9
Quantum Discord and Nonclassical Correlations Beyond Entanglement

Gerardo Adesso Marco Cianciaruso and Thomas R. Bromley

University of Nottingham, Centre for the Mathematics and Theoretical Physics of Quantum Non‐Equilibrium Systems, School of Mathematical Sciences, University Park, Nottingham, NG7 2RD, UK

9.1 Introduction

What is quantum? As researchers of quantum physics, we are constantly bombarded with attributes such as “nonclassical” and “superclassical.” We strive to track down the elusive quantum–classical boundary, to understand what makes quantum mechanics so powerful yet counterintuitive. But for this purpose, we must first have a firm understanding of the classical world and the laws that classical mechanics imposes.

There are, in fact, many ways to think about classicality. One facet of the classical world is that any system is always in a fixed and predetermined state. Take, for example, a bit: it can be either 0 or 1. How does this compare with what is predicted from the rulebook of quantum mechanics? Here, we can have systems existing in a superposition of both 0 and 1, called quantum bits or qubits. This form of nonclassicality is known as quantum coherence (1).

It is also interesting to consider systems of spatially separated parties and the correlations between them. We can try to identify the states that are describable by classical mechanics and the states that are not. You are probably now thinking that this sounds a lot like entanglement (2) and that the classically correlated states are just separable states. However, things are not so simple: it turns out that even separable mixed states can exhibit some quantumness in their correlations!

In this chapter, we explore these manifestations of quantum correlations beyond entanglement (35). We begin by introducing and motivating the classically correlated states and then showing how to quantify the quantum correlations using an entropic approach, arriving at a well‐known measure called the quantum discord (6,7). Quantum correlations and discord are then operationally linked with the task of local broadcasting (8). We conclude by providing some alternative perspectives on quantum correlations and how to measure them.

Finally, before proceeding, it is important to note that there are many layers of quantumness in composite systems. As well as entanglement and discord‐type quantum correlations, one can identify for example, steering and Bell nonlocality. For pure composite states, all of these signatures of quantumness become equivalent, yet for mixed states they are different, showing a strict hierarchy. Each form of quantumness is of independent interest, but in this chapter, we focus on the most general form of quantum correlations, leaving the interested reader to consult Chapter 8 for more information on entanglement and Refs. (9,10) for steering and nonlocality.

9.2 Quantumness Versus Classicality (of Correlations)

Generally, quantumness can represent any of the counterintuitive phenomena that we encounter when investigating microscopic systems such as atoms, electrons, photons, and many others. In particular, the quantumness of correlations manifests itself when two such microscopic systems interact with each other, and stands as one of the weirdest of all quantum features. In order to really appreciate any sort of quantumness, we first need to thoroughly understand how the classical world works, that is, we first need to agree on what exactly “intuitive” means, and only afterward benchmark quantumness against such a standard. This is the purpose of this section.

Let us set the stage for our comparison of the classical and the quantum. From a minimalistic point of view, both classical and quantum systems can be described by resorting to the following four ingredients: the set of states, the set of observables, a real number associated with any pairing of a state and observable, which is the predicted result of a measurement of the given observable when the system is in the given state, and a family of mappings describing the dynamics of the system. However, in the following, we will focus only on the first three ingredients; we will also specialize to discrete variable systems for the sake of simplicity.

The state of a discrete variable classical system, whose phase space images is formed by images points that we label by images , can be described by a probability distribution images defined on images , that is, any set of images numbers that are nonnegative, images , and normalized, images . An observable of such a system is instead any real function images on images , that is, images , while what we actually observe by measuring the observable images when the system is in the state images is the corresponding expectation value, that is, images .

We say that a classical system is in a pure state when we have the best possible knowledge about it, that is, we know with certainty what point of the phase space is occupied by the system. In fact, pure states of classical systems are nothing but Kronecker deltas images , with images being the point in the phase space occupied by the system, that is, images if images while images if images . Moreover, when a classical system is in a pure state images , we can predict with certainty that the result of the measurement of an arbitrary observable images is the value images , where images is the point of the phase space occupied by the system. Interestingly, every state of a classical system that is not pure can be obtained in a unique way as a convex combination of pure states, and it is thus called a mixed state.

Our ignorance about the state images of a classical system can be quantified by resorting to its Shannon entropy,

9.1 equation

which is indeed zero for pure states and reaches its maximum for the so‐called maximally mixed state. The latter is such that images for any images and thus entails that we have the least possible knowledge about which one of the points of the phase space is actually occupied by the system, as all such points equally probable.

When considering two discrete variable classical subsystems images and images , with phase spaces given by images and images , respectively, it happens that the phase space images corresponding to the composite system images is the Cartesian product of the ones corresponding to the two subsystems, that is, images , whose points are given by the images ordered pairs images . The state of a bipartite classical system can be thus described by a joint probability distribution images defined on images , while the states of the subsystems images and images can be characterized by the corresponding marginal probability distributions, that is, images and images , respectively.

In particular, pure states of bipartite classical systems are given by products of Kronecker deltas, images , where images is the point of the phase space occupied with certainty by the bipartite system, that is, images if images while images if images . Again, every state of a bipartite classical state that is not pure can be written in a unique way as a convex combination of pure bipartite states, that is, as a classical mixture of products of Kronecker deltas. Furthermore, quite interestingly, when a bipartite classical system is in a pure state, then also the subsystems are necessarily in a pure state, indeed one can easily see that the marginal distributions of images are images and images . In other words, within the classical world, if we have the best possible knowledge of the state of a composite system, then we necessarily have the best possible knowledge of the states of both its subsystems.

