5
Channels and Maps

M. Keyl1 and R. F. Werner2

1 TU München, Zentrum Mathematik, Bolzmannstraße 3, D‐85748 Garching, Germany

2 Institut für Theoretische Physik, Leibniz Universtität Hannover, Appelstraße 2, D‐30167 Hannover, Germany

5.1 Introduction

Consider a typical quantum system such as a string of ions in a trap. To “process” the quantum information they carry, we have to perform, in general, many different processing steps such as free time evolution (including unwanted but unavoidable interactions with the environment), controlled time evolution (e.g., the application of a “quantum gate” in a quantum computer), preparations and measurements. This lecture aims at providing a unified framework for describing all these different operations.

5.2 Completely Positive Maps

The basic idea is to interpret each processing step as a channel, which transforms the system's initial state images into the output state images the system attains, after completion of the processing. Occasionally, we will represent this picture graphically as in Figure 5.1. To get a mathematical description, consider the two Hilbert spaces images , images (subsequently called the “initial” and the “target” Hilbert space) with (finite) dimensions1 images and images and the algebras images , respectively, images of (bounded) operators on them. Input and output states are described by density operators on images and images , which we denote by images and images again. Using this notation, we can regard a channel as a map images , which transforms the input state images into the output state images .

Graphical illustration of a channel, transforming the system's initial state ρin into the final state ρout.

Figure 5.1 Graphical representation of a channel, transforming the system's initial state images into the final state images .

Each physically reasonable operation should obey the mixing of states, that is, if images , images are transformed into images , the mixture images , images is mapped to images . This implies that images can be extended to a linear map

5.1 equation

and since images maps density matrices to density matrices it must be positive

5.2 equation

and trace preserving

5.3 equation

Hence, each channel can be described by a positive and trace‐preserving linear map images .

However, this picture is incomplete because we can apply a channel not only to the overall system but also to subsystems. A typical example arises, if Alice and Bob share a bipartite system in an entangled state images and Alice applies a local quantum operation to her subsystem (and Bob does nothing). Again, the crucial point is that the overall system shared by Alice and Bob end up in a valid quantum state images . In other words, the combination of “quantum operation performed by Alice” and “doing nothing by Bob” can be interpreted as a valid channel applied to the bipartite system (cf. Figure 5.2). If images and images are density matrices on images or images , respectively, the channel applied by Alice can be described by a positive, trace‐preserving linear map images , while “doing nothing” on Bob's system is just represented by the identity images . Hence, the output state can be written as images . Obviously, images is linear and trace preserving; but positivity of images is not sufficient for positivity of images . The most prominent example where this fails is the transposition. Although the transpose of a positive matrix is positive, the partial transpose is in general not. To describe a physically realizable operation, the map images has to satisfy therefore in addition to 5.2 and 5.3 the condition

5.4 equation

and because Bob's system can be arbitrary, this should hold for any dimension of images . Let us summarize the discussion up to now in the following definition:

Scheme for Channels applied to subsystems even if the overall system is in an entangled state.

Figure 5.2 Channels can be applied to subsystems even if the overall system is in an entangled state.

A channel is represented in the Schrödinger picture by a trace‐preserving cp‐map. To get the Heisenberg picture representation, we have to introduce the dual of images . It is the map images uniquely defined by

5.5 equation

It is easy to see that images is completely positive, if images is, and that images is unital if images is trace preserving (Problem 5.1). If images is an effect, that is, an operator with images , representing a yes/no measurement2, its image images is an effect as well (since images is positive and unital). It should be regarded as the effect we get, if we first apply the channel images to the system and then measure the effect images (cf. Figure 5.3). Some typical examples of channels are given as follows:

  • Unitary time evolution. The most simple example is time evolution, described by a unitary operator on images . The corresponding channel is described by images .
  • Expansion. Another elementary example arises if we expand a given quantum system by a second one (described by the Hilbert spaces images and images , respectively). Hence, initial and target Hilbert spaces are images and images , and if the images ‐system is in the state images the channel images becomes images .
  • Restriction. The inverse operation arises, if we discard a subsystem; that is, the initial Hilbert space is now images , the target Hilbert space is images and images is given by images , where images denotes the partial trace over images .
  • Noisy time evolution. The composition of channels is again a channel. Hence, we can combine the three examples just given: First, expand the system, then let it evolve unitarily, and finally discard the system added in the first step: images and
    5.6 equation

    with a unitary images on images and a density matrix images on images . Physically, this type of channel describes the influence of noise caused by interaction with the environment (represented by the images ‐system): images is the initial state of the environment and images represents the joint evolution of the system and the environment; cf. Figure 5.4. We will see in Section 5.4 that each channel can be written this way.

