal observable space can be obtained from and by unitary transformations and : and . Select and so that is the singular value decomposition of . Then the new observables can be written as , . We have Substituting above relation to (3.14), one gets (3.15).

Bound (3.15) does not depend on the choice of local orthonormal observables. It can be easily applied and realized by direct measurements in experiments. It is in accord with the result in [46] where optimization of entanglement witness based on local uncertainty relation has been taken into account. As an example, let us consider the bound entangled state [76]

where is the identity matrix, , , , , and . We simply choose the local orthonormal observables to be the normalized generators of . Formula (3.13) gives . Formula (3.14) gives [64], while formula (3.15) yields a better lower bound .

If we mix the bound entangled state (3.17) with , that is,

then it is easily seen that (3.15) gives a better lower bound of concurrence than formula (3.13) (Figure 1).

3.1.2. Lower Bound of Concurrence from “Two-Qubit” Decomposition

In [67] the authors derived an analytical lower bound of concurrence for arbitrary bipartite quantum states by decomposing the joint Hilbert space into many dimensional subspaces, which does not involve any optimization procedure and gives an effective evaluation of entanglement together with an operational sufficient condition for the distill ability of any bipartite quantum states.

(1) Lower Bound of Concurrence for Bipartite States
The lower bound of concurrence for bipartite states has been obtained in [67]. For a bipartite quantum state in , the concurrence satisfies where with being the square roots of the four nonzero eigenvalues, in decreasing order, of the non-Hermitian matrix with , while and are the generators of .

The lower bound in (3.19) in fact characterizes all two qubits’ entanglement in a high dimensional bipartite state. One can directly verify that there are at most nonzero elements in each matrix . These elements constitute a matrix , where is the Pauli matrix, the matrix is a submatrix of the original :

, and , with subindices and associated with the first space, and with the second space. The two-qubit submatrix is not normalized but positive semidefinite. are just the concurrences of these states (3.20).

The bound provides a much clearer structure of entanglement, which not only yields an effective separability criterion and an easy evaluation of entanglement, but also helps one to classify mixed-state entanglement.

(2) Lower Bound of Concurrence for Multipartite States
We first consider tripartite case. A general pure state on is of the form with or equivalently where are the reduced density matrices of .

Define , , and , where and of (resp., , resp., ) stand for the subindices of associated with the subspaces , and (resp., , and 2; resp., , and ). Let denote the generators of group associated to the subsystems . Then for a tripartite pure state (3.21), one has where , , and .

Theorem 3.2. For an arbitrary mixed state in , the concurrence satisfies [112] where is a lower bound of where are the square roots of the four nonzero eigenvalues, in decreasing order, of the non-Hermitian matrix with . and are defined in a similar way to .

Proof. Set , , , and . We have, from Minkowski inequality,
Noting that for nonnegative real variables , , and given that , , and , by using Lagrange multipliers, one obtains that the following inequality holds: Therefore we have
The values of , , and can be calculated by using the similar procedure in [53]. Here we compute the value of in detail. The values of and can be obtained analogously.
Let and be eigenvalues and eigenvectors of , respectively. Any decomposition of can be obtained from a unitary matrix , . Therefore one has , where the matrix is defined by . Namely, , which has an analytical expression [53], that , where are the square roots of the eigenvalues of the positive Hermitian matrix , or equivalently the non-Hermitian matrix , in decreasing order. Here as the matrix has rows and columns that are identically zero, the matrix has a rank not greater than 4, that is, for . From (3.29) we have (3.25).

Theorem 3.2 can be directly generalized to arbitrary multipartite case.

Theorem 3.3. For an arbitrary -partite state , the concurrence defined in (4.1) satisfies [112], where is the lower bound of , stands for the summation over all possible combinations of the indices of , , and , , are the square roots of the four nonzero eigenvalues, in decreasing order, of the non-Hermitian matrix where .

Lower Bound and Separability
An N-partite quantum state is fully separable if and only if there exist with , and pure states such that

It is easily verified that, for a fully separable multipartite state , . Thus indicates that there must be some kinds of entanglement inside the quantum state, which shows that the lower bound can be used to recognize entanglement.

