superfidelity and trace distance

4
Physics Letters A 376 (2012) 1025–1028 Contents lists available at SciVerse ScienceDirect Physics Letters A www.elsevier.com/locate/pla Superfidelity and trace distance Yuan Li College of Mathematics and Information Science, Shaanxi Normal University, Xi’an 710062, PR China article info abstract Article history: Received 7 October 2011 Received in revised form 27 January 2012 Accepted 27 January 2012 Available online 1 February 2012 Communicated by P.R. Holland Keywords: Superfidelity Trace distance Quantum operation Trace distance and superfidelity play an important role in quantum information theory. The aim of this Letter is to consider an inequality involving trace distance and superfidelity in infinite dimension and give a necessary and sufficient condition for equality of this inequality. In addition, some related results involving trace distance and superfidelity are obtained. © 2012 Elsevier B.V. All rights reserved. 1. Introduction In quantum information theory, an essential task is to distin- guish two quantum states. One of the important tools is the trace metric, another tool is quantum fidelity [1–9]. Both are widely used by the quantum information community and have found ap- plications in a number of problems such as quantum cryptography [10] and quantum phase transitions [11]. The trace distance is also related to the von Neumann entropy and relative entropy [3]. The mathematical description of a quantum mechanical system is based on a complex Hilbert space H, with bounded quantum observables being represented by bounded self-adjoint operators A acting on H. Let T (H) be the set of all trace class operators on H, P (H) the set of all orthogonal projections and S (H) the set of all density operators, i.e., the trace class positive operators on H of unit trace. The elements ρ S (H) represent the states of a quantum system, and the probability that a quantum effect A occurs in the state ρ is given by P ρ ( A) = tr(ρ A). If A is a quantum observable, then we define | A|= ( A 2 ) 1 2 , A + = | A|+ A 2 and A = | A|− A 2 , where ( A 2 ) 1 2 is the unique positive square root of A 2 . For ρ , σ S (H), the trace distance and the fidelity are defined respectively by D(ρ , σ ) = 1 2 tr |ρ σ | and F (ρ , σ ) = ( tr ρ 1 2 σρ 1 2 ) 2 . In [12], a new fidelity, called superfidelity, was defined as E-mail address: [email protected]. G (ρ , σ ) = tr ρσ + 1 tr ρ 2 1 tr σ 2 . It was shown that when ρ and σ are qubit states, the superfidelity G(ρ , σ ) coincides with fidelity F (ρ , σ ). The most interesting fea- ture of superfidelity is that it gives an upper bound for fidelity [12], that is F (ρ , σ ) G (ρ , σ ), for ρ , σ S (H), where H is a finite-dimensional space. Very re- cently, the properties of superfidelity have been established and studied by many authors [6,12–16]. In [16], the superfidelity has been applied to quantum circuits to enable a measure of distance between two quantum channels, and in [15] a new probability measure was introduced in terms of superfidelity. In particular, if H is a finite-dimensional space, the inequality 1 G (ρ , σ ) D(ρ , σ ) (1) was obtained in [14]. This inequality was also conjectured in [13] and verified numerically for small dimensions. The purpose of this Letter is to consider inequality (1) in an infinite-dimensional space and give a necessary and sufficient con- dition for saturation of inequality (1). In addition, we consider the perturbational upper and lower bound of GA (ρ ), ρ ) over all states of a quantum system, where φ A is a trace preserving, unital quantum operation. 2. Saturation of inequality (1) The following result is well known. For the reader’s conven- ience, we give a simple proof. 0375-9601/$ – see front matter © 2012 Elsevier B.V. All rights reserved. doi:10.1016/j.physleta.2012.01.042

Upload: yuan-li

Post on 11-Sep-2016

213 views

Category:

Documents


0 download

TRANSCRIPT

Physics Letters A 376 (2012) 1025–1028

Contents lists available at SciVerse ScienceDirect

Physics Letters A

www.elsevier.com/locate/pla

Superfidelity and trace distance

Yuan Li

College of Mathematics and Information Science, Shaanxi Normal University, Xi’an 710062, PR China

a r t i c l e i n f o a b s t r a c t

Article history:Received 7 October 2011Received in revised form 27 January 2012Accepted 27 January 2012Available online 1 February 2012Communicated by P.R. Holland

Keywords:SuperfidelityTrace distanceQuantum operation

Trace distance and superfidelity play an important role in quantum information theory. The aim of thisLetter is to consider an inequality involving trace distance and superfidelity in infinite dimension andgive a necessary and sufficient condition for equality of this inequality. In addition, some related resultsinvolving trace distance and superfidelity are obtained.