On the other hand, the state of a discrete‐variable quantum system, whose Hilbert space images has a finite dimension images , can be described by a density operator images acting on images , that is, any linear operator on images that is positive semidefinite, images , and normalized, images . An observable of such a system is instead any Hermitian operator images on images , that is, images , while what we actually observe by measuring the observable images when the system is in the state images is the corresponding expectation value, that is, images .

Again, we say that a quantum system is in a pure state when we have the best possible knowledge about it, that is, we know with certainty what normalized vector of the Hilbert space is occupied by the system. Pure states of quantum systems are thus described by projectors images onto normalized vectors images of images . Moreover, when a quantum system is in a pure state images , we can predict with certainty the result of the measurement of any observable images having images between its eigenvectors, without perturbing the state of the system whatsoever. However, contrary to what happens in the classical world, this is no longer the case when we measure any other kind of observable, whose eigenvectors are different from images . More precisely, if we measure a generic observable images with eigenvectors images when the quantum system is in the state images , it happens that the state of the system can collapse onto any of the eigenstates images of images with probability images . This is not due to our ignorance about the state of the system, but rather due to an intrinsic indeterminism manifested by nature at the microscopic level, a fact which stands as one of the most striking features of quantumness. This phenomenon is mathematically taken into account by the fact that in the quantum setting we have that states and observables are no longer commuting real functions but rather possibly noncommuting Hermitian operators.

Yet there is another striking quantum feature that manifests itself in single quantum systems, as we have already alluded to: the celebrated quantum superposition, or coherence. It arises from the fact that in the quantum setting we are not only allowed to consider classical mixtures of pure states, that is, images , also called simply mixed states, but rather we can also construct coherent superpositions of pure states that give rise to other pure states, that is, images . However, particular mention has to be given to superpositions and mixtures of elements of an orthonormal basis images of images . Indeed, one can easily appreciate that, due to the perfect distinguishability of orthogonal states, any quantum state of the form images can be simulated by the classical state images . Therefore, such states represent a stereotype of classicality within the quantum world and are called incoherent states.

Our classical ignorance about the state images of a quantum system can be quantified by resorting to its von Neumann entropy,

9.2 equation

which is indeed zero for pure states and reaches its maximum for the maximally mixed state, images , with images being the identity on images .

When considering two discrete‐variable quantum systems images and images , with Hilbert spaces given by images and images , respectively, it happens that the Hilbert space images corresponding to the composite system images is the tensor product of the ones corresponding to the two subsystems, that is, images , which is a images ‐dimensional Hilbert space whose vectors are spanned by the orthonormal product basis images , with images and images being orthonormal bases of images and images , respectively. The state of a bipartite quantum system can be thus described by a density operator images acting on images , while the states of the subsystems images and images can be characterized by the corresponding marginal density operators, that is, images and images , respectively, where images is the partial trace over the Hilbert space of subsystem images .

In particular, pure states of bipartite quantum systems are given by projectors onto normalized vectors of images . Here comes one of the most amazing features of quantum mechanics, which is attributed to quantum correlations. Due to both the superposition principle and the tensorial structure of the Hilbert space of the composite system, it happens that a pure bipartite quantum state is not necessarily factorizable in the tensor product of two pure states of the subsystems, that is, images cannot be written in general in the form images , with images and images . An immediate consequence of the nonfactorizability of a pure bipartite state images is the fact that the corresponding subsystems' states are necessarily nonpure. In other words, within the quantum world, the best possible knowledge of the state of a composite system does not imply the best possible knowledge of the states of the two subsystems. This is in stark contrast with what happens in the classical world and, as Schrödinger said, stands as “not one but rather the characteristic trait of quantum mechanics, the one that enforces its entire departure from classical line of thought” (11). This phenomenon was baptized entanglement by Schrödinger, but it is nowadays more broadly known as quantum correlations for pure states. Overall, for pure bipartite quantum states images , we get two possibilities: either images is a product state, images , for some images and images , which is separable and does not manifest any quantum correlations; or images is not factorizable, in which case it is entangled and hence manifests quantum correlations. This is the whole story as far as pure states are concerned: entanglement entirely captures every aspect of quantum correlations.

9.2.1 Identifying Classically Correlated States

For bipartite quantum mixed states, however, the story becomes more complicated than that, as there are many paradigms that we can adopt in order to define what a classically correlated state is. One paradigm identifies the classically correlated states with the states that can be described by a local realistic model. According to this paradigm, only a restricted aristocracy of quantum states are not classically correlated, the so‐called nonlocal states (10). Another paradigm is the one corresponding to entanglement, wherein classically correlated states can be written as convex combinations of tensor product of pure states, so‐called separable states (2), that is,

9.3 equation

with images being a probability distribution, images and images . Separable states remind us of what happens in the classical setting, wherein all joint probability distributions can be written as a convex combination of products of Kronecker deltas, which are indeed the classical pure states. According to the entanglement paradigm, the right of being quantumly correlated is extended from the restricted aristocracy of nonlocal states to the broader bourgeoisie of nonseparable quantum states. Finally, we get to the paradigm representing the focus of this chapter, which goes even beyond entanglement, thus allowing the right of being quantumly correlated to almost all the population of quantum states.