As an example, we consider a tripartite quantum state [79] , where is the identity matrix, and is the tripartite W state. Select an entanglement witness operator to be , where is to be the tripartite GHZ-state. From the condition , the entanglement of is detected for in [79]. In [97] the authors have obtained the generalized correlation matrix criterion which says that if an N-qubit quantum state is fully separable, then the inequality must hold, where , is a kind of matrix unfold of defined by , and stands for the Pauli matrix. Now using the generalized correlation, matrix criterion the entanglement of is detected for . From Theorem 3.2, we have the lower bound for . Therefore the bound (3.71) detects entanglement better than these two criteria in this case. If we replace with GHZ state in , then the criterion in [97] detects the entanglement of for , while detects, again better, the entanglement for .

Nevertheless for PPT states , we have , which can be seen in the following way. A density matrix is called PPT if the partial transposition of over any subsystem(s) is still positive. Let denote the partial transposition with respect to the th subsystem. Assume that there is a PPT state with . Then at least one term in (3.25), say , is not zero. Define . By using the PPT property of , we have

Noting that both and are projectors to two-dimensional subsystems, can be considered as a density matrix, while a PPT density matrix must be a separable state, which contradicts with .

Relation between Lower Bounds of Bi- and Tripartite Concurrence
is basically different from as characterizes also genuine tripartite entanglement that can not be described by bipartite decompositions. Nevertheless, there are interesting relations between them.

Theorem 3.4. For any pure tripartite state (3.21), the following inequality holds [112]: where is the lower bound of bipartite concurrence (3.19), is the lower bound of tripartite concurrence (3.25), and , , , and .

Proof. Since for , , and , we have where we have used the similar analysis in [67, 113] to obtain the equalities , , and . The last equality is due to the fact that is a pure state.

In fact, the bipartite entanglement inside a tripartite state is useful for distilling maximally entangled states. Assume that there are two of the qualities larger than zero; say and . According to [67], one can distill two maximal entangled states and which belong to and , respectively. In terms of the result in [114], one can use them to produce a GHZ state.

3.1.3. Estimation of Multipartite Entanglement

For a pure -partite quantum state , , , the concurrence of bipartite decomposition between subsystems and is defined by

where is the reduced density matrix of by tracing over subsystems . On the other hand, the concurrence of is defined by (3.9).

For a mixed multipartite quantum state , the corresponding concurrences of (3.35) and (3.9) are then given by the convex roof

and (3.12). We now investigate the relation between these two kinds of concurrences.

Lemma 3.5. For a bipartite density matrix , one has where are the reduced density matrices of .

Proof. Let be the spectral decomposition, where . Then , . Therefore

This lemma can be also derived in another way [46, 115].

Theorem 3.6. For a multipartite quantum state with , the following inequality holds [116]: where the maximum is taken over all kinds of bipartite concurrence.

Proof. Without loss of generality, we suppose that the maximal bipartite concurrence is attained between subsystems and .
For a pure multipartite state , . From (3.37) we have that is, .
Let attain the minimal decomposition of the multipartite concurrence. One has

Corollary 3.7. For a tripartite quantum state , the following inequality holds: where the maximum is taken over all kinds of bipartite concurrence.

In [46, 64], from the separability criteria related to local uncertainty relation, covariance matrix, and correlation matrix, the following lower bounds for bipartite concurrence are obtained:

where the entries of the matrix , , , stand for the normalized generator of , that is, and . It is shown that the lower bounds (3.43) and (3.44) are independent of (3.13).

Now we consider a multipartite quantum state as a bipartite state belonging to with the dimensions of the subsystems and being and , respectively. By using Corollary 3.7, (3.13), (3.43), and (3.44), one has the following lower bound.

Theorem 3.8. For any N-partite quantum state [116], where ’s are all possible bipartite decompositions of , and

In [46, 106, 107, 117], it is shown that the upper and lower bounds of multipartite concurrence satisfy

In fact one can obtain a more effective upper bound for multipartite concurrence. Let , where ’s are the orthogonal pure states and . We have

The right side of (3.49) gives a new upper bound of . Since the upper bound obtained in (3.49) is better than that in (3.48).