© 2012 Elsevier B.V. All rights reserved.

1. Introduction

In quantum information theory, an essential task is to distin-guish two quantum states. One of the important tools is the tracemetric, another tool is quantum fidelity [1–9]. Both are widelyused by the quantum information community and have found ap-plications in a number of problems such as quantum cryptography[10] and quantum phase transitions [11]. The trace distance is alsorelated to the von Neumann entropy and relative entropy [3].

The mathematical description of a quantum mechanical systemis based on a complex Hilbert space H, with bounded quantumobservables being represented by bounded self-adjoint operatorsA acting on H. Let T (H) be the set of all trace class operatorson H, P(H) the set of all orthogonal projections and S(H) theset of all density operators, i.e., the trace class positive operatorson H of unit trace. The elements ρ ∈ S(H) represent the statesof a quantum system, and the probability that a quantum effect Aoccurs in the state ρ is given by Pρ(A) = tr(ρ A). If A is a quantumobservable, then we define

|A| = (A2) 1

2 , A+ = |A| + A

2and A− = |A| − A

2,

where (A2)12 is the unique positive square root of A2. For ρ,σ ∈

S(H), the trace distance and the fidelity are defined respectivelyby

D(ρ,σ ) = 1

2tr |ρ − σ | and F (ρ,σ ) = (

tr

√ρ

12 σρ

12)2

.

In [12], a new fidelity, called superfidelity, was defined as

E-mail address: [email protected].

0375-9601/$ – see front matter © 2012 Elsevier B.V. All rights reserved.doi:10.1016/j.physleta.2012.01.042

G(ρ,σ ) = trρσ +√

1 − trρ2√

1 − trσ 2.

It was shown that when ρ and σ are qubit states, the superfidelityG(ρ,σ ) coincides with fidelity F (ρ,σ ). The most interesting fea-ture of superfidelity is that it gives an upper bound for fidelity[12], that is

F (ρ,σ ) � G(ρ,σ ),

for ρ,σ ∈ S(H), where H is a finite-dimensional space. Very re-cently, the properties of superfidelity have been established andstudied by many authors [6,12–16]. In [16], the superfidelity hasbeen applied to quantum circuits to enable a measure of distancebetween two quantum channels, and in [15] a new probabilitymeasure was introduced in terms of superfidelity. In particular, ifH is a finite-dimensional space, the inequality

1 − G(ρ,σ ) � D(ρ,σ ) (1)

was obtained in [14]. This inequality was also conjectured in [13]and verified numerically for small dimensions.

The purpose of this Letter is to consider inequality (1) in aninfinite-dimensional space and give a necessary and sufficient con-dition for saturation of inequality (1). In addition, we considerthe perturbational upper and lower bound of G(φA(ρ),ρ) over allstates of a quantum system, where φA is a trace preserving, unitalquantum operation.

2. Saturation of inequality (1)

The following result is well known. For the reader’s conven-ience, we give a simple proof.

1026 Y. Li / Physics Letters A 376 (2012) 1025–1028

Lemma 2.1. Let ρ,σ ∈ S(H). Then 0 � D(ρ,σ ) � 1, and D(ρ,σ ) = 1if and only if ρσ = 0.

Proof. As in the proof for the finite-dimensional case, we getD(ρ,σ ) = max{tr(P (ρ − σ)): P ∈ P (H)}. Thus 0 � D(ρ,σ ) � 1.

If D(ρ,σ ) = 1, then there exists a projection P1 ∈ P (H) suchthat tr(P1ρ) − tr(P1σ) = 1. It follows from 0 � tr(P1ρ) � 1 and0 � tr(P1σ) � 1 that tr(P1ρ) = 1 and tr(P1σ) = 0, which yieldsP1σ = 0 and ρ P1 = ρ . Thus ρσ = ρ P1σ = 0. The inverse implica-tion is evident. �

The following is the main result of this section.