As we have already mentioned, the embedding of a state of a classical system into the quantum state space is the corresponding classical mixture of elements of an orthonormal basis. However, when embedding the state of a classical composite system, imposing just the orthonormality of the basis is not enough, as one also needs to impose that such a basis is factorizable in order for the corresponding classical mixture to be entirely simulated by a classical bipartite state. This gives rise to a so‐called classical–classical state, that is,

9.4 equation

where images is a joint probability distribution, while images and images are orthonormal bases of images and images , respectively. One can indeed easily see that the marginal states of a classical–classical state are still classical states, that is, classical mixtures of elements of an orthonormal basis: images and images , where we have that images and images are exactly the marginal probability distributions of the joint probability distribution images .

Furthermore, one can also define the embedding of a classical state of only subsystem images into the quantum state space of a bipartite quantum system images by considering what is known as a classical–quantum state, that is,

9.5 equation

with images being a probability distribution, images an orthonormal basis of images and images arbitrary states of subsystem images . In this case, one can easily see that in general only the marginal state of subsystem images is still a classical state, while the marginal state of subsystem images could be in principle any quantum state, that is, images while images .

An analogous definition holds when considering the embedding of a classical state of only subsystem images into the state space of a bipartite quantum system images , also called a quantum–classical state, that is,

9.6 equation

with images being a probability distribution, images an orthonormal basis of images and images arbitrary states of subsystem images .

Classical–classical, classical–quantum, and quantum–classical states, which we may collectively refer to as classically correlated states, form nonconvex sets of measure zero and nowhere dense in the space of all bipartite quantum states images (12). This is in stark contrast with the set of separable states, which is convex and occupies a finite volume in the state space instead (2).

9.3 Quantifying Quantum Correlations – Quantum Discord

As mentioned in the introduction, and as will be shown in more detail in the following sections, quantum correlations beyond entanglement can represent a resource for some operational tasks and allow us to achieve them with an efficiency that is unreachable by any classical means. The quantification of this type of quantumness is thus necessary to gauge the quantum enhancement when performing such tasks.

Let us start from the quantification of quantum correlations for pure states. We have already mentioned that in this case the entire amount of quantum correlations is captured by entanglement. This can be, in turn, described by the fact that, when dealing with pure bipartite quantum states that are not factorizable, the best possible knowledge of a whole does not include the best possible knowledge of all its parts, as the corresponding marginal states are necessarily mixed. Such a loss of information on the pure state of the whole system when accessing only part of it, as quantified for example, by the von Neumann entropy of any of the marginal states, captures exactly the entanglement, and thus the whole quantum correlations, between the two parties01:

9.7 equation

The pure state entanglement quantifier images is also known as entropy of entanglement.

Let us now move on to the quantification of quantum correlations beyond entanglement for mixed states. Both adopting an entropic viewpoint and a thorough comparison with the classical setting will turn out to be crucial at this stage, as happened in the previous section when addressing the characterization of quantum correlations. When a bipartite classical system images is in a mixed state images , then we have some ignorance about it that can be quantified by its strictly positive Shannon entropy images . At the same time, quite intuitively, it turns out that the overall ignorance about the marginal states images and images of the two subsystems images and images treated separately, which is quantified by the quantity images , is necessarily higher than or equal to the ignorance about the state of the combined bipartite system, which is instead quantified by images . In other words, there is, in general, a loss of information on the state of the whole system when accessing only its parts. This can be quantified by the so‐called mutual information:

9.8 equation

Such a loss of information when accessing a composite system locally is attributed to underlying correlations between the subsystems, so that the mutual information stands as a fully fledged quantifier of correlations. We can think of two correlated subsystems images and images as two accomplices. If the policemen interrogate them separately, the more the two accomplices are correlated, the less information the policemen will manage to gain regarding what images did together, with their mutual information representing exactly the amount of information that the two accomplices are hiding to the policemen. Clearly, for pure bipartite classical states we always get a zero mutual information, as both the composite system state and the marginal states are pure and so their Shannon entropies are all zero and there is no loss of information in accessing the composite system locally. This entails that it is impossible to have correlations between classical systems sharing a pure state, contrary to what happens within the quantum world where we can have entanglement for pure states. More generally, the mutual information is equal to zero if, and only if, the bipartite classical state images is factorizable, that is, images for any images and images , which is indeed the paradigmatic form of probability distribution that does not manifest any correlations at all.

Yet there is another equivalent perspective from which we can look at correlations in the classical setting. Let us first define images as the conditional probability distribution of subsystem images after we know that subsystem images occupies exactly the point images of its phase space. Analogously, we define images as the conditional probability distribution of subsystem images after we know that subsystem images occupies exactly the point images of its phase space. Then, one can prove that the mutual information of the bipartite state images is equal to the following quantity:

9.9 equation

The above equivalent expressions of the mutual information tell us that the more two subsystems images and images are correlated, the more the ignorance about one subsystem decreases on average when we know the state of the other subsystem. On the other hand, if images and images are not correlated at all, then gaining some information about one subsystem does not help us in gaining any information about the other subsystem.

Now the question is: how can we translate such a machinery into the quantum setting in order to quantify quantum correlations beyond entanglement? Clearly, we can start by defining the quantum mutual information in order to quantify the totality of correlations of bipartite quantum states images as follows:

9.10 equation

where images here denotes the von Neumann entropy. In analogy with the classical case, the quantum mutual information is equal to zero if, and only if, images is factorizable, that is, images , and, thus, there are no correlations whatsoever, not even classical ones, between images and images . However, in order to fully answer our question, we need to find out how to discern the portion of the total correlations that is purely quantum from the one that can be regarded as mere classical correlations, a problem that was rigorously addressed for the first time by Henderson and Vedral (7).