3.1.4. Bounds of Concurrence and Tangle

In [68], a lower bound for tangle defined in (3.8) has been derived as

where denotes the Frobenius or Hilbert-Schmidt norm. Experimentally measurable lower and upper bounds for concurrence have been also given by Mintert et al. in [106, 107] and Zhang et al. in [46]:

Since the convexity of , we have that always holds. For two-qubit quantum systems, tangle is always equal to the square of concurrence [58, 113], as a decomposition achieving the minimum in (3.3) has the property that . For higher dimensional systems we do not have similar relations. Thus it is meaningful to derive valid upper bound for tangle and lower bound for concurrence.

Theorem 3.9. For any quantum state , one has [118] where is the reduced matrix of , and is the correlation matrix of defined in (3.44).

Proof. We assume that for convenience. By the definition of , we have that for a pure state . Let be the optimal decomposition such that We get
Note that, for pure state [68], Using the inequality for any , we get Now let be the optimal decomposition such that We get which ends the proof.

The upper bound (3.53), together with the lower bounds (3.54), (3.43), (3.44), (3.51), and (3.52), can allow for estimations of entanglement for arbitrary quantum states. Moreover, since the upper bound is exactly the value of tangle for pure states, the upper bound can be a good estimation when the state is very weakly mixed.

3.2. Concurrence and Tangle of Two Entangled States Are Strictly Larger Than Those of One

In this subsection we show that although bound entangled states cannot be distilled, the concurrence and tangle of two entangled states will be always strictly larger than those of one, even if the two entangled states are both bound entangled.

Let and be two quantum states shared by subsystems and . We use to denote the state of the whole system.

Lemma 3.10. For pure states and , the inequalities always hold, and “” in the two inequalities hold if and only if at least one of is separable.

Proof. Without loss of generality we assume that . First note that Let and be the spectral decomposition of and , with and , respectively. By using (3.61) one obtains that while
Now using the definition of concurrence and the normalization conditions of and , one immediately gets If one of is separable, say , then the rank of must be one, which means that there is only one item in the spectral decomposition in . Using the normalization condition of , we obtain . Then inequality (3.64) becomes an equality.
On the other hand, if both and are entangled (not separable), then there must be at least two items in the decomposition of their reduced density matrices and , which means that is strictly larger than .
The inequality (3.60) also holds because, for pure quantum state , .

From the lemma, we have, for mixed states the following.

Theorem 3.11. For any quantum states and , the inequalities always hold, and the “’’ in the two inequalities hold if and only if at least one of is separable, that is, if both and are entangled (even if bound entangled), then and always hold [118].

Proof. We still assume that for convenience. Let and be the optimal decomposition such that By using the inequality obtained in Lemma 3.10, we have Case 1. Now let one of be separable, say , with ensemble representation , where , and is the density matrix of separable pure state. Suppose that is the optimal decomposition of such that . Using Lemma 3.10, we have The inequalities (3.66) and (3.67) show that if is separable, then .Case 2. If both and are inseparable, that is, there is at least one pure state in the ensemble decomposition of (and , resp.), using Lemma 3.10, then we have
The inequality for tangle can be proved in a similar way.

Remark 3.12. In [119] it is shown that any entangled state can enhance the teleportation power of another state . This holds even if the state is bound entangled. But if is bound entangled, then the corresponding must be free entangled (distillable). By Theorem 3.11, we can see that even if two entangled quantum states and are bound entangled, their concurrence and tangle are strictly larger than those of one state.

3.3. Subadditivity of Concurrence and Tangle

We now give a proof of the subadditivity of concurrence and tangle, which illustrates that concurrence and tangle may be proper entanglement measurements.

Theorem 3.13. Let and be quantum states in , the one has [118]

Proof. We first prove that the theorem holds for pure states, that is, for and in : Assume that and are the spectral decomposition of the reduced matrices and . One has
Now we prove that (3.69) holds for any mixed-quantum states and . Let and be the optimal decomposition such that and . We have
The inequality for can be derived in a similar way.