Theorem 2.2. Let ρ,σ ∈ S(H). Then 1 − D(ρ,σ ) � G(ρ,σ ), withequality if and only if one of the following hypotheses holds:

(i) ρ = σ ;(ii) ρσ = σρ and ρ or σ is a pure state.

Proof. Let P+ denote the projection onto the subspaceR[(ρ − σ)+] and P− = I − P+ , where R[(ρ − σ)+] is the clo-sure of the range (ρ − σ)+ . Then P+(ρ − σ) = (ρ − σ)+ , andP−(ρ − σ) = −(ρ − σ)− . By a similar calculation as in [14], weobtain that the inequalities (21)–(24) and Eq. (25) in [14] alsohold in an infinite-dimensional system (for details, see [14, The-orem 1]). Thus the inequality 1 − D(ρ,σ ) � G(ρ,σ ) holds in aninfinite-dimensional system as well.

In the following, we show a necessary and sufficient conditionfor the equality 1 − D(ρ,σ ) = G(ρ,σ ).

The “if” part: If ρ = σ , then it is clear that 1 − D(ρ,σ ) =G(ρ,σ ). In the following, we assume that ρσ = σρ and ρ is apure state. Suppose ρ = |φ〉〈φ| and the subspace H1 is spanned bythe state |φ〉. Then

ρ =(

1 00 0

): H1 ⊕H⊥

1 → H1 ⊕H⊥1 . (2)

Let

σ =(

a1 σ1σ ∗

1 σ2

): H1 ⊕H⊥

1 → H1 ⊕H⊥1 , (3)

where σ1 is an operator from H⊥1 into H1 and σ2 is an operator

from H⊥1 into H⊥

1 . It is obvious that σ ∈ S(H) implies that a1 � 0,σ2 � 0, and a1 + tr(σ2) = 1. By a direct calculation, we obtain fromρσ = σρ that σ1 = 0 and σ ∗

1 = 0. Thus

|ρ − σ | =(

1 − a1 00 σ2

): H1 ⊕H⊥

1 → H1 ⊕H⊥1 ,

which means D(ρ,σ ) = 1 − a1. By Eqs. (2) and (3), we havetr(ρσ ) = a1, so 1 − D(ρ,σ ) = G(ρ,σ ).

The “only if” part: As in the proof of [14, Eq. (25)] for the finite-dimensional case, we have

D(ρ,σ ) = 1

2(tr P+ρ − tr P+σ + tr P−σ − tr P−ρ)

= 1 − tr P+σ − tr P−ρ. (4)

By the assumption that 1 − D(ρ,σ ) = G(ρ,σ ), we have (for a de-tailed calculation see [14, Eq. (26)])√

1 − trρ2√

1 − trσ 2 = tr P+(I − ρ)σ + tr P−ρ(I − σ). (5)

It is clear that tr[P+(I − ρ)(ρ − σ)] = tr[(I − ρ)(ρ − σ)+]� 0 im-plies

tr[

P+(I − ρ)ρ]� tr

[P+(I − ρ)σ

]. (6)

Similarly, we obtain that

tr[

P−ρ(I − ρ)]� tr

[P−ρ(I − σ)

]. (7)

Furthermore, as in the finite-dimensional case [14, inequalities(22), (23)], we also get

1 − trρ2 � tr[

P+(I − ρ)σ] + tr

[P−ρ(I − σ)

](8)

and

1 − trσ 2 � tr[

P+(I − ρ)σ] + tr

[P−ρ(I − σ)

]. (9)

Case 1. Suppose that 1 − trσ 2 > 0 and 1 − trρ2 > 0. Then combin-ing (5), (8) and (9), we have

1 − trρ2 = tr[

P+(I − ρ)σ] + tr

[P−ρ(I − σ)

] = 1 − trσ 2, (10)

which means G(ρ,σ ) = trρσ +1− trρ2, so 1− D(ρ,σ ) = G(ρ,σ )

implies that tr |ρ −σ |+2 trρσ −2 trρ2 = 0. Thus tr[|ρ −σ |− (ρ −σ)2] = 0.