To this purpose, it will be crucial to translate in the quantum setting also the quantity images , which in the classical setting represents just an equivalent expression for the mutual information. We thus need to define also in the quantum setting the conditional state of one subsystem given that we have gained some information about the other subsystem. The most intuitive way to gain information about a single quantum subsystem, say images , is to measure a local observable of the form images , where images is a Hermitian operator on images while images is the identity operator on images . As we have already mentioned, the result of such a measurement is in general uncertain and can map the system, with probability images , into the state images , where the rank‐one projectors images are the eigenstates of images . Therefore, the conditional state of subsystem images after such a local measurement has been performed on images and the result images has been obtained is images . We can thus define the decrease, on average, in the entropy of images given that we have performed the local measurement on images described by the rank‐one projectors images on images as

9.11 equation

Some remarks are now in order. Firstly, contrary to the classical case, we can define different versions of conditional states images of images given images , and so different versions of the quantity images , just by varying the local measurement images that has been performed on images . Secondly, one can even consider more general kinds of local measurements (described by positive operator‐valued measures), but we restrict to rank‐one projective measurements here for the sake of simplicity.

Finally, the correlations underlying such a gain of information about subsystem images , when accessing locally subsystem images after the local measurement images , can be considered classical from the perspective of subsystem images , as they are nothing but the correlations that are left into the postmeasurement state images , which is clearly a classical–quantum state. In other words, one can see that the following equality holds:

9.12 equation

Therefore, if one wants to extract from the total correlations images of the bipartite state images the purely quantum portion of correlations from the perspective of subsystem images , that is, the amount of mutual information of images and images that can be never classically extracted via a local measurement on images , not even by performing a maximally informative one, then one can consider the following quantity:

9.13 equation

where the maximization is over all rank‐one local projective measurements on images . images is the celebrated quantifier of quantum correlations beyond entanglement from the perspective of subsystem images that goes under the name of quantum discord and was introduced by Ollivier and Zurek (6). The complementary quantity

9.14 equation

quantifies the classical correlations from the perspective of subsystem images as formalized by Henderson and Vedral (7). In this way, quantum discord images and classical correlations images add up to the total correlations quantified by the mutual information images , and we have addressed the original question posed in this section, by finding a meaningful way to separate the quantum from the classical portion of correlations in a state images , from the perspective of subsystem images .

Analogous definitions hold when measuring locally subsystem images , by swapping the roles of images and images . In particular, the quantum discord from the perspective of subsystem images can be defined as

9.15 equation

where the maximization is over all rank‐one local projective measurements on images .

A further couple of remarks are in order before concluding this section. Firstly, a fundamental asymmetry arises between how the quantum correlations between images and images are perceived by each subsystem, because in general images is different from images . Quantum discord is, in fact, a one‐sided measure of quantumness of correlations. However, such an asymmetry can be bypassed by considering the action of local joint measurements on both images and images and defining accordingly symmetric (or two‐sided) quantifiers of quantum and classical correlations from the perspective of either images or images within the same entropic framework adopted in this section ( 8,13,14). More details on these quantifiers, which may be denoted, respectively, by images and images , as well as their interplay with one‐sided measures, are available in ( 5,15).

Secondly, by using both the fact that classical–quantum states images can be left invariant by at least one local projective measurement images on images , that is, images , and the fact that the result of such a measurement applied to any state is always a classical–quantum state, that is, images for any images (see Exercise 9.1 at the end of the chapter), one can easily show that images if, and only if, images is classical–quantum. An analogous result holds for quantum correlations with respect to images , that is, images if, and only if, images is quantum–classical. This cements the paradigm adopted in this chapter, according to which almost all quantum bipartite states, and not only entangled states, manifest genuinely quantum features that can be attributed to nonclassical correlations.

9.4 Interpreting Quantum Correlations – Local Broadcasting

We have identified the classically and quantumly correlated states and provided an entropic way to measure quantum correlations in terms of the discord. It is now time for us to place what we have learnt in more concrete terms by understanding the role of quantum correlations in an operational task: local broadcasting ( 8,16).

Let us first consider copying of information. This happens all the time in the classical realm: from hard drives to mobile telephones – our modern world relies on the ability to freely copy information. In stark contrast, general copying of information is expressly prohibited in quantum mechanics by the no‐broadcasting theorem (17), which is a generalization of the well‐known no‐cloning theorem (18,19). Think of a quantum system images in one of two states images or images . We attach an ancilla images in the state images to get the composite state images with images . The goal is to perform some transformation images to the composite state to get images such that images for both images . In other words, we want to be able to copy two arbitrary quantum states images and images . However, it turns out this is only possible if images and images commute, which effectively reduces to copying of classical information.

The objective of local broadcasting is similar (8). Consider now a composite state images shared between two subsystems images and images . We give each subsystem an ancilla images and images so that the joint state is images and ask if there exists a local operation images so that we get the state images obeying the relation images . More generally, we can consider the task of simply distributing the (total) correlations images of images , and ask if there are local operations such that images . This is what we mean by local broadcasting, and it was shown in (8) that such a process can only take place perfectly if images is classical–classical, otherwise we lose correlations during our attempt at local broadcasting.

A similar one‐sided version of local broadcasting has also been proposed in (16). Here, we just give subsystem images their ancilla images and ask if there is a local operation images yielding images such that images . As you might have guessed, this version of local broadcasting can occur only if images is classical–quantum.