4. Fidelity of Teleportation and Distillation of Entanglement

Quantum teleportation is an important subject in quantum information processing. In terms of a classical communication channel and a quantum resource (a nonlocal entangled state like an EPR pair of particles), the teleportation protocol gives ways to transmit an unknown quantum state from a sender traditionally named “Alice" to a receiver “Bob" who are spatially separated. These teleportation processes can be viewed as quantum channels. The nature of a quantum channel is determined by the particular protocol and the state used as a teleportation resource. The standard teleportation protocol proposed by Bennett et al. in 1993 uses Bell measurements and Pauli rotations. When the maximally entangled pure state is used as the quantum resource, it provides an ideal noiseless quantum channel . However in realistic situation, instead of the pure maximally entangled states, Alice and Bob usually share a mixed entangled state due to the decoherence. Teleportation using mixed state as an entangled resource is, in general, equivalent to having a noisy quantum channel. An explicit expression for the output state of the quantum channel associated with the standard teleportation protocol with an arbitrary mixed-state resource has been obtained [120, 121].

It turns out that, by local quantum operations (including collective actions over all members of pairs in each lab) and classical communication (LOCC) between Alice and Bob, it is possible to obtain a number of pairs in nearly maximally entangled state from many pairs of nonmaximally entangled states. Such a procedure proposed in [7377] is called distillation. In [73] the authors give operational protocol to distill an entangled two-qubit state whose single fraction , defined by , is larger than . The protocol is then generalized in [77] to distill any -dimensional bipartite entangled quantum states with . It is shown that a quantum state violating the reduction criterion can always be distilled. For such states, if their single fraction of entanglement is greater than , one can distill these states directly by using the generalized distillation protocol, otherwise a proper filtering operation has to be used at first to transform to another state so that .

4.1. Fidelity of Quantum Teleportation

Let be a -dimensional complex vector space with computational basis , . The fully entangled fraction (FEF) of a density matrix is defined by

under all unitary transformations , where is the maximally entangled state and is the corresponding identity matrix.

In [8, 9], the authors give an optimal teleportation protocol by using a mixed entangled quantum state. The optimal teleportation fidelity is given by which solely depends on the FEF of the entangled resource state .

In fact the fully entangled fraction is tightly related to many quantum information processing such as dense coding [10], teleportation [57], entanglement swapping [1418], and quantum cryptography (Bell inequalities) [1113]. As the optimal fidelity of teleportation is given by FEF [8, 9], experimentally measurement of FEF can be also used to determine the entanglement of the nonlocal source used in teleportation. Thus an analytic formula for FEF is of great importance. In [122] an elegant formula of FEF for two-qubit system is derived analytically by using the method of Lagrange multipliers. For high-dimensional quantum states the analytical computation of FEF remains formidable and less results have been known. In the following we give an estimation on the values of FEF by giving some upper bounds of FEF.

Let , , be the generators of the algebra. A bipartite state can be expressed as where , , and . Let denote the correlation matrix with entries .

Theorem 4.1. For any , the fully entangled fraction satisfies [123] where stands for the transpose of and is the Ky Fan norm of .

Proof. First, we note that where . By definition (4.1), one obtains
Since is a traceless Hermitian operator, it can be expanded according to the generators as Entries define a real matrix . From the completeness relation of generators one can show that is an orthonormal matrix. Using (4.7), we have

For the case , we can get an exact result from (4.4).

Corollary 4.2. For two-qubit system, one has that is, the upper bound derived in Theorem 4.1 is exactly the FEF.

Proof. We have shown in (4.7) that, given an arbitrary unitary , one can always obtain an orthonormal matrix . Now we show that in two-qubit case, for any orthonormal matrix , there always exits unitary matrix such that (4.7) holds.
For any vector with unit norm, define an operator where ’s are Pauli matrices. Given an orthonormal matrix , one obtains a new operator .
and are both Hermitian traceless matrices. Their eigenvalues are given by the norms of the vectors and , respectively. As the norms are invariant under orthonormal transformations , they have the same eigenvalues: . Thus there must be a unitary matrix such that . Hence the inequality in the proof of Theorem 4.1 becomes an equality. The upper bound (4.4) then becomes exact at this situation, which is in accord with the result in [122].

Remark 4.3. The upper bound of FEF (4.4) and the FEF (4.10) depend on the correlation matrices and . They can be calculated directly according to a given set of generators , . As an example, for , if we choose then we have Nevertheless the FEF and its upper bound do not depend on the choice of the generators.