On the other hand, as −I � ρ − σ � I , we have |ρ − σ | � I ,so (ρ − σ)2 = |ρ − σ |2 � |ρ − σ |. Then |ρ − σ | = |ρ − σ |2, whichmeans that |ρ − σ | is a projection. Hence (ρ − σ)+ and (ρ − σ)−are projections. By Lemma 2.1, we know that

1 � D(ρ,σ ) = tr(ρ − σ)+ = tr(ρ − σ)−;hence we conclude that

tr(ρ − σ)+ = tr(ρ − σ)− = 0 or

tr(ρ − σ)+ = tr(ρ − σ)− = 1.

Thus D(ρ,σ ) = 0 or D(ρ,σ ) = 1, so Lemma 2.1 implies that ρ = σor ρσ = 0.

Moreover, if ρσ = 0, then it is clear that 1 − D(ρ,σ ) = G(ρ,σ )

implies√

1 − trρ2√

1 − trσ 2 = 0, so 1 − trρ2 = 0 or 1 − trσ 2 = 0.This contradiction with the assumption that 1 − trσ 2 > 0 and 1 −trρ2 > 0 shows that ρ = σ .

Case 2. Suppose that 1 − trσ 2 = 0 or 1 − trρ2 = 0. Without lossof generality, we assume that 1 − trρ2 = 0; then trρ2 = 1 impliesthat ρ is a pure state. Furthermore, combining (5) and (8), we get

1 − trρ2 = tr[

P+(I − ρ)σ] + tr

[P−ρ(I − σ)

].

Combining (6) and (7), we have

tr[

P+(I − ρ)ρ] = tr

[P+(I − ρ)σ

]and

tr[

P−ρ(I − ρ)] = tr

[P−ρ(I − σ)

],

which yields

(I − ρ)(ρ − σ)+ = 0

and

ρ(ρ − σ)− = 0.

Then

(ρ − σ)+ − ρ(ρ − σ) = (I − ρ)(ρ − σ)+ + ρ(ρ − σ)− = 0,

which means

ρ(ρ − σ) = (ρ − σ)+ = (ρ − σ)ρ,

so ρσ = σρ . �

Y. Li / Physics Letters A 376 (2012) 1025–1028 1027

Remark. Theorem 2.2 shows that the equation 1 − D(ρ,σ ) =G(ρ,σ ) implies the equation F (ρ,σ ) = G(ρ,σ ). However, the re-verse implication does not hold. A simple example comes from thetwo-dimensional Hilbert space, in which for any two states the su-perfidelity is equal to quantum fidelity [12].

For superfidelity, we can define the following function:

C(ρ,σ ) := √1 − G(ρ,σ ), for ρ,σ ∈ S(H).

We recall that the Hilbert–Schmidt norm of ρ ∈ S(H) is defined

as ‖ρ‖HS := (trρ2)12 . It is shown in [13] that C(ρ,σ ) is a genuine

metric. The following two results show that the topologies inducedby C(ρ,σ ) and DHS(ρ,σ ) are the same on S(H) if H is a finite-dimensional space.

Proposition 2.3. Let ρ,σ ∈ S(H). Then

(i) ‖ρ − σ‖HS �√

2C(ρ,σ ),(ii) ‖ρ − σ‖HS = √

2C(ρ,σ ) if and only if trρ2 = trσ 2 .

Proof. (i)

‖ρ − σ‖HS �√

2√

1 − G(ρ,σ )

⇐⇒ 2 − 2√

1 − trρ2√

1 − trσ 2 � trρ2 + trσ 2

⇐⇒ (1 − trρ2) + (

1 − trσ 2)� 2√

1 − trρ2√

1 − trσ 2.

(ii) The conclusion is clear, because the equality between geo-metric and arithmetic means is satisfied only in the case of equalfactors in the proof of (i). �Proposition 2.4. Let H be a finite-dimensional space. Then the topol-ogy induced by C(ρ,σ ) and DHS(ρ,σ ) is the same on S(H), whereDHS(ρ,σ ) := ‖ρ − σ‖HS is called the Hilbert–Schmidt distance.

Proof. It is sufficient to show that

C(ρn,ρ) −→ 0 ⇐⇒ DHS(ρn,ρ) −→ 0, n −→ ∞.