We thus have a very intuitive characterization of classical–classical states and classical–quantum states: they are exactly the states that can be locally broadcast. So can we use this concept of local broadcasting to quantify the quantum correlations present in a state? Now let us imagine that images wants to distribute their correlations with images to images ancillae images using local operations images (20). If we define the reduced state of each pair images and images after such local operations as

9.16 equation

we know from the above analysis that correlations will never increase, that is, images , with equality only if images is classical–quantum. Let us suppose that images is not classical–quantum, but we want to distribute our correlations in an efficient way, that is, losing the least possible amount of correlations. We can consider the loss of correlations images for each ancilla. Averaging this quantity over all ancillae then gives a good figure of merit for our redistribution of correlations. By further minimizing this figure of merit over all possible local operations, we get

9.17 equation

This quantity is zero if images is classical–quantum, and positive otherwise. Can its value quantify the quantum correlations of images ? Remarkably, in the limit of infinitely many ancillae, it has been proven in (20) that the quantity in Eq. 9.17 reproduces exactly the quantum discord given by Eq. 9.13 [see Figure 9.1a]:

9.18 equation

This relation provides a striking operational understanding of quantum discord as the minimum average loss of correlations if one attempts to redistribute the correlations between images and images in the state images to infinitely many ancillae on images 's side: paraphrasing the words of (21), “quantum correlations cannot be shared.” We note that additional operational interpretations for the quantum discord in quantum information theory and thermodynamics have been discovered, as reviewed in ( 3 5).

9.5 Alternative Characterizations of Quantum Correlations

So far we have focused on the characterization of classically correlated states and the quantification of quantum correlations in an entropic setting, using the quantum discord. One property that we have pointed out along the way is that the classically correlated states are insensitive to a local complete rank‐one projective measurement, a hallmark feature of the classical world. It has also been shown that classically correlated states are the only ones that are locally broadcastable, another intuitive property arising from the inability to copy general quantum states. It turns out that there is a whole raft of equivalent defining properties for the classically correlated states, and that with each property comes another way to quantify the quantum correlations (5). The quantum discord accounts for the loss of correlations due to local measurements, but it is just one of many ways to measure the quantum correlations of a state. We will outline two more key properties of classically correlated states in the following, along with the corresponding method of measuring quantum correlations.

Illustration of Operational interpretations and quantification of quantum correlations: Local broadcasting of correlations.; Image described by caption and surrounding text.

Figure 9.1 Operational interpretations and quantification of quantum correlations. (a) Local broadcasting of correlations [Section 9.4]. Two quantum systems images and images are initially in an arbitrary bipartite state images with generally classical and quantum correlations. If a local channel images is applied to images which redistributes it into asymptotically many fragments images , then the only correlations remaining on average between each fragment images and subsystem images are classical ones, while quantum correlations, quantified by the quantum discord images , cannot be shared. This can be seen as a manifestation of quantum Darwinism (Brandão et al. (2015) (20) Copyright 2014, Nature Publishing Group.). (b) Scheme of a premeasurement interaction acting on subsystem images of a bipartite system images , described as a local unitary images on images , followed by a generalized control‐NOT operation with an ancilla images (which plays the role of a measurement apparatus). Provided images is initialized in a pure state images , the output premeasurement state images is always entangled along the images split if and only if the initial state images of the system is not classical–quantum, that is, contains general quantum correlations from the perspective of subsystem images (Streltsov et al. (2011) (22) and Piani et al. (2011) (23). Copyright 2014, American Physical Society.). The minimum entanglement images between images and images in the premeasurement state, where the minimization is over all the local bases on images specified by images , quantifies the quantum correlations images in the input bipartite state images , according to the entanglement activation paradigm [Section 9.5.2]. (c) Graphical legend for the different types of correlations appearing in panels (a) and (b).

9.5.1 Local Coherence

Recall that we define the incoherent states with respect to a reference basis images as those diagonal in this basis, that is, states that can be written as images for some probability distribution images . Any state that is not diagonal in this basis is called coherent ( 1,24). Now let us consider a bipartite quantum system images with local reference bases images in images and images in images . States incoherent with respect to the product basis images can be written as

9.19 equation

for some joint probability distribution images , while states incoherent in the local reference basis images are written as

9.20 equation

for some probability distribution images and with arbitrary states images of subsystem images . We can say that these locally incoherent states are incoherent–incoherent and incoherent–quantum, respectively. Take a look back at Eqs. 9.4 and 9.5 describing the classically correlated states. You would be forgiven for thinking that they are identical to the above equations! However, there is a subtlety here: the locally incoherent states are diagonal in a fixed local basis, while the classically correlated states are diagonal in some local basis. This analogy then provides us with another characterization of the classically correlated states, that is classical–classical states are incoherent–incoherent for some product basis on images and images , while classical–quantum states are incoherent–quantum for some local basis on images (5).