Image described by caption and surrounding text.

Figure 5.3 If images is an effect (i.e. a yes/no measurement) and images a channel, we can construct an effect images by first applying the channel images to the system and then performing an images measurement. The map images represents the channel images in the Heisenberg picture.

Scheme for noisy channel arising from interaction with the environment.

Figure 5.4 A noisy channel arising from interaction with the environment.

5.3 The Choi–Jamiolkowski Isomorphism

The subject of this section is a relation between completely positive maps and states of bipartite systems first discovered by Choi (1) and Jamiolkowski (2), which is very useful in establishing several fundamental properties of cp‐maps.

The idea is based on the setup already discussed in Figure 5.2: Alice and Bob share a bipartite system in a maximally entangled state

5.7 equation

(where images denotes an orthonormal basis of images ) and Alice applies to her subsystem a channel images while Bob does nothing. At the end of the processing, the overall system ends up in a state

5.8 equation

Mathematically, Eq. 5.8 makes sense, if images is only linear but not necessarily positive or completely positive (but then images isn't positive either). If we denote the space of all linear maps from images into images by images we therefore get a map

5.9 equation

which is easily shown to be linear (i.e., images for all images and all images ). Furthermore, this map is bijective, hence a linear isomorphism.

The proof of this theorem is left as an exercise to the reader (Problem 5.2). From the definition of images in Eq. 5.8, it is obvious that images is positive, if images is completely positive. To see that the converse is also true, is not as trivial, because a transposition (which is not completely positive) is involved in the definition of images 5.11. It is therefore useful to rewrite Eq. 5.11 in terms of a purification of images . Hence, consider an auxiliary Hilbert space images and images such that images . Note that the existence of such a images requires positivity of images , but not normalization. If images denotes an orthonormal basis, we can define an operator images by

5.12 equation

Now we get with Eq. 5.11

5.13 equation
5.14 equation
5.15 equation

Let us summarize this result for later reference in the following lemma.

Note that the definition of images in terms of images from Eq. 5.12 depends on the choice of the basis images but not on the images (the images are fixed already by the choice of images in 5.8). However, this ambiguity does not affect the expression images , because all operators images arising from different bases in images are related by unitary operators on images .

From Eq. 5.16, we see immediately that images is completely positive, if images is positive. Together with Theorem 5.2 this leads to

As an immediate consequence of this theorem, we can simplify the original characterization of complete positivity in Definition 5.1. To this end, let us define for each images a map images to be images positive if images is positive (where images denotes as in Definition 5.1 the identity on images ). Note that this is in general a weaker condition than complete positivity, because images is completely positive, if images is images ‐positive for each images . In the finite‐dimensional case, however, it is sufficient to have images ‐positivity for sufficiently large images .

5.4 The Stinespring Dilation Theorem

At the end of section 5.2, we have claimed that each channel can be written in terms of an ancilla as in Eq. 5.6. We are now prepared to prove this statement. The following theorem, which goes back to Stinespring (3), is the central structure theorem about completely positive maps.

Let us consider now the uniqueness of Stinespring representations. Obviously, we can always enlarge the dilation space images by adding extra dimensions (i.e., replacing images by images and leaving images untouched). Hence, Stinespring representations are not unique. But what happens, if we assume that images is “as small as possible,” that is, if the dimension of images cannot be reduced by discarding “superfluous” components? This situation is characterized by the condition

5.18 equation

Now we have the following theorem:

This lemma shows that we get a minimal Stinespring representation if we define images and images in terms of 5.19 with a minimal purification images of images . Its uniqueness follows from the uniqueness (up to unitary equivalence) of the minimal purification.

Let us consider now two alternative representation theorems, which can be derived directly from the Stinespring Theorem. The first is the ancilla form of a channel, which we have encountered already in Eq. 5.6.

Even if the Stinespring representation images used in the proof is the minimal one, there is a lot of freedom to define the unitary images , because it depends on the choice of images and of many matrix elements, which in the end drop out of all results. This is a disadvantage of the ancilla approach in practical computations.

Let us come back now to a general (i.e., not necessarily trace preserving) cp‐map images and consider a Stinespring representation images of it. If we choose vectors images with images we can define a family of operators images by

5.27 equation

In terms of these operators, Eq. 5.17 can be rewritten as follows (cf. Problem 5.4 and ( 1,4)).

Finally, let us state a third result, which is closely related to the Stinespring theorem. It characterizes all decompositions of a given completely positive map into completely positive summands. It shows in particular that all “Kraus representations” of a given cp‐map (i.e., Eq. 5.28 with appropriate operators images ) can be derived as in 5.27. By analogy with results from measure theory, we will call it a Radon–Nikodym theorem (cf. (5))

The properties of completely positive maps we have just discussed are only the most elementary ones. For a much more complete, in‐depth presentation of this subsection, we would like to refer the reader to the book of Paulsen (6).

5.5 Classical Systems as a Special Case

Up to now we have only treated pure quantum systems, for which the possible observables are given by all bounded operators on a Hilbert space. Classical systems can be understood as a special case, with a constraint on what we can measure: namely only those observables, which are diagonal in some fixed basis. Since diagonal matrices commute, this is the same as choosing a commutative subalgebra of observables.

The transition from a quantum system to a classical subdescription is made by a particular channel images , which simply kills all off‐diagonal terms, sometimes called “interference terms.” When images denotes the particular orthonormal basis in which we want to go classical, we set

5.40 equation

This is also called a complete von Neumann measurement: the images term in this sum is the corresponding basis state, multiplied with the probability images for obtaining the result images . It is easily verified that the formula for the Heisenberg picture of this channel is exactly the same as 5.40.

Clearly, the specification of elements images in the classical observable algebra require only images rather than images real parameters, as in the quantum case. Therefore, channels with one classical input or output can also be described by fewer parameters. For example, a channel images with classical input has the property that images : its output depends only on the diagonal matrix elements of the input matrix. Hence, it can be written as

5.41 equation

where the images are arbitrary states of the final system, which characterize images . The input state merely selects the weights in a convex combination of these states. Dually, channels images with classical output are of the form

5.42 equation

where the images are positive operators adding up to the identity operator. Thus, images is an observable, or positive operator‐valued measure.

An important special case is also the channels whose output is the tensor product of a classical and a quantum output. If images , images is classical basis in images , the general form of such a channel is images , with

5.43 equation

where each of the images is completely positive. Such a channel is called an instrument (7). Since there are two outputs, we get two “marginals,” that is, the channels obtained by ignoring either output: If we do not look at the quantum output, we get an observable images in the sense of 5.42 by images . On the other hand, if we do not select according to the results images , we get the channel images .

5.6 Channels with Memory

During a realistic communication process, the same channel is used many times in succession, which raises the question in which way each invocation can depend on the previous ones. A mathematical analysis of this problem leads to the concept of a channel with memory, which is described below. Since this is a very large field, we can only give a very brief overview. A detailed discussion can be found in (8) and (9) and the references therein.

The most simple case is a memoryless channel transmitting images ‐level systems. It is described by a trace‐preserving cp‐map images with images . If Alice uses images to send images systems in the joint state images to Bob, images is invoked images ‐times independently. The images invocation is given by the tensor product

5.44 equation

such that the overall operation becomes the concatenation

5.45 equation

Hence, Bob receives the output systems in the joint state images . This is the appropriate model for situations where memory effects are not present or can be ignored.

If in contrast memory effects have to be taken into account, we have to replace images by a trace‐preserving cp‐map

5.46 equation

Here, images is a (finite‐dimensional) Hilbert space, which describes the memory, and images is called a channel with memory. If Alice transmits one system in a state images with the memory in the initial state images , the output system received by Bob is in general correlated with the memory and the joint output state is images . If Bob is not interested in the memory (or cannot access it), we have to trace images away such that the real output state becomes images . To send an images ‐fold system in the state images , we have to invoke the channel images times in succession. The images invocation is again a tensor product, but now the memory has to be taken into account such that we get

5.47 equation

which is a map of the form

5.48 equation

Note that the images factor is shifted here from the images to the images position. This allows us to write the overall operation as in 5.45 as a concatenation

5.49 equation
5.50 equation

If the memory is ignored at the end, Bob receives the images ‐fold system as above in the final state images . Note that in contrast to images we cannot write images as a tensor product and even if the input state images is a product state the output state in general is not.

The scheme just constructed describes a channel that can act on an arbitrary number images of systems (via the concatenations images ). Furthermore, it satisfies the natural causality condition that the images invocation depends on the images previous ones but not on the images that will take place in the future. It can be shown that any channel that is causal in this way can be written as a concatenation images of a memory channel images ; cf. (9).