Given Proposition 2.3, it remains to show that if DHS(ρn,ρ) −→ 0,then C(ρn,ρ) −→ 0, n −→ ∞. Assume that DHS(ρn,ρ) −→ 0, so√

tr[(ρn − ρ)2] −→ 0, n −→ ∞. Using the Cauchy–Schwarz in-equality, we get tr(|ρn − ρ|) � √

tr[(ρn − ρ)2]√m, where m is thedimension of H, which yields tr(|ρn −ρ|) −→ 0, n −→ ∞. Thus bythe formula C(ρ,σ ) �

√D(ρ,σ ), we know that

C(ρn,ρ) �√

1

2tr

(|ρn − ρ|) −→ 0, for n −→ ∞. �3. Characterization of inf{G(φA(ρ),ρ): ρ ∈ S(H)}

In this section, let B(H) be the set of all bounded linear op-erators on a finite-dimensional Hilbert space H. Recall that theoperator-sum representation is a key result of the quantum oper-ation formalism. Namely, according to a well-known result [17],a general completely positive quantum operation is a bounded lin-ear operator defined on B(H) which has the form

φA(B) =n∑

i=1

Ai B A∗i , (11)

where Ai ∈ B(H) (1 � i � n) are arbitrary bounded operators. If∑ni=1 A∗

i Ai = I , then φA is said to be trace preserving. If φA(I) = I(equivalently

∑ni=1 Ai A∗

i = I), then the quantum operation φA iscalled unital.

Quantum measurements that have more than two values aredescribed by quantum effect valued measures, that is, for i =1,2, . . . ,n, 0 � Ai � I satisfy

∑ni=1 Ai = I . In this case, the Lüders

operation (see [18]) is defined as a bounded linear map ΛA :T (H) → T (H) that satisfies

ΛA(S) =n∑

i=1

A12i S A

12i .

In the following, we consider the perturbational upper and lowerbound of G(φA(ρ),ρ) over all states of a quantum system.

Theorem 3.1. Let φA be a trace preserving, unital quantum operation.Then

(i) sup{G(φA(ρ),ρ): ρ ∈ S(H)} = 1,(ii) inf{G(φA(ρ),ρ): ρ ∈ S(H)} = 0 if and only if there exists

|φ〉 ∈H, such that tr(Ai |φ〉〈φ|) = 0, for i = 1,2, . . . ,n.

Proof. Since H is a finite-dimensional Hilbert space, we know thatS(H) is a compact set in the norm topology. It is clear that thefunction G(φA(ρ),ρ) of ρ is a continuous function in the normtopology. Thus sup and inf can be replaced by max and min, re-spectively.

(i) By Schauder’s fixed point theorem [19, p. 150], we get thatthere exists a state σ such that φA(σ ) = σ , as S(H) is a com-pact convex set. Then it is easy to see that sup{G(φA(ρ),ρ):ρ ∈ S(H)} = 1.

(ii) If inf{G(φA(ρ),ρ): ρ ∈ S(H)} = 0, then there exists ρ0 ∈S(H), such that

φA(ρ0)ρ0 = 0 and√

1 − trρ20

√1 − tr

[φA(ρ0)2

] = 0.

Case 1. If 1 − trρ20 = 0, then ρ0 is a pure state. Let ρ0 = |φ〉〈φ|. It

follows from φA(ρ0)ρ0 = 0 that

n∑i=1

(〈φ|A∗i |φ〉Ai

)|φ〉〈φ| =n∑

i=1

(Ai|φ〉〈φ|A∗

i

)|φ〉〈φ| = 0,

so

n∑i=1

(∣∣〈φ|Ai|φ〉∣∣2) = 0,

which means 〈φ|Ai|φ〉 = 0, for i = 1,2, . . . ,n. Hence tr(Ai |φ〉〈φ|) =0, for i = 1,2, . . . ,n.

Case 2. If 1 − tr[φA(ρ0)2] = 0, then φA(ρ0) is a pure state. Let

φA(ρ0) = |ψ〉〈ψ |. It is easy to see that

1 = 〈ψ |n∑

i=1

Aiρ0 A∗i |ψ〉 =

n∑i=1

〈ψ |Aiρ0 A∗i |ψ〉

=n∑

i=1

∥∥ρ 12

0 A∗i |ψ〉∥∥2 �

(n∑

i=1

∥∥A∗i |ψ〉∥∥2

)∥∥ρ 12

0

∥∥2.

As

n∑i=1

∥∥A∗i |ψ〉∥∥2 = 〈ψ |

n∑i=1

Ai A∗i |ψ〉 = 1,

we have ‖ρ0‖ = ‖ρ12

0 ‖2 � 1, then trρ0 = 1 implies that ρ0 is a purestate. The remaining part of the proof is the same as in Case 1. �

1028 Y. Li / Physics Letters A 376 (2012) 1025–1028

Corollary 3.2. Let ΛA be a Lüders operation. Then inf{G(ΛA(ρ),ρ):ρ ∈ S(H)} > 0.

Proof. Assume the opposite, i.e., that

inf{

G(ΛA(ρ),ρ

): ρ ∈ S(H)

} = 0.

By Theorem 3.1, we know that there exists |φ〉 ∈ H, such thattr(Ai |φ〉〈φ|) = 0, for i = 1,2, . . . ,n, so Ai |φ〉〈φ| = 0. ThenA2

i |φ〉〈φ| = 0, for i = 1,2, . . . ,n, which yields

|φ〉〈φ| =n∑

i=1

A2i |φ〉〈φ| = 0.

This is a contradiction. �4. Conclusion

As an extension of quantum fidelity, a new fidelity, called su-perfidelity is defined in [12]. In this Letter, we consider an im-portant inequality involving superfidelity and trace metric in aninfinite-dimensional space and give a necessary and sufficient con-dition for saturation of this inequality. Our main result showsthat this inequality is saturated only in an extreme condition.On the other hand, in the finite-dimensional case we also obtaina topology structure involving superfidelity and the upper andlower perturbation bound between a state and its transforma-tion by a quantum operation. All results in this Letter showthat superfidelity can be used to infer the distinguishability ofstates.

Acknowledgements

The author would like to express his thanks to the anony-mous referee for the valuable comments. Thanks also are extendedto P. Busch for inspiring discussions. This work was supportedby NSF of China (Nos. 11001159, 11171197) and also supportedby the Fundamental Research Funds for the Central Universities(GK200902049, GK201002012, GK201001002).

References

[1] A. Uhlmann, Rep. Math. Phys. 9 (1976) 273.[2] M.A. Nielsen, I.L. Chuang, Quantum Computation and Quantum Information,

Cambridge Univ. Press, Cambridge, 2000.[3] I. Bengtsson, K. Zyczkowski, Geometry of Quantum States: An Introduction to

Quantum Entanglement, Cambridge Univ. Press, Cambridge, UK, 2006.[4] J.C. Hou, X.F. Qi, arXiv:1107.0354v1.[5] Z.H. Ma, F.L. Zhang, D.L. Deng, J.L. Chen, Phys. Lett. A 373 (2009) 1616.[6] Z.H. Chen, Z.H. Ma, F.L. Zhang, J.L. Chen, Cent. Eur. J. Phys. 9 (2011) 1036.[7] A. Gilchrist, N.K. Langford, M.A. Nielsen, Phys. Rev. A 71 (2005) 062310.[8] A.E. Rastegin, J. Phys. A 40 (2007) 9533.[9] D. Markham, J.A. Miszczak, Z. Puchała, K. Zyczkowski, Phys. Rev. A 77 (2008)

042111.[10] C.A. Fuchs, J. van de Graaf, IEEE Trans. Inform. Theory 45 (1999) 1216.[11] X.-M. Lu, Z. Sun, X. Wang, P. Zanardi, Phys. Rev. A 78 (2008) 032309.[12] J.A. Miszczak, Z. Puchała, P. Horodecki, A. Uhlmann, K. Zyczkowski, Quantum

Inf. Comput. 9 (2009) 0103.[13] P.E.M.F. Mendonca, R.d.J. Napolitano, M.A. Marchiolli, C.J. Foster, Y.C. Liang, Phys.

Rev. A 78 (2008) 052330.[14] Z. Puchała, J.A. Miszczak, Phys. Rev. A 79 (2009) 024302.[15] Z. Puchała, J.A. Miszczak, J. Phys. A: Math. Theor. 44 (2011) 405301.[16] Z. Puchała, J.A. Miszczak, P. Gawron, B. Gardas, Quantum Inf. Proc. 10 (2011) 1.[17] Man-Duen Choi, Linear Algebra Appl. 10 (1975) 285.[18] P. Busch, J. Singh, Phys. Lett. A 249 (1998) 10.[19] J.B. Conway, A Course in Functional Analysis, Springer, New York, 1989.