On the other hand, quantumly correlated states are coherent in every local basis. Can we then use measures of coherence to inform us on the amount of quantum correlations? Consider the observable images diagonal in a fixed reference basis images . One way to measure the coherence of a state images with respect to the reference basis, or more precisely its asymmetry with respect to translations generated by the observable images , is by means of the quantum Fisher information images (25,26). This quantity plays a fundamental role in quantum metrology (27) and indicates the ultimate precision achievable using a quantum probe state images to estimate a parameter encoded in a unitary evolution generated by the observable images . Let us now fix a family of local observables images on subsystem images with fixed nondegenerate spectrum images . Defining the minimum of images over all local observables images with spectrum images gives a measure of quantum correlations (28):

9.21 equation

Such a measure embodies the worst‐case scenario sensitivity of a bipartite state images when a parameter is imprinted onto subsystem images by any of the observables images : a process that is fundamentally linked to quantum interferometry and hence motivates the naming of images as the interferometric power (28). While there are many other good measures of quantum coherence (1), from which one can define corresponding measures of quantum correlations (by minimization over local reference bases) (5), the interferometric power is one of the most compelling as it brings together quantum coherence, quantum correlations, and metrology. Another advantage of this measure is that images admits a computable formula for any state images whenever images is a qubit (28) (see Exercise 9.2 at the end of the chapter), while no such analytical formula is presently available for the quantum discord images of general two‐qubit or qubit–qudit states.

9.5.2 Entanglement Activation

Let us now examine more closely the workings of a local projective measurement images with local projectors images acting on subsystem images of a bipartite state images . According to von Neumann's model (29), this measurement can be realized in two steps. First, subsystem images is allowed to interact with an ancilla images , initialized in a fiducial pure state images , through a unitary images . The unitary acts in the following way:

9.22 equation

and can be realized by the combination images of a local unitary images , which sets the basis of measurement, followed by a generalized controlled‐NOT gate images , whose action on the computational basis images of images is images , with images denoting addition modulo images . The resultant state after applying the unitary images to images , images , and images is

9.23 equation

which is known as the premeasurement state. Next, the local projective measurement is completed by partial tracing over subsystem images , which is achieved by a readout of the ancilla images in its eigenbasis, so that images .

During this process, the ancilla images can become entangled with images and images due to the unitary images , which means that the premeasurement state images may not be separable along the bipartition images . However, sometimes images remains separable along such a cut. It turns out this is the case only when images is initially incoherent–quantum, of the form in Eq. 9.20. It thus becomes clear that we can characterize the classical–quantum states of Eq. 9.5 as exactly all and only the states for which there exists a local basis images such that the premeasurement state images is separable along the split images (22).

Similarly, if we consider a local projective measurement images on both images and images in the bases images and images , we can also introduce an ancilla images for images and a corresponding premeasurement state images . A similar line of thought can then be applied whereby we find that the classical–classical states of Eq. 9.4 are all and only the states for which the premeasurement state is separable along the split images for some local bases images and images (23).

From this analysis, it can be said that the classical correlations are not always activated into entanglement during a premeasurement, while the quantum correlations always are. Such a conversion of nonclassical resources due to a premeasurement interaction has been demonstrated experimentally in (30). Naturally, one can then aim to quantify the quantum correlations of images by measuring the entanglement of the corresponding premeasurement state, via some chosen entanglement measure images , minimized over all local bases.

For every suitable images , we can then define a corresponding (one‐sided or two‐sided) measure of quantum correlations ( 22, 23) as follows [see Figure 9.1b]

9.24 equation

One of the most remarkable features of this approach is that the measures so defined capture quantitatively the hierarchy of quantum correlations, as one has images for any valid entanglement measure images and any bipartite state images , with equalities on pure states images .

For instance, one may choose the relative entropy of entanglement (31)

9.25 equation

as our entanglement measure, where images is the relative entropy and the minimization is over all separable states images of the form in Eq. 9.3. The corresponding measures of quantum correlations, obtained by specifying images as images in Eqs. 9.24, are known, respectively, as relative entropy of discord (one‐sided) and relative entropy of quantumness (two‐sided). Interestingly, these measures have been proven equivalent to the following expressions ( 22, 23):

9.26 equation

with minimizations over the classical–quantum states of Eq. 9.5 and the classical–classical states of Eq. 9.4, respectively. This enriches the quantification of quantum correlations as potential resources for entanglement creation, with an additional geometric interpretation in terms of the distance02 from the set(s) of classically correlated states. In turn, such a geometric approach can be used a priori to quantify quantum correlations adopting different distance functionals, as reviewed in ( 5,32,33).

9.6 General Desiderata for Measures of Quantum Correlations

We have identified several alternative, yet equivalent characterizations of the classically correlated states, in particular providing links with other fundamental elements of quantum mechanics such as coherence (1) and entanglement (2). With each characterization of the classically correlated states comes another way to measure quantum correlations. Given such a catalog of measures (5), it is sensible to wonder what makes a good measure of quantum correlations. This question is typically answered by imposing a number of requirements that any such good measure should obey. Let us consider a one‐sided measure images , defined by a real nonnegative function acting on quantum states images . One natural requirement is that

  • images ,

that is, that our measure is zero for classically correlated states. We should also expect that quantum correlations are not dependent upon the local bases of images and images , which manifests as invariance under local unitaries images on images and images on images ,

  • images .

As we have already pointed out, entanglement and quantum correlations become the same phenomenon for pure states images ; hence, it is sensible to require that a measure of quantum correlations should reduce to a measure of entanglement for pure states,

  • images for some entanglement measure images .

Similar desiderata can be imposed for two‐sided measures of quantum correlations images . However, so far we have not specified how our measure of quantum correlations should behave under dynamics of the system. In the case of entanglement, it is typically required that a measure should not increase under local operations and classical communication (LOCC) images , that is, images ( 2, 31). In other words, one should not be able to generate entanglement by LOCC, the archetypal operations that spatially separated laboratories are limited to. This requirement is typically called monotonicity, and finding a comparable one for quantum correlations is tricky. For one‐sided measures, it can be required that any local operation on subsystem images should not be able to increase the quantum correlations from the perspective of subsystem images (22), that is,

  • images for any local operation images on images .