Let us change our point of view now slightly and look at the final state images of the memory while the transmitted system is ignored, that is, images is given by

5.51 equation

The interesting question is how much information about the initial state images is still contained in images . The most extreme case arises if for some images (and therefore for all images , as well) it does not depend on images at all. Channels of this type are called forgetful (since after at least images invocations the initial state is completely “forgotten”). A simple example for a forgetful channel is the “shift Channel” given by (with images )

5.52 equation

which exchanges the input with the memory (note the flipped positions of the Hilbert spaces at the output side). Hence, the memory is completely overridden after only one invocation. In contrast to this, the identity channel images (taking again into account the flipped Hilbert spaces) is not forgetful, since the memory is passed unchanged. Forgetful channels play a special role since they can be treated in many respects (in particular if channel capacities are discussed) in the same way as memoryless channels.

5.7 Examples

5.7.1 The Ideal Quantum Channel

The simplest possible channel is the description of “doing nothing” to a system of type images , denoted above by images , that is, the identity map on images . This is the channel that we try to achieve when we talk about the transmission of quantum information. All practical ways of sending quantum information introduce noise, which is the same as saying that they are described by channels images . However, by suitable steps of quantum error correction (applied to multiple instances of images ), we can reduce the noise and, in the limit, get a better realization of images .

It is easy to construct the minimal Stinespring dilation of images : We take images , so that images , and images . This simple observation, combined with the Radon–Nikodym Theorem has a very profound consequence, namely that in quantum mechanics there is no measurement without disturbance. Indeed, suppose we have an instrument as in Eq. 5.43, such that the overall state change is images . That is to say, if we perform any further measurements after the measurement by images , we will always find the same expectations as if we had not applied images . Then, by the Radon–Nikodym Theorem, all decompositions of images into completely positive summands are parameterized by operators in the dilation space images , which is, however, one dimensional. Therefore, all images must be proportional to images , say images , for some probability distribution images on the outcomes. But then the observable associated to the instrument will be images , which is to say that the probabilities for the outcomes do not depend at all on the input state. Hence, they do not give any information about the system, and it is fair to say that this is not a measurement at all.

5.7.2 Depolarizing Channel

At the opposite extreme is a channel that destroys all input information, replacing it by a completely chaotic output state images , where images . Slightly more generally, we can look at the channel that does this with probability images , and otherwise ideally transmits the input:

5.53 equation

Here, we have included the trace factor (which is 1 for input states), so that images becomes a linear map. This channel is often used as a noise model, usually with a small depolarization probability images . Interestingly, this channel is completely positive even for some images . For qubits, this has a quite intuitive interpretation in terms of transformations of the Poicaré sphere: as images increases, the set of output states shrinks, until at images it coincides with the origin. Increasing images further means that the Poincaré sphere becomes inverted. For images , we would get a complete inversion, the so‐called Universal‐NOT operation, which sends every pure state to its orthogonal complement. This map is positive, but not completely positive, so it is an impossible operation. Its best approximation by completely positive channels is obtained by taking images as large as possible (images for qubits),

The Kraus decompositions of the fully depolarizing channel (images ) are characterized by the equation

5.54 equation

This can be solved for any images by operators images , which are unitary up to a factor. Such orthogonal sets of unitaries play a central role in teleportation and dense coding schemes.

5.7.3 Entanglement Breaking Channels

Can quantum information be transmitted via classical channels? This would mean to first make a measurement images , transmit the results via a classical channel, and to let the receiver try to reconstruct the quantum input state by a repreparation images , which depends on the results of the measurement. The form of such a channel is images , where images is the von Neumann measurement for the intermediate classical channel. When images and images are given as in 5.41 and 5.42, respectively, and the classical signals transmitted are labeled by images , this gives a channel of the form

5.55 equation

It turns out that these channels are characterized by the property that images turns every entangled state into a separable state, that is, they destroy all entanglement (Problem 5.6).

5.7.4 Covariant Channels

Many channels of interest have a simple characterization in terms of symmetries. For example, the depolarizing channels 5.53 are the only ones that do not distinguish any basis in Hilbert space, in the sense that a basis change by a unitary operator images does not change the action of the channel: images . More general characterizations of symmetries involve subgroups of unitary operators, which may differ for initial and target space:

5.56 equation

where images is some abstract group and images and images are unitary representations of this group on the initial and target Hilbert space, respectively. Channels satisfying this condition are called covariant.

Since the minimal Stinespring representation is unique up to unitary equivalence, the covariance of the channel is also reflected at that level, and this often allows us to give concise formulas for all channels satisfying 5.56, given images and the representations. Let images be the Stinespring isometry. Then, for every images , images is again a dilation, which means that this dilation must be connected with images by a unitary of the form images . In other words, we find the condition

5.57 equation

One readily verifies that images must be a unitary representation of images on images . In the language of group representation theory, this relation says that images must be an intertwining operator between the representations of images , and there is a highly developed formalism to determine such operators. Let us consider two cases:

When the group is images , the irreducible representations are labelled by the spin parameter images . Let us take both input and output representations to be irreducible with spin images and images , respectively. This fixes the dimensions to be images and images . Now it is easy to see that decomposing the representation images into irreducibles corresponds to a convex decomposition of images . Therefore, to find the extremal covariant channels, we can assume images to be irreducible, as well, and hence to be fixed by a spin parameter images . Then, the Clebsch–Gordan theory of adding angular momenta tells us that a nonzero intertwiner images exists if and only if images , and images is integer. Moreover, the intertwiner in these cases is a unique isometry, whose matrix elements are the well‐known Clebsch–Gordan coefficients.

For example, when images , images gives the ideal channel. For images we can also define the channel

5.58 equation

where images denotes the angular momentum operators of the spin images representation. This corresponds precisely to images , because the angular momenta are the components of a vector operator, transforming with the spin 1 representation. images gives the depolarizing channel.

Another interesting group for constructing covariant channels are the phase space translations or, more precisely, the Heisenberg group, consisting of the phase space translations and the multiples of images . The phase space displacement by the phase space vector images is then given by the Weyl operators images , and we assume these to act irreducibly, so that there are no further degrees of freedom. By the canonical commutation relations, the Weyl operators are also characterized as the eigenvectors of the action of phase space translations on operators: that is, images for all images implies that images must be proportional to a Weyl operator images , and images contains an exponential factor characterizing images . Inserting this condition into the covariance equation 5.56, one readily finds that (in the Heisenberg picture) a phase space covariant channel must take Weyl operators to multiples of Weyl operators:

5.59 equation

Moreover, images is a channel if and only if images is the Fourier transform of a probability measure, and images acts by making a random phase space translation, selected according to this measure.

The theory applies also, however, when the Weyl systems on the input and output sides are different, and the displacement parameters are connected by some linear map between input and output phase space. For example, we could take images , with some positive factor images . This corresponds to the amplification or attenuation of a quantum optical light field (depending on whether images or images ). In this case, the complete positivity condition for images is a bit more difficult to write down. It forces images to contain some noise, as is expected from the no‐cloning theorem. The ancilla form of the dilation is particularly instructive: any such channel can be represented by coupling an ancillary system in a specified state to the input, making a symplectic transformation (any interaction, which is quadratic in positions and momenta), and then tracing out a part of the system. In particular, when the initial state of the ancilla is Gaussian, the channel is Gaussian as well, which means that the factor images has Gaussian form.

Problems

  1. 5.1 Show that the dual images of a completely positive map images is completely positive and that images is unital iff images is trace preserving.
  2. 5.2 Give a proof of Theorem 5.1.
  3. 5.3 Give a proof of Lemma 5.2. Hint: Assume that Equation 5.18 does not hold and consider a vector images orthogonal to the span of images .
  4. 5.4 Derive the Kraus form (Corollary 5.3) from the Stinespring form (Theorem 5.3).
  5. 5.5 Find a Kraus decomposition for the depolarizing channel.
  6. 5.6 Show that the channels define in Equation 5.55 are entanglement‐breaking, that is, images turns every entangled state into a separable state. Hint: Use the Jamiolkowski isomorphism.

References

  1. 1 Choi, M.‐D. (1975) Completely positive linear maps on complex matrices. Linear Algebra Appl., 10, 285–290.
  2. 2 Jamiolkowski, A. (1972) Linear transformations which preserve trace and positive semidefiniteness of operators. Rep. Math. Phys., 3, 275–278.
  3. 3 Stinespring, W.F. (1955) Positive functions on C*‐algebras. Proc. Am. Math. Soc., 6, 211–216.
  4. 4 Kraus, K. (1983) States Effects and Operations, Springer‐Verlag, Berlin.
  5. 5 Arveson, W. (1969) Subalgebras of C*‐algebras. Acta Math., 123, 141–224.
  6. 6 Paulsen, V.I. (2002) Completely Bounded Maps and Dilations, Cambridge University Press, Cambridge.
  7. 7 Davies, E.B. (1976) Quantum Theory of Open Systems, Academic Press, London.
  8. 8 Caruso, F., Giovannetti, V., Lupo, C., and Mancini, S. (2014) Quantum channels and memory effects. Rev. Mod. Phys., 86, 1203.
  9. 9 Kretschmann, D. and Werner, R.F. (2005) Quantum channels with memory. Phys. Rev. A, 72, 062323.

Notes

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

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