The usefulness of the bound depends on detailed states. In the following we give two new upper bounds, which is different from Theorem 4.1. These bounds work for different states.

Let and be matrices such that , , with . We can introduce linear-independent matrices , which satisfy

One can also check that satisfy the condition of bases of the unitary operators in the sense of [124], that is,

where is the identity matrix. form a complete basis of matrices, namely, for any matrix , can be expressed as

From , we can introduce the generalized Bell states

where are all maximally entangled states and form a complete orthogonal normalized basis of .

Theorem 4.4. For any quantum state , the fully entangled fraction defined in (4.1) fulfills the following inequality: where s are the eigenvalues of the real part of matrix , is a matrix with entries , and are the maximally entangled basis states defined in (4.16) [125].

Proof. From (4.15), any unitary matrix can be represented by where . Define Then the unitary matrix can be rewritten as . The necessary condition for the unitary property of implies that . Thus we have where is defined in the theorem. One can deduce that from the hermiticity of .
Taking into account the constraint with an undetermined Lagrange multiplier , we have Accounting to (4.20) we have the eigenvalue equation
Inserting (4.22) into (4.19) results in where is the corresponding eigenvalues of the real part of the matrix .

Example 4.5. Horodecki gives a very interesting bound entangled state in [31] as One can easily compare the upper bound obtained in (4.17) and that in (4.4). From Figure 2 we see that, for , the upper bound in (4.17) is larger than that in (4.4). But for the upper bound in (4.17) is always lower than that in (4.4), which means that the upper bound (4.17) is tighter than (4.4).

In fact, we can drive another upper bound for FEF which will be very tight for weakly mixed-quantum states.

Theorem 4.6. For any bipartite quantum state , the following inequality holds [125]: where is the reduced matrix of .

Proof. Note that in [77] the authors have obtained the FEF for pure state as where is the reduced matrix of .
For mixed state , we have Let be the real and nonnegative eigenvalues of the matrix . Recall that for any function subjected to the constraints with being real and nonnegative, the inequality holds, from which it follows that which ends the proof.

4.2. Fully Entangled Fraction and Concurrence

The upper bound of FEF has also interesting relations to the entanglement measure concurrence. As shown in [122], the concurrence of a two-qubit quantum state has some kinds of relation with the optimal teleportation fidelity. For quantum state with high dimension, we have the similar relation between them too.

Theorem 4.7. For any bipartite quantum state , one has [118]

Proof. In [126], the authors show that, for any pure state , the following inequality holds: where denotes the set of dimensional maximally entangled states.
Let be the optimal decomposition such that . We have which ends the proof.

Inequality (4.29) has demonstrated the relation between the lower bound of concurrence and the fully entangled fraction (thus the optimal teleportation fidelity), that is, the fully entangled fraction of a quantum state is limited by its concurrence.

We now consider tripartite case. Let be a state of three-qubit systems denoted by , , and . We study the upper bound of the FEF, , between qubits and , and its relations to the concurrence under bipartite partition and . For convenience we normalize to be

Let denote the concurrence between subsystems and .

Theorem 4.8. For any triqubit state , satisfies [123]

Proof. We first consider the case that is pure, . By using the Schmidt decomposition between qubits , and , can be written as for some orthonormalized bases , of subsystems , , respectively. The reduced density matrix has the form where is a diagonal matrix with diagonal elements , is a unitary matrix, and denotes the conjugation of .
The FEF of the two-qubit state can be calculated by using formula (4.10) or the one in [122]. Let be the matrix constituted by the four Bell bases. The FEF of can be written as where stands for the maximal eigenvalues of the matrix .
For pure state (4.34) in bipartite partition and , we have From (4.32), (4.37), and (4.38) we get
We now prove that the above inequality (4.39) also holds for mixed state . Let be the optimal decomposition of such that . We have where and .

From Theorem 4.8 we see that the FEF of qubits and are bounded by the concurrence between qubits , , and qubit . The upper bound of FEF for decreases when the entanglement between qubits , and increases. As an example, we consider the generalized state defined by , . The reduced density matrix is given by The FEF of is given by while the concurrence of has the form . We see that (4.33) always holds. In particular for and , the inequality (4.33) is saturated (see Figure 3).

4.3. Improvement of Entanglement Distillation Protocol

The upper bound can give rise to not only an estimation of the fidelity in quantum information processing such as teleportation, but also an interesting application in entanglement distillation of quantum states. In [77] a generalized distillation protocol has been presented. It is shown that a quantum state violating the reduction criterion can always be distilled. For such states if their single fraction of entanglement is greater than , then one can distill these states directly by using the generalized distillation protocol. If the FEF (the largest value of single fraction of entanglement under local unitary transformations) is less than or equal to , then a proper filtering operation has to be used at first to transform to another state so that . For , one can compute FEF analytically according to the corollary. For our upper bound (4.4) can supply a necessary condition in the distillation.

Theorem 4.9. For an entangled state violating the reduction criterion, if the upper bound (4.4) is less than or equal to , then the filtering operation has to be applied before using the generalized distillation protocol [123].

As an example, we consider a state

where . It is direct to verify that violates the reduction criterion for , as has a negative eigenvalue . Therefore the state is distillable. From Figure 4, we see that for and , the fidelity is already greater than ; thus the generalized distillation protocol can be applied without the filtering operation. However for , even the upper bound of the fully entangled fraction is less than or equal to ; hence the filtering operation has to be applied first, before using the generalized distillation protocol.

Moreover, the lower bounds of concurrence can be also used to study the distillability of quantum states. Based on the positive partial transpose (PPT) criterion, a necessary and sufficient condition for the distillability was proposed in [127], which is not operational in general. An alternative distillability criterion based on the bound in (3.19) can be obtained to improve the operationality.

Theorem 4.10. A bipartite quantum state is distillable if and only if for some number [67].

Proof. It was shown in [127] that a density matrix is distillable if and only if there are some projectors , that map high-dimensional spaces to two-dimensional ones and some number such that the state is entangled [127]. Thus if , then there exists one submatrix of matrix , similar to (3.20), which has nonzero and is entangled in a space; hence is distillable.

Corollary 4.11. (a) The lower bound is a sufficient condition for the distillability of any bipartite state .
(b) The lower bound is a necessary condition for separability of any bipartite state .

Remark 4.12. Corollary 4.11 directly follows from Theorem 4.10 and this case is referred to as one distillable [128]. The problem of whether non-PPT (NPPT) nondistillable states exist is studied numerically in [128, 129]. By using Theorem 4.10, although it seems impossible to solve the problem completely, it is easy to judge the distillability of a state under condition that is one distillable.

The lower bound , PPT criterion, separability, and distillability for any bipartite quantum state have the following relations: if , then is entangled. If is separable, then it is PPT. If , then is distillable. If is distillable, then it is NPPT. From the last two propositions it follows that if is PPT, then , that is, if , then is NPPT.

Theorem 4.13. For any pure tripartite state in arbitrary dimensional spaces, bound satisfies [67] where , , and .

Proof. Since , one can derive the inequality where . Note that and , where . By using the similar analysis in [113], one has that the right-hand side of (4.45) is equal to . Taking into account that for a pure state, one obtains inequality (4.44).
Generally for any pure multipartite quantum state , one has the following monogamy inequality:

5. Summary and Conclusion

We have introduced some recent results on three aspects in quantum information theory. The first one is the separability of quantum states. New criteria to detect more entanglements have been discussed. The normal forms of quantum states have been also studied, which helps in investigating the separability of quantum states. Moreover, since many kinds of quantum states can be transformed into the same normal forms, quantum states can be classified in terms of the normal forms. For the well-known entanglement measure concurrence, we have discussed the tight lower and upper bounds. It turns out that, although one cannot distill a singlet from many pairs of bound entangled states, the concurrence and tangle of two entangled quantum states are always larger than those of one, even if both of two entangled quantum states are bound entangled. Related to the optimal teleportation fidelity, upper bounds for the fully entangled fraction have been studied, which can be used to improve the distillation protocol. Interesting relations between fully entangled fraction and concurrence have been also introduced. All these related problems in the theory of quantum entanglement have not been completely solved yet. Many problems remain open concerning the physical properties and mathematical structures of quantum entanglement, and the applications of entangled states in information processing.