Unfortunately, this cannot be the only monotonicity requirement, since it only specifies the local operations on images . Identifying the most meaningful and complete set of operations under which a good measure of quantum correlations should be monotone is currently an open question. We point the reader to (5) for a deeper explanation.

9.7 Outlook

We are going to be relying increasingly on the quantum world as technologies evolve during the twenty first century, so it is certainly worthwhile to develop a good understanding of the quantum–classical boundary. In this chapter, we focused on the most general type of quantum correlations between spatially separated parties. Although a promising topic, it is still very much in its infancy, with a plethora of interesting and open questions yet to be answered. From the theoretical side, perhaps the most pressing question is to identify a physically motivated set of “free operations” under which to impose monotonicity for measures of quantum correlations. This can be achieved by treating quantum correlations as a resource, within the framework of resource theories (34). Experimentally, we have yet to witness compelling evidence for the practical role of quantum correlations beyond entanglement in relevant quantum technologies, even though proof‐of‐principle demonstrations, for example, in the context of quantum metrology, are particularly promising (28). In this respect, while the number of insightful operational interpretations for measures of quantum correlations has grown substantially in recent years ( 5,35), killer applications are perhaps still waiting to be devised. It is hoped that by raising the awareness of these concepts within the wider quantum information community, we can begin to truly appreciate the foundational role and power of nonclassical correlations beyond entanglement.

There are still many topics within the study of quantum correlations that we have not had the opportunity to cover here. Foremost among which is the extensive research on their dynamics in open quantum systems, which shows that quantum correlations are generally more resilient than entanglement to the effects of typical sources of noise and decoherence (36,37), a promising feature for any quantum technology. We have also neither discussed the role of quantum correlations in quantum computing (38,39) and cryptography (40), nor the quantification of quantum correlations among more than two parties (41) or in continuous variable systems (42). Nevertheless, there is a wealth of resources available to fill these gaps ( 3 5,32, 33). We hope to have passed on to the reader our enthusiasm for this young and blossoming field at the very core of quantum mechanics and look forward to future progress.

Exercises

  1. 9.1 Classically correlated states and quantum discord
    1. Show the equivalence of the two versions of classical mutual information in Eqs. 9.8 and 9.9.
    2. Verify that the quantum discord Eq. 9.13 reduces to the entropy of entanglement Eq. 9.7 for pure bipartite states images .
    3. For a bipartite system images , show that classical–quantum states of Eq. 9.5 can be left invariant by at least one local rank‐one projective measurement on images .
    4. For a bipartite system images , show that the postmeasurement state of a local rank‐one projective measurement on images is always a classical–quantum state.
  2. 9.2 Alternative characterizations of quantum correlations
    1. For two‐qubit systems, the interferometric power of images in Eq. 9.21 can be computed by finding the smallest eigenvalue of the images matrix images with entries
      equation
      where images and images are the eigenvalues and eigenvectors of images and images are the Pauli matrices. Calculate the interferometric power for the Werner states
      equation

      with images and images .

      Compare it with an entanglement measure, for example, the squared concurrence.

    2. The relative entropy is contractive under any quantum channel images , that is, images . Using this property and Eq. 9.26, show that the one‐sided activation‐based measure of quantum correlations images in Eq. 9.24 can never increase under local operations on subsystem images .

References

  1. 1 Streltsov, A., Adesso, G., and Plenio, M.B. (2017) Colloquium: quantum coherence as a resource, Rev. Mod. Phys. 89, 041003.
  2. 2 Horodecki, R., Horodecki, P., Horodecki, M., and Horodecki, K. (2009) Quantum entanglement. Rev. Mod. Phys., 81, 865.
  3. 3 Modi, K., Brodutch, A., Cable, H., Paterek, T., and Vedral, V. (2012) The classical‐quantum boundary for correlations: discord and related measures. Rev. Mod. Phys., 84, 1655.
  4. 4 Streltsov, A. (2015) Quantum Correlations Beyond Entanglement and their Role in Quantum Information Theory, SpringerBriefs in Physics, Springer International Publishing.
  5. 5 Adesso, G., Bromley, T.R., and Cianciaruso, M. (2016) Measures and applications of quantum correlations, J. Phys. A: Math. Theor., 49, 473001.
  6. 6 Ollivier, H. and Zurek, W.H. (2001) Quantum discord: a measure of the quantumness of correlations. Phys. Rev. Lett., 88, 017 901.
  7. 7 Henderson, L. and Vedral, V. (2001) Classical, quantum and total correlations. J. Phys. A: Math. Gen., 34, 6899.
  8. 8 Piani, M., Horodecki, P., and Horodecki, R. (2008) No‐local‐broadcasting theorem for multipartite quantum correlations. Phys. Rev. Lett., 100, 090 502.
  9. 9 Cavalcanti, D. and Skrzypczyk, P. (2016) Quantum steering: a short review with focus on semidefinite programming. Rep. Prog. Phys., 80 (2), 024001.
  10. 10 Brunner, N., Cavalcanti, D., Pironio, S., Scarani, V., and Wehner, S. (2014) Bell nonlocality. Rev. Mod. Phys., 86, 419.
  11. 11 Schrödinger, E. (1935) Discussion of probability relations between separated systems, in Mathematical Proceedings of the Cambridge Philosophical Society, vol. 31, Cambridge University Press, pp. 555–563.
  12. 12 Ferraro, A., Aolita, L., Cavalcanti, D., Cucchietti, F., and Acín, A. (2010) Almost all quantum states have nonclassical correlations. Phys. Rev. A, 81, 052 318.
  13. 13 Wu, S., Poulsen, U.V., Mølmer, K. et al. (2009) Correlations in local measurements on a quantum state, and complementarity as an explanation of nonclassicality. Phys. Rev. A, 80, 032 319.
  14. 14 DiVincenzo, D.P., Horodecki, M., Leung, D.W., Smolin, J.A., and Terhal, B.M. (2004) Locking classical correlations in quantum states. Phys. Rev. Lett., 92, 067 902.
  15. 15 Lang, M.D., Caves, C.M., and Shaji, A. (2011) Entropic measures of non‐classical correlations. Int. J. Quantum Inf., 9, 1553.
  16. 16 Luo, S. (2010) On quantum no‐broadcasting. Lett. Math. Phys., 92, 143.
  17. 17 Barnum, H., Caves, C.M., Fuchs, C.A., Jozsa, R., and Schumacher, B. (1996) Noncommuting mixed states cannot be broadcast. Phys. Rev. Lett., 76, 2818.
  18. 18 Wootters, W.K. and Zurek, W.H. (1982) A single quantum cannot be cloned. Nature, 299, 802.
  19. 19 Dieks, D. (1982) Communication by EPR devices. Phys. Lett. A, 92, 271.
  20. 20 Brandão, F.G.S.L., Piani, M., and Horodecki, P. (2015) Generic emergence of classical features in quantum Darwinism. Nat. Commun., 6, 7908.
  21. 21 Streltsov, A. and Zurek, W.H. (2013) Quantum discord cannot be shared. Phys. Rev. Lett., 111, 040 401.
  22. 22 Streltsov, A., Kampermann, H., and Bruß, D. (2011) Linking quantum discord to entanglement in a measurement. Phys. Rev. Lett., 106, 160 401.
  23. 23 Piani, M., Gharibian, S., Adesso, G., Calsamiglia, J., Horodecki, P., and Winter, A. (2011) All nonclassical correlations can be activated into distillable entanglement. Phys. Rev. Lett., 106, 220 403.
  24. 24 Baumgratz, T., Cramer, M., and Plenio, M.B. (2014) Quantifying coherence. Phys. Rev. Lett., 113, 140 401.
  25. 25 Girolami, D. (2014) Observable measure of quantum coherence in finite dimensional systems. Phys. Rev. Lett., 113, 170 401.
  26. 26 Marvian, I. and Spekkens, R.W. (2014) Extending Noether's theorem by quantifying the asymmetry of quantum states. Nat. Commun., 5, 3821.
  27. 27 Braunstein, S.L. and Caves, C.M. (1994) Statistical distance and the geometry of quantum states. Phys. Rev. Lett., 72, 3439.
  28. 28 Girolami, D., Souza, A.M., Giovannetti, V., Tufarelli, T., Filgueiras, J.G., Sarthour, R.S., Soares‐Pinto, D.O., Oliveira, I.S., and Adesso, G. (2014) Quantum discord determines the interferometric power of quantum states. Phys. Rev. Lett., 112, 210 401.
  29. 29 von Neumann, J. (1932) Mathematical Foundations of Quantum Mechanics, Springer‐Verlag, Berlin.
  30. 30 Adesso, G., D'Ambrosio, V., Nagali, E., Piani, M., and Sciarrino, F. (2014) Experimental entanglement activation from discord in a programmable quantum measurement. Phys. Rev. Lett., 112, 140 501.
  31. 31 Vedral, V., Plenio, M.B., Rippin, M.A., and Knight, P.L. (1997) Quantifying entanglement. Phys. Rev. Lett., 78, 2275.
  32. 32 Spehner, D. (2014) Quantum correlations and distinguishability of quantum states. J. Math. Phys., 55, 075 211.
  33. 33 Roga, W., Spehner, D., and Illuminati, F. (2016) Geometric measures of quantum correlations: characterization, quantification, and comparison by distances and operations. J. Phys. A: Math. Theor., 49, 235 301.
  34. 34 Horodecki, M. and Oppenheim, J. (2013) (Quantumness in the context of) resource theories. Int. J. Mod. Phys. B, 27, 1345 019.
  35. 35 Georgescu, I. (2014) Quantum technology: the golden apple. Nat. Phys., 10, 474.
  36. 36 Maziero, J., Céleri, L.C., Serra, R.M., and Vedral, V. (2009) Classical and quantum correlations under decoherence. Phys. Rev. A, 80, 044 102.
  37. 37 Mazzola, L., Piilo, J., and Maniscalco, S. (2010) Sudden transition between classical and quantum decoherence. Phys. Rev. Lett., 104, 200 401.
  38. 38 Datta, A., Shaji, A., and Caves, C.M. (2008) Quantum discord and the power of one qubit. Phys. Rev. Lett., 100, 050 502.
  39. 39 Merali, Z. (2011) Quantum computing: the power of discord. Nature, 474, 24.
  40. 40 Pirandola, S. (2014) Quantum discord as a resource for quantum cryptography. Sci. Rep., 4, 6956.
  41. 41 Rulli, C.C. and Sarandy, M.S. (2011) Global quantum discord in multipartite systems. Phys. Rev. A, 84, 042 109.
  42. 42 Adesso, G. and Datta, A. (2010) Quantum versus classical correlations in Gaussian states. Phys. Rev. Lett., 105, 030 501.

Notes

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset