UNIVERSIDAD NACIONAL AUTÓNOMA DE MÉXICO POSGRADO EN CIENCIAS FÍSICAS INSTITUTO DE FÍSICA Divisibility classes of qubit maps and singular Gaussian channels TESIS QUE PARA OPTAR POR EL GRADO DE: DOCTOR EN CIENCIAS (FÍSICA) PRESENTA: DAVID DÁVALOS GONZÁLEZ Director de Tesis: Dr. Carlos Francisco Pineda Zorrilla Instituto de Fı́sica Codirector de Tesis: Dr. Mario Ziman Instituto de Fı́sica, Academia Eslovaca de Ciencias Miembros del comité tutor: Dr. Carlos Francisco Pineda Zorrilla Instituto de Fı́sica Dr. Luis Benet Fernández Instituto de Ciencias Fı́sicas Thomas Henry Seligman Schurch Instituto de Ciencias Fı́sicas Ciudad Universitaria, Ciudad de México, Enero 2020 UNAM – Dirección General de Bibliotecas Tesis Digitales Restricciones de uso DERECHOS RESERVADOS © PROHIBIDA SU REPRODUCCIÓN TOTAL O PARCIAL Todo el material contenido en esta tesis esta protegido por la Ley Federal del Derecho de Autor (LFDA) de los Estados Unidos Mexicanos (México). El uso de imágenes, fragmentos de videos, y demás material que sea objeto de protección de los derechos de autor, será exclusivamente para fines educativos e informativos y deberá citar la fuente donde la obtuvo mencionando el autor o autores. Cualquier uso distinto como el lucro, reproducción, edición o modificación, será perseguido y sancionado por el respectivo titular de los Derechos de Autor. UNIVERSIDAD NACIONAL AUTÓNOMA DE MÉXICO POSGRADO EN CIENCIAS FÍSICAS Divisibility classes of qubit maps and singular Gaussian channels TESIS Que para obtener el grado de: Doctor en Ciencias (Fı́sica) Presenta: David Dávalos González Director de tesis: Dr. Carlos Francisco Pineda Zorrilla Codirector: Dr. Mario Ziman Miembros del Comité Tutoral: Dr. Carlos Pineda, Dr. Luis Benet, y Dr. Thomas H. Seligman México D. F., 2019 UNIVERSIDAD NACIONAL AUTÓNOMA DE MÉXICO POSGRADO EN CIENCIAS FÍSICAS Divisibility classes of qubit maps and singular Gaussian channels THESIS To obtain the degree: Doctor en Ciencias (Fı́sica) Presents: David Dávalos González Director: Dr. Carlos Francisco Pineda Zorrilla Co-director: Dr. Mario Ziman Members of the Tutorial Committee: Dr. Carlos Pineda, Dr. Luis Benet, and Dr. Thomas H. Seligman México D. F., 2019 Truth is ever to be found in simplicity, and not in the multiplicity and confusion of things. Isaac Newton Gracias La idea que nació cuando cursaba la secundarı́a, la de convertirme algún dı́a en cientı́fico, no habrı́a sido posible de no haber nacido en el seno de una familia estable, funcional y de sólidos valores. Por eso les agradezco infinitamente a mis queridos Padres. A mi Madre, Sara, por dedicarme tanto de su tiempo y energı́a, por ser una Madre muy amorosa, llena de valores y por ser la persona mas paciente del mundo. Le agradezco a mi Padre, Juan Manuel, por siempre estar atento a que fuera una persona de principios y un buen ciudadano, por ser un padre amoroso y por darme su confianza. Le agradezco que siempre se haya preocupado por tener una computadora en casa y por ser un entusiasta de la tecnologı́a, eso aportó fundamentalmente a quien soy hoy. A mis hermanas y hermanos por preocuparse por mi y por regalarme tantas veces su tiempo. A mi compañera de vida, a mi esposa Lorena, gracias por tenerme tanta paciencia, por creer en mi y por quer- erme tanto. A mi tutor y amigo, Carlos, le agradezco su paciencia, sus valiosas enseñanzas, su gran apoyo y su amistad. A Thomas Seligman y Luis Benet por siempre apoyarme. Le agradezco a mis amigos Luis Juarez, Arturo Carranza, Thomas Gorin, Mario Ziman, Diego Wisniacki, Ignacio Garcı́a Mata, Juan Diego Urbina, Peter Rapčan, Tomáš Rybár, Edgar Aguilar, David Amaro, Alvaro Dı́az, Miguel Cardona, Chayo Camarena, Antonio Rosado, Sergio Sánchez, Nephtalı́ Garrido, Sergio Pallaleo, Samuel Rosalio. A mis gatitos Lola y Dalı́ por hacerme feliz el poco tiempo que estuvieron en este mundo. A mi querida gatita Lulú por hacer de mi hogar siempre un lugar feliz. Les agradezco a todos los seres queridos que hicieron de esta parte de mi trayectoria algo memorable. Agradezco el apoyo brindado por los proyectos PAPIIT número IG100518 y CONA- CYT CB-285754. SYNOPSIS We present two projects concerning the main part of my PhD work. In the first one we study quantum channels, which are the most general operations mapping quantum states into quantum states, from the point of view of their divisibility properties. We introduced tools to test if a given quantum channel can be imple- mented by a process described by a Lindblad master equation. This in turn defines channels that can be divided in such a way that they form a one-parameter semi- group, thus introducing the most restricted studied divisibility type of this work. Using our results, together the study of other types of divisibility that can be found in the literature, we characterized the space of qubit quantum channels. We found interesting results connecting the concept of entanglement-breaking channel and infinitesimal divisibility. Additionally we proved that infinitely divisible channels are equivalent to the ones that are implementable by one-parameter semigroups, opening this question for more general channel spaces. In the second project we study the functional forms of one-mode Gaussian quantum channels in the position state representation, beyond Gaussian functional forms. We perform a black-box characterization using complete positivity and trace preserving condi- tions, and report the existence of two subsets that do not have a functional Gaus- sian form. The study covers as particular limit the case of singular channels, thus connecting our results with the known classification scheme based on canonical forms. Our full characterization of Gaussian channels without Gaussian func- tional form is completed by showing how Gaussian states are transformed under these operations, and by deriving the conditions for the existence of master equa- tions for the non-singular cases. Keywords: divisibility, qubit channels, open quantum systems. RESUMEN En esta tesis se presentan dos proyectos realizados durante mis estudios de doctor- ado. En el primero se estudian los canales cuánticos, que son las operaciones más generales que transforman estados cuánticos en estados cuánticos, desde el punto de vista de sus propiedades de divisibilidad. Introducimos herramientas para pro- bar si un canal cuántico dado puede ser implementado por un proceso descrito por una ecuación maestra de Lindblad. Ésto a su vez define a los canales que pueden ser divididos de tal manera que ellos forman semigrupos de un parámetro, introduciendo entonces el tipo más restringido de divisibilidad estudiado de este trabajo. Usando nuestros resultados, junto con el estudio de otros tipos de divis- ibilidad que pueden ser encontrados en la literatura, caracterizamos el espacio de canales cuánticos de un qubit. Encontramos resultados interesantes que conectan el concepto de canales que rompen el entrelazamiento (del sistema con cualquier sistema auxiliar) y el de divisibilidad infinitesimal. Además probamos que el conjunto de canales infinitamente divisibles es equivalente al de los canales im- plementables por semigrupos de un parámetro. Ésto abre la pregunta sobre si esto sucede para espacios de canales más generales. En el segundo proyecto estudi- amos las formas funcionales de canales Gaussianos de un solo modo, más allá de la forma funcional Gaussiana. Se hace una caracterización de caja negra uti- lizando las condiciones de completa positividad y preservación de la traza, y se reporta la existencia de dos subconjuntos que no poseen forma funcional Gaus- siana. El estudio cubre en particular el lı́mite de los canales singulares, conectando entonces nuestros resultados con la la clasificación basada en formas canónicas. Nuestra caracterización de canales Gaussianos sin forma funcional Gaussiana es completada mostrando como los estados Gaussianos se transforman bajo esas op- eraciones, ası́ como al derivar las condiciones para la existencia de ecuaciones maestras para los casos no singulares. Contents 1 Introduction 1 2 Open quantum systems and quantum channels 5 2.1 Introduction to the scheme of open quantum systems . . . . . . . 5 2.2 Quantum channels . . . . . . . . . . . . . . . . . . . . . . . . . 15 2.3 Quantum channels of continuous variable systems . . . . . . . . . 23 3 Representations of quantum channels 27 3.1 Kraus representation . . . . . . . . . . . . . . . . . . . . . . . . 27 3.2 Choi-Jamiołkowski representation . . . . . . . . . . . . . . . . . 29 3.3 Operational representations . . . . . . . . . . . . . . . . . . . . . 30 3.4 Qubit channels . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 3.5 Representation of Gaussian quantum channels . . . . . . . . . . . 40 4 Divisibility of quantum channels and dynamical maps 47 4.1 Divisibility of quantum maps . . . . . . . . . . . . . . . . . . . . 47 4.2 Characterization of L-divisibility . . . . . . . . . . . . . . . . . . 56 4.3 Divisibility of unital qubit channels . . . . . . . . . . . . . . . . . 57 4.4 Non-unital qubit channels . . . . . . . . . . . . . . . . . . . . . . 68 4.5 Divisibility transitions and examples with dynamical processes . . 72 5 Singular Gaussian quantum channels 79 5.1 Allowed singular forms . . . . . . . . . . . . . . . . . . . . . . . 79 5.2 Existence of master equations . . . . . . . . . . . . . . . . . . . 84 6 Summary and conclusions 89 7 Appendices 93 A Exact dynamics with Lindblad master equation 95 ix x CONTENTS B On Lorentz normal forms of Choi-Jamiolkowski state 97 C Articles 99 C.1 Article: Divisibility of qubit channels and dynamical maps . . . . 99 C.2 Article: Gaussian quantum channels beyond the Gaussian func- tional form: full characterization of the one-mode case . . . . . . 114 C.3 Article: Positivity and Complete positivity of differentiable quan- tum processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 C.4 Article: Positivity and Complete positivity of differentiable quan- tum processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 Bibliography 148 Chapter 1 Introduction In questions of science, the authority of a thousand is not worth the humble reasoning of a single individual. Galileo Galilei The advent of quantum technologies opens questions aiming for deeper under- standing of the fundamental physics beyond the idealized case of isolated quantum systems. Also the well established Born-Markov approximation used to describe open quantum systems (e.g. relaxation process such as spontaneous decay and decoherence) is of limited use and a more general framework of open system dy- namics is required. Recent efforts in this area have given rise to relatively novel research subjects - non-markovianity and divisibility. A central object of study in quantum information theory and open quantum sys- tems are quantum channels, also called quantum operations. They describe, for instance, the noisy communication between Alice and Bob or the changes that an open quantum system undergoes at some fixed time. They can also be seen as the basic building blocks of time-dependent quantum processes (also called quantum dynamical maps). Conversely, families of quantum channels arise naturally given a quantum dynamical map. Given a quantum channel, for instance an spin flip or the approximation of the universal NOT gate, one can wonder about how it can be implemented. The lat- ter in the sense of, being quantum channels discrete operations, can we find a continuous time-dependent processes that at some time it implements the given channel?, or, does process such that we “just wait for a relaxation of the physical system” implements such channel? It turns that this question is related with the 1 2 Chapter 1. Introduction one of finding simpler operations such that their concatenation equals the given quantum channel [WC08]. Such operations are simpler in the sense that they are closer to the subset of unitary operations, or even “smaller” in the sense that they are closer to the identity channel. This thesis encompasses the results of two works developed during my PhD. The first and the most extended one was devoted to study the divisibility proper- ties of quantum channels (discrete evolutions of quantum systems), for the par- ticular case of qubits. We revise the divisibility types introduced in the seminal paper by Wolf et al. [WC08] and derived several useful relations to decide each type of divisibility. In particular, we characterize channels that can be divided in such a way that they belong to one-parameter semigroups (dynamics described by Lindblad master equations), and extended the analysis of [WECC08] for channels with negative eigenvalues. We did this using the results by Evans et al. [EL77] and Culver [Cul66]. Beyond the mentioned characterization tools, the principal aim of the work was to understand the forms of non-markovianity standing behind the observed quan- tum channels. The non-markovianity character describe the back-action of the system’s environment on the system’s future time evolution. Such phenomena is identified as emergence of memory effects [ARHP14, VSL+11, PGD+16]. On the other side, divisibility questions the possibility of splitting a given quantum channel into a concatenation of other quantum channels. In this work we will in- vestigate the relation between these two notions. Thus, we related features of of continuous time evolutions of quantum systems, and the concept of divisibility of quantum maps, which are discrete evolutions. A very first example of this is the well known identification of one-parameter semigroups with Linbladian dynam- ics [Lin76]. The second project is devoted to representation theory of continuous-variable quantum systems, which is a central topic of study given its role in the descrip- tion of physical systems like the electromagnetic field [CLP07], solids and nano- mechanical systems [AKM14] and atomic ensembles [HSP10]. In this theory the simplest states, both from a theoretical an experimental point of view, are the so- called Gaussian states. An operation that transforms such family of states into itself is called a Gaussian quantum channel (GQC). Even though Gaussian states and channels form small subsets among general states and channels, they have proven to be useful in a variate of tasks such as quantum communication [GVAW+03], quantum computation [LB99] and the study of quantum entanglement in sim- ple [BvL05] and complicated scenarios [LRW+18]. In this project we study the possible functional forms that one-mode Gaussian quantum channels can have 3 in the position state representation, and characterize the particular case of sin- gular channels. Although they are already characterized by their action on the first and second moments of Gaussian states [Hol07, WPGP+12], we connect our framework to such known results. Additionally we give an insight of the possible functional forms of, for instance, Gaussian unitaries. The thesis is organized as follows: In chapter 2 we discuss the most widely adopted scheme to study open quantum systems, introducing the formalism of bipartite systems and useful tools for it. Later on we present the general setting for system plus reservoir dynamics and its formal solution. As a paradigmatic ex- ample of open system dynamics, we present briefly the microscopic derivation of the Lindblad master equation using the well known Born-Markov approximation, and discuss the properties of the generator of the dynamics. Subsequently we in- troduce the formalism of quantum channels, being the most general operations over quantum systems (excluding post-selection), by introducing some useful mathematical definitions and contrasting with its classical analog. Additionally we discuss briefly the concept of local operations and classical communications (LOCC), also known as filtering operations. Finally we give a very brief intro- duction to continuous variable systems, giving special attention to Gaussian states and channels. In chapter 3 we discuss the different available representations for quantum chan- nels and their relation with the concept of complete positivity. In particular we in- troduce the well known Kraus representation and discuss the Choi-Jamiołkowski theorem which in turn defines a very useful representation to study quantum channels and their divisibility properties. Later on we introduce various matrix representations of quantum channels, paying special attention to hermitian and traceless bases types (without taking into account the component proportional to identity). Furthermore we introduce useful decompositions of qubit channels into unitary conjugations and one-way stochastic local operations, and classical com- munication, both being analogous to the well known singular value decomposi- tion. Finally we give an introduction to representations of Gaussian channels and a detailed derivation of the position-state representations for Gaussian channels without Gaussian functional form. In chapter 4 we give the definition of divisible quantum channel, as well the def- inition of various subclasses of divisible channels concerning additional proper- ties. In particular we discuss the concepts of infinitesimal and infinitely divisible channels and some relations and inclusions between them. Among infinitesimal divisible channels we identify two subclasses, being the set of infinitesimal divis- ible channels in complete positive and positive (but not complete positive) maps. 4 Chapter 1. Introduction Later on we introduce the concept of L-divisible channels, defining the set of channels which are members of one-parameter semigroups. We show that the set of infinitely divisible channels is the same of the L-divisible Pauli channels. In chapter 5 we study one-mode Gaussian quantum channels in continuous-variable systems by performing a black-box characterization using complete positivity and trace preserving conditions, and report the existence of two subsets that do not have a functional Gaussian form. Our study covers as particular limit the case of singular channels, thus connecting our results with their known classification scheme based on canonical forms. Our full characterization of Gaussian channels without Gaussian functional form is completed by showing how Gaussian states are transformed under these operations, and by deriving the conditions for the existence of master equations for the non-singular cases. In chapter 6 we give a summary of the two projects introduced in this work and conclusions. Finally, in the appendix A we prove that the exact reduced dynamics of an open quantum system never follow a Lindblad master equation unless they are unitary, given a bounded global Hamiltonian. In appendix B we give an example that shows that the set of Lorentz normal forms introduced in the literature, is incom- plete. In the appendix C we attach copies of the articles produced during my PhD. Chapter 2 Open quantum systems and quantum channels When we talk mathematics, we may be discussing a secondary language built on the primary language of the nervous system. John Von Neumann In this chapter we introduce the usual scheme to study open quantum systems, the widely known Born-Markov approximation and the concept of CP-divisibility. Later on and based on the idea of (classical) stochastic map, we discuss the ax- iomatic formulation of quantum channels and its connection with the usual con- struction of open quantum systems. Finally, for continuous variable systems, we discuss the paradigmatic example of Gaussian channels. 2.1 Introduction to the scheme of open quantum systems The most widely used scheme to study open quantum systems is based on the idea of study a closed system composed by the central system and its environment, see fig. 2.1 for an schematic explanation. Thus, concepts as bipartite Hilbert spaces, density matrix and partial trace are useful tools to study open systems. In what follows we give a brief review of them. Bipartite Hilbert space. Consider a bipartite closed quantum system described by a Hilbert space with the structure H = HS ⊗HE, where HS is the Hilbert 5 6 Chapter 2. Open quantum systems and quantum channels space of the open system and HE is the Hilbert space of the environment. If {|φ S i 〉} dim(HS) i=1 and {|φ E i 〉} dim(HE ) i=1 are basis for the spaces HS and HE, respec- tively, a basis for H is simply {|φ S i 〉⊗ |φ E j 〉} dim(HS),dim(HE) i=1, j=1 . It is typical that for finite dimensional systems one has that dim(HE)≫ dim(HS) as the environment is usually “bigger” than the central system. To describe the states of open quantum systems it is necessary to model the igno- rance that the observer has with respect to the open system. Since the experimen- talist cannot access the degrees of freedom of the environment, they are simply ignored. To do this we need the two following concepts. Density matrix. Let a quantum system that has probability pi to be in the state |φi〉, and let the operator A an observable over such system. Using the average formula 〈A〉= ∑i pi〈φi|A|φi〉 it is straightforward to show that 〈A〉= tr(Aρ) with ρ = ∑ i pi|φi〉〈φi|, (2.1) and ∑i pi = 1. ρ is called density operator or density matrix. Note that ρ is a positive-semidefinite operator given that pi ≥ 0, and the states |φi〉 don’t need to be orthogonal. Also note that since ρ is hermitian, together with positive- semidefiniteness, implies that we can always write any density matrix as in eq. (2.1) with the states |φi〉 being orthogonal. Thus, every operator ρ acting on a Hilbert space H , fulfilling ρ ≥ 0, ρ = ρ† and tr(ρ) = 1 is a density matrix. The set of density matrices will be denoted along this work as S (H ). Comparing the notion of density matrices with the notion of state vectors in the Hilbert space |ψ〉 ∈ H , density matrices describe physical systems where the observer has an incomplete knowledge of the system’s state. Thus, while state vectors are naturally equipped with intrinsic or quantum probabilities, density operators are additionally equipped with classical probabilities. The density ma- trices enjoying the form ρ = |ψ〉〈ψ|, or equivalently ρ2 = ρ , i.e. projectors, are pure states. It is clear that in this case the system is prepared in the state |ψ〉 with probability one. A useful quantity to characterize quantum states is the purity, defined as P(ρ) = tr ( ρ2 ) . (2.2) It ranges from dim(H )−1 to 1; 1 is obtained for pure states and dim(H )−1 for the complete mixture ✶/dim(H ). 2.1. Introduction to the scheme of open quantum systems 7 Additionally the set S is convex, i.e. any convex combination of density matrices is another density matrix, in the same way as classical distributions do. In fact, mixed states (P(ρ) < 1) can be written always as convex combinations of pure states, see eq. (2.1). Furthermore the set S (H ) is a subset of the bigger set of trace-class operators, T (H ), defined as the ones containing operators with finite trace norm. The latter is defined as |∆|tr = tr √ A†A. This set is in turn a subset of the set of bounded operators B(H ), containing operators with finite operator norm, defined as |A|op = sup|ψ〉 |A|ψ〉|, where |A|ψ〉| = √ (〈Aψ|Aψ〉), i.e. the standard Hilbert space norm, with normalized vectors |ψ〉. It is worth to note that for the finite dimensional, bounded operators always have finite trace norm and vice versa, thus T (H ) = B(H ). But the identification of such sets is relevant for infinite dimensional systems, where counter-examples of the non-equivalence of such sets exist [HZ12]. Additionally B(H ) is the dual space of T (H ) under the Hilbert-Schmidt product, defined as 〈A,B〉 = tr(A†B) [Hol01]. Now, to ignore the degrees of freedom of the unaccessible part of the system, we have to perform an operation in a very analogous way as computing marginal distributions in classical probability theory. For density operators this introduces the concept of partial trace. Partial trace. Let ρ ∈ S (HA ⊗HB) and HA,B the Hilbert spaces of systems A and B. Thus, ρ describes a state of a bipartite system composed by A and B. If we want to know the state of the system A alone, one performs a partial trace over B defined as ρA = trB(ρAB) = dB ∑ i=1 ( ✶⊗〈φ B i | ) ρAB ( ✶⊗|φ B i 〉 ) , where {|φ B i 〉}dB i=1 is a complete orthonormal basis on HB. The resulting operator ρA is a density matrix describing the state of the system A alone. It is trivial to show that it is a density operator. A similar formula holds for ρB. An alternative definition is trB (A⊗B) = A tr(B) plus linearity. In general for composite systems, in a pure state, knowing the reduced states (for instance for bipartite systems, ρA and ρB) is in general not enough to know the whole state of a system. This captures the non-local nature of quantum corre- lations, demanding simultaneous measurements on both parts of the system. In such case we say that the subsystems A and B are entangled. To see this, consider the example of the Bell state |Ω〉= 1/ √ 2(|00〉+ |11〉), where {|0〉, |1〉} is an or- thogonal basis of a qubit system. It is trivial to show that |Ω〉 cannot be written 8 Chapter 2. Open quantum systems and quantum channels as |φ〉 ⊗ |ψ〉, a factorizable state, prohibiting the observer to know the state of the whole system only by non-simultaneous measurements on A and B (described by reduced density matrices). In fact it is easy to show that ρA,B = ✶/2 are the reduced density matrices, appearing also when the total state is ρAB = ✶/4. For composite systems in mixed states the situation is quite different. In this case si- multaneous measurements are needed to access classical correlations. To see this consider the state ρAB = ∑ i piρ i A ⊗ρ i B, (2.3) being a convex combination of factorizable mixed states. This state is a mixed separable state [HHHH09], i.e. subsystems A and B are not entangled. Notice now that performing only local non-simultaneous measurements, the accessible reduced states are ρ ′ A,B = ∑ piρ i A,B. This state also arise when the total system is in the factorizable state ρ ′ A ⊗ρ ′ B. Therefore local simultaneous measurements are needed. 2.1.1 System plus reservoir dynamics The most widely used scheme to study open quantum systems is to consider a bipartite system, where the central system S, is interacting with its environment, E. The full system S+E undergoes a closed system evolution, i.e. Hamiltonian dynamics, see fig. 2.1. The total Hamiltonian H, describing the whole system, has the following general structure H = HS +HE +V, (2.4) where HS,E are the free Hamiltonians of the central system and the environment, respectively, and V is the interaction Hamiltonian among them. Now let ρSE(0) be the state of the total system at the time t = 0. Thus, the state of the system S at the time t is simply: ρS(t) = trE ( U(t)ρSE(0)U †(t) ) , (2.5) where U(t) = e−iHt (taking h̄ = 1) and trE is the partial trace over the environ- mental degrees of freedom. Note that for a general initial state ρSE(0), where one allows classical and quantum correlations, ρSE(t) depends on general on ini- tial information about the environment and its correlations with the central sys- tem S. Thus, to compute the dynamics of the central system such that we end up to universal reduced dynamics, i.e. the same for every initial state and inde- pendent of the initial information in the environment, we take a factorized initial 2.1. Introduction to the scheme of open quantum systems 9 S+E |ψ〉 ρS = trE |ψ〉〈ψ| Figure 2.1: Diagram of the scheme to study open quantum systems. The let- ters S and E state for the open (or central) system and environment parts of the total closed system, S+E. The latter is described (typically) by a pure state |ψ〉 ∈ HS ⊗HE and the central system is described by the reduced state computed using the partial trace over the environmental degrees of freedom, see main text. state ρSE(0) = ρS(0)⊗ ρE [BP07, RH12]. We don’t write explicitly the time- dependence of the environmental state since one is not usually interested on its evolution. With the choice of a factorizable total initial state and using equa- tion eq. (2.5), we have the following expression for the evolution of the central system, ρS(t) = trE [ U(t)(ρS(0)⊗ρE)U†(t) ] . (2.6) Therefore we have that the dynamics over S only depend on the total Hamiltonian H and the environmental initial state ρE , whereas ρS(t) depends only on its initial condition. Hence the equation eq. (2.6) defines a dynamical map, Et , parametrized by t. . Thus, we have Et [ρ(0)] = trE [ U(t)(ρS(0)⊗ρE)U†(t) ] . (2.7) Such map possesses all the information concerning the dynamics of the system S, thus knowing Et one can know entirely the evolution of the system S. The map 10 Chapter 2. Open quantum systems and quantum channels ρS(0)⊗ρE ρSE(t) ρS(0) ρS(t) trE(·) U(t)·U†(t) trE(·) Et Figure 2.2: Scheme of the equivalences between the concept of dynamical map and the theory of open quantum systems. Et can be obtained numerically or experimentally (depending on the context) by measuring only the system S by quantum process tomography [NC11]. In fig. 2.2 we present a schematic description of the two equivalent schemes under which the system S evolves, and their connection throughout trE. Eq. (2.7) can be reduced, by writing ρE = ∑ j pE j |φ E j 〉〈φ E j |, in the following way, Et [ρS(0)] = ∑ i, j K(t)i, jρS(0)K(t)† i, j, (2.8) where the operators K(t)i j = √ pE j 〈φ E i |U(t)|φ E j 〉 are called Kraus operators and act upon the system S alone [RH12]. The expression of eq. (2.8) is called sum represention, also called Kraus representation of the map Et , this will be retaken on chapter 3. Now let us discuss the differential equation for the density matrix of an open quan- tum system. The total state of the system evolves according to the Von Neumann equation [BP07], dρSE dt =−i[H,ρSE], (2.9) which is the analog of the Liouville equation describing the evolution of a classical distribution in the phase space. Taking the partial trace on both sides of eq. (2.9) one arrives to the following: dρS dt =−i tr[H,ρSE] = Lt [ρS], (2.10) where Lt is the generator of the master equation of the system S. Integrating time in both sides from τ = 0 to τ = t, we arrive to the equivalent integral equation: ρS(t) = ρS(0)+ ∫ t 0 dτLτ [ρS(t)]. (2.11) 2.1. Introduction to the scheme of open quantum systems 11 To compute the formal solution of this equation, we use the method of succes- sive approximations. This consists on substituting the whole expression for ρS(t) defined by the right hand side of eq. (2.11). A first iteration leads to ρS(t) = ρS(0)+ ∫ t 0 dτ1Lτ1 [ρS(0)]+ ∫ t 0 dτ1 ∫ t 0 dτ2Lτ1 [Lτ2 [ρS(t)]]. (2.12) Repeating this procedure infinite times, i.e. substituting ρS(t) defined by the right hand side of the last equation in its second integrand several times, we arrive to a power series solution for ρS(t) (powers of Lt). This leads to the well known Dyson series for Lt . Compactly, ρ(t) =~Texp ( ∫ t 0 dsLs ) ρ(0) (2.13) with~T the time-ordering operator, defined as ~T[H(τ1)H(τ2)] = θ(τ1 − τ2)H(τ1)H(τ2)+θ(τ2 − τ1)H(τ2)H(τ1), with θ(x) the Heaviside step function. Eq. (2.13) constitutes the formal solution to the Von Neumann equation with generator Lt , and we can easily identify Et = ~Texp ( ∫ t 0 dsLs ) . 2.1.2 Born-Markov approach: microscopic derivation In general the form of the generator Lt , given a global Hamiltonian, can be quite involved [BP07], but in the limit of weak coupling and short memory we can perform the very widely known Born-Markov approximation. A brief discussion is presented in this subsection. The Born-Markov approximation leads to the Lindblad master equation. We will briefly overview its usual textbook derivation. The first step is to use the inter- action picture, hence the total Hamiltonian becomes HI(t) = eiH0tHe−iH0t , where H0 = HS +HE is the free Hamiltonian. Assuming that the dimension of HE is big compared with the dimension of HS, the weak coupling limit leads to negligible changes in the environmental state. Thus, at time t we can approximate ρSE(t)≈ ρS(t)⊗ρE . In other words, the state of the total system is left always approximately uncor- related, while the state of the environment is never updated. Therefore the envi- ronment forgets any information about the central system, while the state of the 12 Chapter 2. Open quantum systems and quantum channels latter undergoes a non-trivial evolution. Additionally to simplify the derivation we choose ρE a stationary state of HE, i.e. [HE,ρE] = 0 [RH12]. ρE is typically cho- sen as a thermal state of the environmental Hamiltonian, ρE ∝ exp(−βHE), with β = 1/(kBT ), kB the Boltzmann constant and T the environment temperature. Now, in the interaction picture the Von Neumann equation becomes dρS dt =−i trE[VI(t),ρS], (2.14) where VI(t) = eiH0tVe−iH0t and the state ρS(t) are now written in the interaction picture. Inserting ρS(t) from its integral equation eq. (2.11) in the differential equation (2.14) and assuming trE[VI(t),ρS ⊗ρE] = 0 [BP07], we obtain dρS dt =− ∫ t 0 dτ trE[VI(t), [VI(τ),ρS(τ)⊗ρE]]. (2.15) If we assume that the dynamics of the state of the central system does not depend on its past, we can change ρS(τ) to ρS(t), this is called the Markovian approxi- mation. Additionally doing the variable change τ ′ = t − τ , we arrive to dρS dt =− ∫ t 0 dτ ′ trE[VI(t), [VI(t − τ ′),ρS(t)⊗ρE]], (2.16) this equation is known as Redfield equation [Red65] and it is local in time [BP07]. Assuming that the time scale on which the central system varies appreciably is much larger than the time on which the correlations of the environment decay (say τE), the integrand decays to zero rapidly for τ ′ ≫ τE. Then we can safely replace t by ∞ in the integrand limits, obtaining dρS dt =− ∫ ∞ 0 dτ ′ trE[VI(t), [VI(t − τ ′),ρS(t)⊗ρE]]. (2.17) Up to this point, eq. (2.17) has in general fast oscillating terms coming from the explicit dependence on VI(t), this in turn can bring a generator that leads to a quantum process that violates complete positivity [ARHP14, RH12]. In order to get rid of such fast oscillations, one uses the aforementioned assumption that the environment is initialized in a stationary state, and perform the so called secular approximation [ARHP14]. A detailed derivation is outside of the scope of this thesis, but it can be consulted on references [BP07, RH12]. After performing the Markov, Born and secular approximations and changing back to the Schrödinger 2.1. Introduction to the scheme of open quantum systems 13 picture, the resulting master equation can be written in the following forms dρS dt = i[ρS, H̃S]+ d2 S−1 ∑ i, j=1 Gi j ( FiρSF † j − 1 2 {F † j Fi,ρS} ) , (2.18) = i[ρS, H̃S]+ d2 S−1 ∑ j=1 γ j ( A jρSA † j − 1 2 {A † jA j,ρS} ) , (2.19) = L[ρS]. (2.20) Fj ( j = 0, · · · ,d2 S − 1) are operators acting on the central system that addition- ally form an orthonormal basis under Hilbert-Schmidt inner product, such that F0 = ✶/ √ dS and trFj = 0 ∀ j > 0 (this will be revised in subsection 3.3.1); the matrix G is called dissipator matrix. In the second inequality we have used the singular value decomposition of matrix G, thus operators Ai are linear combina- tions of Fi. The scalars γ j > 0 are called relaxation rates and the operator H̃S is the shifted free Hamiltonian of the central system. The first term on both equa- tions, the commutator, is called Hamiltonian part, while the second, the super- operator defined with the summations, is called dissipator. Note that if γ j = 0 ∀ j (uncoupled limit), one recovers the Hamiltonian dynamics over the system S. The operator L is called Lindblad operator or Lindbladian and eq. (2.20) is called Lindblad master equation. We will use along the work the notation L for Lindblad operators. Note that L is independent of time, hence the formal solution of the master equa- tion eq. (2.20) equation is simply the exponentiation of L [see eq. (2.13)], i.e. ρS(t) = eLtρS(0). (2.21) Therefore the dynamics is homogeneous in time and, together with the fact that Et = exp(Lt), we have Et+s = EtEs, i.e. the quantum process Et resulting from a Lindblad master equation forms a one-parameter semigroup. In fact, Lindblad has proven the converse, including the case of infinite dimension [Lin76]: Theorem 1 (One-parameter quantum semigroups). Let Et with E0 = id and t ≥ 0 a quantum process, it is a one-parameter quantum semigroup if and only if it has a generator with the form presented in eq. (2.20). It is worth to point out that starting from global dynamics governed by a bounded Hamiltonian, the reduced dynamics are never of Lindblad form. This can be stated as the following, 14 Chapter 2. Open quantum systems and quantum channels Theorem 2 (Exact dynamics with Lindblad master equation). Let Et = etL a quan- tum process generated by a Lindblad operator L. The equation Et [ρ] = trE [ e−iHt (ρ ⊗ρE)eiHt ] , where H has finite dimension, holds if and only if Et is an unitary conjugation for every t. A proof made jointly with Sergey Filippov is given in the appendix A. It was made using an specific matrix representation for operators that will be introduced in the next chapter. But a more general proof can be found in [Exn85]. Let us point out that this is not the case for global unbounded Hamiltonians, they can lead to Lindblad master equations for the reduced dynamics. This is shown below together other illustrative examples. Examples. To illustrate Lindblad dynamics we present several examples. The first one, depolarizing dynamics, is constructed via a continuous and monotonic contraction of the Bloch sphere. The second one corresponds to a system for which the exact reduced dynamics have Lindblad generator. Example 1 (Dephasing dynamics). Let ρ(0) = ( ρ00 ρ01 ρ∗ 01 ρ11 ) be the initial state, written in a basis called decoherence basis, of a system that undergoes depolar- izing dynamics. This is, only coherence terms (in this basis) are modified in the following way: Et : ρ(0) 7→ ( ρ00 ρ01e−γt ρ∗ 01e−γt ρ11 ) =: ρ(t), with γ > 0. It is trivial to check that Et is a one-parameter semigroup with E0 = id. For t → ∞, we get ρ(0) → diag(ρ00,ρ11). For this process it is easy to prove, by taking 0 < t ≪ 1, that its generator is L[ρ] = γ/2(σzρσz −ρ), which has Lindblad form. It has null Hamiltonian part and only one operator A0 = σz and one relaxation ration, γ/2. Example 2 (Dynamics from global unbounded Hamiltonian). Consider a bipar- tite system composed by a qubit interacting with a particle in a line, with global Hamiltonian H = σz⊗ x̂, where x̂ is position operator. Notice that H is unbounded since the configuration space of the particle is the entire real line. Initializing the environment in the state |ψ〉 with 〈x|ψ〉= √ γ π 1 x+ iγ , 2.2. Quantum channels 15 it can be shown that the exact reduced dynamics for the qubit, without any ap- proximation, is L[ρ] = γ/2(σzρσz −ρ) [AHFB15]. The same generator as in the first example. 2.2 Quantum channels In this section we give a brief introduction to classical stochastic processes, this motivates the definition of quantum channel. We first give an overview of stochas- tic processes; based on this we review the construction steps of quantum channels and discuss several of their properties. Additionally we introduce the simplest example of local operations and classical communication. Later on one we dis- cuss the definition of CP-divisible processes based on the definition of classical Markovianity. Finally we give a brief revision of Gaussian quantum states and channels. 2.2.1 A classical analog The classical analog of quantum channels are the widely known stochastic ma- trices or stochastic maps which propagate classical probability distributions. To introduce them consider, for sake of simplicity, a finite dimensional stochastic system whose state xt (at time t) is described by the probability distribution (or probability vector) ~p(t), i.e. xt ∼ ~p(t) [with ∑i pi(t) = 1 and pi(t)≥ 0]. Note that probability vectors form a convex space in the very same way that density matri- ces do. The distribution ~p(t) is the classical analogous object to density matrices. They serve as the tool to model the accessible information of the observer about the state of the classical stochastic system. Consider now the most general linear transformation on probability vectors that takes, for instance ~p(0) to ~p(t) and let us write it explicitly as a matrix multi- plication, ~p(t) = Λ(t,0)~p(0). We have to impose further constrictions over Λ(t,0) in order to preserve the normalization of ~p(t) and the non-negativity of its el- ements. Since pi(t) = ∑ j ( Λ(t,0) ) i j ~p j(0), simple algebra leads us to note that ∑i ( Λ(t,0) ) i j = 1 ∀ j and ( Λ(t,0) ) i j ≥ 0. Matrices that fulfill these conditions are widely known as stochastic matrices, and form a convex set following the con- vexity of the space of probability distributions. A remarkable property of stochastic maps is that they are contractive with respect to the Kolmogorov distance. 16 Chapter 2. Open quantum systems and quantum channels Theorem 3 (Contractivity of stochastic maps). The matrix Λ is a stochastic matrix if and only if DK (Λ~p,Λ~q)≤ DK (~p,~q) , (2.22) where DK (~p,~q) = ∑k |pk −qk| is the Kolmogorov distance. It is worth to note Kolmogorov distance is a measure of distinguishability be- tween classical distributions. A detailed proof of this theorem can be found in Ref. [ARHP14]. A particular and interesting class of stochastic matrices are bistochastic matri- ces. They are defined as the transformations that leave invariant the probabil- ity distribution with maximum entropy, given by ~m = (1/N, . . . ,1/N)T , where N is the number that the system can have. Therefore a bistochastic matrix fulfills ~m = Λ(0,t)~m. Doing simple algebra lead us to note that bistochastic matrices ad- ditionally fulfill ∑ j ( Λ(t,0) ) i j = 1 ∀i. This implies that they are also stochastic matrix acting from the right, i.e. mapping row probability vectors. This is also the origin of the name bistochastic. In the previous section we have introduced the concept of Markovianity in the context of open quantum systems, the so called Markovian approximation. It consisted on assuming that the system ’forgets the information about its previous states’. This concept comes from the theory of classical stochastic processes. Let us introduce the following definition [BP07, ARHP14], Definition 1 (Classical Markovian process). Let xt be the state of a stochastic system where t ∈ [0,τ], and χ = {t0, . . . tn} any ordered set of times such that 0 < t0 < t1 < · · ·< tn < τ , the process is Markovian if P(xtn , tn|xtn−1 , tn−1; . . . ;xt0 , t0) = P(xtn , tn|xtn−1 , tn−1) ∀n > 0, (2.23) where P(·|·) denotes conditional probability. According to this definition, the conditional probability of the system to be at the state xtn at the time tn, given the history of events {xtn−1 , tn−1; . . . ;xt0 , t0}, depends only on the previous state. This definition captures the memoryless character of Markovian processes. Consider now a stochastic process and {Λ(t,0)}t∈χ its set of stochastic matrices given some ordered set of times χ . If the process is Markovian then the matrices Λ(tm,tn) are stochastic matrices for any χ , where tm > tn ∈ χ . The converse is not true [ARHP14, BP07]. This condition implies that the map Λ(t,0) is divisible in 2.2. Quantum channels 17 the sense that it can always be written as Λ(t1,t0) = Λ(t1,s)Λ(s,t0) ∀ t1 > s > t0, (2.24) with Λ(t1,s), Λ(s,t0) and Λ(t1,t0) stochastic matrices, the latter two by definition. In- termediate maps can be constructed as Λ(t1,s) = Λ(t1,t0)Λ −1 (s,t0) if Λ−1 (s,t0) exists. Note that theorem 3 implies that Markovian stochastic processes do not increase the Kolmogorov distance. 2.2.2 Construction of quantum channels The concept of quantum channel, also known as quantum operation, captures the idea of stochastic map in the quantum setting. Thus, being the density matrices the analogous objects to probability vectors, we seek for linear operations that transform density matrices into density matrices. The operations that do such job are defined as follows: Definition 2 (Positive and trace preserving linear operations (PTP)). A linear op- eration E : T (H )→ T (H ) is positive and trace preserving if, for all ∆ ∈ H , we have the following • E [∆]≥ 0 ∀∆ ≥ 0, • tr(E [∆]) = tr(∆). A remarkable property of linear positive maps is that they are contractive respect to the trace norm [ARHP14]. This is leads to a decrease of the distinguishability of quantum states, similar to the classical case. Theorem 4 (Contractivity of positive maps). A linear map E is PTP if and only if |E [∆]|tr ≤ |∆|tr ∀∆† = ∆ ∈ B(H ). A simple proof for the finite dimensional case can be found in Ref. [ARHP14]. Now, given that any hermitian operator can be written as ∆ = (tr∆)Hp, for tr∆ 6= 0, ∆ = tr∆+ (ρ1 −ρ2) , for tr∆ = 0, where Hp = pρ1 − (1− p)ρ2 a Helstrom matrix and p ∈ [0,1], by theorem 4 the generalized trace distance defined as Dp(ρ1,ρ2) = |pρ1 − (1− p)ρ2|tr decreases after the application of a positive map E , i.e. Dp (E [ρ1],E [ρ2])≤ Dp (ρ1,ρ2) . 18 Chapter 2. Open quantum systems and quantum channels It is worth to point out that this is directly related to the two-state discrimination problem where we have, for instance, probability p of erroneously identify ρ1 with ρ2 [NC11, ARHP14]. In this setting the probability of failing with such identification is Perr = 1−Dp(ρ1,ρ2) 2 . Therefore if the distance is zero, the probability of correctly identify ρ1 is the same as choosing randomly between ρ1 and ρ2, but if it is 1, we identify ρ1 from ρ2 with certainty. For p = 1/2 we recover the standard unbiased trace distance. It is well known that any quantum system can be entangled with another, for instance a central system can be entangled with its environment. Thus, in the context of quantum operations we must handle this fact carefully. Let us define the following: Definition 3 (k-positive operations). A linear map E is k-positive if idk ⊗E [∆̃]≥ 0 ∀∆̃ ≥ 0 ∈ B(Hk ⊗H ), with k a positive integer, being the dimension of Hk and idk the identity map in that space. Therefore a positive map is k-positive if the expended map idk ⊗E is positive, the trace preserving of k-positive maps follows immediately from the trace preserving of E . Such maps transform properly density matrices of the extended system (with ancilla of dimension k) into density matrices, apart from the fact that they transform properly the density matrices of the system, hence handling quantum entanglement correctly for this ancilla. Since the dimension of any other quantum system is arbitrary, being for example the rest of the universe, one must have that quantum maps must transform quan- tum states for every positive integer k. Therefore one defines complete positive and trace preserving linear maps as the following, Definition 4 (Complete positive and trace preserving operations (CPTP)). A trace preserving linear operation E : T (H )→ T (H ) is complete positive if idk ⊗E [∆̃]≥ 0 ∀∆̃ ≥ 0 ∈ B(Hk ⊗H ),∀k ∈ Z + 0 , where Z + is the set of the positive integers. It will be shown later in chapter 3, that deciding complete positivity is straightfor- ward using the so called Choi matrix. 2.2. Quantum channels 19 It is trivial to check that unitary operations, U [ρ] =UρU†, are CPTP maps as ex- pected. Additionally they leave invariant the maximally mixed state, ✶/dim(H ). In fact, unitary operations belong to a wider class of CPTP maps called unital quantum maps, similar to its classical counterpart. The set of unital channels is defined simply as the one containing CPTP maps E that additionally fulfill E [✶] = ✶. Additionally notice that due to the trace preserving property the adjoint operator of E is always unital. The adjoint is defined in the usual way, 〈A,E [B]〉= 〈E ∗[A],B〉, (2.25) where the inner product is the Hilbert-Schmidt product and A ∈ T (H ) and B ∈ B(H ) [Hol01, HZ12]. Now, ∀∆ ∈ T (H ) we write the trace preserving condition as tr∆ = trE [∆] = 〈✶,E [∆]〉= 〈E ∗[✶],∆〉, therefore E ∗[✶] = ✶. Let us now illustrate the connection of the concept of quantum channel with the scheme of open quantum systems introduced above. Consider the following the- orem [Sti06]: Theorem 5 (Stinespring dilation theorem). Let E a CPTP map, there exist an environmental Hilbert space HE and ρE ∈ S (HE) such that E [ρ] = trE [ U (ρ ⊗ρE)U† ] , with the unitary matrix U : H ⊗HE → H ⊗HE. The unitary U and the state ρE are not unique [HZ12]. Stinespring theorem is an important result given that one can always understand a CPTP operation as a Hamiltonian evolution in a bigger space, such that we recover the given operation at some fixed time and by performing a partial trace over the environmental de- grees of freedom. Later in this chapter we will discuss an important implication of this theorem for Markovian processes. Along the work we will also denote the set of CPTP linear maps simply as C. A remarkable property of C is its convexity. To show this consider the following convex combination of CPTP maps: E = pE1+(1− p)E2, acting upon the density matrix ρ0. By linearity we have E [ρ0] = pE1[ρ0] + (1− p)E2[ρ0]. Defining the density matrices ρi = Ei[ρ0] ∈ S (H ), it follows from the convexity of S (H ) that E is another CPTP map. Therefore the set C is convex. Unitary maps are extremal channels of C i.e. they cannot be written as convex combinations of other channels, but they can be used to construct other maps, 20 Chapter 2. Open quantum systems and quantum channels id U1 U2 U3 Figure 2.3: The figure shows an schematic slice of CPTP maps, one can see the identity map and other extremal channels. The straight lines are convex combinations of those channels, the curve contains channels in the boundary that cannot be written as convex combinations of unitary channels. see fig.2.3. For instance consider a simple convex combination of unitary maps E [ρ] = ∑i piUρU†, with ∑i pi = 1 and pi ≥ 0. This channel is a more general example of a unital channel, in fact it turns that every unital qubit channel has such form. This can be shown easily using the Ruskai’s decomposition that will be introduced in the next chapter. Convex combinations of unitary channels can be implemented in the laboratory, for instance choosing unitaries randomly by tossing a die. Regarding the algebraic properties of the set C, it enjoys the structure of a semi- group. It is closed under the composition operation, i.e. E1E2 ∈ C, ∀E1,E2 ∈ C, and is associative, (E1E2)E3 = E1 (E2E3). Additionally it contains an identity ele- ment. C does not contain the inverse elements, this captures the irreversible char- acter of general quantum operations, being only the unitaries the ones their inverse elements in C. Furthermore, C contains another remarkable convex structure, Definition 5 (Entanglement-breaking channels). A map E ∈ C is entanglement- breaking if it breaks the entanglement of the system with any ancilla, i.e. ∀k ∈ Z + and ∀σ ∈ S (Hk ⊗H ), the state (idk ⊗E ) [σ ] is separable. This set is convex given that convex combinations of separable states is separa- ble [HHHH09]. Quantum channels can be seen as the basic building of time-dependent quantum processes, also called quantum dynamical maps. Definition 6 (Quantum dynamical maps). A continuous family of channels {Et ∈ C : t≥ 0,E0 = id} is called quantum dynamical map. 2.2. Quantum channels 21 id U1 U2 U3Et Figure 2.4: Scheme of an smooth dynamical map inside a slice of the set C. Given some interval I , if the family is smooth respect to t ∈ I and invertible, it admits a master equation ρ̇(t) = At [ρ(t)] with At = ĖtE −1 t . An schematic description is shown in fig. 2.4. Note that the standard scheme open quantum systems, introduced at the beginning of this chapter, leads to quantum dynamical maps. 2.2.3 Non-Linear CPTP operations Notice that the set C does not contain everything that can be performed on a quan- tum system; it contains only linear operations. Therefore C do not contain post- seletion procedures, i.e. updating the state once a measurement is done and the result is known. For instance, let ρ the state of some system and {Mi} a collection of measurement operators over it, where the index i refers to the measurement outcome. The probability of measuring i is p(i) = tr ( MiρM † i ) , while the opera- tion performed over the state is ρ 7→ MiρM † i tr ( MiρM † i ) . This operation is explicitly non-linear but it is trivially complete positive and trace preserving. Note that if the action of the measurement apparatus is performed but the experimentalist do not read the outcome, or it is simply forgotten, the resulting map belongs to C [NC11]. This is shown by noting that the operation MiρM † i is applied with probability p(i), then the performed operation is ∑i p(i)MiρM † i and it is linear and CPTP by construction. Complete positivity follows immediately 22 Chapter 2. Open quantum systems and quantum channels from the complete positivity of ρ 7→ MiρM † i and the trace preserving property from the weighted summation. A more general set of operations including measurements, postselection and ex- change of classical information will be introduced in the next subsection. 2.2.4 Local operations and classical communication Several types of quantum operations can be found and studied, in particular in Ref. [HZ12] there is a classification mainly based on its locality. A paradigmatic and widely studied type are the so called local operations and classical com- munication [HHHH09]. A surprising feature of this operations is that they can increase the entanglement of entangled states of a system (at the cost of throw- ing some members of the ensemble), but cannot create them from non-entangled ones [VDD01, HHHH09]. In this work we are particularly interested in one-way stochastic local operations and classical communication channels (1wSLOCC). Consider a bipartite system where one part is controlled by Alice and the other by Bob. Alice performs an operation which includes measurements with postselection, and then she commu- nicates its outcome to Bob. Then Bob performs a local operation that can be again a measurement with postselection, finishing the protocol. The stochasticity comes from the fact that this operation, for each particular set of measurement outcomes, has a certain probability generally less than 1 of occurrence. And the one-way comes from the fact that no feedback is given to Alice and no more operations and classical communications are performed. This operations can be written in the following way: ρ 7→ ρ ′ = (X ⊗Y )ρ (X ⊗Y )† tr [ (X ⊗Y )ρ (X ⊗Y )† ] . (2.26) Additionally we will consider detX 6= 0 and detY 6= 0, this is the usual choice as projective measurements destroy entanglement [VDD01]. This operations are complete positive and trace preserving, but non-linear unless X and Y are uni- taries. Additional notice that given ρ and ρ ′, the matrices X and Y can always be chosen such that detX = detY = 1 (for the invertible case). Therefore for two- level systems it is enough to consider X ,Y ∈ SL(2,C) [Tun85], where the latter is the special linear group of 2× 2 matrices with complex entries. Furthermore notice that the operation ρ 7→ (X ⊗Y )ρ (X ⊗Y )† (2.27) 2.3. Quantum channels of continuous variable systems 23 preserves the determinant, i.e. detρ = detρ ′. In the next chapter we will ex- ploit this to show that there is a correspondence between 1wSLOCC and Lorentz transformations. We use this to introduce a decomposition analogous to the sin- gular value decomposition, but using the Lorentz metric instead of the Euclidean, enjoying an useful physical meaning. 2.3 Quantum channels of continuous variable systems Many of the definitions and tools introduced in the previous sections are also rele- vant for the infinite dimensional case. Although we can always choose countable basis for the Hilbert space as long it is separable [HZ12], it is often of interest to consider non-countable bases, typically phase-space variables. This introduces the theory of continuous variable systems. It is a central topic of study given that they appear naturally in the description of many physical systems. A few examples are the electromagnetic field [CLP07], solids and nano-mechanical sys- tems [AKM14] and atomic ensembles [HSP10]. In particular, in this section we introduce and discuss a set of continuous variable channels called Gaussian quan- tum channels. 2.3.1 Gaussian quantum states To introduce the definition of Gaussian quantum channel, consider first the sim- plest state type of quantum states in continuous variable, both from a theoretical an experimental point of view, the so-called Gaussian states. The operations that transform such family of states into itself are called Gaussian quantum channels (GQC). Even though Gaussian states and channels are small subsets of all possi- ble states/channels, they have proven to be useful in a very wide variate of tasks such as quantum communication [GVAW+03], quantum computation [LB99] and the study of quantum entanglement in simple [BvL05] and complicated scenar- ios [LRW+18]. Gaussian states are defined as those having Gaussian Wigner function. In partic- ular, for one-mode the Wigner function is W (~u) = 1 2π √ detσ e− 1 2(~u−~d) T σ−1(~u−~d), (2.28) where ~u = (q, p)T [EW07]. The mean vector ~d and the covariance matrix σ are 24 Chapter 2. Open quantum systems and quantum channels the first and second moments, respectively. They are given by σ = ( 〈q̂2〉−〈q̂〉2 1 2 〈q̂p̂+ p̂q̂〉−〈q̂〉〈 p̂〉 1 2 〈q̂p̂+ p̂q̂〉−〈q̂〉〈 p̂〉 〈 p̂2〉−〈 p̂〉2 ) , ~d = (〈q̂〉,〈p̂〉)T . The observables q̂ and p̂ are the standard canonical conjugate position and mo- mentum variables. As for any other Gaussian variable, Gaussian quantum states are characterized completely by first and second probabilistic moments. Therefore a Gaussian state S can be denoted as S = S ( σ , ~d ) . 2.3.2 Gaussian quantum channels To start with, we recall the following definition [WPGP+12]: Definition 7 (Gaussian quantum channels). A quantum channel is Gaussian (GQC) if it transforms Gaussian states into Gaussian states. This definition is strictly equivalent to the statement that any GQC, say A , can be written as A [ρ] = trE [ U (ρ ⊗ρE)U† ] (2.29) where U is a unitary transformation, acting on a combined global state obtained from enlarging the system with an environment E, that is generated by a quadratic bosonic Hamiltonian (i.e. U is a Gaussian unitary) [WPGP+12]. The environ- mental initial state ρE is a Gaussian state and the trace is taken over the environ- mental degrees of freedom. Following definition 7, a GQC is fully characterized by its action over Gaussian states, and this action is in turn defined by affine transformations [WPGP+12]. Specifically, A = A (T,N,~τ) is given by a tuple (T,N,~τ) where T and N are 2×2 real matrices with N = NT [WPGP+12] acting on Gaussian states according to A (T,N,~τ) [ S ( σ , ~d )] = S ( TσTT +N,T~d +~τ ) . In the particular case of closed systems we have N = 0 and T is a symplectic matrix. The particular form and properties of Gaussian quantum channels in the continuous variable representations, as well their connection the mentioned affine transformations, will be given in chapter 3. 2.3. Quantum channels of continuous variable systems 25 In this work we explore GQCs without Gaussian functional form in the position state representation. In particular we study channels that can arise when singu- larities on the coefficients of Gaussian forms GF occur (they will be denoted by δGQC). Such channels can lead immediately to singular Gaussian operations. Thus, we characterize which forms in δGQC lead to valid quantum channels, and under which conditions singular operations lead to valid singular Gaussian quantum channels (SGQC). Let us note that although channels with Gaussian form trivially transform Gaus- sian states into Gaussian states, the definition goes beyond GF. We will use the typical difference and sum coordinates, x = q2 − q1 and r = (q1 + q2)/2, respec- tively. Defining ρ(x,r) = 〈 r− x 2 ∣ ∣ ρ̂ ∣ ∣r+ x 2 〉 , a quantum channel in this represen- tation is defined such that ρ f (x f ,r f ) = ∫ R2 dxidriJ(x f ,xi;r f ,ri)ρi (xi,ri) , (2.30) where ρ̂i and ρ̂ f are the initial and final states, respectively, and J(x f ,xi;r f ,ri) is the representation of the quantum channel in the aforementioned variables. An example of a channel without GF can be constructed from the general form of Gaussian quantum channel with GF [MP12]: JG(x f ,xi;r f ,ri) = b3 2π exp [ ı ( b1x f r f +b2x f ri +b3xir f +b4xiri + c1x f + c2xi ) −a1x2 f −a2x f xi −a3x2 i ] , (2.31) where all coefficients are real and no quadratic terms in ri, f are allowed. Choosing an = αnε−1 + ãn and bn = βnε−1/2 + b̃n, with ε > 0, αn,βn, ãn, b̃n ∈ R ∀n and b̃3 = 0. Taking the limit ε → 0 and using the formula δ (x) = lim ε→0 1 2 √ πε e −x2 4ε , (2.32) we arrive to lim ε→0 JG(x f ,xi;r f ,ri) = N δ (αx f −βxi)e Σ′(x f ,xi;r f ,ri), (2.33) where α , β ∈R and Σ′(x f ,xi;r f ,ri) is a quadratic form that now admits quadratic terms in ri, f , arising from the completion of the square of the exponent of eq. (2.31) 26 Chapter 2. Open quantum systems and quantum channels to take the limit of eq. (2.32). This is the first example of a δGQC. This channel is still a GQC according to the definition. A physical, but complicated realization occurs in the system of one Brownian quantum particle with harmonic potential and linearly coupled to the bath. In such system, channels with the functional form of eq. (2.33) are realized at isolated points in time, see equations 6.71-75 of Ref. [GSI88]. Since the form of eq. (2.33) admits quadratic terms in ri, f in the exponent, it suggest that a form with two deltas exist and can be defined using the same limit, see eq. (2.32). In fact, the identity map is a particular case; it is realized setting J(x f ,xi;r f ,ri) = δ (x f − xi)δ (r f − ri). In any case, to avoid working with such limits, it is convenient to perform a black-box characterization of general forms involving Dirac’s deltas, which will be done in the next chapter. This will lead to explicit relations between position state representation and affine representations of Gaussian quantum channels without Gaussian functional form. Chapter 3 Representations of quantum channels Simplicity is the ultimate sophistication. Leonardo da Vinci In this chapter we introduce several and useful representations of quantum chan- nels for the finite dimensional case. We start with the Kraus representation, al- ready mentioned in the previous chapter, but additionally we will show that quan- tum channels always have this form. Later on we introduce Choi’s theorem (and the so called Choi-Jamiołkowski representation) which is cornerstone tool to study many properties of quantum channels. We also discuss operational representations by introducing two types of basis. These representations are useful to prove sev- eral results in this work. Next, we apply the introduced tools to the qubit case. Additionally we discuss two decomposition of qubit channels, leading to two nor- mal forms that are essential to study divisibility properties of quantum channels. 3.1 Kraus representation In the previous chapter we have shown that starting from the usual scheme of open quantum systems, we arrive to the Kraus representation, see eq. (2.8). Later on, using the Stinespring dilation theorem, see Theorem 5, we show that CPTP maps can always fit in the scheme of open quantum systems for some global 27 28 Chapter 3. Representations of quantum channels unitary evolution. Since the latter scheme always has a Kraus representation, one concludes that CPTP maps always have a Kraus representation. It turns out that the converse also holds [KBDW83]. Theorem 6 (Kraus). A linear operation E : T (H ) → T (H ) belongs to C if and only if there exist a set of bounded operators {Ki} such that E [∆] = ∑ i Ki∆K† i ∀∆ ∈ T (H ), with ∑i K † i Ki = ✶. Proof. The ’only if’ part is already commented in the main text and follows the logic: every E ∈ C has a dilation such that it has the familiar form of the open quantum systems dynamics, i.e. there exists U and ρE such that E [ρ] = trE [ U (ρ ⊗ρE)U† ] . We already showed that writing ρE in terms of its eigenbasis, the latter expression leads to the Kraus representation, see eq. (2.8). To prove the ’if’ part, we only have to construct the extended map to test its complete positivity. Let k > 0 ∈Z and τk = (idk ⊗E ) [∆̃k], where ∆̃k ∈B(Hk ⊗H ) and ∆̃k ≥ 0, using Kraus decomposition and evaluating 〈φ |τk|φ〉 with |φ〉 ∈ Hk ⊗H , one arrives to 〈φ |τk|φ〉= ∑ i 〈φ |(✶k ⊗Ki) ∆̃k ( ✶⊗K † i ) |φ〉 = ∑ i 〈φi|∆̃k|φi〉 ≥ 0. The latter follows immediately from the positive-semidefinitiveness of ∆̃k, i.e. 〈φi|∆̃k|φi〉 ≥ 0. The condition ∑i K † i Ki = ✶ comes from the trace-preserving of E and the cyclic property of the trace, trE [∆] = ∑ i tr [ Ki∆K † i ] = ∑ i tr [ K † i Ki∆ ] = tr [( ∑ i K † i Ki ) ∆ ] = tr∆, 3.2. Choi-Jamiołkowski representation 29 Therefore ∑i K † i Ki = ✶. It is worth to note that Kraus operators are not unique. Defining a new set of operators, Ak = ∑l uklKl , it is easy to show that ∑i Ki∆K † i = ∑k Ak∆A † k if and only if ukl are the components of an unitary matrix. Therefore different Kraus representations are related by unitary conjugations. 3.2 Choi-Jamiołkowski representation The Choi-Jamiołkowski representation arises as part of a very useful theorem in quantum information theory, the so called Choi’s theorem [Cho75, HZ12]. Theorem 7 (Choi). Let E : Cn×n → C m×m be a linear map. The following state- ments are equivalent: i) E is n−positive. ii) The matrix CE = n ∑ i, j=1 |ϕi〉〈ϕ j|⊗E [|ϕi〉〈ϕ j|] ∈ C n×m ⊗C n×m is positive-semidefinite with {|ϕi〉}n i=1 an orthonormal basis in C n. iii) E is completely positive. Proof. The proof of iii) → i) is trivial, if E is completely positive then it is n−positive. The implication i) → ii) can be proved easily by noticing that nor- malizing CE → CE /n =: τE , where τE can be obtained as the application τE = (idn ⊗E ) [ω], where ω = |Ω〉〈Ω| with |Ω〉 = 1/ √ n∑ n i |ϕi〉⊗ |ϕi〉 a Bell state be- tween two copies of Cn. Therefore, by the n−positivity of E it follows that τE is positive-semidefinite. What remains to prove is ii) → iii). To do this observe that the space C n×m is isomorphic to the direct sum of n copies of Cm, i.e. Cn×m ∼= C m 1 ⊕C m 2 ⊕·· ·⊕C m n , and define the projector into the kth copy as Pk = 〈ϕk| ⊗✶, such that PkCE Pl = E [|ϕk〉〈ϕl|]. Now, given that CE is positive-semidefinite, it can be written as CE = ∑ nm i |Ψi〉〈Ψi|, where |Ψi〉 ∈ C n×m are generally unnormalized vectors. Thus, we have that E [|ϕk〉〈ϕl|] = ∑i Pk|Ψi〉〈Ψi|Pl , where Pk|Ψi〉 ∈ C m k . Defining the opera- tors {Ki : Cn →C m}i via the equation Pk|Ψi〉= Ki|ϕk〉, where choosing for exam- ple |ϕk〉 as the canonical basis, the columns of Ki contain the n projections of |Ψi〉 into the copies of Cm. Finally we arrive to E [|ϕk〉〈ϕl|] = ∑i Ki|ϕk〉〈ϕl|K† i ∀k, l = 1, . . . ,n. In conclusion, since {|ϕk〉〈ϕl|}k,l is a complete basis of Cn×n, by linearity and by theorem 6, the map E is completely positive. 30 Chapter 3. Representations of quantum channels The matrix CE is commonly known as Choi matrix and τE as Choi-Jamiołkowski state. Both define the Choi-Jamiołkowski representation, in this work labeled as τE since it is normalized. Choi’s theorem provides a simple test of complete positivity, which I find beau- tiful. For instance, if we want to know if a given PTP map E is a valid quantum map, we just have to consider two copies of our system in only one state, the Bell state, then apply E to one of the copies and check if the result, τE = (idn ⊗E ) [ω], is a density matrix. The Choi-Jamiołkowski representation enjoys other useful properties, if E pre- serves the trace, the matrix τE = 1 n    E [|ϕ1〉〈ϕ1|] . . . E [|ϕ1〉〈ϕn|] ... . . . ... E [|ϕn〉〈ϕ1|] . . . E [|ϕn〉〈ϕn|]    (3.1) has blocks of trace 1/n and 0, since trE [|ϕi〉〈ϕ j|] = δi j. This property additionally means that not every density matrix in C n×m ⊗C n×m has a corresponding CPTP map. The matrix rank of τE coincides with the so called Kraus rank, i.e. the number of linearly independent Kraus operators required to write the channel. This can be shown easily noticing that computing τE from the Kraus sum, one arrives to the equality |Ψi〉= 1/ √ n✶⊗Ki|Ω〉, therefore the linear independence of {|Ψi〉}i follows immediately from the linear independence of {Ki}i. Therefore the max- imum Kraus rank is mn and the minimum 1. Channels with Kraus rank equal to 1 are trivially unitary channels given that E [∆] = K∆K† with K†K = ✶. Channels with the maximum rank are called full Kraus rank channels. Another interesting property is that if τE is separable (i.e. not entangled) , then E is entanglement-breaking, see definition 5. For qubit channels it is enough to test that the concurrence is zero [RFZB12]. 3.3 Operational representations It has been shown that the Choi-Jamiołkowski representation is useful to test sev- eral properties of quantum channels. In this section we will introduce other repre- sentations, this time with operational meanings. They are basically operator basis that give matrix and vector forms to channels and density matrices, respectively. 3.3. Operational representations 31 The vectorization of density matrices can be achieved simply “making them flat”, this is,    ρ11 . . . ρ1d ... . . . ... ρd1 . . . ρdd    7→      ρ11 ρ12 ... ρdd      =:~ρ. Using this mapping, the matrix form of operators acting on T (H ) is build using the simple rule [GTW09] AρB 7→ ( A⊗BT ) ~ρ. (3.2) For instance applying this rule to a commutator, [H,ρ] 7→ ( H ⊗✶−✶⊗HT ) ~ρ . This representation is useful to prove various results involving operators acting on the space of density matrices, see for instance the appendix A. Additionally it is simple to prove that the Hilbert-Schmidt inner product is mapped to 〈γ,ρ〉 7→~γ†~ρ . One can use other operator basis accordingly to our purposes. In general we have the following, consider {Ai}i an orthonormal operator basis in the space T (H ), the components of the density matrix are αi = 〈Ai,ρ〉= tr [ A † i ρ ] , so ρ = ∑ i αiAi. Correspondingly, the components of operators acting on B(H ), for instance E , are simply Êi j = 〈Ai,E [A j]〉= tr [ A † i E [A j] ] . Using this equation it is easy to prove that the representation of the adjoint opera- tor of E , see eq. (2.25), is simply Ê ∗ = Ê †. 3.3.1 Hermitian and traceless basis Two types of basis are specially useful in this work, the first one are the hermitian basis. This is, every orthonormal basis {Ai}i that fulfills Ai = A † i , ∀i. To show the utility of this kind of basis, let us introduce the following definition, Definition 8 (Hermiticity preserving operators). A linear operator E : B(H )→ B(H ) preserves hermiticity if E [∆]† = E [∆†], ∀∆ ∈ B(H ). 32 Chapter 3. Representations of quantum channels Using the Kraus representation is trivial to prove that linear CPTP maps preserve hermiticity, using complete positivity. Furthermore, hermiticity preserving maps enjoy an hermitian Choi-Jamiołkowski representation, i.e. τE = τ† E [Wol11]. Using an hermitian basis it is straightforward to prove the following, Proposition 1 (Representation with real entries). Let E be a linear and hermiticity preserving map. Its matrix representation using an hermitian basis {Ai} has real entries. Proof. Let Êi j = tr [AiE [A j]], where the line over denotes complex conjugation. Distributing the latter inside the argument of the trace and using the hermiticity of Ai, we get Êi j = tr [ AiE [A j] † ] , finally stressing that E [A j] † = E [A† j ] = E [A j], we arrive to Êi j = Êi j. This simple property will be used later to prove the equivalence of the problem of finding channels that can be written as E = exp(L), with L a Lindblad operator. The second useful type of basis are the so called traceless bases. They are defined as follows. Let {Fi}d2−1 i=0 be an orthogonal basis, where we have indicated the dimension of the space T (H ) as d2 with d = dim(H ), it is traceless if F0 = ✶/ √ n and trFi = 0 ∀i > 0. The traceless property comes from the fact that only one element has non-zero trace, it is easy to prove that it must be proportional to the identity, given that one can write the identity matrix using such basis. This basis is useful to prove that generators of quantum dynamical maps, Lt , de- fined with E(t+ε,t)[ρ] = ρ + εLt [ρ]+O(ε2), have the following specific structure. Theorem 8 (Specific form of generators of dynamical maps). Let L : T (H )→ T (H ) be a linear operator fulfilling L[∆]† = L[∆†] and tr [L[∆]] = 0 (or equiva- lently L∗[✶] = 0), then it has the following form, L[ρ] = i[ρ,H]+ d2−1 ∑ i, j=1 Gi j ( FiρF † j − 1 2 {F † j Fi,ρ} ) , (3.3) where d = dim(H ), H ∈C d×d and G∈C(d2−1)×(d2−1) are hermitian, and {Fi}d2−1 i=0 is an orthonormal traceless basis of B(H ). Notice that Lindblad operators enjoy such form with the additional condition that G ≥ 0, see eq. (2.20). A proof of this is given in Ref. [EL77] for the infinite dimensional case using technicalities beyond this work. Here we will prove it for 3.3. Operational representations 33 the finite dimensional case, using the notation of an incomplete proof given in Ref. [WECC08]. Proof. Since L preserves hermiticity, it has an hermitian Choi-Jamiołkowski ma- trix, τL ∈ C d2×d2 . We can write such matrix always as τL = τφ −|Ψ〉〈Ω|− |Ω〉〈Ψ|, (3.4) where |Ψ〉 = −ω⊥τL|Ω〉− (λ/2) |Ω〉, λ = 〈Ω|τL|Ω〉, ω⊥τLω⊥ = ω⊥τφ ω⊥ = τφ and ω⊥ = ✶−ω . Observe that choosing the traceless operator basis {Fi}d2−1 i=0 , it is simple to prove that the matrix τφ can be understood also as the Choi- Jamiołkowski matrix of the following operator: φ [ρ] = d2−1 ∑ i, j=1 Gi jFiρF † j , (3.5) with G hermitian, i.e. τφ = (idd2 ⊗φ) [ω]. This can be shown noticing that the summation ∑ d2−1 i, j=1 goes over only traceless operators, therefore the projections into the one-dimensional space of |Ω〉 of the Choi matrix of φ are null, ωτφ = d2−1 ∑ i, j=1 Gi j 1 d tr(Fi)ω ( ✶⊗F † j ) = 0 , and similarly for τφ ω = 0 . For the second and third terms of Eq. 3.4, it is easy to show that the corresponding operator is simply ρ 7→ −κρ −ρκ†, where we identify |Ψ〉= (✶⊗κ) |Ω〉. Up to now we have shown that hermiticity preserving generators have the form L [ρ] = φ [ρ]−κρ −ρκ†. (3.6) Using the condition L∗[✶] = 0, we have that κ +κ† = φ ∗[✶], i.e. the hermitian part of κ is given by 1 2 ∑ d2−1 i, j=1 Gi jF † j Fi. Simply writing the anti- hermitian part as iH we end up with κ = iH + 1 2 d2−1 ∑ i, j=1 Gi jF † j Fi. Substituting this expression and eq. (3.5), in eq. (3.6), we arrive to the desired form, see 4.5. 34 Chapter 3. Representations of quantum channels Notice that the operator φ is completely positive if and only if G≥ 0 [HZ12], thus, G ≥ 0 ⇔ τφ ≥ 0. In such case L has exactly a Lindblad form. This condition will be introduced later as conditional complete positivity [EL77, WECC08]. The following is a central and useful result for our work. Proposition 2 (Conditional complete positivity). An hermiticity preserving linear operator L : T (H )→ T (H ) fulfilling tr [L∗[✶]] = 0, has Lindblad form if and only if ω⊥τLω⊥ ≥ 0. Additionally choosing an arbitrary basis of the Hilbert space to write operators {Fi}d2 i=1, it is easy to prove that G and ω⊥τLω⊥ are related by an unitary conjuga- tion [CDG19]. 3.4 Qubit channels We shall devote some time to the most simple but non-trivial quantum system, the qubit. This case turns to be rich enough to use and test the tools provided by the literature and the ones developed here, in the context of divisibility. We recall a particular representation and a couple of decomposition theorems for qubit channels. 3.4.1 Pauli representation and Ruskai’s decomposition In the case of qubit channels we can have at the same time an hermitian, traceless and unitary basis, it is the simple Pauli basis 1√ 2 {1,σx,σy,σz}. This induces a simple 4×4 representation with real entries given by Ê = ( 1 ~0T ~t ∆ ) , (3.7) where ∆ is a 3×3 matrix with real entries and~t a column vector. This describes the action of the channel in the Bloch sphere picture in which the points ~r are identified with density matrices ρ~r = 1 2 (✶+~r ·~σ) [RSW02]. Therefore the action of the channel is described by E (ρ~r) = ρ∆~r+~t . 3.4. Qubit channels 35 In order to study qubit channels with simpler expressions, we will consider a decomposition in unitaries such that E = U1DU2. (3.8) This can be achieved using Ruskai’s decomposition [RSW02], which can be per- formed by decomposing ∆ in rotation matrices, i.e. ∆ = R1DR2, where D = diag(λ1,λ2,λ3) is diagonal and the rotations R1,2 ∈ SO(3) (of the Bloch sphere) correspond to the unitary channels U1,2. Notice that as D is not required to be positive-semidefinite, Ruskai’s decomposition must not be confused with the sin- gular value decomposition. The latter allows decompositions that include total re- flections. Such operations do not correspond to unitaries over a qubit, in fact they are not CPTP. An example of this is the universal NOT gate defined by ρ 7→ ✶−ρ , it is PTP but not CPTP. The resulting form from Ruskai’s decomposition is stated in the following theorem, Theorem 9 (Special orthogonal normal form). For any qubit channel E , there exist two unitary conjugations , U1 and U2, such that E = U1DU2, where D has the following form in the Pauli basis, D̂ = ( 1 ~0T ~γ D ) , (3.9) and is called special orthogonal normal form of E . Here, RT 1 ∆RT 2 = D and ~γ = RT 1 ~t. The latter describes the shift of the center of the Bloch sphere under the action of D . The parameters ~λ determine the length of semi-axes of the Bloch ellipsoid, being the deformation of Bloch sphere under the action of E . In particular detD̂ = det Ê = λ1λ2λ3. To develop geometric intuition in the space determined by the possible values of the three parameters of~λ , consider the Choi-Jamiołkowski representation of the special orthogonal normal form of an arbitrary channel in the basis that diago- nalises D, τD = 1 4     γ3 +λ3 +1 γ1 − iγ2 0 λ1 +λ2 γ1 + iγ2 −γ3 −λ3 +1 λ1 −λ2 0 0 λ1 −λ2 γ3 −λ3 +1 γ1 − iγ2 λ1 +λ2 0 γ1 + iγ2 −γ3 +λ3 +1     . (3.10) Complete positivity is determined by the non-negativity of its eigenvalues, given that it is hermitian, but it turns that for the general case they have complicated expressions. To overcome this problem we use the fact that if D is a channel, then 36 Chapter 3. Representations of quantum channels its unital part, defined by taking ~γ =~0, is a channel too [Wol11]. Therefore the set of the possible values of~λ for the general case is contained in the set arising from the unital case. The complete positivity conditions for the latter are 1+λi ± (λ j +λk)≥ 0, (3.11) with i, j and k all different, this implies that the possible set of lambdas lives inside the tetrahedron with corners (1,1,1), (1,−1,−1), (−1,1,−1) and (−1,−1,1), see fig. 3.1. For unital channels, all points in the tetrahedron are allowed. The corner ~λ = (1,1,1) corresponds to the identity channel, ~λ = (1,−1,−1) to σx ~λ = (−1,1,−1) to σy and~λ = (−1,−1,1) to σz (Kraus rank 1 operations). Points in the edges correspond to Kraus rank 2 operations, points in the faces to Kraus rank 3 operations and in the interior of the tetrahedron to Kraus rank 4 operations. In particular, this tetrahedron defines the set of Pauli channels, which are defined to have diagonal special orthogonal normal form. Definition 9 (Pauli channels). A qubit channel E is a Pauli channel if E [ρ] = 3 ∑ i=0 piσiρσi, (3.12) with σ0 := ✶, pi ≥ 0 and ∑ 3 i=0 pi = 1. For non-unital channels more restrictive conditions arise, an example will be given later. 3.4.2 1wSLOCC and singular value decomposition using the Lorentz metric There is another parametrization for qubit channels called Lorentz normal decom- position [VDD01] which is specially useful to characterize infinitesimal divisi- bility CInf. To introduce it, let us resort to chapter 2 where we discussed local operations and classical communication. For the two-qubit case, the operations that Alice and Bob apply for their reduced states are ρA 7→ XρAX† ρB 7→ Y ρBY †, (3.13) where we have shown that it is enough to consider X ,Y ∈ SL(2,C) for X and Y invertible, see chapter 2 Now we are going to show that such operations can be 3.4. Qubit channels 37 λ1 λ3 λ2 id ρ 7→ σxρσx ρ 7→ σzρσz Figure 3.1: Set of the possible values of~λ . This set has the shape of a tetra- hedron where the corners are the Pauli unitaries (✶, σx and σz are indi- cated in the figure, while σy lies behind). The rest of the body contains convex combinations of Pauli unitaries. Unital qubit channels can be obtained by concatenating Pauli channels with unitary conjugations, see theorem 9. 38 Chapter 3. Representations of quantum channels understood as proper orthochronous Lorentz transformations in the Pauli repre- sentation. Consider an arbitrary hermitian operator ∆ and its representation in the Pauli basis, ∆ = (✶ tr∆+~r∆ ·~σ)/2 with det∆ = (tr∆)2 − |~r∆|2. Now observe that det∆ can be understood as the squared Lorentz norm of the four-vector r∆ = (tr∆,~r∆) T , lying in the Minkowski vector space, denoted as (R4,η), where η = diag(1,−1,−1,−1) is the Lorentz metric. Therefore we have det∆ = |r∆|2Lorentz = 〈r∆,ηr∆〉, (3.14) with 〈·〉 the standard inner product. Then tr∆ is a time-like component and ~r∆ space-like components. Given that operations shown in eq. (3.13) preserve the determinant, they are isome- tries in the Minkowski space. That is, they preserve the norm shown in eq. (3.14) for any vector r∆ (with ∆ hermitian). Additionally, due to linearity of eq. (3.13), these operations belong to SO(3,1) (the Lorentz group). In fact, due to the positivity of quantum operations, they do not change the sign of the trace (the time-like component); therefore the transformations are orthochronous. Also no- tice that SL(2,C) contains the identity transformation, therefore the set of one- way stochastic local operations and classical communication is identified with the proper orthochronous Lorentz group, SO+(1,3) [Wol11, Tun85]. However, since −X and X give the same result, see eq. (3.13), and both belong to SL(2,C), one says that the latter is a double cover of SO+(1,3). This map is also called spinor map. Given this map, it is expected that the operations mentioned in eq. (3.13) are explicitly Lorentz matrices, when writing them in the Pauli basis. Also notice that unitary conjugations are particular cases of them [RSW02], therefore one can think of a different decomposition using the Lorentz metric instead of the three dimensional Euclidean metric, used in Ruskai’s decomposition. The Lorentz normal form was introduced first for two-qubit states by writing them as τ = 1 4 ∑ i j Ri jσi ⊗σ j, (3.15) where we have used the notation of Choi-Jamiołkowski states for convenience. This decomposition is derived from the theorem 3 of Ref. [VDD01], which essen- tially states that the matrix R can be decomposed as R = L1ΣLT 2 . (3.16) 3.4. Qubit channels 39 Here L1,2 are proper orthochronous Lorentz transformations and Σ is either Σ = diag(s0,s1,s2,s3) with s0 ≥ s1 ≥ s2 ≥ |s3|, or Σ =     a 0 0 b 0 d 0 0 0 0 −d 0 c 0 0 −b+ c+a     . (3.17) Note that Σ corresponds to an unnormalized state, with trace trΣ = a. Thus, the normalization constant is α = a−1. To introduce the Lorentz normal decomposition of qubit channels, let us first in- troduce the following. Let E a qubit channel and Ê its matrix representation using the Pauli basis. The latter is related with the matrix R, which defines its Choi-Jamiołkowski state, see eq. (3.15), Ê ΦT = R, (3.18) where ΦT = diag(1,1,−1,1). This can be shown by defining a generic Pauli chan- nel, computing its Choi matrix and extracting R using the Hilbert-Schmidt inner product with the basis {σi⊗σ j}i, j. Now, defining the decomposition for channels throughout decomposing the Choi-Jamiołkowski state, we can easily compute the corresponding Lorentz transformations using equations (3.18) and (3.16): RΦT = αL1ΣLT 2 ΦT Ê = αL1ΣLT 2 ΦT (3.19) = L1 (αΣΦT)ΦTLT 2 ΦT (3.20) = L1 ˆ̃ E L̃T 2 , (3.21) where ˆ̃ E = αΣΦT is the Lorentz normal form of Ê . Also notice that L̃T 2 = ΦTLT 2 ΦT is proper and orthochronous, given that its determinant is positive and ΦT is proper. Therefore the possible Lorentz normal forms for channels are ˆ̃ E = diag(s0,s1,−s2,s3) with s0 ≥ s1 ≥ s2 ≥ |s3|, or Σ =     a 0 0 b 0 d 0 0 0 0 d 0 c 0 0 −b+ c+a     . (3.22) Using this, the authors of Ref. [VV02] introduced a theorem (theorem 8 of the reference) defining the Lorentz normal form for channels by forcing b = 0, in order to have normal forms proportional to trace-preserving operations. The latter 40 Chapter 3. Representations of quantum channels is equivalent to say that the decomposition of Choi-Jamiołkowski states leads to states that are also Choi-Jamiołkowski. We didn’t find a good argument to justify such assumption, and found a counterexample that shows that Lorentz normal forms with b 6= 0 exist (see appendix B). Therefore in general we can find a Σ with form of eq. (3.17) with b 6= 0. The consequence of this is that the theorem 8 of Ref. [VV02] is incomplete, but given that form of eq. (3.17) is Kraus rank deficient (it has rank three for b 6= c and two for b = c), the full Kraus rank case is still useful. Thus, we propose a restricted version of their theorem: Theorem 10 (Restricted Lorentz normal form for qubit quantum channels). For any full Kraus rank qubit channel E there exist rank-one completely positive maps T1,T2 such that T = T1E T2 is proportional to ( 1 ~0T ~0 Λ ) , (3.23) where Λ = diag(s1,s2,s3) with 1 ≥ s1 ≥ s2 ≥ |s3|. The channel T is called the Lorentz normal form of the channel E . For unital qubit channels D coincides with Λ. 3.5 Representation of Gaussian quantum channels In this section we start from two ansätze, that put together with the Gaussian functional form considered in Ref. [MP12], lead to the complete set of functional forms in position state representation of one-mode Gaussian channels. We will show that only two possible forms of δGQC hold according to trace pre- serving (TP) and hermiticity preserving (HP) conditions. The one corresponding to eq. (2.33) is one of these, as expected. Later on we will impose complete posi- tivity in order to have valid GQC, i.e. complete positive and trace preserving (C) Gaussian operations. Following definition 7, those channels can be characterized by how they act over Gaussian states. It is well known that the action of GQCs on Gaussian states is described by affine transformations [WPGP+12]. Let A be a GQC defined by a tuple such that A = A (T,N,~τ), where T and N are 2× 2 real matrices with N = NT [WPGP+12]. The transformation acts on Gaussian states according to A (T,N,~τ) [ S ( σ , ~d )] = S ( TσTT +N,T~d +~τ ) . 3.5. Representation of Gaussian quantum channels 41 In the particular case of closed systems, where the system is governed by a Gaus- sian unitary, we have that N = 0 and T is a symplectic matrix. 3.5.1 Possible functional forms of δGQC operations Let us introduce the ansätze for the possible forms of GQC in the position rep- resentation, to perform the black-box characterization. Following eq. (2.29) and taking the continuous variable representation of difference and sum coordinates, the trace becomes an integral over position variables of the environment. Then we end up with a Fourier transform of a multivariate Gaussian. Since the Fourier transform of a Gaussian is again a Gaussian (unless there are singularities in the coefficients, as in the example of eq. (2.33)), the result of the Fourier transform for one mode can have the following structures: a Gaussian form [eq. (2.31)], a Gaussian form multiplied with one-dimensional delta or a Gaussian form multi- plied by a two-dimensional delta. No more deltas are allowed given that there are only two integration variables when applying the channel, see eq. (2.30). Thus, in order to start with the black-box characterization, we shall propose the following general Gaussian operations with one and two deltas, respectively JI(x f ,r f ;xi,ri) = NIδ (~α T~v f +~β T~vi)e Σ(x f ,xi;r f ,ri), (3.24) JII(x f ,r f ;xi,ri) = NIIδ (A~v f −B~vi)e Σ(x f ,xi;r f ,ri), (3.25) with ~vi, j = (ri, j,xi, j), and NI,II are normalization constants. Coefficient arrays A, B, ~α , and ~β have real entries since initial and final coordinates must be real. Finally, the exponent reads: Σ(x f ,xi;r f ,ri) = ı ( b1x f r f +b2x f ri +b3xir f +b4xiri + c1x f + c2xi ) −a1x2 f −a2x f xi −a3x2 i − e1r2 f − e2r f ri − e3r2 i −d1r f −d2ri. They provide, together with eq. (2.31) all possible ansätze for GQC. 3.5.2 Hermiticity and trace preserving conditions Before studying CPTP conditions it is useful to simplify expressions of equations (3.24) and (3.25). To do this we use the fact that linear CPTP operations preserve hermiticity and trace. For channels of continuous variable systems in the position 42 Chapter 3. Representations of quantum channels state representation, J(q f ,q ′ f ;qi,q ′ i), HP condition is derived as follows, ρ f (q ′ f ,q f ) ∗ = ∫ R2 dqidq′iJ(q ′ f ,q f ;qi,q ′ i) ∗ρi(qi,q ′ i) ∗ = ∫ R2 dqidq′iJ(q ′ f ,q f ;q′i,qi) ∗ρi(qi,q ′ i) = ρ f (q f ,q ′ f ), (3.26) where the last equality holds if J(q f ,q ′ f ;qi,q ′ i) = J(q′f ,q f ;q′i,qi) ∗. Using sum and difference coordinates, HP becomes J(−x f ,r f ;−xi,ri) = J(x f ,r f ;xi,ri) ∗. (3.27) Following this equation and comparing exponents of the both sides of the last equations, it is easy to note that the coefficients an, bn, cn, en and dn must be real. Concerning the delta factors, in eq. (3.27) we end up with expressions like δ (α1x f +α2xi +β1r f +β2ri) = δ (−α1x f −α2xi +β1r f +β2ri) for both cases. Therefore the equality holds for eq. (3.24) only for two possible combinations of variables: i) δ (αx f − βxi) and ii) δ (αr f − β ri). For the case of eq. (3.25), equality holds only for iii) δ (γr f −ηri)δ (αx f −βxi). Let us now analyze the trace preserving condition (TP), since the trace of ρ f in sum and dif- ference coordinates is trρ f = ∫ R dr′f ρ f (x f = 0,r′f ) = ∫ R dr′f dridxiJ(x f = 0,r′f ;xi,ri)ρi(xi,ri) = ∫ R driρi(xi = 0,ri). To fulfill the last equality, the following must be accomplished ∫ R dr′f J(x f = 0,r′f ;xi,ri) = δ (xi). (3.28) This condition immediately discards ii) from the above combinations of deltas, thus we end up with cases i) and iii). For case i) TP reads: NI ∫ dr f δ (−βxi)e Σ = NI |β | √ π e1 δ (xi)e ( e2 2 4e1 −e3 ) r2 i , (3.29) 3.5. Representation of Gaussian quantum channels 43 thus, the relation between the coefficients assumes the form e2 2 4e1 − e3 = 0,d1 = 0,d2 = 0, (3.30) and the normalization constant NI = |β | √ e1 π with β 6= 0 and e1 > 0. For case iii) the trace-preserving condition reads NII ∫ dr f δ (γr f −ηri)δ (−βxi)e Σ = NII |βγ|δ (xi)e − ( e1( η γ ) 2+e2 η γ +e3 ) r2 i − ( d1 η γ +d2 ) ri . Thus, the following relation between en and dn coefficients must be fulfilled: e1 ( η γ )2 + e2 η γ + e3 = 0, d1 η γ +d2 = 0, (3.31) with γ,β 6= 0 and NII = |βγ|. In the particular case of η = 0, eq. (3.31) is reduced to e3 = d2 = 0. As expected from the analysis of limits above, we showed that δGQC’s admit quadratic terms in ri, j. 3.5.3 Complete positivity conditions Up to this point we have hermitian and trace preserving Gaussian operations; to derive the remaining CPTP conditions, it is useful to write its Wigner’s function and Wigner’s characteristic function. The representation of the Wigner’s charac- teristic function reads χ(~k) = tr [ ρD(~k) ] = exp [ −1 2 ~kT ( ΩσΩT ) ~k− ı(Ω〈x̂〉)T~k ] (3.32) and its relation with Wigner’s function: W (x) = ∫ R2 d~xe−ı~xTΩ~kχ ( ~k ) (3.33) = ∫ R eıpxdx 〈 r− x 2 ∣ ∣ ∣ ρ̂ ∣ ∣ ∣r+ x 2 〉 , (3.34) where ~k = (k1,k2) T , ~x = (r, p)T and h̄ = 1 (we are using natural units). Using the previous equations to construct Wigner and Wigner’s characteristic functions of the initial and final states, and substituting them in the equation 2.30, it is 44 Chapter 3. Representations of quantum channels straightforward to get the propagator in the Wigner’s characteristic function rep- resentation: J̃ ( ~k f ,~ki ) = ∫ R6 dΓK(~l)J(~v f ,~vi), (3.35) where the transformation kernel reads K(~l) = 1 (2π)3 e[ı(k f 2 r f −k f 1 p f −ki 2ri+ki 1 pi−pixi+p f x f )], with dΓ = d p f d pidx f dxidr f dri and~l = (p f , pi,x f ,xi,r f ,ri) T . By elementary in- tegration of eq. (3.35) one can show that for both cases J̃I,III ( ~k f ,~ki ) = δ ( ki 1 − α β k f 1 ) δ ( ki 2 −~φ T I,III ~k f ) ePI,III(~k f ), (3.36) where PI,III(~k f ) = ∑ 2 i, j=1 P (I,III) i j k f i k f j +∑ 2 i=1 P (I,III) 0i k f i with P (I,III) i j = P (I,III) ji . For case i) we obtain P (I) 11 =− ( ( α β )2( a3 + b2 3 4e1 ) + α β ( a2 + 1 2 b1b3 e1 ) +a1 + b2 1 4e1 ) , P (I) 12 =− ( α β b3 2e1 + b1 2e1 ) , P (I) 22 =− 1 4e1 . (3.37) For case iii) we have P (III) 11 =− ( ( α β )2 a3 + α β a2 +a1 ) , P (III) 12 = P (III) 22 = 0. (3.38) And for both cases we have P (I,III) 01 = ı ( α β c2 + c1 ) and P (I,III) 02 = 0. Vectors ~φ are given by ~φI = ( α β ( b4 − b3e2 2e1 ) − b1e2 2e1 +b2,− e2 2e1 )T , ~φIII = ( α β η γ b3 + α β b4 + η γ b1 +b2, η γ )T . (3.39) 3.5. Representation of Gaussian quantum channels 45 We are now in position to write explicitly the conditions for complete positivity. Having a Gaussian operation characterized by (T,N,~τ), the CP condition can be expressed in terms of the matrix C = N+ ıΩ− ıTΩTT, (3.40) where Ω = ( 0 1 −1 0 ) is the symplectic matrix. An operation A (T,N,~τ) is CP if and only if C ≥ 0 [Lin00, WPGP+12]. Applying the propagator on a test characteristic function, eq. (3.32), it is easy compute the corresponding tuples. For both cases we get: NI,III = 2 ( −P22 P12 P12 −P11 ) , ~τI,III = ( 0, ıP (I,III) 01 )T , (3.41) while for case i) matrix T is given by TI = ( e2 2e1 0 ~φI,1 −α β ) , (3.42) where ~φI,1 denotes the first component of vector ~φI, see eq. (3.39). The complete positive condition is given by the inequalities raised from the eigenvalues of ma- trix eq. (3.40): ± √ α2e2 2 +4αβe2e1 +4β 2e2 1 ( 4P (I) 12 2 + ( P (I) 11 −P (I) 22 )2 +1 ) 2βe1 ≥ P (I) 11 +P (I) 22 . (3.43) For case iii) matrix T is TIII = ( −η γ 0 ~φIII,1 −α β ) , (3.44) and complete positivity conditions read: ± √ (βγ −αη)2 +β 2γ2P (III) 11 2 βγ −P (III) 11 ≥ 0. (3.45) Note that in both cases the complete positivity conditions do not depend on ~φ . Chapter 4 Divisibility of quantum channels and dynamical maps Wine is sunlight, held together by water. Galileo Galilei In this chapter we introduce the formal definition of divisibility of quantum chan- nels, inspired by questioning how can we implement a given quantum channel via the concatenation of simpler channels. Later on we define further types of divisibility by adding extra conditions, such as channels being infinitesimal di- visible and channels belonging to one-parameter semigroups. These types are physically relevant since both lead to Markovian dynamical maps [ARHP14]. We additionally prove three theorems, which are the central contributions of this part of the work. Finally, a complete characterization of channels belonging to one- parameter semigroups that is given. 4.1 Divisibility of quantum maps A quantum channel E is said to be divisible if it can be expressed as the concate- nation of two non-trivial channels, Definition 10 (Divisibility). A linear map E ∈ C is divisible if there exists a de- composition, E = E2E1, (4.1) such that E1 and E2 are non-unitary channels, or E is unitary. 47 48 Chapter 4. Divisibility of quantum channels and dynamical maps Notice that this definition ensures that unitary channels are divisible, and that non- unitary channels must be divisible in non-unitary channels. This prevents one to consider simple changes of basis as a “division” of a given quantum operation. This type of divisibility, which is the most general and less restrictive one, defines a set that will be denoted by Cdiv. The set of indivisible channels is the comple- ment of Cdiv in C, therefore it will be denoted as Cdiv. Notice that this definition is different to the one given in Ref. [WC08] where unitary channels are excluded to be divisible. The concept of indivisible channels resembles the concept of prime numbers, uni- tary channels play the role of unity (which are not indivisible/prime), i.e. a com- position of indivisible and a unitary channel results in an indivisible channel. We now introduce three results from Ref. [WC08] that shall be used later. We only give the proof for the second for the sake of brevity. Theorem 11 (Full Kraus rank channels). Let E : T (H )→T (H ) be a quantum channel. If it has full Kraus Kraus rank, i.e. d2 with d = dim(H ), then it is divisible. An example of full Kraus rank channel is the total depolarizing channel ρ 7→ ✶/dim(H ), which maps every state into the maximal mixed one. Theorem 12 (Indivisible channels). Consider the set Cd of channels acting on the space of density matrices of d×d, i.e. E : T (H )→ T (H ) with d = dim(H ). The channel with minimal determinant, E0 ∈ Cd, is indivisible. Proof. To prove this we use the fact that channels with negative determinant ex- ist [WC08] (two examples are given below), and the property of monotonicity of the determinant. Let E ∈ C with detE < 0 and E = E2E1 an arbitrary division of E with E1,E2 ∈ C. The monotonicity of the determinant implies the following, |det(E2E1) |= |detE2||detE1| ≤ |detE1|. Assuming, without loss of generality that detE1 < 0 and detE2 > 0, we have that detE1 detE2 ≤ detE1. Multiplying both sides by −1 we arrive to |detE2||detE1| ≥ |detE1|. 4.1. Divisibility of quantum maps 49 Therefore, by monotonicity of the determinant, we have |detE2||detE1|= |detE1|, which implies that detE2 = 1, i.e. E2 is an unitary conjugation [WC08] and E has minimum determinant. By definition 10 E is indivisible. Two examples for the qubit case are the approximate NOT and the approximate transposition maps: ρ 7→ tr(ρ)✶+ρT 3 (approximate transposition), ρ 7→ tr(ρ)✶−ρ 3 (approximate NOT gate), (4.2) both have minimal determinant corresponding to −1/27, which can be computed from their matrix representation. Theorem 13 (Unital Kraus rank three channels). A unital qubit channel is indi- visible if and only if it has Kraus rank equal to three. This is a restricted version of theorem 23 of Ref. [WECC08], where authors proved the theorem for any qubit channel instead of only unital ones. Since their proof rely on the validity of the Lorentz normal decomposition for channels, we have written here a restricted version, where Lorentz normal form is equivalent to the special orthogonal normal form (see theorem 10 and its discussion). These results can be used immediately to identify the divisibility character of uni- tal qubit channels, see fig. 4.5. The faces of the tetrahedron (without edges) corre- spond to indivisible channels, in particular the center of every face corresponds to channels with minimal determinant. The body (full Kraus rank channels) contain divisible channels. 4.1.1 Subclasses of divisible maps Divisibility of quantum dynamical maps We motivate the extra conditions to define new types of divisibility on the con- cept of Markovian process. In subsection 2.2.1 we have introduced the definition of Markovian process and its consequences at the level of propagators of one- point probabilities, see eq. (2.24). Based on this, we introduce the concept of 50 Chapter 4. Divisibility of quantum channels and dynamical maps CP-divisibility of quantum dynamical maps, which is often used as definition of Markovianity in the quantum realm [ARHP14]. Definition 11 (CP-divisible quantum dynamical maps). Consider a quantum dy- namical map E(t,0) : T (H ) → T (H ) with t ∈ R +. It is CP-divisible in the interval [0, t]⊂ R + if for every decomposition of the form E(t,0) = E(t,s)E(s,0), E(t,s) is a quantum channel for every s ∈ (0, t). A remarkable theorem on CP-divisible maps is the following [Kos72b, Kos72a, Gor76, Lin76, ARHP14], Theorem 14 (Gorini-Kossakowski-Susarshan-Lindblad). An operator Lt is the generator of a CP-divisible process if and only if it can be written in the following form: Lt [ρ] =−i[H(t),ρ]+∑ i, j Gi j(t) ( Fi(t)ρF † j (t)− 1 2 {F † j (t)Fi(t),ρ} ) , (4.3) where G is hermitian and positive semidefinite, H(t),Fk(t)∈C d×d are time-dependent operators acting on H , with H(t) hermitian for every t ∈R +, and d = dim(H ). In Ref. [RH12] a proof is given starting from the Kraus representation of quantum dynamical maps and the definition of CP-divisibility. Here we will give a simpler proof resorting to theorem 8. Proof. Notice that for each time t we can define the “instant” map E(t+ε,t)[ρ] = ρ +εLt [ρ]+O(ε2), with ε > 0, therefore the hermiticity preserving of Lt follows from the hermiticity preserving of E(t,0). Also note that we can always choose the same traceless basis, {Fi}d2−1 i=0 , to write eq. (4.3), such that the time dependence is dropped only in G(t) ∈ C d2×d2 and H(t). By theorem 8, Lt has the form stated in eq. (4.3), the only thing that remains to prove is that G(t) ≥ 0 for every t. To do this we construct the Choi-Jamiołkowski matrix of the instant map, τt,ε = ω+ε (idd2 ⊗Lt) [ω]+O(ε2). We remind the reader that ω = |Ω〉〈Ω|, where |Ω〉 is the Bell state between two copies of Cd . Now we test positive-semidefinitiveness of τt,ε , 〈ϕ|τt,ε |ϕ〉= 〈ϕ|Ω〉〈Ω|ϕ〉+ ε〈ϕ|(idd2 ⊗Lt) [|Ω〉〈Ω|]|ϕ〉+O(ε2)≥ 0, 4.1. Divisibility of quantum maps 51 ∀|ϕ〉 ∈ C d2 . The inequality always holds for any 〈ϕ|Ω〉 6= 0 and ε > 0. For 〈ϕ|Ω〉 = 0 we have that for ε > 0 the inequality 〈ϕ|(idd2 ⊗Lt) [|Ω〉〈Ω|]|ϕ〉 ≥ 0 must be accomplished, i.e. ω⊥τLω⊥ ≥ 0 (conditional complete positivity). There- fore by proposition 2, one has that G(t)≥ 0. Analogously to CP-divisible processes, if we relax the condition of the interme- diate maps to be PTP (and not necessarily CPTP), we arrive to the following definition: Definition 12 (P-divisible quantum dynamical maps). Consider a quantum dy- namical map E(t,0) : T (H )→T (H ) with t ∈R +. It is P-divisible in the interval [0, t]⊂ R + if for every decomposition of the form E(t,0) = E(t,s)E(s,0), E(t,s) belongs to PTP for every s ∈ (0, t). Unfortunately, to the best of our knowledge, there doesn’t exist a statement similar to theorem 14, nor a simple test of P-divisibility. But for certain types of genera- tors of dynamical maps, conditions for P-divisibility were derived in Ref. [CDG19]. Divisibility of quantum channels Let us discuss these two types of divisibility but now from a statical point of view. First notice that instant operations E(t+ε,t) are arbitrarily close to the identity map as ε → 0+, for both P-divisible and CP-divisible processes. In other words, they are infinitesimal. Consider now the idea of quantum channels divisible in infinitesimal parts, i.e. what is given this time is a quantum channel instead of a dynamical map. This idea motivates the following definition [WC08], Definition 13 (Infinitesimal divisible channels in CPTP). Let LCP be the set con- taining operations E ∈ C with the property that ∀ε > 0 there exist a finite number of channels Ei ∈ C such that |Ei− id|< ε and E = ∏i Ei, see fig. 4.1. It is said that a channel is infinitesimal divisible if it belongs to the closure of LCP. This set is denoted as CCP. The necessity of the closure can be motivated using the following example. Con- sider the qubit channel defined as follows: E∞ : ( ρ00 ρ01 ρ∗ 01 ρ11 ) 7→ ( ρ00 0 0 ρ11 ) . (4.4) 52 Chapter 4. Divisibility of quantum channels and dynamical maps E ∈ C CP E = E 1 E 2 E Nε id E Ei Figure 4.1: Diagrammatic decomposition of channels belonging to LCP whose closure is CCP, see definition 13. We show the circuit repre- senting the decomposition of E into channels (left) arbitrarily close to the identity map (right). This channel is singular, i.e. does not belong to LCP. Now observe that using the dynamical process, Et , given in example 1, one can get arbitrarily close to E∞ when t → ∞, i.e. E∞ = limt→∞ Et . Note that Et ∈ LCP for every t ∈ R +, see theorem 14, therefore E∞ is an accumulation point of LCP. Thus, the closure is taken to define infinitesimal divisible channels, to include channels such as E∞. Up to this point we have shown that CP-divisible processes are infinitesimal di- visible, i.e. CP-divisible processes parametrize families of channels belonging to CCP. In Ref. [WECC08], authors have shown that channels in CCP can always be implemented with CP-divisible processes. This can be roughly shown as follows. Since C is connected, we can understand infinitesimal channels as the ending point of an arbitrarily small curve parametrized by t, i.e. channels Ei in definition 13 can be written approximately as Ei ≈ id+Li ≈ exp(Li). We have shown that Li has Lindblad form, see theorem 14. Therefore we have that if E ∈ CCP, it can be written as E = ∏ i eLi . Therefore E can be implemented using a CP-divisible dynamical processes. Bounds of the convergence ratio using channels of the form exp(Li) instead of general in- finitesimal channels, are computed in Ref. [WECC08]. Analogous to infinitesimal divisible channels in C and its relation with CP-divisible processes, one can also define the following set involving PTP maps. Definition 14 (Infinitesimal divisible channels in PTP). Let LP be the set con- 4.1. Divisibility of quantum maps 53 E ∈ C P E = E 1 E 2 E Nε id E Ei CPTP PTP Figure 4.2: Diagramatic decomposition of channels belonging to LP which closure is CP, see definition 14. At the left we show the circuit rep- resenting the decomposition of E into channels arbitrarily close to the identity map, see figure at the right. In contrast to figure 4.1, note that infinitesimal channels can be outside the set of CPTP maps, but inside PTP. taining operations E ∈ C with the property that ∀ε > 0 there exist a finite number of channels Ei ∈ PTP such that |Ei − id|< ε and E = ∏i Ei, see fig. 4.2. It is said that a channel is infinitesimal divisible in PTP if it belongs to the closure of LP. This set is denoted as CP. Infinitesimal divisibility in PTP maps is interesting since this kind of maps can arise in settings where the system is initially correlated with its surroundings, or if the operation is correlated with the initial state [CTZ08]. Infinitesimal divisible (either in CPTP and PTP) channels have non-negative de- terminant due to its continuity [WECC08]. To see this note that channels arbitrar- ily close to the identity map have positive determinant; and by its multiplicative property, the channel resulting from the concatenation of infinitesimal channels has non-negative determinant. Proposition 3 (Determinant of infinitesimal divisible channels). If a quantum map E belongs either to CP or CCP, then detE ≥ 0. It turns out that a non-negative determinant is a sufficient condition for a channel to be infinitesimal divisible in PTP, see theorem 25 of Ref. [WECC08]. Other interesting type of divisibility that in turn forms a subset of CCP is the fol- lowing [WECC08, Den89]. 54 Chapter 4. Divisibility of quantum channels and dynamical maps E ∈ C ∞ E = En En En n times Figure 4.3: Diagrammatic decomposition of channels belonging to C∞, see definition 15. This set contains channels for which every n-root exist and is a valid quantum channel, denoted in the circuit as En. Definition 15 (Infinitely divisible channels). A quantum channel E is infinitely divisible if ∀n ∈ Z + ∃En ∈ C such that E = (En) n . This set is denoted as C∞, see fig. 4.3. This set contains channels for which every n-root exists and is a valid quantum channel. Denisov has shown in [Den89] that infinitely divisible channels can be written as E = E0 exp(L), with L a Lindblad generator, and an E0 idempotent operator that fulfills E0LE0 = E0L. In this work we will prove that every infinitely divisible Pauli channel has the simple form exp(L). Let us now introduce the most restricted type of divisibility studied in this work, Definition 16 (Channels belonging to one-parameter semigroups (L-divisibility)). Let LL be the set containing non-singular operations E ∈ C, such that there exist at least one logarithm, denoted as L = logE , such that L[ρ] = i[ρ,H]+∑ i, j Gi j ( FiρF † j − 1 2 {F † j Fi,ρ} ) , (4.5) where H and G are hermitian with G ≥ 0, and {Fi}i are bounded operators acting on T (H ). It is said that a channel is L-divisible if it belongs to the closure of LL. This set is denoted as CL. Analogous to the relation of CP-divisible dynamical maps and its relations with CCP, time-independent Markovian processes form families of L-divisible chan- nels. The converse is true by definition. One of the principal objectives of this work is to construct a test to check whether a given channel belongs to CL or not. 4.1. Divisibility of quantum maps 55 C C L C div C P C CP C ∞ C P ∩ Cdiv Unitary channels Figure 4.4: Scheme illustrating the different sets of quantum channels for a given dimension. In particular, the inclusion relations presented in eq. (4.6) are depicted. 4.1.2 Relation between channel divisibility classes Let us summarize the introduced divisibility sets and the relations between them. Since channels belonging to CCP can be implemented with time-dependent Lind- blad master equations, and time-independent ones are a particular case of time de- pendent ones, we have CL ⊂CCP. Now, since infinitely divisible channels have the form E0 exp(L), channels with form exp(L) are a particular case of C∞, therefore CL ⊆ C∞. Also, given that CPTP maps are also PTP, then CCP ⊂ CP. Finally, ev- ery set except CP is subset of Cdiv, given that an infinitesimal divisible channels in PTP is not necessarily divisible in CPTP channels. In summary we have [WC08], C∞ ⊂ CCP ⊂ Cdiv ⊆ CL ⊂ CCP ⊂ CP . (4.6) The intersection of CP and Cdiv is not empty since CCP ⊂ Cdiv and CCP ⊂ CP, later on we will investigate if CP ⊆ Cdiv or not. A scheme of the inclusions is given in fig. 4.4. 56 Chapter 4. Divisibility of quantum channels and dynamical maps 4.2 Characterization of L-divisibility Deciding L-divisibility is equivalent to proving the existence of a hermiticity pre- serving generator, which additionally fulfills the ccp condition, see proposition 2. To prove hermiticity preserving we recall that every HP operator has a real ma- trix representation when choosing an hermitian basis, see subsection 3.3.1. Since quantum channels preserve hermiticity, the problem is reduced to find a real loga- rithm log Ê given a real matrix Ê , where the hat means that E is written using an hermitian basis. This problem was already solved by Culver [Cul66] who charac- terized completely the existence of real logarithms of real matrices. In this work we restrict the analysis to diagonalizable channels. The results can be summarized as follows. Theorem 15 (Existence of hermiticity preserving generator). A non-singular matrix with real entries Ê has a real generator (i.e. a log Ê with real entries) if and only if the spectrum fulfills the following conditions: i) negative eigenval- ues are even-fold degenerate; ii) complex eigenvalues come in complex conjugate pairs. We now discuss the multiplicity of the solutions of log Ê and its parametriza- tion, as finding an appropriate one is essential to test for the ccp condition. If Ê has positive degenerate, negative, or complex eigenvalues, its real logarithms are not unique, and are spanned by real logarithm branches [Cul66]. The latter are defined using the real quaternion, which coincides with iσy, using the fact that ✶ = exp(iσy2πk), with k ∈ Z. In case of having negative eigenvalues, it turns out that real logarithms always have a continuous parametrization, in addition to real branches due to the freedom of the Jordan normal form transformation matri- ces [Cul66]. To compute the logarithm given a real representation of E , i.e. Ê , we calculate its Jordan normal form, J, such that Ê = wJw−1 = w̃Jw̃−1, where w = w̃K and K belongs to a continuum of matrices that commute with J [Cul66]. In the case of diagonalizable matrices, if there are no degeneracies, K commutes with log(J). In the case of having degeneracies, matrix K is responsible of the continuous parametrization of the logarithm. We compute explicitly the logarithms for the case of Pauli channels in section 4.3.4. 4.3. Divisibility of unital qubit channels 57 4.3 Divisibility of unital qubit channels We will apply various of the results from the literature [WECC08] to decide if a given unital qubit channel belongs to CL, CCP and/or CP. The non-unital case will be discussed later. Before starting with the characterization let us point out the following. From the definition of divisibility, the concatenation of a given channel with unitary conjugations (which are infinitesimal divisible) do not change its divisibility char- acter, except for L-divisibility. In addition to this, since unitary conjugations are infinitesimal divisible, they do not change the infinitesimal divisible character ei- ther. We can summarize this in the following, Proposition 4 (Divisibility of special orthogonal normal forms). Let E a qubit quantum channel and D its special orthogonal normal form, E belongs to CX if and only if D does, where X = {“Div”, “P”, “CP”}. This proposition is in fact a consequence of theorem 17 of Ref. [WECC08]. No- tice that this result does not apply for CL since conjugating with unitaries breaks the implementability by means of time-independent Lindblad master equations. Thus, if a channel belongs to CL, unitary conjugations can bring it to CInf \CL and vice versa. Therefore, by proposition 4 and the theorem 9, to study CP and CCP of unital qubit channels, it is enough to study Pauli channels. 4.3.1 Channels belonging to C div Divisibility in CPTP of unital qubit channels is completely characterized by means of theorem 13. Therefore the only indivisible channels lie in the faces of the tetrahedron (without the edges), see fig. 4.5. 4.3.2 Channels belonging to C P Recalling that all unital qubit channels belonging to CP have non-negative deter- minant [WC08], and using special orthogonal normal forms, see theorem 9, the condition in terms of its parameters is given by λ1λ2λ3 ≥ 0. (4.7) 58 Chapter 4. Divisibility of quantum channels and dynamical maps This set is the intersection of the tetrahedron with the octants where the product of all λ s is positive. In fact, it consists of four triangular bipyramids starting in each vertex of the tetrahedron and meeting in its center, see fig. 4.5. Let us study the intersection of this set with the set of unital entanglement-breaking (EB) channels [ZB05], see definition 5. In the case of unital qubit channels, the set entanglement-breaking channels is an octahedron that lie inside the tetrahedron of unital qubit channels, see fig. 4.6. The inequalities that define such octahedron are the following, λi ± (λ j +λk)≤ 1, (4.8) with i, j and k all different [ZB05], together eq. (3.11). It follows that unital qubit channels that are not achieved by P-divisible dynamical maps are necessarily entanglement-breaking (see fig. 4.6 and fig. 4.9). In fact this holds for general qubit channels, see section 4.4. 4.3.3 Channels belonging to C CP To characterize CP-divisible channels it is useful to consider the Lorentz normal form for channels, see theorem 10. A remarkable property of the Lorentz normal decomposition is that it preserves the infinitesimal divisible character of E , see Corollary 13 of [WC08]. To use it, we resort to theorem 24 of Ref. [WC08]. Due to the mentioned drawback of Lorentz normal forms, see appendix B, we must modify such to theorem to a restricted class of channels. Theorem 16 (Restricted characterization of channels belonging to CCP). A qubit channel E with diagonal Lorentz normal form belongs to CCP if and only if i) the rank of the form is smaller than three or ii) s2 min ≥ s1s2s3 > 0, where smin is the smallest of s1, s2 and s3, see theorem 10. For non-unital Kraus deficient channels, the pertinent theorems are based on non- diagonal Lorentz normal forms [VV02, WC08]. According to our appendix B such results should be reviewed and are out of the scope of this work. Notice that the Lorentz normal form coincides with the special orthogonal nor- mal form for unital qubit channels. Therefore, by theorem 16, unital channels belonging to CCP are non-singular with 0 < λ1λ2λ3 ≤ λ 2 min , (4.9) or singular with a matrix rank less than three. They determine a body that is symmetric with respect to permutation of Pauli unitary channels (i.e. in λ j), hence, 4.3. Divisibility of unital qubit channels 59 λ1 λ2 λ3 Figure 4.5: Tetrahedron of Pauli channels, see fig. 3.1. The bipyramid in blue correspond to channels with λi > 0 ∀i, i.e. channels of the positive octant belonging to CP. The whole set CP includes other three bipyramids corresponding to the other vertexes of tetrahedron. This implies that CP enjoys the symmetries of the tetrahedron, see eq. (4.7). The faces of the bipyramids matching the corners of the tetrahedron are subsets of the faces of the tetrahedron, i.e. contain Kraus rank three channels. Such channels are both CP and Cdiv, showing that the intersection shown in fig. 4.4 is not empty. 60 Chapter 4. Divisibility of quantum channels and dynamical maps λ1 λ2 λ3 Figure 4.6: Tetrahedron of Pauli channels with the octahedron of entangle- ment breaking channels shown in red, see eq. (4.8). The blue pyramid inside the octahedron is the intersection of the bipyramid shown in 4.5, with the octahedron. The complement of the intersections of the four bipyramids forms the set of divisible but not infinitesimal divisible chan- nels in PTP. Thus, a central feature of the figure is that the set Cdiv\CP is always entanglement-breaking, but the converse is not true. 4.3. Divisibility of unital qubit channels 61 λ1 λ2 λ3 Figure 4.7: Tetrahedron of Pauli channels with part of the set of CP-divisible, see eq. (4.9), but not L-divisible channels (CCP\CL) shown in purple. The whole set CCP is obtained applying the symmetry transformations of the tetrahedron to the purple volume. the set of CCP of Pauli channels possesses the symmetries of the tetrahedron. The set CCP\CL is plotted in fig. 4.7. 4.3.4 L-divisible unital qubit channels We restrict our analysis of L-divisibility for two particular sets of unital channels, Pauli channels and a family with complex eigenvalues that will be introduced later. Pauli channels with non-degenerate positive eigenvalues First let us now derive the conditions for L-divisibility of Pauli channels with pos- itive eigenvalues λ1,λ2,λ3 (λ0 = 1). The logarithm of D , induced by the principal logarithm of its eigenvalues is L = Kdiag(0, logλ1, logλ2, logλ3)K −1 , (4.10) 62 Chapter 4. Divisibility of quantum channels and dynamical maps which is real (hermiticity preserving). In case of no-degeneration the dependency on K vanishes and L is unique. In such case the ccp conditions, see theorem 2, read logλi − logλ j − logλk ≥ 0 ⇒ λi λ jλk ≥ 1 (4.11) for all combinations of mutually different i, j,k. This set (channels belonging to CL with positive eigenvalues) forms a three dimensional manifold, see fig. 4.8. Pauli channels with degenerate positive eigenvalues In case of degeneration, let us label the eigenvalues η , λ and λ . In this case, the real solution for L is not unique and is parametrized by real branches in the degenerate subspace and by the continuous parameters of K [Cul66]. Let us study the principal branch with K = ✶. Eq. (4.11) is then reduced to λ 2 ≤ η ≤ 1 . (4.12) Therefore, if this inequalities are fulfilled, the generator has Lindblad form. If not, then a priori other branches can fulfill ccp condition and consequently have a Lindblad form. Thus, Eq. (4.12) provides a sufficient condition for the channel to be in CL. We will prove it is also necessary. The complete positivity condition requires η ,λ ≤ 1, thus, it remains to verify only the condition λ 2 ≤ η . It holds trivially for the case λ ≤ η . If η ≤ λ , then this condition coincides with the CP-divisibility condition from eq. (4.9). Since CL implies CCP the proof is completed. In conclusion, the condition in eq. (4.11) is a necessary and sufficient for a given Pauli channel with positive eigenvalues to belong to CL. Let us stress that the obtained subset of L-divisible channels does not possess the tetrahedron symmetries. In fact, composing D with a σz rotation Uz = diag(1,−1,−1,1) results in the Pauli channel D ′ = diag(1,−λ1,−λ2,λ3). Clearly, if λ j are pos- itive (D is L-divisible), then D ′ has non-positive eigenvalues. Moreover, if all λ j are different, then D ′ does not have any real logarithm, therefore, it cannot be L-divisible. In conclusion, the set of L-divisible unital qubit channel is not symmetric with respect to tetrahedron symmetries. 4.3. Divisibility of unital qubit channels 63 Pauli channels with negative eigenvalues In what follows we will investigate the case of negative eigenvalues. Theorem 15 implies that that eigenvalues have the form (modulo permutations) η ,−λ ,−λ , where η ,λ > 0. The corresponding Pauli channels are Dx = diag(1,η ,−λ ,−λ ), Dy = diag(1,−λ ,η ,−λ ), Dz = diag(1,−λ ,−λ ,η), thus forming three two-dimensional regions inside the tetrahedron. Take, for in- stance, Dz that specifies a plane (inside the tetrahedron) containing I, σz and com- pletely depolarizing channel N = diag(1,0,0,0). The real logarithms for this case are given by L = K     0 0 0 0 0 log(λ ) (2k+1)π 0 0 −(2k+1)π log(λ ) 0 0 0 0 log(η)     K−1, (4.13) where k ∈ Z and K, as mentioned above, belongs to a continuum of matrices that commute with Dz. Note that L is always non-diagonal. For this case (similarly for Dx and Dy) the ccp condition reduces again to conditions specified in Eq. (4.12). Using the same arguments one arrives to more general conclusion: Theorem 17 (L-divisibility of Pauli channels). Let E be a non-singular Pauli channel. It belongs to CL if and only if its non-trivial eigenvalues fulfill λi λ jλk ≥ 0 (4.14) for i, j and k mutually different. This is one of the central results of this work, and it implies that for testing L- divisibility of Pauli channels, it is enough to consider the principal real logarithm branch and K = ✶. The singular cases are included in the closure of channels fulfilling eq. (4.14). The set of L-divisible Pauli channels is illustrated in fig. 4.8. To get a detailed picture of the position and inclusions of the divisibility sets, we illustrate in fig. 4.9 two slices of the tetrahedron where different types of divisi- bility are visualized. Notice the non-convexity of the considered divisibility sets. 64 Chapter 4. Divisibility of quantum channels and dynamical maps λ1 λ2 λ3 Figure 4.8: Tetrahedron of Pauli channels with the set of L-divisible channels (or equivalently infinitely divisible, see Theorem 19) shown in green, see equations (4.11) and (4.12). The solid set corresponds to channels with positive eigenvalues, and the 2D sets correspond to the negative eigenvalue case. The point where the four sets meet correspond to the total depolarizing channel. Notice that this set does not have the sym- metries of the tetrahedron. 4.3. Divisibility of unital qubit channels 65 λ1 λ2λ3  CL  CCP\CL  CP\CCP  Cdiv\CP  Cdiv  EB boundary Figure 4.9: We show two slices of the unitary tetrahedron (figure in the left) determined by ∑i λi = 0.4 (shown in the center) and ∑i λi = −0.4 (shown in the right). The non-convexity of the divisibility sets can be seen, including the set of indivisible channels. The convexity of sets C and entanglement breaking channels can also be noticed in the slices. A central feature is that the set Cdiv \CP is always inside the octahedron of entanglement breaking channels. 66 Chapter 4. Divisibility of quantum channels and dynamical maps Family of unital channels with complex eigenvalues To give an insight to the case of complex eigenvalues, consider the following family of channels with real logarithm, written in the Pauli basis, Ecomplex =     1 0 0 0 0 c 0 0 0 0 a −b 0 0 b a     . (4.15) The latter has complex eigenvalues a± ib and a real one c > 0, together with the trivial eigenvalue equal to 1. Its real logarithm is given by, L = K log ( Ecomplex ) k K−1 = K     0 0 0 0 0 log(c) 0 0 0 0 log(|z|) arg(z)+2πk 0 0 −arg(z)−2πk log(|z|)     K−1 with z = a+ ib. The non-diagonal block of the logarithm has the same structure of Ecomplex, so K also commutes with log(Ecomplex)k, leading to a countable paramet- ric space of hermitian preserving generators. The ccp condition, see proposition 2, is reduced to a2 +b2 ≤ c ≤ 1. (4.16) Note that it does not depend on k and the second inequality is always fulfilled for CPTP channels. The set containing them is shown in fig. 4.10. 4.3.5 Relation of L-divisibility with other divisibility classes Consider a Pauli channel with 0 < λmin = λ1 ≤ λ2 ≤ λ3 < 1, thus the condition λ1λ2 ≤ λ3 trivially holds. Since λ1λ2 ≤ λ1λ3 ≤ λ2λ3 ≤ λ2, it follows that λ1λ3 ≤ λ2, thus, two (out of three) L-divisibility conditions hold always for Pauli chan- nels with positive eigenvalues. Moreover, one may observe that CP-divisibility condition eq. (4.9) reduces to one of L-divisibility conditions λ2λ3 ≤ λ1. In con- clusion, the conditions of CP-divisibility and L-divisibility for Pauli channels with positive eigenvalues coincide, thus, in this case CCP implies CL. Concatenating (positive-eigenvalues) L-divisible Pauli channels with Dx,y,z, one can generate the whole set of CCP Pauli channels. In other words, Dx,y,z brings the body (with vertex in id) shown in fig. 4.8 to the bodies shown in fig. 4.7 (with vertexes x,y,z). Therefore we can formulate the following theorem: 4.3. Divisibility of unital qubit channels 67 λ1 λ2 λ3 Figure 4.10: Tetrahedron of Pauli channels, with qubit unital L-divisible channels of the form Êcomplex (see main text). Note that the set does not have the symmetries of the tetrahedron. Theorem 18 (Infinitesimal divisible unital channels). Let E CP unital be an arbitrary infinitesimal divisible unital qubit channel. There exists at least one L-divisible Pauli channel Ẽ , and two unitary conjugations U1 and U2, such that E CP unital = U1Ẽ U2 . Notice that if E CP unital is invertible, Ẽ = eL. Let us continue with another equivalence relation valid for Pauli channels. In general, CL ⊂ C∞; however, for Pauli channels these two subsets coincide. Theorem 19 (Infinitely divisible Pauli channels). The set of L-divisible Pauli channels is equivalent to the set of infinitely divisible Pauli channels. Proof. A channel is infinitely divisible if and only if it can be written as E0eL, where E0 is an idempotent channel satisfying E0LE0 = E0L and L has Lindblad form, see definition 15. The only idempotent qubit channels are contractions of the Bloch sphere into single points, diagonalization channels Ediag transforming Bloch sphere into a line connecting a pair of basis states, and the identity channel. Among the single-point contractions, the only one that is a Pauli channel is the 68 Chapter 4. Divisibility of quantum channels and dynamical maps contraction of the Bloch sphere into the complete mixture; let us call it N . Notice that E = N eL = N for all L. The channel N belongs to the closure of CL, because a sequence of channels eLn with L̂n = diag(0,−n,−n,−n) converges to ˆN in the limit n → ∞. For the case of E0 being the identity channel we have E = eL, thus, trivially such infinitely divisible channel E is in CL too. It remains to analyze the case of diagonalization channels. First, let us note that the matrix of eL̂ is necessarily of full rank, since detÊ 6= 0. It follows that the matrix Ê = ÊdiageL̂ has rank two as Êdiag is a rank two matrix, thus, it takes one of the following forms Ê λ x = diag(1,λ ,0,0), Ê λ y = diag(1,0,λ ,0), Ê λ z = diag(1,0,0,λ ). The infinitely divisibility implies λ > 0 in order to keep the roots of λ real. In what follows we will show that Êz belongs to (the closure of) CL. Let us define the channels Ê λ ,ε z = diag(1,ε,ε,λ ) with ε > 0. The complete positivity and ccp conditions translate into the inequalities ε ≤ 1+λ 2 and ε2 ≤ λ , respectively; therefore one can always find an ε > 0 such that Ê λ ,ε z is a L-divisible channel. If we choose ε = √ λ/n with n ∈ Z +, the channels Êz,n = diag ( 1, √ λ/n, √ λ/n,λ ) form a sequence of L-divisible channels converging to Ê λ z when n → ∞. The analogous reasoning implies that Ê λ x , Ê λ y ∈ CL too. Let us note that one parameter family Ez are convex combinations of the complete diagonalization channel Ê 1 z = diag(1,0,0,1) and the complete mixture contraction ˆN . This completes the proof. Finally, let us remark that using the theorem 13 we conclude that the intersection CP ∩Cdiv depicted in fig. 4.4 is not empty. To show this, notice that there are channels with positive determinant inside the faces (i.e. CP but not Cdiv), for example diag ( 1, 4 5 , 4 5 , 3 5 ) . Therefore we conclude that up to unitaries, CP ∩Cdiv corresponds to the union of the four faces faces of the tetrahedron minus the faces of the octahedron that intersect with the faces of the tetrahedron, see fig. 4.6. We have to remove such intersection since it corresponds to channels with negative determinant, and thus not in CP. 4.4 Non-unital qubit channels Similar to unital channels, using theorem 4 we are able to characterize Cdiv, CP and CCP by studying special orthogonal normal forms. Such channels are char- acterized by ~λ and ~τ , see eq. (3.9). Thus, we can study if a channel is Cdiv by computing the rank of its Choi matrix, see theorem 11. For this case algebraic equations are in general fourth order polynomials. In fact, in Ref. [RPZ18] a con- dition in terms of the eigenvalues and ~τ is given. For special cases, however, we 4.4. Non-unital qubit channels 69 can obtain compact expressions, see fig. 4.11. The characterization of CP is given by again by the condition λ1λ2λ3 ≥ 0. CCP is tested, for full Kraus rank non-unital channels, using theorem 16, the calculation of si’s is done using the algorithm pre- sented in Ref. [VDD01]. For the characterization of CL we use theorem 15 and evaluate numerically the cpp condition. We can plot illustrative pictures even though the whole space of qubit channels has 12 parameters. This can be done using special orthogonal normal forms and fixing ~τ , exactly in the same way as the unital case. Recall that unitaries only modify CL, leaving the shape of other sets unchanged. CPTP channels are represented as a volume inside the tetrahedron presented in fig. 4.5, see fig. 4.11. In the later figure we show a slice corresponding to ~τ = (1/2,0,0)T . Indeed, it has the same structure of the slices for the unital case, but deformed, see fig. 4.9. A difference with respect to the unital case is that L-divisible channels with negative eigenvalues (up to unitaries) are not completely inside CP-divisible channels. A part of them are inside the CP channels. A central feature of Figs. 4.9 and 4.11 is that the set Cdiv \CP is inside the convex slice of the set of entanglement breaking channels (deformed octahedron). Indeed, we can proof the following theorem. Theorem 20 (Entanglement-breaking channels and divisibility). Consider a qubit channel E . If det Ê < 0, then E is entanglement-breaking, i.e. all qubit channels outside CP are entanglement breaking. Before introducing the proof, let us first show that the proper orthochronous Lorentz transformations present in the Lorentz normal decomposition for chan- nels, see sec. 3.4.2, correspond to 1wSLOCC at the level of their Choi-Jamiołkowski state. Consider a channel E and its Lorentz normal form Ẽ given by Ẽ = αF2E F1, (4.17) where Fi : ρ 7→ XiρX † i , with Xi ∈ SL(2,C), i = 1,2, and α is a constant that must be included for Ẽ to be trace preserving, We showed already that SL(2,C) is a double cover of SO+(3,1), i.e. Fi’s correspond to the proper orthochronous Lorentz transformations of the decomposition. Now let us compute the Choi-Jamiołkowski state of Ẽ , τ̃ , using the Kraus decom- 70 Chapter 4. Divisibility of quantum channels and dynamical maps λ1 λ2 λ3 Figure 4.11: (left) Set of non-unital unital channels up to unitaries, defined by ~τ = (1/2,0,0), see eq. (3.9). This set lies inside the tetrahedron. For this particular case the CP conditions reduce to the two inequali- ties 2± 2λ1 ≥ √ 1+4(λ2 ±λ3)2. A cut corresponding to ∑i λi = 0.3 is presented inside and in the right, see fig. 4.9 for the color coding. The structure of divisibility sets presented here has basically the same structure as for the unital case except for CL. A part of the channels with negative eigenvalues belonging to CL lies outside CCP \CL, see green lines. As for the unital case a central feature is that the chan- nels in Cdiv \CP are entanglement breaking channels. Channels in the boundary are not characterized due to the restricted character of Theo- rem 10. 4.4. Non-unital qubit channels 71 position of E [Wol11], τ̃ = α (id2 ⊗F2E F1) [ω] = α ∑ i (✶⊗X2)(✶⊗Ki)(✶⊗X1) |Ω〉〈Ω| ( ✶⊗X † 1 )( ✶⊗K † i )( ✶⊗X † 2 ) = α ∑ i ( XT 1 ⊗X2 ) (✶⊗Ki) |Ω〉〈Ω|) ( ✶⊗K † i )( X1 ⊗X † 2 ) = α(XT 1 ⊗X2)τ(X T 1 ⊗X2) †, (4.18) where {Ki}i are a choice of Kraus operators of E and τ = ∑ i (✶⊗Ki) |Ω〉〈Ω| ( ✶⊗K † i ) its Choi-Jamiołkowski matrix. Here, |Ω〉 is the Bell state between two copies of the system, in this case a qubit, for which the identity A⊗ ✶|Ω〉 = ✶⊗ AT|Ω〉 holds. It can be observed that eq. (4.18) has exactly the form of the normalized 1wSLOCC scheme, where α turns to be the normalization constant, see eq. (2.27), i.e. tr τ̃ = 1. That’s why we have introduced it at the first place. Now let us proceed with the proof of theorem 20, Proof. Let E be a qubit channel with negative determinant and Ê its matrix rep- resentation using the Pauli basis, see eq. (3.7). Recall that the matrix R defining the Choi-Jamiołkowski state of E , τE = 1 4 3 ∑ jk R jkσ j ⊗σk, and Ê are related by R = Ê ΦT, where ΦT = diag(1,1,−1,1). It follows immediately that R has positive determi- nant, detR =−det Ê > 0, since detΦT =−1. Using the aforementioned Lorentz normal decomposition for matrix R, we have R = LT 1 R̃L2 where detL1,2 > 0 and det R̃ > 0. Stressing that transformations L1,2 correspond to 1wSLOCC (see eq. (4.18)), then R̃ parametrizes an unnormalized two-qubit state. 72 Chapter 4. Divisibility of quantum channels and dynamical maps Let us first discuss the case when R̃ is diagonal. The channel corresponding to R̃ (in the Pauli basis) is Ĝ = R̃ΦT/R̃00, where R00 = tr R̃ = trτG . Since R̃ is diagonal, then G is a Pauli channel with det Ĝ < 0. A Pauli channel has a negative determinant, if either all λ j are negative, or exactly one of them is negative. In Ref. [ZB05] it has been shown that the set of channels with λ j < 0 ∀ j are entanglement breaking channels. Now, using the symmetries of the tetrahedron, one can generate all channels with negative determinant by concatenating this set with the Pauli rotations. Therefore every Pauli channel with negative determinant is entanglement breaking, thus, τG is separable. Given that LOCC operations can not create entanglement [HHHH09], we have that τE is separable, therefore E is entanglement breaking. The case when R̃ is non-diagonal corresponds to Kraus deficient channels (the matrix rank of 3.17 is at most 3). This case can be analyzed as follows. Since the neighborhood of any Kraus deficient channel with negative determinant con- tains full Kraus rank channels, by continuity of the determinant such channels have negative determinant too. The last ones are entanglement breaking since full Kraus rank channels have diagonal Lorentz normal form. Therefore, by continu- ity of the concurrence [ZB05], Kraus deficient channel with negative determinant are entanglement breaking. 4.5 Divisibility transitions and examples with dynamical processes The aim of this section is to use illustrative examples of quantum dynamical pro- cesses to show transitions between divisibility types of the instantaneous channels. From the slices shown above (see figures 4.9 and 4.11) it can be noticed that every transition between the studied divisibility types is permitted. This is due to the existence of common borders between all combinations of divisibility sets; we can think of any continuous line inside the tetrahedron [FPMZ17] as describing some quantum dynamical map. We analyze two examples. The first is an implementation of the approximate NOT gate, ANOT throughout a specific collision model [RFZB12]. The second is the well known setting of a two-level atom interacting with a quantized mode of an optical cavity [HR06]. We define a simple function that assigns a particular value 4.5. Divisibility transitions and examples with dynamical processes 73 to a channel Et according to divisibility hierarchy, i.e. δ [E ] =        1 if E ∈ CL , 2/3 if E ∈ CCP \CL , 1/3 if E ∈ CP \CCP , 0 if E ∈ C\CP . (4.19) A similar function can be defined to study the transition to/from the set of entanglement- breaking channels, i.e. χ[E ] = { 1 if E is entanglement breaking , 0 if E if not. (4.20) The quantum NOT gate is defined as NOT : ρ 7→ ✶−ρ , i.e. it maps pure qubit states to its orthogonal state. Although this map transforms the Bloch sphere into itself it is not a CPTP map, and the closest CPTP map is ANOT : ρ 7→ (✶−ρ)/3. This is a rank-three qubit unital channel, thus, it is indivisible [WC08]. Moreover, detANOT = −1/27 implies that this channel is not achievable by a P-divisible dynamical map. It is worth noting that ANOT belongs to Cdiv. A specific collision model was designed in Ref. [RFZB12] simulating stroboscop- ically a quantum dynamical map that implements the quantum NOT gate ANOT in finite time. The dynamical map is given by Et(ρ) = cos2(t)ρ + sin2(t)ANOT(ρ)+ 1 2 sin(2t)F (ρ) , (4.21) where F (ρ) = i 1 3 ∑ j[σ j,ρ]. It achieves the desired gate ANOT at t = π/2. Let us stress that this dynamical map is unital, i.e. Et(✶) = ✶ for all t, thus, its special orthogonal normal form can be illustrated inside the tetrahedron of Pauli channels, see fig. 4.12. In fig. 4.13 we plot δ [Et ], χ[Et ] and the value of the detEt . We see the transitions CL → CP \CCP → Cdiv \CP → Cdiv and back. Notice that in both plots the trajectory never goes through the CCP \CL region. This means that when the parametrized channels, up to rotations, belong to CL, so do the original ones. The transition between P-divisible and divisible channels, i.e. CP\CCP and Cdiv\CP, occurs at the discontinuity in the yellow curve in fig. 4.12. Let us note that this discontinuity only occurs in the space of ~λ ; it is a consequence of the special orthogonal normal decomposition, see eq. (3.9). The complete channel is continuous in the full convex space of qubit CPTP maps. The transition from CP \ Cdiv and back occurs at times π/3 and 2π/3. It can also be noted that the transition to entanglement breaking channels occurs shortly before the channel enters in the 74 Chapter 4. Divisibility of quantum channels and dynamical maps λ1 λ2 λ3 Figure 4.12: (top left) Tetrahedron of Pauli channels with the trajectory, up to rotations, of the quantum dynamical map eq. (4.21) leading to the ANOT gate, as a yellow curve. (right) Cut along the plane that con- tains the trajectory; there one can see the different regions where the channel passes. For this case, the characterization of the CL of the channels induced gives the same conclusions as for the corresponding Pauli channel, see eq. (3.9). The discontinuity in the trajectory is due to the reduced representation of the dynamical map, see eq. (4.21); the trajectory is continuous in the space of channels. See fig. 4.9 for the color coding. 4.5. Divisibility transitions and examples with dynamical processes 75 0 π 3 π 2 2 π 3 π 0 1 3 2 3 1 δ [E t ], d et E t , χ [E t ] t C P\CCP C CP\CL C L C div\CP Figure 4.13: Evolution of divisibility, determinant, and entanglement break- ing properties of the map induced by eq. (4.21), see eq. (4.19) and eq. (4.20). Notice that the channel ANOT, implemented at t = π/2, has minimum determinant. The horizontal gray dashed lines show the image of the function δ , with the divisibility types in the right side. It can be seen that the dynamical map explores the divisibility sets as CL → CP \CCP → Cdiv \CP → Cdiv and back. The channels are entan- glement breaking in the expected region. 76 Chapter 4. Divisibility of quantum channels and dynamical maps Cdiv \CP region; likewise, the channel stops being entanglement breaking shortly after it leaves the Cdiv \CP region, see theorem 20. Consider now the dynamical map induced by a two-level atom interacting with a mode of a boson field. This model serves as a workhorse to explore a great variety of phenomena in quantum optics [GKL13]. Using the well known rotating wave approximation one arrives to the Jaynes-Cummings model [JC63], whose Hamiltonian is H = ωa 2 σz +ω f ( a†a+ 1 2 ) +g ( σ−a† +σ+a ) . (4.22) By initializing the environment in a coherent state |α〉, one gets the familiar collapse and revival setting. Considering a particular set of parameters shown in fig. 4.14, we constructed the channels parametrized by time numerically, and studied their divisibility and entanglement-breaking properties. In the same figure we plot functions δ [Et ] and χ[Et ], together with the probability of finding the atom in its excited state pe(t), to study and compare the divisibility properties with the features of the collapses and revivals. The probability pe(t) is calculated choos- ing the ground state of the free Hamiltonian ωa/2σz of the qubit, and it is given by [KC09]: pe(t) = 〈σz(t)〉+1 2 , (4.23) where 〈σz(t)〉=− ∞ ∑ n=0 Pn ( ∆2 4Ω2 n + ( 1− ∆2 4Ω2 n ) cos(2Ωnt) ) , with Pn = e−|α|2 |α|2n/n!, Ωn = √ ∆2/4+g2n and ∆ = ω f −ωa the detuning. The divisibility indicator function δ exhibits an oscillating behavior, roughly at the same frequency of pe(t), see inset in fig. 4.14. The figure shows fast periodic tran- sitions between CP \CCP and CCP \CL occurring in the region of revivals. There are also few transitions among CCP \CP and CL in the second revival. Respect to the entanglement breaking and the function χ , there are no fast transitions in the former, and during revivals, channels are not entanglement breaking. We also observe that channels belonging to Cdiv \CP are entanglement breaking, which agrees with theorem 20 for the non-unital case. 4.5. Divisibility transitions and examples with dynamical processes 77 3.85 3.9 3.95 0 2 4 6 8 10 0 1 3 2 3 1 δ [E t ], p e (t ), χ [E t ] t C P\CCP C CP\CL C L C div\CP Figure 4.14: Black and red curves show functions δ and χ of the chan- nels induced by the Jaynes-Cummings model over a two-level system, see eq. (4.22) with the environment initialized in a coherent state |α〉. The blue curve shows the probability of finding the two-level atom in its excited state, pe(t). The figure shows that the fast oscillations in δ occur roughly at the same frequency as the ones of pe(t), see the inset. Notice that there are fast transitions between CP \CCP and CCP \CL occurring in the region of revivals, with a few transitions between CCP \CP and CL in the second revival. The function χ shows that dur- ing revivals channels are not entanglement breaking, but we find that channels belonging to Cdiv \CP are always entanglement breaking, in agreement with theorem 20. The particular chosen set of parameters are α = 6, g = 10, ωa = 5, and ω f = 20. Chapter 5 Singular Gaussian quantum channels Self-education is, I firmly believe, the only kind of education there is. Isaac Asimov In this chapter we derive the conditions for δGQC to be singular, see sec. 3.5.1. In particular we will show that only the functional form involving one Dirac delta can be singular, together with the Gaussian form. Additionally we derive, for the non-singular cases, the conditions for the existence of master equations that parametrize channels that have always the same functional form. We do this by letting the channels parameters to depend on time. 5.1 Allowed singular forms There are two classes of Gaussian singular channels. Since the inverse of a Gaus- sian channel A (T,N,~τ) is A ( T−1,−T−1NT−T ,−T−1~τ ) , its existence rests on the invertibility of T. Therefore, studying the rank of the latter we are able to ex- plore singular forms. We are going to use the classification of one-mode channels developed by Holevo [Hol07]. For singular channels there are two classes charac- terized by its canonical form [Hol08], i.e. any channel can be obtained by apply- ing Gaussian unitaries before and after the canonical form. The class called “A1” corresponds to singular channels with Rank(T) = 0 and coincide with the family of total depolarizing channels. The class “A2” is characterized by Rank(T) = 1. 79 80 Chapter 5. Singular Gaussian quantum channels Both channels are entanglement-breaking [Hol08]. Before analyzing the functional forms constructed in this work, let us study chan- nels with GF. The tuple of the affine transformation, corresponding to the prop- agator JG, eq. (2.31), were introduced in Ref. [MP12] up to some typos. Our calculation for this tuple, following eq. (3.35), is: TG = ( −b4 b3 1 b3 b1b4 b3 −b2 −b1 b3 ) , NG =   2a3 b2 3 a2 b3 − 2a3b1 b2 3 a2 b3 − 2a3b1 b2 3 −2 ( −a3b2 1 b2 3 + a2b1 b3 −a1 )   , ~τG = ( − c2 b3 , b1c2 b3 − c1 )T . (5.1) It is straightforward to check that for b2 = 0, TG is singular with Rank(TG) = 1, i.e. it belongs to class A2. Due to the full support of Gaussian functions, it was surprising that Gaussian channels with GF have singular limit. In this case the singular behavior arises from the lack of a Fourier factor for x f ri, see eq. (2.31). This is the only singular case for GF. Now we analyze functional forms derived in sec. 3.5.1. The complete positivity conditions of the form J̃III, presented in eq. (3.45), have no solution for α → 0 and/or γ → 0, thus, this form cannot lead to singular channels. This is not the case for J̃I, eq. (3.36), which leads to singular operations belonging to class A2 for αe2 = 0, (5.2) and to class A1 for e2 = α = b2 = 0. (5.3) For the latter, the complete positivity conditions, see eq. (3.40), read: e1 ≤ a1. (5.4) By using an initial state characterized by σi and ~di we can compute the explicit dependence of the final states on the initial parameters. The final states for chan- nels of class A2 with the functional form involving one delta, see eq. (3.25), and 5.1. Allowed singular forms 81 with e2 = 0, are (σ f )11 = 1 2e1 , (σ f )22 = ( α β )2( b2 3 2e1 +2a3 ) + α β ( 2a2 + b1b3 e1 ) +2a1 + b2 1 2e1 + s1, (σ f )12 =−α β b3 2e1 − b1 2e1 , ~d f (s3) = ( 0,−α β c2 − c1 + s2 )T , (5.5) where s1 = ( b2 2 +2 α β b2b4 + ( α β )2 b2 4 ) (σi)11 −2 ( α β b2 + ( α β )2 b4 ) (σi)12 + ( α β )2 (σi)22 , s2 = ( α β b4 +b2 ) (di)1 − α β (di)2. (5.6) For the same functional form but now with α = 0, the final states are (σ f )11 = e2 2 4e2 1 (σi)11 + 1 2e1 , (σ f )12 = ( b2e2 2e1 − b1e2 2 4e2 1 ) (σi)11 − b1 2e1 , (σ f )22 = 2a1 + ( b2 − b1e2 2e1 )2 (σi)11 + b2 1 2e1 , (5.7) and ~d f = ( e2 2e1 ( ~di ) 1 , ( b2 − b1e2 2e1 ) ( ~di ) 1 − c1 )T . (5.8) The explicit formulas of the final states for channels of class A2 with Gaussian 82 Chapter 5. Singular Gaussian quantum channels form are (σ f )11 (s1) = 2a3 b2 3 + s1, (σ f )12 (s1) = a2 b3 − 2a3b1 b2 3 −b1s1, (σ f )22 (s1) = b1 (b3 (b1b3s1 −2a2)+2a3b1) b2 3 +2a1, ~d f (s2) = ( s2 − c2 b3 ,b1 ( c2 b3 − s2 ) − c1 )T , (5.9) where s1 = b2 4 b2 3 (σi)11 − 2b4 b2 3 (σi)12 + 1 b2 3 (σi)22 , s2 = 1 b3 (di)2 − b4 b3 (di)1 . (5.10) See fig. 5.1 for an schematic description of the final states. From such combina- tions it is obvious that we cannot solve for the initial state parameters given a final state as expected; this is because the parametric space dimension is reduced from 5 to at most 3. The channel belonging to A1 [see eq. (3.42) with e2 = α = b2 = 0 and eq. (5.4)] maps every initial state to a single one characterized by σ f = N and ~d f = (0,−c1) T , see fig. 5.2 for a schematic description. According to our ansätze [see equations (3.24) and (3.25)], we conclude that one- mode SGQC can only have the functional forms given in eq. (2.31) and eq. (3.24). This is the central result of this chapter and can be stated as: Theorem 21 (One-mode singular Gaussian channels). A one-mode Gaussian quan- tum channel is singular if and only if it has one of the following functional forms in the position space representation: 1. b3 2π exp [ ı ( b1x f r f +b3xir f +b4xiri + c1x f + c2xi ) −a1x2 f −a2x f xi −a3x2 i ] , 2. |β | √ e1/πδ (αx f −βxi)exp [ −a2x f xi −a1x2 f −a3x2 i +ı ( b2x f ri +b3r f xi +b1r f x f +b4rixi +c1x f +c2xi ) −e1r2 f −e2r f ri − e2 2r2 i 4e1 ] , with e2α = 0. Corollary 1 (Singular classes). A one-mode singular Gaussian channel belongs to class A1 if and only if its position representation has the following form: √ e1/πδ (xi)exp [ −a1x2 f + ı ( b2x f ri +b1r f x f + c1x f ) − e1r2 f ] . 5.1. Allowed singular forms 83 Otherwise the channel belongs to class A2. Since channels on each class are connected each other by unitary conjugations [Hol07], a consequence of the theorem and the subsequent corollary is that the set of al- lowed forms must remain invariant under unitary conjugations. To show this we must know the possible functional forms of Gaussian unitaries. They are given by the following lemma for one mode: Lemma 1 (One-mode Gaussian unitaries). Gaussian unitaries can have only GF or the one given by eq. (3.25). Proof. Recalling that for a unitary GQC, T must be symplectic (TΩTT = Ω) and N = 0. However, an inspection to eq. (3.37) lead us to note that N 6= 0 unless e1 diverges. Thus, Gaussian unitaries cannot have the form JI [see eq. (3.24)]. An inspection of matrices T and N of GQC with GF [see eq. (5.1)] and the ones for JII [see equations (3.38) and (3.44)] lead us to note the following two observations: (i) in both cases we have N = 0 for an = 0 ∀n; (ii) the matrix T is symplectic for GF when b2 = b3, and when αη = βγ for JII. In particular the identity map has the last form. This completes the proof. One can now compute the concatenations of the SGQCs with Gaussian unitaries. This can be done straightforward using the well known formulas for Gaussian in- tegrals and the Fourier transform of the Dirac delta. Given that the calculation is elementary, and for sake of brevity, we present only the resulting forms of each concatenation. To show this compactly we introduce the following abbreviations: Singular channels belonging to class A2 with form JI and with α = 0, e2 = 0 and α = e2 = 0, will be denoted as δ α A2 , δ e2 A2 and δ α,e2 A2 , respectively; singular channels belonging to the same class but with GF will be denoted as AA2 ; channels belong- ing to class A1 will be denoted as δA1 ; finally Gaussian unitaries with GF will be denoted as AU and the ones with form JII as δU . Writing the concatenation of two channels in the position representation as J(f)(x f ,r f ;xi,ri) = ∫ R2 dx′dr′J(1) ( x f ,r f ;x′,r′ ) J(2) ( x′,r′;xi,ri ) , (5.11) the resulting functional forms for J(f) are given in table 5.1. As expected, the table shows that the integral has only the forms stated by our theorem. Additionally it shows the cases when unitaries change the functional form of class A2, while for class A1 J(f) has always the unique form enunciated by the corollary. 84 Chapter 5. Singular Gaussian quantum channels J(1) J(2) J(f) δ α A2 AU AA2 AU δ α A2 δ α A2 δ α A2 δU δ α A2 δU δ α A2 δ α A2 δ e2 A2 AU δ e2 A2 AU δ e2 A2 AA2 δ e2 A2 δU δ e2 A2 δU δ e2 A2 δ e2 A2 AU ,δU δ α,e2 A2 δ α,e2 A2 δ α,e2 A2 AU ,δU δ α,e2 A2 δU ,AU δA1 δA1 δA1 δU ,AU δA1 Table 5.1: The first and second columns show the functional forms of J(1) and J(2), respectively. The last column shows the resulting form of the concatenation of them, see eq. (5.11). See main text for symbol coding. 5.2 Existence of master equations In this section we show the conditions under which master equations, associated with the channels derived in sec. 3.5.1, exist. To be more precise, we study if the functional forms derived above parametrize channels belonging to one-parameter differentiable families of GQCs. As a first step, we let the coefficients of forms presented in equations (3.24) and (3.25) to depend on time. Later we derive the conditions under which they bring any quantum state ρ(x,r; t) to ρ(x,r; t + ε) (with ε > 0 and t ∈ [0,∞)) smoothly, while holding the specific functional form of the channel, i.e. ρ(x,r; t + ε) = ρ(x,r; t)+ εLt [ρ(x,r; t)]+O(ε2), (5.12) where both ρ(x,r; t) and ρ(x,r; t + ε) are propagated from t = 0 with channels either with the form JI or JII, and Lt is a bounded superoperator in the state subspace. This is basically the problem of the existence of a master equation ∂tρ(x,r; t) = Lt [ρ(x,r; t)] , (5.13) for such functional forms. Thus, the problem is reduced to prove the existence of the linear generator Lt , also known as Liouvillian. 5.2. Existence of master equations 85 Class A2 r p (σ f )11 (s1 ,s2) (σ f )22 (s1 ,s2) (σ f )12 (s1 ,s2) ~d i 7→ ~d f (s3 ) ( σi , ~di ) Figure 5.1: Schematic picture of the channels belonging to class A2. The explicit dependence of the final state in terms of the combinations s1, s2 and s3 are presented in the appendix. As well the formulas for si depending on the form of the channel. To do this we use an ansatz proposed in Ref. [KG97] to investigate the existence and derive the master equation for GFs, L = Lc(t)+(∂x,∂r)X(t) ( ∂x ∂r ) +(x,r)Y(t) ( ∂x ∂r ) +(x,r)Z(t) ( x r ) (5.14) where Lc(t) is a complex function and X(t) = ( Xxx(t) Xxr(t) Xrx(t) Xrr(t) ) (5.15) is a complex matrix as well as Y(t) and Z(t), whose entries are defined in a similar way as in eq. (5.15). Note that X(t) and Z(t) can always be chosen symmetric, i.e. Xxr = Xrs and Zxr = Zrx. Thus, we must determine 11 time-dependent functions from eq. (5.14). This ansatz is also appropriate to study the functional forms introduced in this work, given that the left hand side of eq. (5.13) only involves quadratic polynomials in x, r, ∂/∂x and ∂/∂ r, as in the GF case. Notice that singular channels do not admit a master equation since its existence implies that channels with the functional form involved can be found arbitrarily 86 Chapter 5. Singular Gaussian quantum channels Class A1 r p 1 2e1 2a1 + b2 1 2e1 − b1 2e1 (0,−c1) T ( σi , ~di ) Figure 5.2: Schematic picture of the class A1. Every channel of this class maps every initial quantum state, in particular GSs characterized by ( σi, ~di ) , to a Gaussian state that depends only on the channel param- eters. We indicate in the figure the values of the corresponding compo- nents of the first and second moments of the final Gaussian state. 5.2. Existence of master equations 87 close from the identity channel. This is not possible for singular channels due to the continuity of the determinant of the matrix T. For the non-singular cases presented in equations (3.24) and (3.25), the condition for the existence of a master equation is obtained as follows. (i) Substitute the ansatz of eq. (5.14) in the right hand side of the eq. (5.13). (ii) Define ρ(x,r; t) using eq. (2.30), given an initial condition ρ(x,r;0), for each functional form JI,II. (iii) Take ρ f (x f ,r f ) → ρ(x,r; t) and ρi(xi,ri) → ρ(x,r;0). Finally, (iv) compare both sides of eq. (5.13). Defining A(t) = α(t)/β (t) and B(t) = γ(t)/η(t), the conclusion is that for both JI and JII, a master equations exist if c(t) ∝ A(t) (5.16) holds, where c(t) = c1(t)+A(t)c2(t). Additionally, for the form JI the solutions for the matrices X(t), Y(t) and Z(t) are given by Xxx = Xxr = Yrx = Zrr = 0, Yxx = Ȧ A , Lc = Yrr = ė1 e1 − ė2 e2 , Xrr = ė1 4e2 1 − ė2 2e1e2 , Yxr = ı ( λ1ė2 e1e2 + λ2Ȧ e2A − λ1ė1 2e2 1 − λ̇2 e2 ) , Zxx = λ 2 1 2 ( ė2 e1e2 − ė1 2e2 1 ) + λ1 e2 ( λ2Ȧ A − λ̇2 ) +2λ3 Ȧ A − λ̇3, Zxr = ı ( Ȧ A ( e1λ2 e2 − λ1 2 ) + λ̇1 2 − λ̇2e1 e2 + λ2 2 ( ė2 e2 − ė1 e1 )) , (5.17) where we have defined the following coefficients: λ1 = b1 +Ab3, λ2 = b2 +Ab4 and λ3 = a1 +Aa2 +A2a3. 88 Chapter 5. Singular Gaussian quantum channels For the form JII the solutions are the following Lc = Xxx = Xxr = Xrr = Zrr = 0, Yrx = Yxr = 0, Yxx = Ȧ A , Yrr = Ḃ B . Zxx = a2(t)Ȧ(t)+ 2a1(t)Ȧ(t) A(t) −A(t)2 −ȧ3(t)−A(t)ȧ2(t)− ȧ1(t), Zxr = ı ( 1 2 λ̇ − λ 2 ( Ȧ A + Ḃ B )) , (5.18) where λ = b1 +Ab3 +B(b2 +Ab4). Chapter 6 Summary and conclusions Living is worthwhile if one can contribute in some small way to this endless chain of progress. Paul A.M. Dirac In this thesis we have introduced two works developed during my PhD. The first one was devoted to study quantum channels from the point of view of their di- visibility properties. We made use of several results from the literature, specially from the seminal work by M. M. Wolf and J. I. Cirac [WECC08], and completed and fixed some results of Ref. [WC08]. This led to the construction of a tool to decide whether a quantum channel can be implemented using time-independent Markovian master equations or not, for the finite dimensional case. We addition- ally proved three theorems relating some of the studied divisibility types. Some of the tools introduced in chapter 3 are results from other paper developed during my PhD, where I am a secondary author, see Ref. [CDG19]. In the second work we have studied one-mode Gaussian channels without Gaussian functional form in the position state representation. We performed a characterization based on the universal properties that quantum channels must fulfill; in particular we studied the case of singular channels. We showed that the transition from unitarity to non-unitarity can correspond directly to a change in the functional form of the channel, in particular it turns out that functional form with one Dirac delta factor do not parametrize unitary channels. Additionally in this project we derived the conditions under which master equations for particular functional forms exist. Let us summarize the results for the first project in more detail. We imple- mented the known conditions to decide the compatibility of channels with time- independent master equations (the so called L-divisibility) for the general diag- 89 90 Chapter 6. Summary and conclusions onalizable case, and a discussion of the parametric space of Lindblad generators was given. We additionally clarified one of the results of the paper [WECC08]. There, the authors arrived to erroneous conclusions for the case of channels with negative eigenvalues. In our work we handled this case carefully. For unital qubit channels it was shown that every infinitesimal divisible map can be written as a concatenation of one L-divisible channel and two unitary conjugations. For the particular case of Pauli channels case, we have shown that the sets of infinitely divisible and L-divisible channels coincide. We made an interesting observation, connecting the concept of divisibility with the quantum information concept of entanglement-breaking channels: we found that divisible but not infinitesimal di- visible qubit channels (in positive but not necessarily completely positive maps) are necessarily entanglement-breaking. We also noted that the intersection of in- divisible and P-divisible channels is not empty. This allows us to implement indi- visible channels with infinitesimal positive and trance preserving maps. Finally, we studied the possibility of dynamical transitions between different classes of divisibility channels. We argued that all the transitions are, in principle, possible, given that every divisibility set appears connected in our plots. We exploited two simple models of dynamical maps to demonstrate that these transitions exist. They clearly illustrate how the channels evolutions change from being implementable by Markovian dynamical maps (infinitesimal divisible in complete positive maps and/or L-divisible) to non-Markovian (divisible but not infinitesimal divisible or infinitesimal divisible in positive but not complete positive maps), and vice versa. For the second project we have critically reviewed the deceptively natural idea that Gaussian quantum channels always admit a Gaussian functional form. To this end, we went beyond the pioneering characterization of Gaussian channels with Gaussian form presented in Ref. [MP12] in two new directions. First we have shown that, starting from their most general definition (a quantum ma that takes Gaussian states to Gaussian states), a more general parametrization of the coordinate representation of the one-mode case exists, that admits non-Gaussian functional forms. Second, we were able to provide a black-box characterization of such new forms by imposing complete positivity (not considered in Ref. [MP12]) and trace preserving conditions. While our parametrization connects with the analysis done by Holevo [Hol08] in the particular cases where besides having a non-Gaussian form the channel is also singular, it also allows the study of Gaus- sian unitaries, thus providing similar classification schemes. We completed the classification of the studied types of channels by deriving the form of the Liou- villian super operator that generates their time evolution in the form of a master equation. Surprisingly, Gaussian quantum channels without Gaussian form can be experimentally addressed by means of the celebrated Caldeira-Legget model for the quantum damped harmonic oscillator [GSI88], where the new types of 91 channels described here naturally appear in the sub-ohmic regime. We are interested in several directions to continue the investigation. From the project of divisibility of quantum channels, an extension of this analysis to larger- dimensional systems could give a deeper sight to the structure of quantum chan- nels. In particular we are interested on proving if the equivalence of infinitely divisible channels and L-divisible channels is present also in the general qubit case. Additionally a plethora of interesting questions are related to design of effi- cient verification procedures of the divisibility classes for channels and dynamical maps. For instance, can we define an extension of the Lorentz normal decompo- sition to systems composed of many qubits?, this would be useful to character- ize infinitesimal divisibility of many particle systems; or Is the non-countable parametrization of channels with negative eigenvalues relevant on deciding L- divisibility?. Finally the area of channel divisibility contains several open struc- tural questions, e.g. the existence of at most n-divisible channels. From the project concerning one-mode Gaussian channels, a natural direction to follow is to extend the analysis for other types of channels (or more modes) by following the classifi- cation introduced by Holevo, see Ref. [Hol07]. The latter is based on the form of a canonical form of one-mode Gaussian channels. Therefore a connection of this classification with ours could be useful to assess quantum information features, in particular for systems for which position state representation is advantageous. Chapter 7 Appendices 93 Appendix A Proof of theorem “Exact dynamics with Lindblad master equation” The theorem announced in chapter 2 is, Theorem 2 (Exact dynamics with Lindblad master equation) Let Et = etL a quan- tum process generated by a Lindblad operator L. The equation Et [ρS] = trE [ e−iHt (ρS ⊗ρE)eiHt ] , where H has finite dimension, holds if and only if Et is a unitary conjugation for every t. Proof. To prove this theorem, we will compute ρS(t) to first order in t, see eq. (2.12). Following the master equation of eq. (2.10) and taking t = ε ≪ 1, we have ρS(ε)≈ ρS + trE ∫ ε 0 dt {i [ρS ⊗ρE,H]} = ρS + trE {i [ρS ⊗ρE,H]}ε = ρS +LExact[ρS]ε. where LExact = trE {i [ρS ⊗ρE,H]}. Since Et is generated by a Lindblad master equation, LExact must coincide with the Lindblad generator since the process is homogeneous in time, i.e. LExact is time-independent. Writing the global Hamil- tonian as H = ∑ k,l=0 hklF (S) k ⊗F (E) l , 95 96 Chapter A. Exact dynamics with Lindblad master equation where hkl ∈ R, and { F (S) k } k and { F (E) k } k are orthogonal hermitian bases of B ( H (S) ) and B ( H (E) ) , respectively, with H (S) and H (E) are the Hilbert spaces of the central system S and the environment E. We have, LExact[ρS] = i trE { ∑ k,l hkl[ρS ⊗ρE,F (S) k ⊗F (E) l ] } = i∑ k,l hkl { ρSF (S) k tr[ρEF (E) l ]−F (S) k ρS tr[F (E) l ρE] } = i∑ k,l hkl tr[F (E) l ρE] { ρSF (S) k −F (S) k ρS } = i[ρS, H̃], where H̃ = ∑k,l hkl tr[F (E) l ρE]F (S) k is an hermitian operator. Therefore LExact is the generator of Hamiltonian dynamics with Hamiltonian H̃, thus Et is unitary for all t. Appendix B On Lorentz normal forms of Choi-Jamiolkowski state In this appendix we compute the Lorentz normal decomposition of a channel for which one gets b 6= 0, supporting our observation that Lorentz normal decom- position does not take Choi-Jamiołkowski states to something proportional to a Choi-Jamiołkowski state. Consider the following Kraus rank three channel and its RE matrix, both written in the Pauli basis: Ê =     1 0 0 0 0 − 1 3 0 0 0 0 −1 3 0 2 3 0 0 1 3     , (B.1) and RE =     1 0 0 0 0 −1 3 0 0 0 0 1 3 0 2 3 0 0 1 3     . (B.2) Using the algorithm introduced in Ref. [VDD01] to calculate RE ’s Lorentz de- composition into orthochronous proper Lorentz transformations we obtain L1 = 1 γ1     4 0 0 1 0 −γ1 0 0 0 0 −γ1 0 1 0 0 4     , (B.3) 97 98 Chapter B. On Lorentz normal forms of Choi-Jamiolkowski state L2 = 1 γ2     89+9 √ 97 0 0 −8 0 −γ2 0 0 0 0 −γ2 0 −8 0 0 89+9 √ 97     , and ΣE = 1 γ3       √ 11+ 109√ 97 0 0 − √ 97+1√ 89 √ 97+873 0 − γ3 3 0 0 0 0 γ3 3 0 √ 1+ 49√ 97 0 0 √ −1+ 49√ 97       with γ1 = √ 15, γ2 = 3 √ 178 √ 97+1746, and γ3 = √ 30. Although the central ma- trix ΣE is not exactly of the form eq. (3.17), it is equivalent. To see this notice that the derivation of the theorem 2 in [VDD01] considers only decompositions into proper orthochronous Lorentz transformations. But to obtain the desired form, the authors change signs until they get eq. (3.17); this cannot be done without chang- ing Lorentz transformations. If we relax the condition over L1,2 of being proper and orthochronous, we can bring ΣE to the desired form by conjugating ΣE with G = diag(1,1,1,−1): G−1ΣE G = 1 γ3       √ 11+ 109√ 97 0 0 √ 97+1√ 89 √ 97+873 0 − γ3 3 0 0 0 0 γ3 3 0 − √ 1+ 49√ 97 0 0 √ −1+ 49√ 97       . In both cases (taking ΣE or G−1ΣE G as the normal form of RE ), the corresponding channel is not proportional to a trace-preserving one since b 6= 0, see eq. (3.17). This completes the counterexample. Appendix C Articles C.1 Article: Divisibility of qubit channels and dynamical maps Quantum 3, 144 (2019). Click to go to the webpage, Click to go to arXiv. 99 Divisibility of qubit channels and dynamical maps David Davalos1, Mario Ziman2,3, and Carlos Pineda1,4 1Instituto de Física, Universidad Nacional Autónoma de México, México, D.F., México 2Institute of Physics, Slovak Academy of Sciences, Dúbravská cesta 9, Bratislava 84511, Slovakia 3Faculty of Informatics, Masaryk University, Botanická 68a, 60200 Brno, Czech Republic 4Faculty of Physics, University of Vienna, 1090 Vienna, Austria May 7, 2019 The concept of divisibility of dynamical maps is used to introduce an analogous con- cept for quantum channels by analyzing the simulability of channels by means of dynami- cal maps. In particular, this is addressed for Lindblad divisible, completely positive divisi- ble and positive divisible dynamical maps. The corresponding L-divisible, CP-divisible and P- divisible subsets of channels are characterized (exploiting the results by Wolf et al. [25]) and visualized for the case of qubit channels. We discuss the general inclusions among divisibil- ity sets and show several equivalences for qubit channels. To this end we study the conditions of L-divisibility for finite dimensional chan- nels, especially the cases with negative eigen- values, extending and completing the results of Ref. [26]. Furthermore we show that tran- sitions between every two of the defined divis- ibility sets are allowed. We explore particular examples of dynamical maps to compare these concepts. Finally, we show that every divisible but not infinitesimal divisible qubit channel (in positive maps) is entanglement-breaking, and open the question if something similar occurs for higher dimensions. 1 Introduction The advent of quantum technologies opens questions aiming for deeper understanding of the fundamental physics beyond the idealized case of isolated quantum systems. Also the well established Born-Markov ap- proximation used to describe open quantum systems (e.g. relaxation process such as spontaneous decay) is of limited use and a more general framework of open system dynamics is demanded. Recent efforts in this area have given rise to relatively novel research sub- jects - non-markovianity and divisibility. Non-markovianity is a characteristic of continuous time evolutions of quantum systems (quantum dy- namical maps), whereas divisibility refers to prop- erties of system’s transformations (discrete quantum David Davalos: davidphysdavalos@gmail.com processes) over a fixed time interval (quantum chan- nels). The non-markovianity aims to capture and de- scribe the back-action of the system’s environment on the system’s future time evolution. Such phenomena is identified as emergence of memory effects [1, 18, 22]. On the other side, the divisibility questions the possi- bility of splitting a given quantum channel into a con- catenation of other quantum channels. In this work we will investigate the relation between these two no- tions. Our goal is to understand the possible forms of the dynamics standing behind the observed quantum channels, specially in regard to their divisibility prop- erties which in turn determine their markovian or non- markovian nature. In particular, we provide charac- terization of the subsets of qubit channels depending on their divisibility properties and implementation by means of dynamical maps. An attempt to charac- terize the set of channels belonging to one-parameter semigroups induced by (time-independent) Lindblad master equations has been already done in Ref. [26]. However, it has drawbacks when dealing with chan- nels with negative determinants. Using the results of Ref. [5] and Ref. [3], we will extend the analysis of [26] also for channels with negative eigenvalues. The paper is organized as follows: In section 2 we give the formal definition of quantum channels and of quantum dynamical maps, and some of their proper- ties. We discuss the meaning of divisibility for each object and discuss the known inclusions and equiva- lences between divisibility types. In section 3 we dis- cuss properties and representations of qubit channels and their divisibility. We introduce a useful theorem to decide L-divisibility, which is in turn valid for any finite dimension. In section 4 we discuss the possi- ble transition that can be occur between divisibility types, and show two examples of dynamical maps and their transitions. Finally in Section 5 we summarize our results and discuss open questions. Accepted in Quantum 2019-04-24, click title to verify 1 2 Basic definitions and divisibility 2.1 Channels and divisibility classes We shall study transformations of a physical system associated with a complex Hilbert space Hd of di- mension d. In particular, we consider linear maps on bounded operators, B(Hd), that for the finite- dimensional case coincides with the set of trace-class operators that accommodate the subset of density op- erators representing the quantum states of the system. We say a linear map E : B(Hd) → B(Hd) is positive, if it maps positive operators into positive operators, i.e. X ≥ 0 implies E [X] ≥ 0. Quantum channels are associated with elements of the convex set C of com- pletely positive trace-preserving linear maps (CPTP) transforming density matrices into density matrices, i.e. E : B(H) → B(H) such that tr(E [X]) = tr(X) for all X ∈ B(H), and all its extensions idn ⊗ E are positive maps for all n > 1, where idn is the iden- tity channel on a n-dimensional quantum system. In general a channel has the form E [X] = ∑ i KiXK† i . The minimum number of operators Ki required in the previous expression is called the Kraus rank of E . Let us introduce two subsets of channels. First, we say a channel is unital if it preserves the identity operator, i.e. E [✶] = ✶. Unital channels have a simple parametrization which will be useful for our purposes. Second, if E [X] = UXU† for some unitary operator U (meaning UU† = U†U = ✶), we say the channel is unitary. A quantum channel E is called indivisible if it can- not be written as a concatenation of two non-unitary channels, namely, if E = E1E2 implies that either E1, or E2, exclusively, is a unitary channel. If the channel is not indivisible, it is said to be divisible. We denote the set of divisible channels by Cdiv and that of in- divisible channels by Cdiv. Following this definition, unitary channels are divisible, because for them both (decomposing) channels E1,2 must be unitary. The concept of indivisible channels resembles the concept of prime numbers: unitary channels play the role of unity (which are not indivisible/prime), i.e. a compo- sition of indivisible and a unitary channel results in an indivisible channel. We now define the set of infinitely divisible channels (C∞) and the set of infinitesimal divisible channels (CInf). Infinitely divisible channels, in some sense op- posite to indivisible channels, are defined as channels E for which there exist for all n = 1, 2, 3, . . . a channel An such that E = (An)n. Now, consider channels E that may be written as products of channels close to identity, i.e. such that for all ǫ > 0 there exists a finite set of channels εj with ||id − εj || ≤ ǫ and E = ∏ j εj . Its closure determines the set of infinitesimal divisible channels CInf. 2.2 Quantum dynamical maps and more divis- ibility classes The next sets of channels are going to be defined using three types of dynamical maps. A quantum dynami- cal map is identified with a continuous parametrized curve drawn inside the set of channels starting at the identity channel, i.e. a one-parametric function t 7→ Et ∈ C for all t belonging to an interval with min- imum element 0 and satisfying the initial condition E0 = id. Let Et,s = E−1 t Es be the linear map describ- ing the state transformations within the time interval [t, s], whenever E−1 t exists. • A given quantum dynamical map is called CP- divisible if for all t < s the map Et,s is a channel. • A given quantum dynamical map is called P- divisible [22] if Et,s is a positive trace-preserving linear map for all t < s. • A given quantum dynamical map is called L- divisible if it is induced by a time-independent Lindblad master equation [7, 14, 16], i.e. Et = etL with L(ρ) = i[ρ, H]+ ∑ α,β Gαβ ( FαρF † β − 1 2 {F † βFα, ρ} ) , where H = H† ∈ B(Hd) is known as Hamilto- nian, {Fα} are hermitian and form an orthonor- mal basis of the operator space B(Hd), and Gαβ constitutes a hermitian positive semi-definite ma- trix. If we allow the Lindblad generator L to depend on time, we recover the set of CP-divisible quan- tum dynamical maps as the resulting dynamical maps Et = T̂e ∫ t 0 L(τ)dτ (T̂ denotes the time-ordering op- erator) are compositions of infinitesimal completely- positive maps [1, 7, 14, 16]. Notice that there is a hierarchy for quantum dynamical maps: L-divisible quantum dynamical maps are CP-divisible which in turn are P-divisible. Using the introduced families of quantum dynam- ical maps we can now classify quantum channels ac- cording to whether they can be implemented by the aforementioned kinds of quantum dynamical maps. We define subsets CL, CCP, and CP of L-divisible, CP-divisible and P-divisible channels, respectively. In particular, we say E ∈ CL if it belongs to the clo- sure of a L-divisible quantum dynamical map. Let us stress that the requirement of the existence of Lind- blad generator L such that E = eL is not sufficient and closure is necessary. For example, the evolution governed by L(ρ) = i[ρ, H]+γ[H, [H, ̺]] results [27] in the diagonalization of states in the energy eigenbasis of H. Such transformation Ediag is not invertible, thus (by definition) L = log Ediag does not exist (contains infinities). Analogously, we say E ∈ CCP (E ∈ CP) Accepted in Quantum 2019-04-24, click title to verify 2 if there exists a CP-divisible (P-divisible) dynamical map Et such that E = Et (with arbitrary precision) for some value of t (including t = ∞). We now recall how to verify whether a channel is L-divisible since we will build upon the method for some of our results. Verifying whether E ∈ CL [26] requires evaluation of the channel’s logarithms, how- ever, the matrix logarithm is defined only for invert- ible matrices and it is not unique. In fact, we need to check if at least one of its branches has the Lind- blad form. It was shown in [5] that E is L-divisible if and only if there exists L such that: exp L = E , is hermitian preserving, trace-preserving and condition- ally completely positive (ccp). Thus, we are looking for logarithm satisfying L(X†) = L(X)† (hermiticity preserving), L∗(✶) = 0 (trace-preserving), and (✶ − ω)(id ⊗ L)[ω](✶ − ω) ≥ 0 (1) (ccp condition), where ω = 1 d ∑d j,k=1 |j ⊗ j〉〈k ⊗ k| is the projector onto a maximally entangled state. To the best of our knowledge, there is no general method to verify if a channel is P or CP divisible, with some exceptions [25]. 2.3 Relation between channel divisibility classes Due to the inclusion set relations between the three kinds of dynamical maps discussed in the previous sections, we can see that CL ⊂ CCP ⊂ CP. Simi- larly, from the earlier definitions, one can easily ar- gue that C∞ ⊂ CInf ⊂ Cdiv. By the definition of L- divisibility it is trivial to see that in general CL ⊂ C∞. Indeed Denisov has shown in [4] that infinitely divis- ible channels can be written as E = E0eL, with L a Lindblad generator, and E0 idempotent operator such that E0LE0 = E0L. Further, it was shown in Ref. [25] that E ∈ Cinf implies det E ≥ 0 and also that E can be approximated by ∏ j eLj , i.e. CCP = Cinf . In other words, the positivity of determinant is necessary for the channel to be (in the closure of) channels attain- able by CP-divisible dynamical maps. In summary, we have the following relations between sets (see also Fig. 1): C∞ ⊂ CInf ⊂ Cdiv ⊆ = CL ⊂ CCP ⊂ CP . (2) The relation between CP and Cdiv is unknown, al- though it is clear that Cdiv ⊂ CP is not possible since channels in CP are not necessarily divisible in CP maps. The intersection of CP and Cdiv is not empty since CCP ⊆ Cdiv and CCP ⊆ CP. Later on we will investigate if CP ⊆ Cdiv or not. C C L C div C P C CP C ∞ C P ∩ Cdiv Unitary channels Figure 1: Scheme illustrating the different sets of quantum channels for a given dimension, discussed in sec. 2. In partic- ular, the inclusion relations presented in Eq. (2) are depicted. 3 Qubit channels 3.1 Representations Using the Pauli basis 1√ 2 {1, σx, σy, σz}, and the stan- dard Hilbert-Schmidt inner product, the real repre- sentation for qubit channels is given by [10, 20]: Ê = ( 1 ~0T ~t ∆ ) . (3) This describes the action of the channel in the Bloch sphere picture in which the points ~r are identified with density operators ̺~r = 1 2 (I + ~r · ~σ). We will write E = (∆,~t) meaning that E(ρ~r) = ρ∆~r+~t. In order to study qubit channels with simpler ex- pressions, we will consider a decomposition in uni- taries such that E = U1DU2. (4) This can be performed by decomposing ∆ in ro- tation matrices, i.e. ∆ = R1DR2, where D = diag(λ1, λ2, λ3) is diagonal and the rotations R1,2 ∈ SO(3) (of the Bloch sphere) correspond to the uni- tary channels U1,2. This decomposition should not be confused with the singular value decomposition. The latter allows decompositions that include, say, total reflections. Such operations do not correspond to unitaries over a qubit, in fact they are not CPTP. Therefore the channel D, in the Pauli basis, is given by D̂ = ( 1 ~0T ~τ D ) , (5) where ∆ = R1DR2 and ~τ = RT 1 ~t. The latter describes the shift of the center of the Bloch sphere under the action of D. The parameters ~λ determine the length of semi-axes of the Bloch ellipsoid, being the defor- mation of Bloch sphere under the action of E . From Accepted in Quantum 2019-04-24, click title to verify 3 now we will call the form D, special orthogonal normal form. We shall develop a geometric intuition in the space determined by the possible values of these three pa- rameters. For an arbitrary channel, complete positiv- ity implies that the possible set of lambdas lives in- side the tetrahedron with corners (1, 1, 1), (1, −1, −1), (−1, 1, −1) and (−1, −1, 1), see Fig. 2. For unital channels, all points in the tetrahedron are allowed, but for non-unital channels more restrictive condi- tions arise. In Fig. 8 we present a visualization of the permitted values of ~λ for a particular nontrivial value of ~τ , and in [2] the steps to study the general case from an algebraic point of view are presented. For the unital case, the corner ~λ = (1, 1, 1) corre- sponds to the identity channel, ~λ = (1, −1, −1) to σx ~λ = (−1, 1, −1) to σy and ~λ = (−1, −1, 1) to σz (Kraus rank 1 operations). Points in the edges corre- spond to Kraus rank 2 operations, points in the faces to Kraus rank 3 operations and in the interior of the tetrahedron to Kraus rank 4 operations. In addition to this decomposition, following the def- inition of divisibility, concatenation with unitaries of a given quantum channel do not change the divisibility character of the latter. Thus orthogonal normal forms are useful to study divisibility since, following also the properties of CP and CCP introduced in sec. 3.1, im- mediately one has: Theorem 1 (Divisibility of special orthogo- nal normal forms). Let E a qubit quantum chan- nel and D its special orthogonal normal form, E be- longs to CX if and only if D does, where X = {“Div”, “P”, “CP”}. There is another another parametrization for qubit channels called Lorentz normal decomposition [23, 24] which is specially useful to characterize infinitesimal divisibility CInf, and geometric aspects of entangle- ment [15]. This decomposition is derived from the theorem 3 of Ref. [24], which essentially states that for a qubit state ρ = 1 4 ∑3 i,j=0 Rijσi⊗σj the matrix R can be decomposed as R = L1ΣLT 2 . Here L1,2 are proper orthochronous Lorentz transformations and Σ is ei- ther Σ = diag (s0, s1, s2, s3) with s0 ≥ s1 ≥ s2 ≥ |s3|, or Σ =     a 0 0 b 0 d 0 0 0 0 −d 0 c 0 0 −b + c + a     . (6) In theorem 8 of Ref. [23] the authors make a simi- lar claim, exploiting the Choi-Jamiołkowski isomor- phism. They forced b = 0 in order to have normal forms proportional to trace-preserving operations in the case of Kraus rank deficient ones, see Eq. (6). The latter is equivalent to saying that the decomposition of Choi-Jamiołkowski states leads to states that are proportional to Choi-Jamiołkowski states. We didn’t find a good argument to justify such an assumption λ1 λ2 λ3 Figure 2: Tetrahedron of Pauli channels. The corners cor- respond to unitary Pauli operations (✶, σx,y,z) while the rest can be written as convex combinations of them. The bipyramid in blue corresponds to channels with λi > 0∀i, i.e. channels of the positive octant belonging to CP. The whole set CP includes three other bipyramids corresponding to the other vertexes of tetrahedron, i.e. CP enjoys the symmetries of the tetrahedron, see Eq. (8). The faces of the bipyramids matching the corners of the tetrahedron are subsets of the faces of the tetrahedron, i.e. contain Kraus rank three chan- nels. Such channels are then CPbut also Cdiv, showing that the intersection shown in Fig. 1 is not empty. Accepted in Quantum 2019-04-24, click title to verify 4 and found a counterexample (see appendix A). Thus we propose a restricted version of their theorem: Theorem 2 (Restricted Lorentz normal form for qubit quantum channels). For any full Kraus rank qubit channel E there exists rank-one completely positive maps T1, T2 such that T = T1ET2 is propor- tional to ( 1 ~0T ~0 Λ ) , (7) where Λ = diag(s1, s2, s3) with 1 ≥ s1 ≥ s2 ≥ |s3|. The channel T is called the Lorentz normal form of the channel E . For unital qubit channels D coincides with Λ, thus in such case the form of (7) holds for any Kraus rank. 3.2 Divisibility In this subsection we will recall the criteria to decide if a qubit channel belongs to Cdiv, CP and CCP follow- ing [25]. We shall start with some general statements, and then focus on different types of channels (unital, diagonal non-unital and general ones). We will also discuss in detail the characterization of CL, which en- tails a higher complexity. It was shown ([25], Theorem 11) that full Kraus rank channels are divisible (Cdiv). This simply means that all points in the interior of the set of channels cor- respond to divisible channels. Moreover, according to Theorem 23 of the same reference, qubit channels are indivisible if and only if they have Kraus rank three and diagonal Lorentz normal form. Notice that since we dispute the theorem upon which such statement is based, the classification might be inaccurate, see the appendix. It follows from the definition that for qubit channels E is divisible if and only if D is divisible. To test if D is divisible, we check that all eigenvalues of its Choi matrix are different from zero. A non-negative determinant of E is a necessary con- dition for a general channel to belong to CP ([25], Proposition 15). For qubits, this is also sufficient ([25], Theorem 25), and given that det D = det E , the condition for qubit channels simply reads det E = λ1λ2λ3 ≥ 0. (8) However, to our knowledge, a simple condition for arbitrary dimension is yet unknown. With respect to testing for CP-divisibility we re- strict the discussion to qubit channels. To character- ize CP-divisible channels it is useful to consider the Lorentz normal form for channels. A full Kraus rank qubit channel E belongs to CCP if and only if it has diagonal Lorentz normal form with s2 min ≥ s1s2s3 > 0 (9) where smin is the smallest of s1, s2 and s3, see theo- rem 2 and [26]. For Kraus deficient channels the per- tinent theorems are based on Kraus deficient Lorentz normal forms that according to our appendix should be reviewed. Deciding L-divisibility, as mentioned above, is equivalent to proving the existence of a hermiticity preserving generator which additionally fulfills the ccp condition. To prove the former we recall that every hermitic- ity preserving operator has a real matrix representa- tion when choosing a hermitian basis. Since quantum channels preserve hermiticity, the problem is reduced on finding a real logarithm log Ê given a real matrix Ê . This problem was already solved by Culver [3] who characterized completely the existence of real loga- rithms of real matrices. For diagonalizable matrices the results can be summarized as follows: Theorem 3 (Existence of hermiticity preserv- ing generator). A non-singular matrix with real en- tries Ê has a real generator (i.e. a log Ê has real en- tries) if and only if the spectrum fulfills the following conditions: (i) negative eigenvalues are even-fold degenerated; (ii) complex eigenvalues come in complex conjugate pairs. Let us examine this theorem for the particular case of qubits. In this case this theorem means that real logarithm(s) of Ê exist if and only if E has either only positive eigenvalues, one positive and two complex, or one positive and two equal non-positive eigenvalues, apart from the trivial eigenvalue equal to one. No- tice that quantum channels with complex eigenvalues will fulfill the last condition immediately since they preserve hermiticity. We now continue discussing the multiplicity of the solutions of log Ê , as finding an appropriate parametrization is essential to test for the ccp con- dition, see Eq. (1). If Ê has positive degenerated, negative, or complex eigenvalues, its real logarithms are not unique, and are spanned by real logarithm branches [3]. In case of having negative eigenval- ues, it turns out that real logarithms always have a non-continuous parametrization, in addition to real branches due to the freedom of the Jordan normal form transformation matrices. Given a real repre- sentation of E , i.e. Ê , the Jordan form is given by Ê = wJw−1 = w̃Jw̃−1, where w = w̃K with K be- longing to a continuum of matrices that commutes with J [3]. In the case of diagonalizable matrices, if there are no degeneracies, K commutes with log(J). Finally let us note that if a channel belongs to CL, unitary conjugations can bring it to CInf \CL and vice versa. 3.3 Unital channels We shall start our study of unital qubit channels, by considering Pauli channels, defined as convex combi- Accepted in Quantum 2019-04-24, click title to verify 5 nations of the unitaries σi: EPauli[ρ] = 3 ∑ i=0 piσiρσi, (10) where σ0 = ✶ and pi ≥ 0 with ∑ i pi = 1. The spe- cial orthogonal normal form of a Pauli channel [see Eqs. (4) and (5)] has U1 = U2 = id and ~τ = ~0. Thus Pauli channels are fully characterized only by ~λ. No- tice that every unital qubit channel can be written as Eunital = U1EPauliU2. (11) This implies that arbitrary unital qubit channels can be expressed as convex combinations of (at most) four unitary channels. Following theorem 1 it is straightforward to note that by characterizing Cdiv, CP and CCP of Pauli channels, the same conclusions hold for general uni- tal qubit channels having the same ~λ. Additionally we can have a one-to-one geometrical view of the di- visibility sets for Pauli channels given they have a one-to-one correspondence to the tetrahedron shown in Fig. 2, defined by the inequalities 1 + λi − λj − λk ≥ 0 (12) 1 + λ1 + λ2 + λ3 ≥ 0 (13) with i, j and k all different [28]. 3.3.1 P-divisibility Let us discuss the divisibility properties of Pauli chan- nels. Divisibility in CPTP (Cdiv) is guaranteed for full Kraus rank channels, i.e. for the interior channels of the tetrahedron. For Pauli channels this is equiva- lent to taking only the inequality of equations (13). The characterization of CP can be done directly using Eq. (8), as it depends only on ~λ. This set is the in- tersection of the tetrahedron with the octants where the product of all λs is positive. In fact, it consists of four triangular bipyramids starting in each vertex of the tetrahedron and meeting in its center, see Fig. 2. Let us study the intersection of this set with the set of unital entanglement-breaking (EB) channels [28], forming an octahedron (being the intersection of the tetrahedron with its space inversion, see Fig. 3). It is defined by the inequalities λ1 + λ2 + λ3 ≤ 1 λi − λj − λk ≤ 1, (14) with i, j and k all different [28], together with Eq. (13). It follows that unital qubit channels that are not achieved by P-divisible dynamical maps are nec- essarily entanglement-breaking (see Fig. 3 and Fig. 7). In fact this holds for general qubit channels, see sec- tion 3.4. λ1 λ2 λ3 Figure 3: Tetrahedron of Pauli channels with the octahe- dron of entanglement breaking channels shown in red, see Eq. (14). The blue pyramid inside the octahedron is the intersection of the bipyramid shown in Fig. 2, with the oc- tahedron. The complement of the intersections of the four bipyramids forms the set of divisible but not infinitesimal di- visible channels in PTP. Thus, a central feature of the figure is that the set Cdiv\CP is always entanglement-breaking, but the converse is not true. 3.3.2 CP-divisibility The subset of CP-divisible Pauli channels, following Eq. (9) and theorem 2, is determined by the inequal- ities 0 < λ1λ2λ3 ≤ λ2 min . (15) They determine a body that is symmetric with respect to permutation of Pauli unitary channels (i.e. in λj), hence, the set of CCP of Pauli channels possesses the symmetries of the tetrahedron. The set CCP\CL is plotted in Fig. 5. Notice that this set coincides with the set of unistochastic qubit channels, see Ref. [17]. 3.3.3 L-divisibility Let us now derive the conditions for L-divisibility of Pauli channels with positive eigenvalues λ1, λ2, λ3 (λ0 = 1). The logarithm of D, induced by the princi- pal logarithm of its eigenvalues, is thus L = Kdiag(0, log λ1, log λ2, log λ3)K−1 , (16) which is real (hermiticity preserving). In case of no- degeneration the dependency on K vanishes and L is unique. In such case the ccp conditions log λj − log λk − log λl ≥ 0 imply λjλk ≤ λl (17) for all combinations of mutually different j, k, l. This set (channels belonging to CL with positive eigenval- ues) forms a three dimensional manifold, see Fig. 6. Accepted in Quantum 2019-04-24, click title to verify 6 In case of degeneration, let us label the eigenvalues η, λ and λ. In this case, the real solution for L is not unique and is parametrized by real branches in the degenerate subspace and by the continuous parame- ters of K [3]. Let us study the principal branch with K = ✶. Eq. (17) is then reduced to λ2 ≤ η ≤ 1 . (18) Therefore, if these inequalities are fulfilled, the gen- erator has Lindblad form. If not, then a priori other branches can fulfill the ccp condition and con- sequently have a Lindblad form. Thus, Eq. (18) pro- vides a sufficient condition for the channel to be in CL. We will see it is also necessary. Indeed, the complete positivity condition requires η, λ ≤ 1, thus, it remains to verify only the condition λ2 ≤ η. It holds for the case λ ≤ η. If η ≤ λ, then this condition coincides with the CP-divisibility con- dition from Eq. (15). Since CL implies CCP the proof is completed. In conclusion, the condition in Eq. (17) is a necessary and sufficient condition for a given Pauli channel with positive eigenvalues to belong to CL. Let us stress that the obtained subset of L-divisible channels does not possess the tetrahedron symme- tries. In fact, composing D with a σz rotation Uz = diag(1, −1, −1, 1) results in the Pauli channel D′ = diag(1, −λ1, −λ2, λ3). Clearly, if λj are positive (D is L-divisible), then D′ has non-positive eigenval- ues. Moreover, if all λj are different, then D′ does not have any real logarithm, therefore, it cannot be L-divisible. In conclusion, the set of L-divisible uni- tal qubit channels is not symmetric with respect to tetrahedron symmetries. In what follows we will investigate the case of non-positive eigenvalues. Theorem 3 implies that that eigenvalues have the form (modulo permuta- tions) η, −λ, −λ, where η, λ ≥ 0. The correspond- ing Pauli channels are Dx = diag(1, η, −λ, −λ), Dy = diag(1, −λ, η, −λ), Dz = diag(1, −λ, −λ, η), thus forming three two-dimensional regions inside the tetrahedron. Take, for instance, Dz that specifies a plane (inside the tetrahedron) containing I, σz and completely depolarizing channel N = diag(1, 0, 0, 0). The real logarithms for this case are given by L = K     0 0 0 0 0 log(λ) (2k + 1)π 0 0 −(2k + 1)π log(λ) 0 0 0 0 log(η)     K−1, (19) where k ∈ Z and K, as mentioned above, belongs to a continuum of matrices that commute with Dz. Note that L is always non-diagonal. For this case (simi- larly for Dx and Dy) the ccp condition reduces again to conditions specified in Eq. (18). Using the same arguments one arrives to a more general conclusion: Eq. (17) provides necessary and sufficient conditions for L − divisibility of a given Pauli channel, and if it is the case, the principal branch with K = ✶ has Lindblad form. The set of L-divisible Pauli channels is illustrated in Fig. 6. In order to decide L-divisibility of general unital channels it remains to analyze the case of complex eigenvalues. The logarithms are parametrized as fol- lows Lk,K = wK log (J)k K−1w−1, (20) where J is the real Jordan form of the (qubit) chan- nel [3]: J = diag (1, c) ⊕ ( a −b b a ) (21) with a ± ib being the complex eigenvalues and c > 0. Let us note that K log (J)k K−1 is reduced to equa- tions (16) and (19) in the case of real eigenvalues. In general, the generator is (up to diagonalization): K log (J)k K−1 = Kdiag (0, log(c)) ⊕ ( log(|z|) arg(z) + 2πk − arg(z) − 2πk log(|z|) ) K−1. with z = a + ib. The non-diagonal block of the loga- rithm has the same structure as the real Jordan form of the channel, so K also commutes with log(J)k, leading to a countable parametric space of hermitian preserving generators. In fact, generators of diagonal- izable channels have continuous parametrizations if and only if they have degenerate eigenvalues; the non- diagonalizable case can be found elsewhere [3]. Since we are dealing with a diagonalization, the ccp condi- tion can be very complicated and depends in general on k, see Eq. (1). But for the complex case we can simplify the condition for qubit channels which have exactly the form presented in Eq. (21), say Êcomplex, i.e. w = ✶. In such case the ccp condition is reduced to a2 + b2 ≤ c ≤ 1. (22) Note that it does not depend on k and the second inequality is always fulfilled for CPTP channels. We can present the conditions for L-divisibility for the case of complex eigenvalues. The orthogonal nor- mal form of Êcomplex is D̂ = diag (1, η, λ, λ) with η = c and λ = sign(ab) √ a2 + b2. The ccp con- dition for degenerated eigenvalues, see Eq. (18), is reduced to Eq. (22) for this case. Therefore, the L-divisible channels with form Êcomplex are also L- divisible, up to unitaries. This also applies for chan- nels arising from changing positions of the 1 × 1 block containing c and the 2 × 2 block containing a and b in Êcomplex, with orthogonal normal forms diag (1, λ, η, λ) and diag (1, λ, λ, η). The set contain- ing them is shown in Fig. 4. 3.3.4 Divisibility relations Consider a Pauli channel with 0 < λmin = λ1 ≤ λ2 ≤ λ3 < 1, thus, the condition λ1λ2 ≤ λ3 triv- ially holds. Since λ1λ2 ≤ λ1λ3 ≤ λ2λ3 ≤ λ2, it Accepted in Quantum 2019-04-24, click title to verify 7 λ1 λ2 λ3 Figure 4: Tetrahedron of Pauli channels, with qubit unital L-divisible channels of the form Êcomplex (see main text). Note that the set does not have the symmetries of the tetrahedron. follows that λ1λ3 ≤ λ2, thus, two (out of three) L- divisibility conditions hold always for Pauli channels with positive eigenvalues. Moreover, one may ob- serve that CP-divisibility condition Eq. (15) reduces to one of L-divisibility conditions λ2λ3 ≤ λ1. In conclusion, the conditions of CP-divisibility and L- divisibility for Pauli channels with positive eigenval- ues coincide, thus, in this case CCP implies CL. Concatenating (positive-eigenvalues) Pauli chan- nels with Dx,y,z one can generate the whole set of CCP Pauli channels. Using the identity CCP = CInf and considering Eq. (11)we can formulate the follow- ing theorem: Theorem 4 (Infinitesimal divisible unital channels). Let ECP unital be an arbitrary infinitesimal divisible unital qubit channel. There exists at least one L-divisible Pauli channel Ẽ, and two unitary conjugations U1 and U2, such that ECP unital = U1ẼU2 . Notice that if ECP unital is invertible, Ẽ = eL. Let us continue with another equivalence relation holding for Pauli channels. Regarding infinitely divis- ibility channels, we know that, in general, CL ⊂ C∞, however, for Pauli channels the corresponding subsets coincide. Theorem 5 (Infinitely divisible Pauli channels). The set of L-divisible Pauli channels is equivalent to the set of infinitely divisible Pauli channels. Proof. A channel is infinitely divisible if and only if it can be written as E0eL, where E0 is an idem- potent channel satisfying E0LE0 = E0L and L has λ1 λ2 λ3 Figure 5: Tetrahedron of Pauli channels with part of the set of CP-divisible, see Eq. (15), but not L-divisible channels (CCP\CL) shown in purple. The whole set CCP is obtained applying the symmetry transformations of the tetrahedron to the purple volume. Lindblad form [4]. The only idempotent qubit chan- nels are contractions of the Bloch sphere into sin- gle points, diagonalization channels Ediag transform- ing Bloch sphere into a line connecting a pair of basis states, and the identity channel. Among the single-point contractions, the only one that is a Pauli channel is the contraction of the Bloch sphere into the complete mixture. In particular, E = N eL = N for all L. The channel N belongs to the clo- sure of CL, because a sequence of channels eLn with L̂n = diag (0, −n, −n, −n) converges to N̂ in the limit n → ∞. For the case of E0 being the identity channel we have E = eL, thus, trivially such infinitely divisi- ble channel E is in CL too. It remains to analyze the case of diagonalization channels. First, let us note that the matrix of eL̂ is necessarily of full rank, since detÊ 6= 0. It follows that the matrix Ê = ÊdiageL̂ has rank two as Êdiag is a rank two matrix, thus, it takes one of the following forms Êλ x = diag (1, λ, 0, 0), Êλ y = diag (1, 0, λ, 0), Êλ z = diag (1, 0, 0, λ). The in- finitely divisibility implies λ > 0 in order to keep the roots of λ real. In what follows we will show that Êz belongs to (the closure of) CL. Let us define the channels Êλ,ǫ z = diag (1, ǫ, ǫ, λ) with ǫ > 0. The com- plete positivity and ccp conditions translate into the inequalities ǫ ≤ 1+λ 2 and ǫ2 ≤ λ, respectively; there- fore one can always find an ǫ > 0 such that Êλ,ǫ z is a L-divisible channel. If we choose ǫ = √ λ/n with n ∈ Z +, the channels Êz,n = diag ( 1, √ λ/n, √ λ/n, λ ) form a sequence of L-divisible channels converging to Êλ z when n → ∞. The analogous reasoning implies that Êλ x , Êλ y ∈ CL too. Let us note that one parame- Accepted in Quantum 2019-04-24, click title to verify 8 λ1 λ2 λ3 Figure 6: Tetrahedron of Pauli channels with the set of L- divisible channels (or equivalently infinitely divisible, see The- orem 5) shown in green, see equations (17) and (18). The solid set corresponds to channels with positive eigenvalues, and the 2D sets correspond to the negative eigenvalue case. The point where the four sets meet corresponds to the total depolarizing channel. Notice that this set does not have the symmetries of the tetrahedron. ter family Ez are convex combinations of the complete diagonalization channel Ê1 z = diag (1, 0, 0, 1) and the complete mixture contraction N̂ . This completes the proof. Finally, let us remark that using theorem 23 of Ref. [25] we conclude that the intersection CP ∩ Cdiv depicted in Fig. 1 is not empty. To show this, notice that applying the mentioned theorem to Pauli chan- nels we get that the faces of the tetrahedron are indi- visible in CPTP channels. However, there are chan- nels with positive determinant inside the faces, for example diag ( 1, 4 5 , 4 5 , 3 5 ) . Therefore we conclude that up to unitaries, CP ∩ Cdiv correspond to the union of the four faces faces of the tetrahedron minus the faces of the octahedron that intersects with the faces of the tetrahedron, see Fig. 3. We have to remove such inter- section since it corresponds to channels with negative determinant, i.e. not in CP. To get a detailed picture of the position and inclu- sions of the divisibility sets, we illustrate in Fig. 7 two slices of the tetrahedron where different types of di- visibility are visualized. Notice the non-convexity of the considered divisibility sets. 3.4 Non-unital qubit channels Similar to unital channels, using theorem 1 we are able to characterize Cdiv, CP and CCP by studying special orthogonal normal forms of non-unital chan- nels. They are characterized by ~λ and ~τ , see Eq. (5). Thus we can study if a channel is Cdiv by computing the rank of its Choi matrix. For this case algebraic equations are in general fourth order polynomials. In fact, in Ref. [19] a condition in terms of the eigenval- ues and ~τ is given. For special cases, however, we can obtain compact expressions, see Fig. 8. The charac- terization of CP is given by Eq. (8) (note that it only depends on ~λ), and CCP is tested, for full Kraus rank non-unital channels, using Eq. (9), see Ref. [24] for the calculation of the si’s. For the characterization of CL we use the results developed at the end of the last section, see Eqs. (18)-(22). We can plot illustrative pictures even though the whole space of qubit channels has 12 parameters. This can be done using orthogonal normal forms and fixing ~τ , exactly in the same way as the unital case. Re- call that unitaries only modify CL, leaving the shape of other sets unchanged. CPTP channels are repre- sented as a volume inside the tetrahedron presented in Fig. 2, see Fig. 8. In the later figure we show a slice corresponding to ~τ = (1/2, 0, 0) T. Indeed, it has the same structure of the slices for the unital case, but deformed, see Fig. 7. A difference with respect to the unital case is that L-divisible channels with negative eigenvalues (up to unitaries) are not completely inside CP-divisible channels. A part of them are inside the CP channels. A central feature of Figs. 7 and 8 is that the set Cdiv \ CP is inside the convex slice of the set of en- tanglement breaking channels (deformed octahedron). Indeed, we can proof the following theorem. Theorem 6 (Entanglement-breaking channels and divisibility). Consider a qubit channel E. If det Ê < 0, then E is entanglement-breaking, i.e. all qubit chan- nels outside CP are entanglement breaking. Proof. Consider the Choi-Jamiołkowski state of a channel E written in the factorized Pauli operator ba- sis [24] τE = 1 4 ∑3 jk Rjkσj ⊗ σk and let Ê be its repre- sentation in the Pauli operator basis. Then the matrix identity R = ÊΦT with ΦT = diag (1, 1, −1, 1) holds. Since det Ê < 0 it follows that det R = − det Ê > 0. Using the aforementioned Lorentz normal decompo- sition R = LT 1 R̃L2 with det L1,2 > 0 and R̃ diagonal, see Ref. [24]. The transformations L1,2 correspond to one-way stochastic local operations and classical com- munications (SLOCC) of τE , thus, R̃ is an unnormal- ized two-qubit state with det R̃ > 0. The channel cor- responding to R̃ (in the Pauli basis) is Ĝ = R̃ΦT/R̃00. Since the latter is diagonal, then G is a Pauli channel with det Ĝ < 0. A Pauli channel has a negative deter- minant, if either all λj are negative, or exactly one of them is negative. In Ref. [28] it has been shown that the set of channels with λj < 0 ∀j are entanglement breaking channels. Now, using the symmetries of the tetrahedron, one can generate all channels with neg- ative determinant by concatenating this set with the Pauli rotations. Therefore every Pauli channel with Accepted in Quantum 2019-04-24, click title to verify 9 λ1 λ2λ3  C L  C CP\CL  C P\CCP  C div\CP  Cdiv  EB boundary Figure 7: We show two slices of the unitary tetrahedron (figure in the left) determined by ∑ i λi = 0.4 (shown in the center) and ∑ i λi = −0.4 (shown in the right). The non-convexity of the divisibility sets can be seen, including the set of indivisible channels. The convexity of sets C and entanglement breaking channels can also be noticed in the slices. A central feature is that the set Cdiv \ CP is always inside the octahedron of entanglement breaking channels. negative determinant is entanglement breaking, thus, τG is separable and given that SLOCC operations can not create entanglement [11], we have that τE is sep- arable, too. This implies that if det Ê < 0, then the qubit channel E is entanglement-breaking. 4 Divisibility transitions and examples with dynamical process The aim of this section is to use illustrative examples of quantum dynamical processes to show transitions between divisibility types of the instantaneous chan- nels. From the slices shown above (see figures 7 and 8) it can be noticed that every transition between the studied divisibility types is permitted. This is due to the existence of common borders between all combi- nations of divisibility sets; we can think of any con- tinuous line inside the tetrahedron [6] as describing some quantum dynamical map. We analyze two examples, the first is an implemen- tation of the approximate NOT gate, ANOT through- out a specific collision model [21]. The second is the well known setting of a two-level atom interacting with a quantized mode of an optical cavity [9]. We de- fine a simple function that assigns a particular value to a channel Et according to divisibility hierarchy, i.e. δ[E ] =        1 if E ∈ CL , 2/3 if E ∈ CCP \ CL , 1/3 if E ∈ CP \ CCP , 0 if E ∈ C \ CP . (23) A similar function can be defined to study the transi- tion to/from the set of entanglement-breaking chan- λ1 λ2 λ3 Figure 8: (left) Set of non-unital unital channels up to uni- taries, defined by ~τ = (1/2, 0, 0), see Eq. (5). This set lies inside the tetrahedron. For this particular case the CP conditions reduce to the two inequalities 2 ± 2λ1 ≥ √ 1 + 4(λ2 ± λ3)2. A cut corresponding to ∑ i λi = 0.3 is presented inside and in the right, see Fig. 7 for the color coding. The structure of divisibility sets presented here has basically the same structure as for the unital case except for CL. A part of the channels with negative eigenvalues belong- ing to CL lies outside CCP \ CL, see green lines. As for the unital case a central feature is that the channels in Cdiv \ CP are entanglement breaking channels. Channels in the bound- ary are not characterized due to the restricted character of Theorem 2. Accepted in Quantum 2019-04-24, click title to verify 10 nels, i.e. χ[E ] = { 1 if E is entanglement breaking , 0 if E if not. (24) The quantum NOT gate is defined as NOT : ρ 7→ ✶ − ρ, i.e. it maps pure qubit states to its orthogo- nal state. Although this map transforms the Bloch sphere into itself it is not a CPTP map, and the clos- est CPTP map is ANOT : ρ 7→ (✶ − ρ)/3. This is a rank-three qubit unital channel, thus, it is indivisible [25]. Moreover, det ANOT = −1/27 implies that this channel is not achievable by a P-divisible dynamical map. It is worth noting that ANOT belongs to Cdiv. A specific collision model was designed in Ref. [21] simulating stroboscopically a quantum dynamical map that implements the quantum NOT gate ANOT in finite time. The model reads Et(̺) = cos2(t)̺ + sin2(t)ANOT(̺) + 1 2 sin(2t)F(̺) , (25) where F(̺) = i 1 3 ∑ j [σj , ̺]. This quantum dynamical map achieves the desired gate ANOT at t = π/2. Let us stress that this dynamical map is unital, i.e. Et(✶) = ✶ for all t, thus, its orthogonal nor- mal form can be illustrated inside the tetrahedron of Pauli channels, see Fig. 9. In Fig. 10 we plot δ[Et], χ[Et] and the value of the det Et. We see the tran- sitions CL → CP \ CCP → Cdiv \ CP → Cdiv and back. Notice that in both plots the trajectory never goes through the CCP \ CL region. This means that when the parametrized channels up to rotations be- long to CL, so do the original ones. The transition be- tween P-divisible and divisible channels, i.e. CP\CCP and Cdiv\CP, occurs at the discontinuity in the yel- low curve in Fig. 9. Let us note that this discontinu- ity only occurs in the space of ~λ; it is a consequence of the orthogonal normal decomposition, see Eq. (5). The complete channel is continuous in the full con- vex space of qubit CPTP maps. The transition from CP \ Cdiv and back occurs at times π/3 and 2π/3. It can also be noted that the transition to entanglement breaking channels occurs shortly before the channel enters in the Cdiv \ CP region; likewise, the chan- nel stops being entanglement breaking shortly after it leaves the Cdiv \ CP region. Consider now the dynamical map induced by a two- level atom interacting with a mode of a boson field. This model serves as a workhorse to explore a great variety of phenomena in quantum optics [8]. Using the well known rotating wave approximation one arrives to the Jaynes-Cummings model [12], whose Hamilto- nian is H = ωa 2 σz + ωf ( a†a + 1 2 ) + g ( σ−a† + σ+a ) . (26) By initializing the environment in a coherent state |α〉, one gets the familiar collapse and revival setting. Considering a particular set of parameters shown λ1 λ2 λ3 Figure 9: (top left) Tetrahedron of Pauli channels with the trajectory, up to rotations, of the quantum dynamical map Eq. (25) leading to the ANOT gate, as a yellow curve. (right) Cut along the plane that contains the trajectory; there one can see the different regions where the channel passes. For this case, the characterization of the CL of the channels induced gives the same conclusions as for the corresponding Pauli channel, see Eq. (5). The discontinuity in the trajec- tory is due to the reduced representation of the dynamical map, see Eq. (25); the trajectory is continuous in the space of channels. See Fig. 7 for the color coding. . 0 π 3 π 2 2 π 3 π 0 1 3 2 3 1 δ [E t ], d et E t , χ [E t ] t C P\CCP C CP\CL C L C div\CP Figure 10: Evolution of divisibility, determinant, and entan- glement breaking properties of the map induced by Eq. (25), see Eq. (23) and Eq. (24). Notice that the channel ANOT, implemented at t = π/2, has minimum determinant. The horizontal gray dashed lines show the image of the function δ, with the divisibility types in the right side. It can be seen that the dynamical map explores the divisibility sets as CL → CP \CCP → Cdiv \CP → Cdiv and back. The channels are entanglement breaking in the expected region. Accepted in Quantum 2019-04-24, click title to verify 11 3.85 3.9 3.95 0 2 4 6 8 10 0 1 3 2 3 1 δ [E t ], p e (t ), χ [E t ] t C P\CCP C CP\CL C L C div\CP Figure 11: Black and red curves show functions δ and χ of the channels induced by the Jaynes-Cummings model over a two- level system, see Eq. (26) with the environment initialized in a coherent state |α〉. The blue curve shows the probability of finding the two-level atom in its excited state, pe(t). The figure shows that the fast oscillations in δ occur roughly at the same frequency as the ones of pe(t), see the inset. Notice that there are fast transitions between CP \ CCP and CCP \ CL occurring in the region of revivals, with a few transitions between CCP \CP and CL in the second revival. The function χ shows that during revivals channels are not entanglement breaking, but we find that channels belonging to Cdiv \CP are always entanglement breaking, in agreement with theorem 6. The particular chosen set of parameters are α = 6, g = 10, ωa = 5, and ωf = 20. in Fig. 11, we constructed the channels parametrized by time numerically, and studied their divisibility and entanglement-breaking properties. In the same fig- ure we plot functions δ[Et] and χ[Et], together with the probability of finding the atom in its excited state pe(t), to study and compare the divisibility properties with the features of the collapses and revivals. The probability pe(t) is calculated choosing the ground state of the free Hamiltonian ωa/2σz of the qubit, and it is given by [13]: pe(t) = 〈σz(t)〉 + 1 2 , (27) where 〈σz(t)〉 = − ∞ ∑ n=0 Pn ( ∆2 4Ω2 n + ( 1 − ∆2 4Ω2 n ) cos (2Ωnt) ) , with Pn = e−|α|2 |α|2n/n!, Ωn = √ ∆2/4 + g2n and ∆ = ωf − ωa the detuning. The divisibility indicator function δ exhibits an os- cillating behavior, roughly at the same frequency of pe(t), see inset in Fig. 11. The figure shows fast peri- odic transitions between CP \ CCP and CCP \ CL oc- curring in the region of revivals. There are also few transitions among CCP \ CP and CL in the second re- vival. Respect to the entanglement breaking and the function χ, there are no fast transitions in the former, and during revivals, channels are not entanglement breaking. We also observe that channels belonging to Cdiv \ CP are entanglement breaking, supporting theorem 6 for the non-unital case. 5 Conclusions We studied the relations between different types of di- visibility of time-discrete and time-continuous quan- tum processes, i.e. channels and dynamical maps, re- spectively. In particular, we investigated classes of channels by means of their achievability by dynamical maps of different divisibility types, and also the divis- ibility of channels occurring during the time evolu- tions. Apart from investigating the relations between these concepts in general, we provided a detailed anal- ysis for the case of qubit channels. We implemented the known conditions to decide CL for the general diagonalizable case, and a discus- sion of the parametric space of Lindblad generators was given (clarifying one of the results of the paper [26]). For unital qubit channels it was shown that every infinitesimal divisible map can be written as a concatenation of one CL channel and two unitary con- jugations. For the particular case of Pauli channels case, we have shown that the sets of infinitely divis- ible and L-divisible channels coincide. We made an interesting observation, connecting the concept of di- visibility with the quantum information paradigm of entanglement-breaking channels. We found that di- visible but not infinitesimal divisible qubit channels, in PTP maps, are necessarily entanglement-breaking. We also noted that the intersection of indivisible and P-divisible channels is not empty. This allows us to implement indivisible channels with infinitesimal PTP maps. Finally, we questioned the existence of dynam- ical transitions between different classes of divisibility channels. We argued that all the transitions are, in principle, possible, and exploited two simple models of dynamical maps to demonstrate these transitions. They clearly illustrate how the channels evolutions change from being implementable by markovian dy- namical maps to non-markovian, and vice versa. There are several directions how to proceed further in investigation of divisibility of channels and dynam- ical maps. Apart from extension of this analysis to larger-dimensional systems, a plethora of interesting questions are related to design of efficient verification procedures of the divisibility classes for channels and dynamical maps. In this paper we question the divisi- bility features of snapshots of the evolution, however, it might be of interest to understand when the time intervals of dynamical maps implemented by non- markovian evolutions, can be simulated by markovian dynamical maps. Also the area of channel divisibility contains several open structural questions, e.g. the existence of at most n-divisible channels. Acknowledgements We acknowledge Thomas Gorin and Tomáš Rybár for useful discussions, as well PAEP and RedTC for finan- cial support. Support by projects CONACyT 285754, Accepted in Quantum 2019-04-24, click title to verify 12 UNAM-PAPIIT IG100518, IN-107414, APVV-14- 0878 (QETWORK) is acknowledged. CP acknowl- edges support by PASPA program from DGAPA- UNAM. MZ acknowledges the support of VEGA 2/0173/17 (MAXAP) and GAČR project no. GA16- 22211S. A On Lorentz normal forms of Choi- Jamiolkowski state In this appendix we compute the Lorentz normal de- composition of a channel for which one gets b 6= 0, supporting our observation that Lorentz normal de- composition does not take Choi-Jamiołkowski states to something proportional to a Choi-Jamiołkowski state. Consider the following Kraus rank three chan- nel and its RE matrix, both written in the Pauli basis: Ê =     1 0 0 0 0 − 1 3 0 0 0 0 − 1 3 0 2 3 0 0 1 3     , (28) and RE =     1 0 0 0 0 − 1 3 0 0 0 0 1 3 0 2 3 0 0 1 3     . (29) Using the algorithm introduced in Ref. [24] to calcu- late RE ’s Lorentz decomposition into orthochronous proper Lorentz transformations we obtain L1 = 1 γ1     4 0 0 1 0 −γ1 0 0 0 0 −γ1 0 1 0 0 4     , (30) L2 = 1 γ2     89 + 9 √ 97 0 0 −8 0 −γ2 0 0 0 0 −γ2 0 −8 0 0 89 + 9 √ 97     , and ΣE = 1 γ3       √ 11 + 109√ 97 0 0 − √ 97+1√ 89 √ 97+873 0 − γ3 3 0 0 0 0 γ3 3 0 √ 1 + 49√ 97 0 0 √ −1 + 49√ 97       with γ1 = √ 15, γ2 = 3 √ 178 √ 97 + 1746, and γ3 =√ 30. Although the central matrix ΣE is not exactly of the form Eq. (6), it is equivalent. To see this no- tice that the derivation of the theorem 2 in [24] con- siders only decompositions into proper orthochronous Lorentz transformations. But to obtain the desired form, the authors change signs until they get Eq. (6); this cannot be done without changing Lorentz trans- formations. If we relax the condition over L1,2 of being proper and orthochronous, we can bring ΣE to the desired form by conjugating ΣE with G = diag (1, 1, 1, −1): G−1ΣEG = 1 γ3       √ 11 + 109√ 97 0 0 √ 97+1√ 89 √ 97+873 0 − γ3 3 0 0 0 0 γ3 3 0 − √ 1 + 49√ 97 0 0 √ −1 + 49√ 97       . In both cases (taking ΣE or G−1ΣEG as the normal form of RE), the corresponding channel is not pro- portional to a trace-preserving one since b 6= 0, see Eq. (6). This completes the counterexample. References [1] Ángel Rivas, Susana F Huelga, and Martin B Plenio. Quantum non-markovianity: characteri- zation, quantification and detection. Rep. Prog. Phys., 77(9):094001, 2014. DOI: 10.1088/0034- 4885/77/9/094001. [2] I. Bengtsson and K. Życzkowski. Geome- try of Quantum States: An Introduction to Quantum Entanglement. Cambridge Univer- sity Press, 2017. ISBN 9781107026254. URL https://books.google.com.mx/books?id= sYswDwAAQBAJ. [3] W. J Culver. On the Existence and Uniqueness of the Real Logarithm of a Matrix. Proceed- ings of the American Mathematical Society, 17 (5):1146–1151, 1966. DOI: 10.1090/S0002-9939- 1966-0202740-6. [4] L. V. Denisov. Infinitely Divisible Markov Map- pings in Quantum Probability Theory. The- ory Prob. Appl., 33(2):392–395, 1989. DOI: 10.1137/1133064. [5] D. E. Evans and J. T. Lewis. Dilations of Irreversible Evolutions in Algebraic Quantum Theory, volume 24 of Communications of the Dublin Institute for Advanced Studies: Theoreti- cal physics. Dublin Institute for Advanced Stud- ies, 1977. URL http://orca.cf.ac.uk/34031/. [6] S. N. Filippov, J. Piilo, S. Maniscalco, and M. Zi- man. Divisibility of quantum dynamical maps and collision models. Phys. Rev. A, 96(3):032111, 2017. DOI: 10.1103/PhysRevA.96.032111. [7] V. Gorini, A. Kossakowski, and E. C. G. Sudar- shan. Completely positive dynamical semigroups of N-level systems. J. Math. Phys., 17(5):821, 1976. DOI: 10.1063/1.522979. [8] A. D. Greentree, J. Koch, and J. Larson. Fifty years of Jaynes–Cummings physics. J. Phy. B, 46(22):220201, 2013. DOI: 10.1088/0953- 4075/46/22/220201. Accepted in Quantum 2019-04-24, click title to verify 13 [9] S. Haroche and J.-M. Raimond. Exploring the Quantum: Atoms, Cavities, and Photons. Ox- ford University Press, USA, 2006. URL http: //www.worldcat.org/isbn/0198509146. [10] T. Heinosaari and M. Ziman. The Mathemati- cal Language of Quantum Theory: From Uncer- tainty to Entanglement. Cambridge University Press, 2012. DOI: 10.1017/CBO9781139031103. [11] R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki. Quantum entanglement. Rev. Mod. Phys., 81(2):865–942, 2009. DOI: 10.1103/RevModPhys.81.865. [12] E. T. Jaynes and F. W. Cummings. Comparison of quantum and semiclassical radiation theories with application to the beam maser. Proc. IEEE, 51:89, 1963. DOI: 10.1109/PROC.1963.1664. [13] A. B. Klimov and S. M. Chumakov. A Group- Theoretical Approach to Quantum Optics: Mod- els of Atom-Field Interactions. Wiley-VCH, 2009. DOI: 10.1002/9783527624003. [14] A. Kossakowski. On quantum statistical me- chanics of non-hamiltonian systems. Rep. Math. Phys., 3(4):247 – 274, 1972. DOI: 10.1016/0034- 4877(72)90010-9. [15] J. M. Leinaas, J. Myrheim, and E. Ovrum. Ge- ometrical aspects of entanglement. Phys. Rev. A, 74:012313, Jul 2006. DOI: 10.1103/Phys- RevA.74.012313. [16] G. Lindblad. On the generators of quantum dy- namical semigroups. Comm. Math. Phys., 48(2): 119–130, 1976. DOI: 10.1007/BF01608499. [17] M. Musz, M. Kuś, and K. Życzkowski. Uni- tary quantum gates, perfect entanglers, and unis- tochastic maps. Phys. Rev. A, 87:022111, Feb 2013. DOI: 10.1103/PhysRevA.87.022111. [18] C. Pineda, T. Gorin, D. Davalos, D. A. Wis- niacki, and I. García-Mata. Measuring and us- ing non-Markovianity. Phys. Rev. A, 93:022117, 2016. DOI: 10.1103/PhysRevA.93.022117. [19] Ł. Rudnicki, Z. Puchała, and K. Zyczkowski. Gauge invariant information concerning quan- tum channels. Quantum, 2:60, April 2018. ISSN 2521-327X. DOI: 10.22331/q-2018-04-11-60. [20] M. B. Ruskai, S. Szarek, and E. Werner. An anal- ysis of completely-positive trace-preserving maps on M2. Lin. Alg. Appl., 347(1):159 – 187, 2002. DOI: 10.1016/S0024-3795(01)00547-X. [21] T. Rybár, S. N. Filippov, M. Ziman, and V. Bužek. Simulation of indivisible qubit channels in collision models. J. Phys. B, 45(15):154006, 2012. DOI: 10.1088/0953- 4075/45/15/154006. [22] B. Vacchini, A. Smirne, E.-M. Laine, J. Pi- ilo, and H.-P. Breuer. Markovianity and non- markovianity in quantum and classical sys- tems. New J. Phys., 13(9):093004, 2011. DOI: 10.1088/1367-2630/13/9/093004. [23] F. Verstraete and H. Verschelde. On quan- tum channels. Unpublished, 2002. URL http: //arxiv.org/abs/quant-ph/0202124. [24] F. Verstraete, J. Dehaene, and B. DeMoor. Lo- cal filtering operations on two qubits. Phys. Rev. A, 64(1):010101, 2001. DOI: 10.1103/Phys- RevA.64.010101. [25] M. M. Wolf and J. I. Cirac. Dividing quantum channels. Comm. Math. Phys., 279(1):147–168, 2008. DOI: 10.1007/s00220-008-0411-y. [26] M. M. Wolf, J. Eisert, T. S. Cubitt, and J. I. Cirac. Assessing non-Markovian quantum dy- namics. Phys. Rev. Lett., 101(15):150402, 2008. DOI: 10.1103/PhysRevLett.101.150402. [27] M. Ziman and V. Bužek. All (qubit) decoher- ences: Complete characterization and physical implementation. Phys. Rev. A, 72:022110, Aug 2005. DOI: 10.1103/PhysRevA.72.022110. [28] M. Ziman and V. Bužek. Concurrence versus purity: Influence of local channels on Bell states of two qubits. Phys. Rev. A, 72(5):052325, 2005. DOI: 10.1103/PhysRevA.72.052325. Accepted in Quantum 2019-04-24, click title to verify 14 114 Chapter C. Articles C.2 Article: Gaussian quantum channels beyond the Gaus- sian functional form: full characterization of the one- mode case The following article was sent to Physical Review A. Click to go to arXiv. Gaussian quantum channels beyond the Gaussian functional form: full characterization of the one-mode case David Davalos,1 Camilo Moreno,2 Juan-Diego Urbina,2 and Carlos Pineda1 1Instituto de F́ısica, Universidad Nacional Autónoma de México, México∗ 2Institut für Theoretische Physik, Universität Regensburg, D-93040 Regensburg, Germany We study one-mode Gaussian quantum channels in continuous-variable systems by performing a black-box characterization using complete positivity and trace preserving conditions, and report the existence of two subsets that do not have a functional Gaussian form. Our study covers as particular limit the case of singular channels, thus connecting our results with their known classi- fication scheme based on canonical forms. Our full characterization of Gaussian channels without Gaussian functional form is completed by showing how Gaussian states are transformed under these operations, and by deriving the conditions for the existence of master equations for the non-singular cases. PACS numbers: 03.65.Yz, 03.65.Ta, 05.45.Mt I. INTRODUCTION Within the theory of continuous-variable quantum sys- tems (a central topic of study given their role in the description of physical systems like the electromagnetic field [1], solids and nano-mechanical systems [2] and atomic ensembles [3]) the simplest states, both from a theoretical an experimental point of view, are the so- called Gaussian states. An operation that transforms such family of states into itself is called a Gaussian quan- tum channel (GQC). Even though Gaussian states and channels form small subsets among general states and channels, they have proven to be useful in a variate of tasks such as quantum communication [4], quantum com- putation [5] and the study of quantum entanglement in simple [6] and complicated scenarios [7]. Writing Gaussian channels in the position state repre- sentation is often of theoretical convenience, for instance for the calculation of position correlation functions. Thus an obvious way to proceed is to characterize the possible functional forms of GQC in such representation. First at- tempts in this direction were given in Ref. [8], but their ansatz is limited to only Gaussian functional forms (de- noted simply by Gaussian forms or GF). Going beyond such restrictive assumption, in the present work we char- acterize another two possible forms that can arise directly from the definition of Gaussian channel in the one-mode case. We thus give a complete characterization of GQC in position state representation, and study the special case of singular Gaussian quantum channels (SGQC), i.e. the operations for which the inverse operation doesn’t exist. The paper is organized as follows. In section II we discuss the definition of GQC and introduce functional forms beyond the GF that emerge from singularities in the coefficients that define a GQC with GF. In section III we give a black-box characterization of such channels, us- ing complete positivity and trace preserving conditions. ∗ davidphysdavalos@gmail.com In section IV we study functional forms that lead to SGQC and derive their explicit form. Finally in section V we derive conditions of existence of master equations and their explicit forms. We conclude in section VI. II. GAUSSIAN QUANTUM CHANNELS Gaussian states are characterized completely by first (mean) and second (correlations) moments encoded in the mean vector d⃗ and the covariance matrix σ. There- fore, a Gaussian state S can be denoted as S = S (σ, d⃗), where for the one-mode case we have σ = ( ⟨q̂2⟩ − ⟨q̂⟩2 1 2 ⟨q̂p̂ + p̂q̂⟩ − ⟨q̂⟩⟨p̂⟩ 1 2 ⟨q̂p̂ + p̂q̂⟩ − ⟨q̂⟩⟨p̂⟩ ⟨p̂2⟩ − ⟨p̂⟩2 ) , and d⃗ = (⟨q̂⟩, ⟨p̂⟩)T with q̂ and p̂ denoting the standard position and momen- tum (quadrature) operators [9]. To start with, we recall the following definition [10]: Definition 1 (Gaussian quantum channels). A quantum channel is Gaussian (GQC) if and only if it transforms Gaussian states into Gaussian states. This definition is strictly equivalent to the statement that any GQC, say G, can be written as G[ρ] = trE [U (ρ⊗ ρE)U †] (1) where U is a unitary transformation, acting on a com- bined global state obtained from enlarging the system with an environment E, that is generated by a quadratic bosonic Hamiltonian (i.e. U is a Gaussian unitary) [10]. The environmental initial state ρE is a Gaussian state and the trace is taken over the environmental degrees of freedom. Following definition 1, a GQC is fully characterized by its action over Gaussian states, and this action is in ar X iv :1 9 0 2 .0 7 1 6 3 v 1 [q u an t- p h ] 1 9 F eb 2 0 1 9 2 turn defined by affine transformations [10]. Specifically,G = G (T,N, τ⃗) is given by a tuple (T,N, τ⃗) where T and N are 2 × 2 real matrices with N = NT [10] acting on Gaussian states according to G (T,N, τ⃗) [S (σ, d⃗)] = S (TσTT +N,Td⃗ + τ⃗). In the particular case of closed systems we have N = 0 and T is a symplectic matrix. Let us note that although channels with Gaussian form trivially transform Gaussian states into Gaussian states, the definition goes beyond GF. Introducing difference and sum coordinates, x = q2 − q1 and r = (q1 + q2)/2, and ρ(x, r) = ⟨r − x 2 ∣ ρ̂ ∣r + x 2 ⟩, a quantum channel ρf (xf , rf) = ∫ R2 dxidriJ(xf , xi; rf , ri)ρi (ri, xi) , (2) maps an initial ρ̂i into a final ρ̂f state linearly through the kernel J(xf , xi; rf , ri). In order to see how a chan- nel without GF can be constructed as a limiting case of a quantum channel with GF, consider the general parametrization of the later as given in [12] JG(xf , xi; rf , ri) = b3 2π exp [ı(b1xfrf + b2xfri + b3xirf + b4xiri + c1xf + c2xi) − a1x2 f − a2xfxi − a3x2 i ], (3) where all coefficients are real and no quadratic terms in ri,f are allowed. Now it is easy to see that if the co- efficients of the quadratic form in the exponent of J in eq. (3) depend on a parameter ǫ such that for ǫ→ 0 they scale as an ∝ ǫ−1 and bn ∝ ǫ−1/2, then lim ǫ→0 JG(xf , xi; rf , ri) =N δ(αxf −βxi)eΣ′(xf ,xi;rf ,ri), (4) where α, β ∈ R and Σ′(xf , xi; rf , ri) is a quadratic form that now admits quadratic terms in ri,f . This is the first example of a δGQC, namely a Gaussian quantum channel that contains Dirac-delta functions in its coordi- nate representation. This particular example is not only of academic interest. Physically, it can be implemented by means of the ubiquitous quantum Brownian motion model for harmonic systems (damped harmonic oscilla- tor) [13]. In such system δGQC occur at isolated points of time, defined in the limit of the antisymmetric position autocorrelation function tending to zero. Since the form of eq. (4) admits quadratic terms in ri,f , it suggest that a form with two deltas can exist and can be defined with an appropriate limit. In order to avoid working with such limits, in this work we provide a black-box characterization of general GQCs without Gaussian form. In particular we study channels that can arise when singularities on the coefficients of Gaussian forms GF occur, that lead immediately to singular Gaus- sian operations. We characterize which forms in δGQC lead to valid quantum channels, and under which condi- tions singular operations lead to valid singular quantum channels (SGQC). We will show that only two possible forms of δGQC hold according to trace preserving (TP) and hermiticity preserving (HP) conditions. The chan- nel of eq. (4) is one of these forms, as expected. Later on we will impose complete positivity in order to have valid GQC, i.e. complete positive and trace preserving (CPTP) Gaussian operations, going beyond previous characteri- zations of GQC [12]. III. COMPLETE POSITIVE AND TRACE-PRESERVING δ−GAUSSIAN OPERATIONS Let us introduce the ansätze for the possible forms of GQC in the position representation, to perform the black-box characterization. Following eq. (1) and taking the continuous variable representation of difference and sum coordinates, the trace becomes an integral over posi- tion variables of the environment. Then we end up with a Fourier transform of a multivariate Gaussian, having for one mode the following structures: a Gaussian form eq. (3), a Gaussian form multiplied with one-dimensional delta or a Gaussian form multiplied by a two-dimensional delta. Thus, in order to start with the black-box charac- terization, we shall propose the following general Gaus- sian operations with one and two deltas, respectively JI(xf , rf ;xi, ri) =NIδ(α⃗Tv⃗f + β⃗Tv⃗i)eΣ(xf ,xi;rf ,ri), (5) JII(xf , rf ;xi, ri) =NIIδ(Av⃗f −Bv⃗i)eΣ(xf ,xi;rf ,ri), (6) where the exponent reads Σ(xf , xi; rf , ri) = ı(b1xfrf + b2xfri + b3xirf + b4xiri + c1xf + c2xi + d1rf + d2ri) −a1x2 f − a2xfxi − a3x2 i − e1r2f − e2rfri − e3r2i , (7) v⃗i,j = (ri,j , xi,j) and NI,II are normalization constants. They provide, together with eq. (3) all possible ansätze for GQC. Note that the coefficients in the exponential of every form must be finite, otherwise the functional form can be modified. Let us study now CPTP conditions, since complete pos- itivity implies positivity and in turn it implies hermiticity preserving (HP). For sum and difference coordinates HP reads J(−xf , rf ;−xi, ri) = J(xf , rf ;xi, ri)∗. (8) Following this equation, it is easy to note that the coef- ficients an, bn, cn, en must be real with dn = 0, as well the entries of matrices (and vectors) A, B, α⃗, β⃗. The factor concerning the delta function of eq. (5), is reduced into two possible combinations variables: i) δ(αxf −βxi) and ii) δ(αrf −βri). For the case of eq. (6), the prefactor concerning the two-dimensional delta is reduced to iii) δ(γrf − ηri)δ(αxf − βxi). Let us now analyze the trace preserving condition (TP), which for continuous variable systems reads ∫ R drfJ(xf = 0, rf ;xi, ri) = δ(xi). (9) 3 This condition immediately discards ii) from the above combinations of deltas, thus we end up with cases i) and iii). For case i) TP reads NI ∫ drfδ(−βxi)eΣ = NI∣β∣ √ π e1 δ(xi)e e2 2 r2 i 4e1 e−e3r 2 i , (10) thus the relation between the coefficients assumes the form e22 4e1 − e3 = 0, (11) and the normalization constant NI = ∣β∣√ e1 π with β ≠ 0 and e1 > 0. For case iii) the trace-preserving condition reads NII ∫ drfδ(γrf − ηri)δ(−βxi)eΣ = NII∣βγ∣δ(xi)e−e1( ηγ )2r2i −e2 η γ r2i −e3r2i . Thus the following relation between en coefficients must be fulfilled e1 ⎛⎝ηγ⎞⎠ 2 + e2 η γ + e3 = 0, (12) with γ, β ≠ 0 and NII = ∣βγ∣. In the particular case of η = 0, eq. (12) is reduced to e3 = 0. As expected from the analysis of limits above, we showed that δGQC admit quadratic terms in ri,j . Up to this point we have hermitian and trace preserv- ing Gaussian operations; to derive the remaining CPTP conditions, it is useful to write Wigner’s function and Wigner’s characteristic function, which we now derive. The representation of the Wigner’s characteristic func- tion reads χ(k⃗) = tr [ρD(k⃗)] = exp [−1 2 k⃗T (ΩσΩT) k⃗ − ı (Ω⟨x̂⟩)T k⃗] (13) and its relation with Wigner’s function W (x) = ∫ R2 dx⃗e−ıx⃗ TΩk⃗χ (k⃗) (14) = ∫ R eıpxdx ⟨r − x 2 ∣ ρ̂ ∣r + x 2 ⟩ . (15) where k⃗ = (k1, k2)T, x⃗ = (r, p)T and h̵ = 1 (we are using natural units). Using the previous equations to construct Wigner and Wigner’s characteristic functions of the ini- tial and final states, and substituting them in eq. (2), it is straightforward to get the propagator in the Wigner’s characteristic function representation J̃ (k⃗f , k⃗i) = ∫ R6 dΓK(l⃗)J(v⃗f , v⃗i), (16) where the transformation kernel reads K(l⃗) = 1(2π)3 e[ı(k f 2 rf−kf 1 pf−ki 2ri+ki 1pi−pixi+pfxf)], with dΓ = dpfdpidxfdxidrfdri and l⃗ = (pf , pi, xf , xi, rf , ri)T . By elementary integration of eq. (16) one can show that for both cases J̃I,III (k⃗f , k⃗i) = δ (ki1 − α β kf1) δ (ki2 − φ⃗T I,IIIk⃗f) ePI,III(k⃗f ), (17) where PI,III(k⃗f) = ∑2 i,j=1 P (I,III) ij kfi k f j +∑2 i=1 P (I,III) 0i kfi with P (I,III) ij = P (I,III)ji . For case i) we obtain P (I) 11 = −((αβ ) 2 (a3 + b23 4e1 ) + α β (a2 + 1 2 b1b3 e1 ) + a1 + b21 4e1 ) , P (I) 12 = −(αβ b3 2e1 + b1 2e1 ) , P (I) 22 = − 1 4e1 . (18) For case iii) we have P (III) 11 = −((α β )2 a3 + α β a2 + a1) , P (III) 12 = P (III) 22 = 0, (19) And for both cases we have P (I,III) 01 = ı (α β c2 + c1) and P (I,III) 02 = 0. Vectors φ⃗ are given by φ⃗I = (α β (b4 − b3e2 2e1 ) − b1e2 2e1 + b2,− e2 2e1 )T , φ⃗III = (α β η γ b3 + α β b4 + η γ b1 + b2, η γ )T . (20) We are now in position to write explicitly the condi- tions for complete positivity. Having a Gaussian opera- tion characterized by (T,N, τ⃗), the CP condition can be expressed in terms of the matrix C =N + ıΩ − ıTΩTT, (21) where Ω = ( 0 1−1 0 ) is the symplectic matrix. An op- eration G (T,N, τ⃗) is CP if and only if C ≥ 0 [10, 14]. Applying the propagator on a test characteristic func- tion, eq. (13), it is easy to compute the corresponding tuples. For both cases we get (TI,III,NI,III, τ⃗I,III): NI,III = 2( −P22 P12 P12 −P11 ) , τ⃗I,III = (0, ıP (I,III)01 )T , (22) 4 while for case i) matrix T is given by TI = ( e2 2e1 0 φ⃗I,1 −α β ) , (23) where φ⃗I,1 denotes the first component of vector φ⃗I, see eq. (20). The complete positive condition is given by the inequalities raised from the eigenvalues of matrix eq. (21) ± √ α2e22 + 4αβe2e1 + 4β2e21 (4P (I)12 2 + (P (I)11 − P (I)22 )2 + 1) 2βe1− (P (I)11 + P (I)22 ) ≥ 0. (24) For case iii) matrix T is TIII = ( − η γ 0 φ⃗III,1 −α β ) , (25) and complete positivity conditions read ± √(βγ − αη)2 + β2γ2P (III) 11 2 βγ − P (III)11 ≥ 0. (26) Note that in both cases complete positivity conditions do not depend on φ⃗. IV. ALLOWED SINGULAR FORMS There are two classes of Gaussian singular channels. Since the inverse of a Gaussian channel G (T,N, τ⃗) isG (T−1,−T−1NT−T ,−T−1τ⃗), its existence rests on the invertibility of T. Thus studying the rank of the lat- ter we are able to explore singular forms. We are going to use the classification of one-mode channels developed by Holevo [15]. For singular channels there are two classes character- ized by its canonical form [16], i.e. any channel can be obtained by applying Gaussian unitaries before and after the canonical form. The class called “A1” correspond to singular channels with Rank (T) = 0 and coincide with the family of total depolarizing channels. The class “A2” is characterized by Rank (T) = 1. Both channels are entanglement-breaking [16]. Before analyzing the functional forms constructed in this work, let us study channels with GF. The tuple of the affine transformation, corresponding to the propagator JG, eq. (3), were introduced in Ref. [12] up to some typos. Our calculation for this tuple is TG = ( − b4 b3 1 b3 b1b4 b3 − b2 − b1 b3 ) , NG = ⎛⎝ 2a3 b2 3 a2 b3 − 2a3b1 b2 3 a2 b3 − 2a3b1 b2 3 −2(−a3b 2 1 b2 3 + a2b1 b3 − a1) ⎞⎠ , τ⃗G = (−c2 b3 , b1c2 b3 − c1)T . (27) It is straightforward to check that for b2 = 0, TG is singu- lar with Rank (TG) = 1, i.e. it belongs to class A2. Due to the full support of Gaussian functions, it was surpris- ing that Gaussian channels with GF have singular limit. In this case the singular behavior arises from the lack of a Fourier factor for xfri. This is the only singular case for GF. Now we analyze functional forms derived in sec. III. The complete positivity conditions of the form J̃III, pre- sented in eq. (26), have no solution for α → 0 and/or γ → 0, thus this form cannot lead to singular channels. This is not the case for J̃I, eq. (17), which leads to sin- gular operations belonging to class A2 for αe2 = 0, (28) and to class A1 for e2 = α = b2 = 0. (29) For the latter, the complete positivity conditions read e1 ≤ a1. (30) By using an initial state characterized by σi and d⃗i we can compute the explicit dependence of the final states on the initial parameters. For channels belonging to class A2 [see eq. (27) with b2 = 0 and eq. (23) with e2α = 0] the final state only depends one combination of the compo- nents of σi, and in one combination of the components of d⃗i, i.e. ∑mn lmn (σi)mn and ∑m nm (d⃗i)m, respectively, where lmn and nm depend on the channel parameters. See the appendix for the explicit formulas and fig. 1 for an schematic description of the final states. From such com- binations it is obvious that we cannot solve for the initial state parameters given a final state as expected; this is because the parametric space dimension is reduced from 5 to 2. The channel belonging to A1 [see eq. (23) with e2 = α = b2 = 0 and eq. (30)] maps every initial state to a single one characterized by σf =N and d⃗f = (0,−c1)T, see fig. 2 for a schematic description. According to our ansätze [see equations (5) and 6)], we conclude that one-mode SGQC can only have the func- tional forms given in eq. (3) and eq. (5). This is the central result our work and can be stated as: Theorem 1 (One-mode singular Gaussian channels). A one-mode Gaussian quantum channel is singular if and only if it has one of the following functional forms in the position space representation 1. b3 2π exp [ı(b1xfrf + b3xirf + b4xiri + c1xf + c2xi) − a1x2 f − a2xfxi − a3x2 i ], 2. ∣β∣√e1/πδ(αxf − βxi)× exp [ − a2xfxi − a1x2 f − a3x2 i +ı(b2xfri + b3rfxi + b1rfxf + b4rixi + c1xf + c2xi) −e1r2f − e2rfri − e22r 2 i 4e1 ], with e2α = 0. 5 Corollary 1 (Singular classes). A one-mode singular Gaussian channel belongs to class A1 if and only if its position representation has the following form: √ e1/πδ(xi) exp [−a1x2 f +ı(b2xfri+b1rfxf +c1xf)−e1r2f ]. Otherwise the channel belongs to class A2. Since channels on each class are connected each other by unitary conjugations [15], a consequence of the the- orem and the subsequent corollary is that the set of al- lowed forms must remain invariant under unitary conju- gations. To show this we must know the possible func- tional forms of Gaussian unitaries. They are given by following lemma for one mode Lemma 1 (One-mode Gaussian unitaries). Gaussian unitaries can have only GF or the one given by eq. (6). Proof. Recalling that for a unitary GQC, Tmust be sym- plectic (TΩTT = Ω) and N = 0. However, an inspec- tion to eq. (18) lead us to note that N ≠ 0 unless e1 di- verges. Thus Gaussian unitaries cannot have the form JI [see eq. (5)]. An inspection of matrices T and N of GQC with GF [see eq. (27)] and the ones for JII [see equations (19) and (25)] lead us to note the following two observa- tions: (i) in both cases we have N = 0 for an = 0 ∀n; (ii) the matrix T is symplectic for GF when b2 = b3, and when αη = βγ for JII. In particular the identity map has the last form. This completes the proof. One can now compute the concatenations of the SGQCs with Gaussian unitaries. This can be done straightforward using the well known formulas for Gaus- sian integrals and the Fourier transform of the Dirac delta. Given that the calculation is elementary, and for sake of brevity, we present only the resulting forms of each concatenation. To show this compactly we intro- duce the following abbreviations: Singular channels be- longing to class A2 with form JI and with α = 0, e2 = 0 and α = e2 = 0, will be denoted as δαA2 , δe2A2 and δα,e2A2 , re- spectively; singular channels belonging to the same class but with GF will be denoted as GA2 ; channels belong- ing to class A1 will be denoted as δA1 ; finally Gaussian unitaries with GF will be denoted as GU and the ones with form JII as δU . Writing the concatenation of two channels in the position representation as J(f)(xf , rf ;xi, ri) = ∫ R2 dx′dr′J (1) (xf , rf ;x ′, r′)J(2) (x′, r′;xi, ri) , (31) the resulting functional forms for J(f) are given in table I. As expected, the table shows that the integral has only the forms stated by our theorem. Additionally it shows the cases when unitaries change the functional form of class A2, while for class A1 J(f) has always the unique form enunciated by the corollary. J(1) J(2) J(f) δαA2 GU GA2 GU δαA2 δαA2 δαA2 δU δαA2 δU δαA2 δαA2 δe2A2 GU δe2A2 GU δe2A2 GA2 δe2A2 δU δe2A2 δU δe2A2 δe2A2 GU , δ α,e2 A2 δ α,e2 A2 δ α,e2 A2 GU GA2 δU , δ α,e2 A2 δ α,e2 A2 δ α,e2 A2 δU δ α,e2 A2 δU ,GU δA1 δA1 δA1 δU ,GU δA1 TABLE I. The first and second columns show the functional forms of J(1) and J(2), respectively. The last column shows the resulting form of the concatenation of them [see eq. (31)]. See main text for symbol coding. V. EXISTENCE OF MASTER EQUATIONS In this section we show the conditions under which master equations, associated with the channels derived in sec. III, exist. To be more precise, we study if the functional forms derived above parametrize channels be- longing to one-parameter differentiable families of GQCs. As a first step, we let the coefficients of forms presented in equations (5) and (6) to depend on time. Later we de- rive the conditions under which they bring any quantum state ρ(x, r; t) to ρ(x, r; t + ǫ) (with ǫ > 0 and t ∈ [0,∞)) smoothly, while holding the specific functional form of the channel, i.e. ρ(x, r; t + ǫ) = ρ(x, r; t) + ǫLt [ρ(x, r; t)] +O(ǫ2), (32) where both ρ(x, r; t) and ρ(x, r; t+ǫ) are propagated from t = 0 with channels either with the form JI or JII, and Lt is a bounded superoperator in the state subspace. This is basically the problem of the existence of a master equa- tion ∂tρ(x, r; t) = Lt [ρ(x, r; t)] , (33) for such functional forms. Thus the problem is reduced to prove the existence of the linear generator Lt, also known as Liouvillian. To do this we use an ansatz proposed in Ref. [17] to investigate the existence and derive the master equation 6 Class A2 r p (σf )11 (s1) (σf )22 (s1) (σf )12 (s1) ~di 7→ ~df (s2 ) ( σi, ~di ) FIG. 1. Schematic picture of the channels belonging to class A2, acting on Wigner functions of Gaussian states. The ex- plicit dependence of the final state in terms of the combina- tions s1 and s2 are presented in the appendix. As well the formulas for si depending on the form of the channel. The pic- tured coordinate system corresponds to the position variable r and its conjugate momentum. for GFs, L = Lc(t) + (∂x, ∂r)X(t)⎛⎝∂x∂r ⎞⎠ + (x, r)Y(t)⎛⎝∂x∂r ⎞⎠ + (x, r)Z(t)⎛⎝xr⎞⎠ (34) where Lc(t) is a complex function and X(t) = ⎛⎝Xxx(t) Xxr(t) Xrx(t) Xrr(t) ⎞⎠ (35) is a complex matrix as well as Y(t) and Z(t), whose entries are defined in a similar way as in eq. (35). Note that X(t) and Z(t) can always be chosen symmetric, i.e. Xxr = Xrs and Zxr = Zrx. Thus we must determine 11 time-dependent functions from eq. (34). This ansatz is also appropriate to study the functional forms introduced in this work, given that the left hand side of eq. (33) only involves quadratic polynomials in x, r, ∂/∂x and ∂/∂r, as in the GF case. Notice that singular channels do not admit a master equation since its existence implies that channels with the functional form involved can be found arbitrarily close from the identity channel. This is not possible for sin- gular channels due to the continuity of the determinant of the matrix T. This fact can be also shown using the ansatz of eq. (34), one finds infinitely Liuville operators, thus the master equation is not well defined. For the non-singular cases presented in equations (5) and (6), the condition for the existence of a master equa- Class A1 r p 1 2e1 2a1 + b 2 1 2e1 − b1 2e1 (0,−c1) T ( σi, ~di ) FIG. 2. Schematic picture of the channels belonging to class A1, acting on Wigner functions of Gaussian states. Every channel of this class maps every initial quantum state, in par- ticular GSs characterized by (σi, d⃗i), to a Gaussian state that depends only on the channel parameters. We indicate in the figure the values of the corresponding components of the first and second moments of the final Gaussian state. The pictured coordinate system corresponds to the position variable r and its conjugate momentum. tion is obtained as follows. (i) Substitute the ansatz of eq. (34) in the right hand side of the eq. (33). (ii) Define ρ(x, r; t) using eq. (2), given an initial condi- tion ρ(x, r; 0), for each functional form JI,II. (iii) Take ρf(xf , rf) → ρ(x, r; t) and ρi(xi, ri) → ρ(x, r; 0). Fi- nally, (iv) compare both sides of eq. (33). Defining A(t) = α(t)/β(t) and B(t) = γ(t)/η(t), the conclusion is that for both JI and JII, a master equations exist if c(t)∝ A(t) (36) holds, where c(t) = c1(t) + A(t)c2(t). Additionally, for the form JI the solutions for the matrices X(t), Y(t) 7 and Z(t) are given by Xxx =Xxr = Yrx = Zrr = 0, Yxx = Ȧ A , Lc = Yrr = ė1 e1 − ė2 e2 , Xrr = ė1 4e21 − ė2 2e1e2 , Yxr = ı(λ1ė2 e1e2 + λ2Ȧ e2A − λ1ė1 2e21 − λ̇2 e2 ) , Zxx = λ2 1 2 ⎛⎝ ė2 e1e2 − ė1 2e21 ⎞⎠ + λ1 e2 ⎛⎝λ2Ȧ A − λ̇2 ⎞⎠ + 2λ3 Ȧ A − λ̇3, Zxr = ı⎛⎝ ȦA⎛⎝e1λ2 e2 − λ1 2 ⎞⎠ + λ̇1 2 − λ̇2e1 e2 + λ2 2 ⎛⎝ ė2e2 − ė1 e1 ⎞⎠⎞⎠ , (37) where we have defined the following coefficients: λ1 = b1 +Ab3, λ2 = b2 +Ab4 and λ3 = a1 +Aa2 +A2a3. For the form JII the solutions are the following Lc =Xxx =Xxr =Xrr = Zrr = 0, Yrx = Yxr = 0, Yxx = Ȧ A , Yrr = Ḃ B . Zxx = a2(t)Ȧ(t) + 2a1(t)Ȧ(t) A(t) −A(t)2 −ȧ3(t) −A(t)ȧ2(t) − ȧ1(t), Zxr = ı⎛⎝12 λ̇ − λ 2 ⎛⎝ ȦA + Ḃ B ⎞⎠⎞⎠ , (38) where λ = b1 +Ab3 +B(b2 +Ab4). VI. CONCLUSIONS In this work we have critically reviewed the deceptively natural idea that Gaussian quantum channels always ad- mit a Gaussian functional form. To this end, we went beyond the pioneering characterization of Gaussian chan- nels with Gaussian form presented in Ref. [12] in two new directions. First we have shown that, starting from their most general definition as mapping Gaussian states into Gaussian states, a more general parametrization of the coordinate representation of the one-mode case exists, that admits non-Gaussian functional forms. Second, we were able to provide a black-box characterization of such new forms by imposing complete positivity (not consid- ered in Ref. [12]) and trace preserving conditions. While our parametrization connects with the analysis done by Holevo [16] in the particular cases where besides hav- ing a non-Gaussian form the channel is also singular, it also allows the study of Gaussian unitaries, thus pro- viding similar classification schemes. We completed the classification of the new types of channels by deriving the form of the Liouvillian super operator that gener- ates their time evolution in the form of a master equa- tion. Surprisingly, Gaussian quantum channels without Gaussian form can be experimentally addressed by means of the celebrated Caldeira-Legget model for the quan- tum damped harmonic oscillator, where the new types of channels described here naturally appear in the sub- ohmic regime. ACKNOWLEDGEMENTS We acknowledge PAEP and RedTC for financial sup- port. Support by projects CONACyT 285754, UNAM- PAPIIT IG100518 is acknowledged. CP acknowledges support by PASPA program from DGAPA-UNAM. CAM and JDU acknowledge financial support from the German Academic Exchange Service (DAAD). We are thankful to the University of Vienna where part of this project was done. Appendix A: Explicit formulas for class A2 The explicit formulas of the final states for channels of class A2 with the form presented in eq. (6) with e2 = 0 are (σf)11 = 1 2e1 , (σf)22 = (αβ ) 2 ( b23 2e1 + 2a3) + α β (2a2 + b1b3 e1 )+ 2a1 + b21 2e1 + s1, (σf)12 = −αβ b3 2e1 − b1 2e1 , d⃗f (s3) = (0,−α β c2 − c1 + s2)T , (A1) where s1 = (b22 + 2αβ b2b4 + (α β )2 b24)(σi)11 − 2(α β b2 + (α β )2 b4)(σi)12 + (αβ ) 2 (σi)22 , s2 = (α β b4 + b2)(di)1 − α β (di)2. (A2) 8 The explicit formulas of the final states for channels of class A2 with the form presented in eq. (6) with α = 0 are (σf)11 = e22 4e21 (σi)11 + 1 2e1 , (σf)12 = (b2e22e1 − b1e 2 2 4e21 )(σi)11 − b1 2e1 , (σf)22 = 2a1 + (b2 − b1e2 2e1 )2 (σi)11 + b21 2e1 , (A3) and d⃗f = ( e2 2e1 (d⃗i)1 ,(b2 − b1e2 2e1 )(d⃗i)1 − c1) T . (A4) The explicit formulas of the final states for channels of class A2 with Gaussian form are (σf)11 (s1) = 2a3 b23 + s1, (σf)12 (s1) = a2 b3 − 2a3b1 b23 − b1s1, (σf)22 (s1) = b1 (b3 (b1b3s1 − 2a2) + 2a3b1) b23 + 2a1, d⃗f (s2) = (s2 − c2 b3 , b1 (c2 b3 − s2) − c1)T , (A5) where s1 = b24 b23 (σi)11 − 2b4 b23 (σi)12 + 1 b23 (σi)22 , s2 = 1 b3 (di)2 − b4 b3 (di)1 . (A6) [1] N. J. Cerf, G. Leuchs, and E. S. Polzik, Quan- tum Information with Continuous Variables of Atoms and Light (Imperial College Press, 2007) https://www.worldscientific.com/doi/pdf/10.1142/p489. [2] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Rev. Mod. Phys. 86, 1391 (2014). [3] K. Hammerer, A. S. Sørensen, and E. S. Polzik, Rev. Mod. Phys. 82, 1041 (2010). [4] F. Grosshans, G. Van Assche, J. Wenger, R. Brouri, N. J. Cerf, and P. Grangier, Nature 421, 238 (2003). [5] S. Lloyd and S. L. Braunstein, Phys. Rev. Lett. 82, 1784 (1999). [6] S. L. Braunstein and P. van Loock, Rev. Mod. Phys. 77, 513 (2005). [7] L. Lami, B. Regula, X. Wang, R. Nichols, A. Winter, and G. Adesso, Phys. Rev. A 98, 022335 (2018). [8] E. A. Martinez and J. P. Paz, Phys. Rev. Lett. 110, 130406 (2013). [9] J. Eisert and M. M. Wolf, in Quantum Information with Continuous Variables of Atoms and Light (Imperial Col- lege Press, 2007) pp. 23–42. [10] C. Weedbrook, S. Pirandola, R. Garćıa-Patrón, N. J. Cerf, T. C. Ralph, J. H. Shapiro, and S. Lloyd, Reviews of Modern Physics 84, 621 (2012). [11] A. Holevo and R. Werner, Physical Review A 63, 032312 (2001). [12] E. A. Martinez and J. P. Paz, (2012). [13] H. Grabert, P. Schramm, and G.-L. Ingold, Physics Re- ports 168, 115 (1988). [14] G. Lindblad, Journal of Physics A: Mathematical and General 33, 5059 (2000). [15] A. S. Holevo, Problems of Information Transmission 43, 1 (2007), arXiv:quant-ph/0607051 [quant-ph]. [16] A. S. Holevo, Problems of Information Transmission 44, 171 (2008), arXiv:arXiv:0802.0235v1. [17] R. Karrlein and H. Grabert, Physical Review E 55, 153 (1997), arXiv:9610001 [physics]. C.3. Article: Positivity and Complete positivity of differentiable quantum processes 123 C.3 Article: Positivity and Complete positivity of differ- entiable quantum processes Physics Letters A 383, 23 (2019). Click to go to the webpage, Click to go to arXiv. JID:PLA AID:25754 /SCO Doctopic: Quantum physics [m5G; v1.260; Prn:5/06/2019; 16:24] P.1 (1-10) Physics Letters A ••• (••••) •••–••• Contents lists available at ScienceDirect Physics Letters A www.elsevier.com/locate/pla Positivity and complete positivity of differentiable quantum processes Gustavo Montes Cabrera a, David Davalos b, Thomas Gorin a,∗ a Departamento de Física, Universidad de Guadalajara, Guadalajara, Jalisco, Mexico b Instituto de Física, Universidad Nacional Autónoma de México, Ciudad de México, Mexico a r t i c l e i n f o a b s t r a c t Article history: Received 7 February 2019 Received in revised form 27 May 2019 Accepted 28 May 2019 Available online xxxx Communicated by M.G.A. Paris Keywords: Quantum process Divisibility Quantum non-Markovianity We study quantum processes, as one parameter families of differentiable completely positive and trace preserving (CPTP) maps. Using different representations of the generator, and the Sylvester criterion for positive semi-definite matrices, we obtain conditions for the divisibility of the process into completely positive (CP-divisibility) and positive (P-divisibility) infinitesimal maps. Both concepts are directly related to the definition of quantum non-Markovianity. For the single qubit case we show that CP- and P- divisibility only depend on the dissipation matrix in the master equation form of the generator. We then discuss three classes of processes where the criteria for the different types of divisibility result in simple geometric inequalities, among these the class of non-unital anisotropic Pauli channels.  2019 Elsevier B.V. All rights reserved. 1. Introduction Non-Markovianity of quantum processes has been a topic of increasing interest during approximately the last ten years [1–3]. Starting with the papers by Breuer et al. [4] and Rivas et al. [5], a definition of quantum Markovianity has been reduced to the ques- tion whether all intermediate quantum maps are physically real- izable; this induces a characterization that is more closely related to the Chapman-Kolmogorov condition than to the full definition of classical Markovianity [6]. For differentiable quantum processes, the question of divisibility into physically realizable quantum maps can be further reduced to the analysis of the time dependent gen- erator of the process. This is the approach taken for this work. The concept of divisibility has been introduced in Refs. [7,8]. In its original form, it refers to the condition that all intermediate maps are completely positive (CP-divisibility). However, one may as well consider P-divisibility, where it is sufficient that the inter- mediate maps are positive [9,10]. Note that it has been shown in Ref. [11] that P-divisibility provides a direct connection to classi- cal Markovianity. If an intermediate map is positive but not com- pletely positive, one may observe information backflow for entan- gled states between system and some ancillary system, but not in the system alone [5,12]. In this work, we derive general criteria for the positivity and the complete positivity (of the local infinitessimal intermediate map). In particular, for single qubit processes we show that both, CP-divisibility and P-divisibility conditions, only depend on the dis- * Corresponding author. E-mail address: thomas.gorin@cucei.udg.mx (T. Gorin). sipation matrix of the master equation. We identify three different classes of single qubit processes, where the criteria for CP- and P-divisibility are reduced to simple explicit geometric inequalities. One of these classes consists of processes where the Choi-matrix has the shape of an X (i.e. all non-zero elements are located on the diagonal or the anti-diagonal). Many examples considered in the literature of quantum non-Markovianity are of this type. A second class consists of those processes, where the Choi-matrix has the shape of an O . The third class is that of the non-unital anisotropic Pauli channels. While criteria applicable to the gener- ators have been studied in the context of CP-divisibility, see for instance Ref. [13,29], this has rarely been done for P-divisibility. The complete positivity of non-unital anisotropic Pauli channels, as such, has been considered in Refs. [30] and [31]. Explicit analytical criteria are valuable for the construction of Markovian approximations to a non-Markovian process as pro- posed in Ref. [8] and more specifically in Ref. [5]. Another area of applications is that of quantum process tomography [14–16], where it is important to identify the independent parameters which are to be determined. Finally, it may be of interest to iden- tify quantum channels, which are P-divisible but not CP-divisible as processes where non-Markovianity may be identified as a gen- uine quantum effect [17]. Our work relies on a few general results which have been de- rived previously: (i) The Kossakowski theorem, which establishes the equivalence between positivity and contractivity (for the do- main of Helstrom matrices), see Ref. [12] (and references therein), (ii) necessary and sufficient criteria which can be applied di- rectly to the time dependent generator of the quantum process, see Ref. [9] for positivity and Ref. [8] for complete positivity, (iii) https://doi.org/10.1016/j.physleta.2019.05.049 0375-9601/ 2019 Elsevier B.V. All rights reserved. JID:PLA AID:25754 /SCO Doctopic: Quantum physics [m5G; v1.260; Prn:5/06/2019; 16:24] P.2 (1-10) 2 G. Montes Cabrera et al. / Physics Letters A ••• (••••) •••–••• Sylvester’s criterion for definite and semi-definite positivity [18, 19]. The paper is organized as follows: In Sec. 2 we discuss the de- scription of quantum processes in terms of their generators and the general conditions for CP- and P-divisibility in terms of the generator. In Sec. 3 we analyze these conditions for general sin- gle qubit processes. In Sec. 4 we present classes of single qubit processes, where the conditions for P- and CP-divisibility can be solved analytically. In Sec. 5 we present our conclusions. 2. Differentiable quantum processes In this section we introduce differentiable quantum processes and the definitions of P-divisibility and CP-divisibility. For both properties, we present criteria which can be applied directly to the generator of the quantum process in question. 2.1. Processes and generators Let us denote a quantum process t , (∀t ∈ R+ 0 ), as a one- parameter family of differentiable (with respect to t) completely positive and trace preserving linear maps (CPTP-maps), with 0 = 1, the identity. For simplicity, we assume that the corresponding Hilbert space is of finite dimension, dim(H) = d < ∞. The quan- tum process t can be defined equivalently by the generator Lt , such that d dt t = Lt t , 0 = 1 . (1) One natural question to ask would be the following: What prop- erties must be fulfilled by Lt in order to produce a valid quan- tum process of CPTP maps? Very recently this question has been addressed in Ref. [20]. In the present work, our objective is differ- ent. Assuming that Lt generates a valid quantum process, we ask whether that process is CP-divisible and/or P-divisible. Note that for a given quantum process t , we can compute its generator as Lt = dt dt −1 t . (2) In what follows we will assume that t is invertible. Where this is not the case (if at all, this typically happens at isolated points in time), one has to proceed with care [21]. In order to derive P-divisibility and CP-divisibility criteria in terms of the generator, we need to relate Lt to the intermediate quantum map, t+δ,t = t+δ −1 t , δ > 0 . (3) This can be achieved by considering an infinitesimal intermediate quantum map, such that Lt = lim δ→0 δ−1 ( t+δ,t − 1 ) . (4) Choi-matrix representation. A direct method to represent linear quantum maps (this includes generators such as Lt ) consists in embedding the state space into the vector space Md×d of com- plex quadratic matrices of dimension d. In that case, the ele- ments { |i〉〈 j| }1≤i, j≤d form an orthonormal basis with respect to the Hilbert-Schmidt scalar product 〈A, B〉 = tr(A† B), and their im- ages under the quantum map uniquely define that map. Arranging these images in a d × d block-matrix results in the Choi matrix representation. Formally, the Choi-matrix representation [22] of a linear map  in Md×d can be defined as C = ∑ i, j |i〉〈 j| ⊗ [ |i〉〈 j| ] . (5) It has the following remarkable properties: (i) C = C †  iff [†] = [] for every bounded operator , (ii) C ≥ 0 iff  is complete positive, (iii) tr (C) = d if  preserves the trace [22,23]. Master equation. The generator obtained in Eq. (4) preserves Her- miticity by construction, thus we can bring it to the following standard form [8,24] (see Appendix E for a detailed derivation): d dt ̺ = Lt[̺] , (6) Lt[̺] = −i [H,̺] + d2−1 ∑ i, j=1 D i j ( F i ̺ F † j − 1 2 { F † j F i , ̺ } ) . In this expression, Planck’s constant h̄ has been absorbed into the Hamiltonian H . The matrix D is Hermitian, and the set {F i}1≤i≤d2 forms an orthonormal basis in the space of operators, such that tr(F † i F j) = δi j . In addition, the operators are chosen such that tr(F i) = 0, except for the last element, which is given by Fd2 = 1/ √ d. In this work, we will use Eq. (6) as one possible representation of the generator Lt , at some arbitrary, fixed time t . We call this representation the “master equation representation” of Lt , and D the “dissipation matrix”. Note that an intermediate quantum pro- cess t+δ,t ≈ 1 + δLt , as defined in Eq. (3), is CPTP if and only if D is positive semidefinite [2,13]. In this case, t becomes a valid CPTP map [11]. Finally, if the generator is time-independent, t becomes a one-parameter semigroup in the space of CPTP maps [25–27]. 2.2. Markovianity: P-divisibility vs. CP-divisibility In this subsection we present the definitions for the P-divis- ibility and the CP-divisibility of quantum processes. We use the term “Markovianity” in cases, where we want to refer to both properties, indistinctively. CP-divisibility. A process t is called CP-divisible if and only if the intermediate map t+δ,t as defined in Eq. (3) is CPTP for all t, δ ∈ R + 0 . Complete positivity of a quantum map  is conve- niently verified using the Choi matrix representation, introduced in Eq. (5). P-divisibility. A process t is called P-divisible if and only if the intermediate map t+δ,t as defined in Eq. (3) is PTP (positivity and trace preserving) for all t, δ ∈ R + 0 . Positivity of a Hermiticity and trace preserving quantum map  is more complicated to verify. In that case, one has to show that [̺] ≥ 0 for all density matrices ̺. In practice, it is sufficient to check the condition for all density matrices representing pure states. Local complete positivity. Following Refs. [8], and [28], let CL be the Choi-matrix representation of the generator Lt , and let C⊥ be a matrix representation of CL in the subspace orthogonal to the Bell state |B〉 = 1√ d ∑ i |ii〉 , (7) where d is the dimension of the Hilbert space. Then, the quantum process t is locally CP at time t , if and only if C⊥ ≥ 0 . (8) JID:PLA AID:25754 /SCO Doctopic: Quantum physics [m5G; v1.260; Prn:5/06/2019; 16:24] P.3 (1-10) G. Montes Cabrera et al. / Physics Letters A ••• (••••) •••–••• 3 Therefore the process t is CP-divisible, if and only if it is locally CP for all t ∈ R + 0 . Note that in Ref. [28], it has been shown that C⊥ is unitarily equivalent to the dissipation matrix D (see Appendix E for a detailed derivation). Local positivity. A quantum process is locally positive at time t , if and only if for all orthogonal states |ψ〉, |φ〉 ∈ H it holds that 〈ψ |Lt[ |φ〉〈φ| ]ψ〉 ≥ 0 . (9) Similar to the CP case, it holds that a quantum process t is P- divisible if and only if it is locally positive for all t ∈ R + 0 [9]. The equivalence between local positivity and P-divisibility follows from Eq. (4): 〈ψ |t+δ,t[ |φ〉〈φ| ] |ψ〉 ≥ 0 ⇔ δ 〈ψ |Lt[ |φ〉〈φ| ] |ψ〉 + 〈ψ |φ〉 〈φ|ψ〉 ≥ 0 . (10) In the limit δ → 0, this can only happen if ψ and φ are orthog- onal, 〈ψ |φ〉 = 0. In fact, if |〈ψ |φ〉|2 > 0, it might very well be that 〈ψ | Lt [ |φ〉〈φ| ] |ψ〉 < 0 even if the process is P-divisible in the neighborhood of that point. To summarize, we may express both properties CP-divisibility and P-divisibility in terms of local conditions which have to be fulfilled by the generator Lt . In what follows, we analyze these in more detail. To avoid overly cumbersome terminology, we denote generators which fulfill Eq. (9) and/or Eq. (8) simply as “positive” and/or “completely positive generators”. 3. Single qubit processes In the case of single qubit processes, the Bloch vector represen- tation is yet another method to represent quantum channels and their generators. In Sec. 3.1 we discuss the following three repre- sentations: (i) the master equation, (ii) the Choi-matrix, and (iii) the Bloch vector representation and how they are related one-to- another. In Sec. 3.2, we derive explicit criteria for local positivity and local complete positivity in terms of the dissipation matrix D . 3.1. Equivalent representations Choi matrix representation. For our purposes, the Choi matrix repre- sentation will be the most useful. A CPTP-map , which belongs to a quantum process, may be parametrized as C = ( [ |0〉〈0| ] [ |0〉〈1| ] [ |1〉〈0| ] [ |1〉〈1| ] ) = ⎛ ⎜ ⎜ ⎝ 1− r1 y∗ 1 x∗ 1− z∗1 y1 r1 z2 −x∗ x z∗2 r2 y∗ 2 1− z1 −x y2 1− r2 ⎞ ⎟ ⎟ ⎠ . (11) The structure of C is due to the fact that  must preserve Her- miticity and the trace. We have chosen the parametrization in such a way that the parameters r1, r2 , y1, y2, x, z1, z2 as functions of time are all zero at t = 0. Note that any intermediate map t+δ,t is at least Hermitic- ity and trace preserving. Therefore, Eq. (4) implies that the Choi- matrix representation of the generator Lt must be Hermitian, and in all blocks, the partial trace must be equal to zero. That leaves us with the following parametrization: CL = ⎛ ⎜ ⎜ ⎝ −q1 Y ∗ 1 X∗ −Z∗ 1 Y1 q1 Z2 −X∗ X Z∗ 2 q2 Y ∗ 2 −Z1 −X Y2 −q2 ⎞ ⎟ ⎟ ⎠ . (12) In general, there is no simple relation between the parametrization used here, and that of Eq. (11). This is because the expression for the generator Lt includes the inverse of t . Master equation representation. Note that every generator Lt of a Hermiticity and trace preserving quantum process, can be written in the form of Eq. (6), with Hermitian matrices H and D . There- fore, we may calculate the Choi-representation of the generator, by inserting ̺ = |i〉〈 j| into the RHS of Eq. (6), and compare the result to the general form in Eq. (12). For the calculation, we choose the following orthonormal operator basis {F i}1≤i≤d2 : F1 = 1√ 2 ( |0〉〈0| − |1〉〈1| ) , F2 = |0〉〈1| , F3 = |1〉〈0| , and F4 = 1/ √ 2 . (13) As a result, we obtain a linear one-to-one correspondence between the parameters used in the master equation representation and those, used in the Choi representation: ⎛ ⎝ q1 q2 ReZ1 ⎞ ⎠ = ⎛ ⎝ 0 0 1 0 1 0 1 1/2 1/2 ⎞ ⎠ ⎛ ⎝ D11 D22 D33 ⎞ ⎠ , Im Z1 = H22 − H11 , Z2 = D32 , ⎛ ⎝ Y1 Y2 X∗ ⎞ ⎠ = ⎛ ⎝ − √ 2/4 √ 18/4 −i − √ 18/4 √ 2/4 i√ 2/4 √ 2/4 i ⎞ ⎠ ⎛ ⎝ D12 D31 H21 ⎞ ⎠ . (14) As one might have expected, the quantity H11 + H22 is irrelevant for the representation of the generator, and may be set equal to zero without loss of generality. Then, Eq. (14) is clearly an invert- ible linear system of equations. Bloch vector representation. Any qubit density matrix can be written in terms of the Pauli matrices and the identity matrix 1 as follows: ̺ = 1 2 ( v0 1+ 3 ∑ j=1 v j σ j ) , (15) where v0 = 1 and v = (v1, v2, v3) is a vector in R3 of norm ‖v‖ ≤ 1. Any Hermiticity and trace preserving quantum map  can then be written as an affine transformation [32]  : v → v ′ = R v +t , (16) where R is a real not necessarily symmetric square matrix and t is a real three-dimensional vector. The coefficients of R and t are given by t j = 1 2 tr ( σ j Lt[1 ] ) , R jk = 1 2 tr ( σ j Lt[σk ] ) . (17) For the generator with the Choi-matrix representation given in Eq. (12), we find R = ⎛ ⎝ Re(Z2 − Z1) Im(Z1 + Z2) Re(Y1 − Y2) Im(Z2 − Z1) −Re(Z1 + Z2) Im(Y1 − Y2) 2Re(X) −2 Im(X) −q1 − q2 ⎞ ⎠ , t = ⎛ ⎝ Re(Y1 + Y2) Im(Y1 + Y2) q2 − q1 ⎞ ⎠ . (18) Again, it is easy to verify that the relation between this Bloch vec- tor representation and the Choi representation is invertible. JID:PLA AID:25754 /SCO Doctopic: Quantum physics [m5G; v1.260; Prn:5/06/2019; 16:24] P.4 (1-10) 4 G. Montes Cabrera et al. / Physics Letters A ••• (••••) •••–••• 3.2. Criteria for positivity and complete positivity Local complete positivity. In order to verify if the Choi-matrix (as a linear transformation) projected onto the orthogonal subspace of |φB〉〈φB|, is positive, we choose the orthonormal states |ψ1〉 = 1√ 2 ⎡ ⎢ ⎢ ⎣ 1 0 0 −1 ⎤ ⎥ ⎥ ⎦ , |ψ2〉 = ⎡ ⎢ ⎢ ⎣ 0 0 1 0 ⎤ ⎥ ⎥ ⎦ , |ψ3〉 = ⎡ ⎢ ⎢ ⎣ 0 1 0 0 ⎤ ⎥ ⎥ ⎦ , (19) to span that subspace. Then we obtain for the matrix representa- tion of the Choi matrix of Lt , projected on that subspace: C⊥ = ⎛ ⎜ ⎜ ⎝ Re(Z1) − q1+q2 2 X∗−Y2√ 2 X+Y ∗ 1√ 2 X−Y ∗ 2√ 2 q2 Z∗ 2 X∗+Y1√ 2 Z2 q1 ⎞ ⎟ ⎟ ⎠ = ⎛ ⎝ D11 D12 D13 D21 D22 D23 D31 D32 D33 ⎞ ⎠ . (20) The second equality is obtained by solving Eq. (14) for the matrix elements D i j . It simply means that C⊥ = D . We may now use the Sylvester criterion to check whether D ≥ 0 or not. A general discussion of that criterion can be found in the text book [18]; the present positive semidefinite case has been treated in Ref. [19]. In that case, the statement is the following: A Hermitian matrix is positive semidefinite if and only if all principal minors are larger or equal to zero. Hence, for D ≥ 0, it must hold: D11, D22, D33 ≥ 0 , D11D33 − |D31|2 ≥ 0 , D11D22 − |D21|2 ≥ 0 , D33D22 − |D32|2 ≥ 0 , D11D22D33 + 2Re(D12D23D31) ≥ D11|D32|2 + D22|D31|2 + D33|D21|2 . (21) Local positivity. According to the criterion in Eq. (9), we need to verify that 〈ψ | L[ |φ〉〈φ| ] |ψ〉 ≥ 0 for all |ψ〉 ⊥ |φ〉. Such general orthonormal states may be written as the column vectors of a uni- tary matrix, taken from the group SU (2). Removing an ineffective global phase we find: |ψ〉 = ( cos(θ/2) eiβ sin(θ/2) ) , |φ〉 = ( − sin(θ/2) eiβ cos(θ/2) ) . Hence, we consider p(θ, β) = 〈ψ | L[ |φ〉〈φ| ] |ψ〉 as a function of θ and β . Therefore, we may say that the Lt is positive at time t , if and only if p(θ, β) ≥ 0 for all θ and β . Using the parametrization of Eq. (12), p(θ, β) may be written as p(θ,β) = q1 + q2 2 cos2 θ + q2 − q1 2 cos θ + A 2 sin2 θ + Re [ (Y1 + Y2)e−iβ ] 2 sin θ (22) + Re [ (Y2 − Y1)e−iβ − 2X eiβ ] 2 sin θ cos θ , where A = Re[Z1 − Z2 e−2iβ ]. In terms of the master equation pa- rameters, we find R = D22 + D33 , Y1 + Y2 = √ 2 (D21 − D13) , S = D33 − D22 , A1 = D11 − D33 + D22 2 − Re D23 , Y2 − Y1 − 2 X∗ = − √ 2 (D21 + D13) , (23) such that 2 p(θ,β) = R + S cos θ + ( D11 − R 2 ) sin2 θ + Re [ − D23 e −2iβ sin θ + √ 2 (D21 − D13)e −iβ − √ 2 (D21 + D13)e −iβ cos θ ] sin θ . (24) This shows that positivity, just as complete positivity, only depends on the dissipation matrix D . In general, one should try to find all minima of this function and verify that those are non-negative. Since the domain of p(θ, β) is a torus without boundaries, it is sufficient to find the critical points where the partial derivatives ∂p/∂θ and ∂p/∂β are both equal to zero. The corresponding equations may be reduced to a root-finding problem for 4’th order polynomials. Thus analytical expressions may be obtained in principle, even so they are proba- bly not very useful. Still, numerical evaluations are pretty straight forward to implement. In Sec. 4 we will discuss different classes of generators, where particularly simple analytical solutions can be found. 4. Examples In this section, we consider three different classes of generators. For each class, the set of positive (completely positive) generators is interpreted as a region in a certain parameter space (a subspace of the 9-dimensional vector space of dissipation matrices). In gen- eral, these regions must be convex, since the respective criteria involve expectation values of some linear matrix which represents the generator. Hence, if we consider the expectation value of any convex combination of two generators, it immediately decomposes into the corresponding convex combination of expectation values. Unless stated otherwise, we analyze the criteria for positivity and complete positivity in terms of the dissipation matrix D . 4.1. X-shaped quantum channels and generators The term “X-shape” refers to the case, where the non-zero ele- ments in the Choi matrix appear to form the letter “X”, that means that Y1 = Y2 = X = 0 in Eq. (12). Hence, CL = ⎛ ⎜ ⎜ ⎝ −q1 0 0 −Z∗ 1 0 q1 Z2 0 0 Z∗ 2 q2 0 −Z1 0 0 −q2 ⎞ ⎟ ⎟ ⎠ . (25) In this case, the X-shape of the generator implies the X-shape of the quantum channel, and vice versa. Many important models lead to quantum channels of that type [2,4]. In terms of the Bloch vec- tor representation, the X-shape implies that the dynamics along the z-axis is independent from that in the (x, y)-plane [13]. According to Eq. (14) the X-shape of the Choi matrix CL im- plies for H and D from the master equation representation in Eq. (6): H12 = 0, D13 = D12 = 0 as well as q1 = D22 , q2 = D33 , Z2 = D23 and Z1 = i (H22 − H11) + D11 + D33 + D22 2 . (26) For the matrix C⊥ we thus obtain: C⊥ = ⎛ ⎝ D11 0 0 0 D22 D23 0 D32 D33 ⎞ ⎠ . (27) Complete positivity. Considering all principal minors of the dissipa- tion matrix in Eq. (27), we find JID:PLA AID:25754 /SCO Doctopic: Quantum physics [m5G; v1.260; Prn:5/06/2019; 16:24] P.5 (1-10) G. Montes Cabrera et al. / Physics Letters A ••• (••••) •••–••• 5 Fig. 1. The parameter space D22, D33 ≥ 0 for visualizing the regions of positivity and complete positivity for the X-shaped generator, for |D32| = 1. For complete positivity, the point (D22, D33) must lie above the black dashed line, while D11 ≥ 0 is required. For positivity, the allowed region for (D22, D33) depends on D11: For D11 ≥ 1, it is the whole quadrant; for D11 = 1/2 the allowed region consists of the dark green and blue areas; for D11 = 0 it consists of the blue areas above the black dashed line; and for D11 = −1/2 it consists of the dark blue area alone. (For interpretation of the colors in the figure(s), the reader is referred to the web version of this article.) D11, D22, D33 ≥ 0 , D22 D33 − |D23|2 ≥ 0 , D11 D22 ≥ 0 , D11 [ D22 D33 − |D23|2 ] ≥ 0 . (28) Removing redundant inequalities, we are left with D11, D22, D33 ≥ 0 , D22 D33 ≥ |D23|2 . (29) Positivity. From Eq. (24) we find: 2p(θ,β) = R + S cos θ + A sin2 θ ≥ 0 , (30) where A = D11 − R 2 − Re [ D23 e −2iβ ] , R = D22+D33 , and S = D33−D22 . This inequality must hold for all values of θ and β , parametrizing the quantum state to which the generator is applied. Thus, we only need to verify if the minimum of this expression is larger than zero. As far as β is concerned, this means that we may replace A by its minimum (as a function of β), which is given by Amin = D11 − R/2 − |D23|. We are then left with the condition ∀ θ : R + S cos θ + Amin sin2 θ ≥ 0 . (31) This condition is further evaluated in Appendix A. As a result, we find that the conditions for positivity become D22, D33 ≥ 0 , (32) and if D11 < |D23|, in addition ∣ ∣ |D23| − D11 ∣ ∣ ≤ √ D22 D33 . (33) In Fig. 1, we show the parameter space D22, D33 ≥ 0 for vi- sualizing the regions of positivity and complete positivity for the X-shaped generator. For complete positivity, the inequalities to ful- fill are given in Eq. (29), which states independent conditions on D11 on the one hand and D22, D33, |D23|2 on the other. For posi- tivity, by contrast, the conditions on D22 and D33 depend on D11 . Here, we observe an interesting behavior: As D11 approaches zero from above, the region of positivity becomes more and more sim- ilar to the region of complete positivity, until they coincide for D11 = 0. When D11 becomes negative, complete positivity is vio- lated while positivity is still maintained sufficiently far away from the black dashed line. 4.2. O-shaped quantum channels Here, we consider another subset of single qubit generators, which also allow for an analytic solution. These are in some sense complementary to the X-shaped channels, These are obtained from the general case by setting q1 = q2 = q, Y1 = −Y2 = Y , and Z2 = 0. The Choi matrix, representing the generator resembles an O , in- stead of an X , that is why we call them O -shaped channels. CL = ⎡ ⎢ ⎢ ⎣ −q Y ∗ X∗ −Z∗ 1 Y q 0 −X∗ X 0 q −Y ∗ −Z1 −X −Y −q ⎤ ⎥ ⎥ ⎦ (34) According to Eq. (14), this implies for the matrix elements of H and D from the master equation (6): q = D33 , Z1 = i (H22 − H11) + D11 + D33 , Y = −i H21 + D13√ 2 , X∗ = i H21 + D13√ 2 , (35) with D22 = D33 , D21 = D13 , and D23 = 0. The matrix for verifying complete positivity reads C⊥ = ⎛ ⎝ D11 D31 D13 D13 D22 0 D31 0 D22 ⎞ ⎠ , (36) Complete positivity. Expressed in terms of the dissipation matrix, considering all principal minors. D11, D22 ≥ 0 , D11D22 − |D13|2 ≥ 0 , D11 D 2 22 − D13D31D22 − D31D22D13 ≥ 0 (37) This can be reduced to D11, D22 ≥ 0 , D11D22 ≥ 2 |D13|2 . (38) Positivity. Under the conditions mentioned above, the function 2 p(θ, β) from Eq. (24) becomes (in terms of the master equation parameters) 2p(θ,β) = 2 D22 + (D11 − D22) sin2 θ − 2 √ 2Re [ D13 e −iβ] sin θ cos θ ≥ 0 . (39) Again, it is possible to derive the conditions for positivity, which do no longer involve the angles θ and β . The respective calculation is outlined in Appendix B, with the result [see Eq. (B.4)] 3 D22 + D11 ≥ 0 , D22 (D11 + D22) ≥ |D13|2 . (40) In Fig. 2, we show the different regions of positivity and com- plete positivity in the parameter space of D22, D11 . We distinguish two qualitatively different cases, |D13| = 0 and |D13| = 1. In both cases, the region of positivity is considerably larger than the region for complete positivity. JID:PLA AID:25754 /SCO Doctopic: Quantum physics [m5G; v1.260; Prn:5/06/2019; 16:24] P.6 (1-10) 6 G. Montes Cabrera et al. / Physics Letters A ••• (••••) •••–••• Fig. 2. The parameter space D22, D11 for visualizing the regions of positivity [Eq. (40)] and complete positivity [Eq. (38)] for the O -shaped generator. For |D13| = 0, the region of complete positivity is simply the positive quadrant D22, D11 ≥ 0, while the region for positivity is the whole colored region. For |D13| = 1, the region of complete positivity is colored in orange, the region of positivity is dark green and orange. 4.3. Non-unital anisotropic Pauli channels Here, t is given as an affine transformation of state vec- tors in the Bloch sphere [32] (see the corresponding paragraph in Sec. 3.1): t : v → v ′ = R v + s , (41) where R is a real diagonal matrix and s a real vector. R = ⎛ ⎝ R11 0 0 0 R22 0 0 0 R33 ⎞ ⎠ , s = ⎛ ⎝ s1 s2 s3 ⎞ ⎠ . (42) Using the general formula, Eq. (2), for constructing the generator, we find LP : v → v ′ = dR dt [ R−1 (v − s) ] + ds dt . (43) Hence, the generator for this Pauli channel is given by the affine transformation v → v ′ = RLP v +tLP , with RLP = ⎛ ⎝ −γ1 0 0 0 −γ2 0 0 0 −γ3 ⎞ ⎠ , tLP = ⎛ ⎝ τ1 τ2 τ3 ⎞ ⎠ , where γ j = −1 R j j dR j j dt , τ j = d s j dt + γ j s j . (44) The Choi matrix representation of LP is obtained by inverting Eq. (18), with the result CLP = 1 2 ⎛ ⎜ ⎜ ⎝ −γ3 + τ3 τ1 − iτ2 0 −γ1 − γ2 τ1 + iτ2 γ3 − τ3 γ2 − γ1 0 0 γ2 − γ1 γ3 + τ3 τ1 − iτ2 −γ1 − γ2 0 τ1 + iτ2 −γ3 − τ3 ⎞ ⎟ ⎟ ⎠ . (45) In what follows, we compute the positivity and the complete pos- itivity condition in terms of the parameters γ j and τ j , since this allows for relatively simple geometric interpretations. In the case of complete positivity, this has been worked out previously, in Ref. [29]. For the parametrization in terms of the master equa- tion (6), we obtain from Eq. (12) and (14): Fig. 3. Parameter space of γ1 , γ2 and γ3 . For the generator LP in Eq. (43) to be positive, all elements γ j must be positive (blue transparent color). For LP to be completely positive, the elements γ j must fulfill the conditions in Eq. (48). The corresponding region is colored in orange. Note that we show a cut through the regions of positivity and complete positivity which really extend towards arbitrary large positive values. D22 = γ3 − τ3 2 , D33 = γ3 + τ3 2 , H22 = H11 , D11 = γ1 + γ2 − γ3 2 , D23 = γ2 − γ1 2 , H12 = 0 , D21 = τ1 + iτ2 2 √ 2 = −D13 . (46) This yields C⊥ = 1 2 ⎛ ⎝ γ1 + γ2 − γ3 w∗ −w w γ3 − τ3 γ2 − γ1 −w∗ γ2 − γ1 γ3 + τ3 ⎞ ⎠ , (47) with w = (τ1 + iτ2)/ √ 2. Complete positivity. The complete derivation can be found in Ap- pendix C. It yields separate conditions for the diagonal elements γ j and the vector τ . For the diagonal elements γ j we find: ∀ i = j = k = i : |γi − γ j| ≤ γk ≤ γi + γ j . (48) The corresponding region in the parameter space of the elements γ j is depicted as a orange region in Fig. 3. Assuming these con- ditions are fulfilled, the vector τ must lie inside the following ellipsoid: τ 2 1 a21 + τ 2 2 a22 + τ 2 3 a23 ≤ 1 , a1 = γ 2 1 − (γ2 − γ3) 2 , a2 = γ 2 2 − (γ1 − γ3) 2 , a3 = γ 2 3 − (γ1 − γ2) 2 . (49) The regions of τ where the generator LP fulfills the conditions of complete positivity are shown in Fig. 4 in orange. Note that in this figure, we consider two particular cases, where γ1 = γ2 such that the resulting ellipsoid as defined above is symmetric with respect to the τ3 axis. Positivity. In the general expression for p(θ, β) in Eq. (24), we re- place the parameters with those from the Pauli channel, given in Eq. (46). This yields 2 p(θ,β) = γ3 cos2 θ + [ γ1 cos2 β + γ2 sin2 β ] sin2 θ + τ3 cos θ + ( τ1 cosβ + τ2 sinβ ) sin θ . (50) We can express the general inequality 2 p(θ, β) ≥ 0 in a geometric form: JID:PLA AID:25754 /SCO Doctopic: Quantum physics [m5G; v1.260; Prn:5/06/2019; 16:24] P.7 (1-10) G. Montes Cabrera et al. / Physics Letters A ••• (••••) •••–••• 7 Fig. 4. Comparison of the region of positivity and complete positivity in the param- eter space of τ for γ1 = γ2 . The black solid line shows the ellipsoid τ = γ er , the orange region shows the region of complete positivity, the green region (including orange) the region of positivity. In panel (a), γ1 = 0.255, γ3 = 0.49 which amounts to an ellipsoid of the shape of a rugby ball, in panel (b) γ1 = 0.495, γ3 = 0.01 where the ellipsoid looks more like a pancake. er = ⎛ ⎝ sin θ cosβ sin θ sinβ cos θ ⎞ ⎠ : er · ( γ er + τ ) ≥ 0 , (51) where γ is the diagonal matrix with elements γ j . The interpretation of this result is easy: −γ er + τ is the image of er under the generator LP . Thus an infinitesimal intermediate map would yield t,t+δ : er → er ′ = er + δ LP[er] . In order to have ‖er ′‖ ≤ 1, the image under the generator must be pointing towards the center of the Bloch sphere, i.e. the scalar product between −γ er + τ and er must be negative. Multiplying the resulting inequality by minus one, we find ∀er : er · ( γ er − τ ) ≥ 0 . This relation is equivalent to the inequality in Eq. (51), as can be seen by replacing er by −er . As shown in Appendix D, the set of τ for which the Pauli gen- erator LP is positive, i.e. the inequality in Eq. (51) holds, is the convex region, which contains the origin and is limited by the sur- face [see Eq. (D.2)] T = {τ (θ,β) = (er · γ er) er − 2γ er } . (52) In Fig. 4, we show the region in τ -space which corresponds to positivity and complete positivity of the Pauli channel generator LP . We consider two cases: γ1 = γ2 = 0.255, γ3 = 0.49 in panel (a), and γ1 = γ2 = 0.495, γ3 = 0.01 in panel (b). In the yellow triangle shown in Fig. 3, these points are located near the up- per horizontal line (a) and near the lower corner (b), respectively. Choosing γ1 = γ2 leads to regions of (complete) positivity, which are symmetric with respect to the τ3-axis, which allows us to show two-dimensional projections. We find that the regions of positivity and complete positivity are always contained in ellipsoid with the parametrization τ (θ, β) = γ er . As required, the region of complete positivity (orange) is fully contained in the region of positivity (olive green). In panel (a), we show a case where the ellipsoid γ er resemble roughly a rugby ball. In that case, the are only rather thin stripes near the border of the ellipsoid, where the generator is not positive any more. In panel (b), the ellipsoid has the shape of a flat pancake, and the region of positivity in the center is much smaller. 5. Conclusions In order to determine whether a given differentiable quantum process is CP-divisible and/or P-divisible, we derive criteria to be applied to the generator of the process. For the single qubit case, we discuss three common representations of the generator and work out the one-to-one mappings between them. We find criteria for CP- and P-divisibility, which can be expressed as inequalities in terms of the elements of the dissipation matrix. In the CP case, we avoid solving an eigenvalue problem by using the principal mi- nor test for semidefinite matrices. In the P case, the corresponding inequality must be fulfilled for a whole two-parameter family of functions, which leads to an optimization problem without explicit general solution. We then discuss three different classes of generators, where our criteria do yield explicit results: the familiar X-shaped channels where the elements of the Choi matrix are non-zero in the diago- nal and the anti-diagonal, only; the so called O -shaped channels, where C23 = 0, C11 = C44 and C12 = −C34; and most importantly the non-unital Pauli channels. Besides its general value, as for instance the positivity criteria for the Pauli channel, we expect our results to prove useful in the area of quantum process tomography and the construction of op- timal P-divisible or CP-divisible approximations to non-Markovian quantum processes. In particular there, the renouncement on the calculation of higher order roots may help to find analytical or semi-analytical solutions. Acknowledgements We gratefully acknowledge Sergey Filippov for useful discus- sions, as well as Carlos Pineda for valuable comments on the manuscript. Appendix A. Positivity of X-shaped generators The condition in Eq. (31) can be expressed equivalently in terms of the variable x = cos θ as follows: ∀ x ∈ [−1,1] : f (x) = (R − Amin) x 2 + S x+ Amin ≥ 0 . (A.1) First note the following obviously necessary conditions f (0) : Amin ≥ 0 and f (±1) : R ± S ≥ 0 ⇔ 0 ≤ |S| ≤ R . (A.2) To find the necessary and sufficient conditions, we will divide the problem in two cases: (i)  = R − Amin ≤ 0 and (ii)  > 0. In case (i) the conditions in Eq. (A.2) are also sufficient as can be seen as follows: f (x) is convex, such that for any x1, x2 and 0 < λ < 1: f (λ x1 + (1 − λ) x2) ≥ λ f (x1) + (1− λ) f (x2) . Choosing x1 = −1 and x2 = 1, we find f (1 − 2λ) ≥ λ f (−1) + (1− λ) f (1) , which implies that f (x) ≥ 0 in the interval (−1, 1). In case (ii)  > 0, the conditions in Eq. (A.2) are not sufficient. In this case, positivity requires that either f (x) has no zeros, or its zeros x1,2 = − S 2 ± √ S2 2 − 4 Amin  , are lying both to the left or both to the right of the interval (−1.1). This can be expressed as |S| ≤ 2 √ Amin  or |S| ≥ 2 + √ S2 − 4Amin  . JID:PLA AID:25754 /SCO Doctopic: Quantum physics [m5G; v1.260; Prn:5/06/2019; 16:24] P.8 (1-10) 8 G. Montes Cabrera et al. / Physics Letters A ••• (••••) •••–••• The inequality to the right is equivalent to |S| ≥ 2 and ( |S| − 2 )2 ≥ S2 − 4Amin  , which is equivalent to |S| ≥ 2 and |S| ≤ R , where |S| ≤ R had already been identified as a necessary condi- tion, previously. Therefore, in case (ii) the necessary and sufficient conditions for positivity are Amin ≥ 0 , 0 ≤ |S| ≤ R and |S| ≤ 2 √ Amin  or |S| ≥ 2 . It turns out that for  ≤ Amin it holds that 2 √ Amin  < 2 such that the two conditions cancel each other, i.e. one of the two conditions is always fulfilled. For  > Amin which is equivalent to 2 > R , by contrast, implies that |S| ≥ 2 cannot hold, such that |S| ≤ 2 √ Amin  must be fulfilled. To summarize, the neces- sary and sufficient conditions for positivity are as follows: • Amin = Re Z1 − |Z2| ≥ 0 , 0 ≤ |S| ≤ R . • In addition, if R > 2Amin: |S| ≤ 2 √ Amin (R − Amin) . For the parametrization in terms of the master equation, we find that positivity only depends on the dissipation matrix D . Since R = D33 + D22 , S = D22 − D33 , (A.3) the condition 0 ≤ |S| ≤ R implies that both D22 and D33 must be larger than or equal to zero. Furthermore, with Amin = D11 + D33 + D22 2 − |D32| , (A.4) the condition R > 2 Amin implies that |D32| > D11 . Finally, Amin(R − Amin) = [D33 + D22 2 − ( |D32| − D11 ) ] × [ D33 + D22 2 + ( |D32| − D11 ) ] = (D33 + D22) 2 4 − ( |D32| − D11 )2 , (A.5) such that |S| ≤ 2 √ Amin (R − Amin) is equivalent to |D22 − D33|2 ≤ (D33 + D22) 2 − 4 ( |D32| − D11 )2 ⇔ ( |D32| − D11 )2 ≤ D33 D22 . (A.6) To summarize, in this parametrization, the conditions for positivity read D22, D33 ≥ 0 , D11 − |D32| + D33 + D22 2 ≥ 0 (A.7) and if D11 < |D32|, in addition ∣ ∣ |D32| − D11 ∣ ∣ ≤ √ D33 D22 . (A.8) Note that this last inequality implies the second inequality of Eq. (A.7), which can therefore be ignored. Appendix B. Positivity of O -shaped generators We start from the condition for positivity in Eq. (39). Us- ing the trigonometric identities 2 sin2 θ = 1 − cos2θ and sin2θ = 2 sin θ cos θ , Eq. (39) becomes 2p(θ,β) = 3D22 + D11 2 + D22 − D11 2 cos 2θ − √ 2Re ( D12 e −iβ ) sin 2θ ≥ 0 . (B.1) This expression is minimized with respect to β , simply by making sure that Re( D12 e−iβ ) = ± |D12|. In other words: 2p(θ, β) ≥ 0 for all β and θ is equivalent to 3D22 + D11 2 + D22 − D11 2 cos 2θ ± √ 2 |D12| sin 2θ ≥ 0 . (B.2) This condition is equivalent to 3D22 + D11 2 ≥ 0 and (3D22 + D11) 2 4 ≥ (D22 − D11) 2 4 + 2|D12|2 (B.3) These two inequalities are equivalent to 3D22 + D11 ≥ 0 and D22 (D22 + D11) ≥ |D12|2 (B.4) Appendix C. Complete positivity of the non-unital anisotropic Pauli channel For the generator LP to be completely positive, the matrix C⊥ given in Eq. (47) must fulfill the inequalities in Eq. (21). In the present case, this yields three sets of inequalities: γ1 + γ2 − γ3 ≥ 0 , γ3 − τ3 ≥ 0 , γ3 + τ3 ≥ 0 , (γ3 − τ3)(γ3 + τ3) − (γ2 − γ1) 2 ≥ 0 , (C.1) (γ1 + γ2 − γ3)(γ3 − τ3) − |w|2 ≥ 0 , (γ1 + γ2 − γ3)(γ3 + τ3) − |w|2 ≥ 0 , (C.2) and (γ1 + γ2 − γ3) [ (γ3 − τ3)(γ3 + τ3) − (γ2 − γ1) 2] − w [ w∗ (γ3 + τ3) + w (γ2 − γ1) ] − w∗ [ w∗ (γ2 − γ1) + w (γ3 − τ3) ] ≥ 0 , (C.3) where w = (τ1 + i τ2)/ √ 2. From Eq. (C.1), we find γ3 ≥ |τ3| ≥ 0 , γ1 + γ2 ≥ γ3 , (γ2 − γ1) 2 ≤ γ 2 3 − τ 2 3 , which yields the following conditions as necessary conditions (since we set τ3 = 0 to arrive there): γ1,γ2,γ3 ≥ 0 , |γ2 − γ1| ≤ γ3 ≤ γ1 + γ2 . (C.4) It is easy to verify that these inequalities are invariant under any permutation of indices; see Fig. 3. The remaining conditions, may be interpreted as conditions for the vector τ . These consist of the inequalities in Eq. (C.2) together with |τ3| ≤ √ γ 2 3 − (γ2 − γ1)2 , and (C.5) (γ1 + γ2 − γ3) (γ2 + γ3 − γ1) (γ3 + γ1 − γ2) ≥ (γ1 + γ2 − γ3) τ 2 3 + (γ2 + γ3 − γ1) τ 2 1 + (γ3 + γ1 − γ2) τ 2 2 . (C.6) In Appendix C.1 we demonstrate that condition (C.6) implies all other conditions for the vector τ , which can therefore be omitted. Reorganizing the terms in Eq. (C.6), we arrive at τ 2 1 a21 + τ 2 2 a22 + τ 2 3 a23 ≤ 1 , a1 = γ 2 1 − (γ2 − γ3) 2 , a2 = γ 2 2 − (γ1 − γ3) 2 , a3 = γ 2 3 − (γ1 − γ2) 2 . (C.7) JID:PLA AID:25754 /SCO Doctopic: Quantum physics [m5G; v1.260; Prn:5/06/2019; 16:24] P.9 (1-10) G. Montes Cabrera et al. / Physics Letters A ••• (••••) •••–••• 9 Appendix C.1. Omissible inequalities for τ In what follows, we demonstrate that Eq. (C.5) as well as Eq. (C.2) follow from Eq. (C.7) such that we may consider Eq. (C.7) as the only condition on τ . To that end note first that setting τ1 = τ2 = 0 we can make the LHS of Eq. (C.7) only smaller which hence implies τ 2 3 ≤ a23 = γ 2 3 − (γ1 − γ2) 2 , which is exactly Eq. (C.5). To show that Eq. (C.7) also implies Eq. (C.2), it is convenient to express τ in elliptical coordinates, τ = λ ⎛ ⎝ a1 sin θ cosϕ a2 sin θ sinϕ a3 cos θ ⎞ ⎠ , such that Eq. (C.7) allows arbitrary values for the angles θ, ϕ and limits λ to the range 0 ≤ λ ≤ 1. The two inequalities in Eq. (C.2) may be combined, and then read γ3 ± λa3 cos θ ≥ λ2 sin2 θ (a21 cos2 ϕ + a22 sin2 ϕ) 2 (γ1 + γ2 − γ3) . Since a21 and a22 have the common factor (γ1 + γ2 − γ3) this in- equality simplifies to (γ3 ± λa3 cos θ) ≥ λ2 sin2 θ 2 [ (γ3 + γ1 − γ2) cos2 ϕ + (γ3 − γ1 + γ2) sin2 ϕ ] = λ2 sin2 θ 2 [ γ3 + (γ1 − γ2) cos(2ϕ) ] (C.8) Due to the conditions in Eq. (C.7), we may assume that γ3 ≥ a3 and γ3 ≥ |γ1 −γ2|. Therefore, in order to show that Eq. (C.2) holds, it is sufficient to prove that γ3 ± λa3 cos θ ≥ λ2 sin2 θ 2 [ γ3 + |γ1 − γ2| ] . For that purpose, we substitute x = cos θ to obtain a quadratic ex- pression: A x2 ± λa3 x+ γ3 − A ≥ 0 , A = λ2 2 [ γ3 + |γ1 − γ2| ] . The LHS describes a parabola. Therefore, the inequality holds, if we can prove that the equation A x2 ± λ a3 x + γ3 − A = 0 has no solution or at most one solution. For that purpose we consider the discriminant and show that it is less or equal to zero. For later convenience, we define g± = λ3 ± |γ1 − γ2|. Then we may write: λ2 a23 − 4A (γ3 − A) ≤ 0 ⇐ a23 − g+ (2γ3 − λ2 g+) ≤ 0 , A = λ2 g+ 2 ⇐ g+ g− − 2g+ γ3 + λ2g2+ ≤ 0 , a23 = g+ g− ⇐ g− − 2γ3 + λ2g+ ≤ 0 ⇐ −g+(1 − λ2) ≤ 0 , g− − 2λ3 = −g+ . This completes the proof. The discriminant is negative semidefi- nite. Therefore the two inequalities in Eq. (C.2) are always fulfilled and can be omitted. Appendix D. Positivity of the non-unital anisotropic Pauli channel We start from the condition, given in Eq. (51), er · (γ er + τ ) ≥ 0 , where er is a unit vector in spherical coordinates, parametrized by the angles θ, β . We aim at constructing the surface T which forms the outer boundary of the region of points τ , where the above inequality holds (note that this region contains the origin τ = o, and that it must be convex1). The condition for τ ∈ T can be cast into the following set of equations: er · (γ er + τ ) = 0 ∂ ∂θ er · (γ er + τ ) = 0 (D.1) ∂ ∂β er · (γ er + τ ) = 0 . The argument is as follows: Consider the LHS of the first equation as a function f (τ , θ, β), then we may compute fmax(τ ) = max θ,β f (τ , θ,β) , by finding the critical points (there may be more than one) (θi, βi), where the last two equalities of Eq. (D.1) hold. Typically, for some fixed but arbitrary point τ , some of the values of { f (τ , θi, βi) } may be positive and others negative; some may correspond to local maxima, others to local minima, and still others may correspond neither to one nor to the other group. However, the global maxi- mum will always be among these points. The calculation of the partial derivatives is simplified by the fact that ∂ er ∂θ = ⎛ ⎝ cos θ cosβ cos θ sinβ − sin θ ⎞ ⎠ = eθ , ∂ er ∂β = ⎛ ⎝ − sin θ sinβ sin θ cosβ 0 ⎞ ⎠ = sin θ eβ , such that {er, eθ , eβ } form a system of orthonormal vectors. There- fore the system of equations in Eq. (D.1) becomes er · (γ er + τ ) = 0 eθ · (γ er + τ ) + er · γ eθ = 0 eβ · (γ er + τ ) + er · γ eβ = 0 , which is equivalent to er · (γ er + τ ) = 0 eθ · (2γ er + τ ) = 0 sin θ eβ · (2γ er + τ ) = 0 . We started by asking for which points τ , there exist a critical point (θi, βi) corresponding to a global maximum such that this set of equations is fulfilled. That point would then fore sure belong to the desired surface T . However, starting from this relation, we may say that it assigns to any pair of angles (θ, β), a unique τ , such that that pair of angles is a critical point (of any nature), while 1 For fixed R , two different quantum generators L1, L2 are given by τ1 and τ2 , and any intermediate generator λ L1 + (1 − λ) L2 is given by λ τ1 + (1 − λ) τ2 . JID:PLA AID:25754 /SCO Doctopic: Quantum physics [m5G; v1.260; Prn:5/06/2019; 16:24] P.10 (1-10) 10 G. Montes Cabrera et al. / Physics Letters A ••• (••••) •••–••• f (τ , θ, β) = 0. That means that for τ ∈ T , it is a necessary but not sufficient condition that it satisfies this equation for some pair of angles (θ, β). Therefore, the surface T must be a subset of the set of solutions τ to this equation. The last to equalities imply that 2γ er + τ = α er for some un- known real parameter α. Inserting this into the first equality, we obtain er · (α er − γ er ) = 0 ⇒ α = er · γ er , and finally τ = (er · γ er) er − 2γ er . (D.2) Appendix E. Canonical form of quantum process generators Here, we prove that the dissipation matrix D , introduced in Eq. (6) is unitarily equivalent to C⊥ , defined in Sec. 2.2. For that purpose, we start from the master equation representation of the generator L of a quantum process, compute its Choi-matrix repre- sentation CL . Finally, we compute C⊥ , the projection of CL onto the subspace orthogonal to the Bell state, given in Eq. (7). We start by rewriting Eq. (6) as L[̺] = φ[̺] − κ ̺ − ̺κ† , where (E.1) φ[̺] = d2−1 ∑ i, j=1 D i j F i ̺ F † j , κ = i H + 1 2 d2−1 ∑ i, j=1 D i j F † j F i . To shorten the notation, we introduce the projector on the Bell state, as ω = |B〉〈B| and the projector on the complementary subspace as ω⊥ = 1 − ω. We find, d (id⊗L)[w] = d (id⊗ φ)[w] − id⊗ κ w − w id⊗ κ† , and therefore ω⊥ CL ω⊥ = d ω⊥ (id⊗ φ)[w] ω⊥ = Cφ , (E.2) the Choi-matrix representation of the map φ[̺]. The latter equality means that Cφ already is orthogonal to w . This can be seen from ω (id⊗ φ)[ω] = d2−1 ∑ i, j=1 D i j ω (1⊗ F i)ω (1⊗ F † j) = d2−1 ∑ i, j=1 D i j tr(F i) ω (1⊗ F † j) = 0 , (E.3) and similarly (id ⊗ φ)[ω] ω = 0, also. Finally, we consider the matrix representation C⊥ of Cφ , with respect to a basis {|φi〉}d 2−1 i=1 orthogonal to |B〉. Then, we prove that C⊥ is related to D by an unitary transformation. For that pur- pose, consider the matrix elements of C⊥ , given by (C⊥)nm = 〈φn|C⊥|φm〉 = d 〈φn| (id⊗ φ[ω]) |φm〉 = d d2−1 ∑ i, j=1 D i j aia j 〈φn|(i)〉〈( j)|φm〉, (E.4) with √ ai |(i)〉 = 1 ⊗ F i |B〉 and ai = ‖(1 ⊗ F i) B‖2 . Now observe that 〈(i)|( j)〉 = 〈B|1⊗ F † i F j|B〉 = 1 d tr ( F † i F j ) = 1 d δi j , thus ai = 1 d , independent of i. Defining Vni = 〈φn|(i)〉 and substi- tuting the value of ai in Eq. (E.4) we end up with C⊥ = V DV † , with V unitary, given that |φn〉 and |(i)〉 are properly normalized quantum states. References [1] H.-P. Breuer, Foundations and measures of quantum non-Markovianity, J. Phys. B, At. Mol. Opt. Phys. 45 (2012) 154001. [2] Á. Rivas, S.F. Huelga, M.B. Plenio, Quantum non-Markovianity: characterization, quantification and detection, Rep. Prog. Phys. 77 (2014) 094001. [3] L. Li, M.J. Hall, H.M. Wiseman, Concepts of quantum non-Markovianity: a hier- archy, Phys. Rep. 759 (2018) 1–51. [4] H.-P. Breuer, E.-M. Laine, J. Piilo, Measure for the degree of non-Markovian behavior of quantum processes in open systems, Phys. Rev. Lett. 103 (2009) 210401. [5] A. Rivas, S.F. Huelga, M.B. Plenio, Entanglement and non-Markovianity of quan- tum evolutions, Phys. Rev. Lett. 105 (2010) 050403. [6] H.-P. Breuer, F. Petruccione, The Theory of Open Quantum Systems, Oxford Uni- versity Press, New York, 2002. [7] M. Wolf, J. Cirac, Dividing quantum channels, Commun. Math. Phys. 279 (2008) 147–168. [8] M.M. Wolf, J. Eisert, T.S. Cubitt, J.I. Cirac, Assessing non-Markovian quantum dynamics, Phys. Rev. Lett. 101 (2008) 150402. [9] F. Benatti, R. Floreanini, M. Piani, Quantum dynamical semigroups and non- decomposable positive maps, Phys. Lett. A 326 (2004) 187–198. [10] D. Chruściński, S. Maniscalco, Degree of non-markovianity of quantum evolu- tion, Phys. Rev. Lett. 112 (2014) 120404. [11] S. Wißmann, H.-P. Breuer, B. Vacchini, Generalized trace-distance measure connecting quantum and classical non-Markovianity, Phys. Rev. A 92 (2015) 042108. [12] D. Chruściński, A. Kossakowski, A. Rivas, Measures of non-markovianity: divis- ibility versus backflow of information, Phys. Rev. A 83 (2011) 052128. [13] M.J.W. Hall, Complete positivity for time-dependent qubit master equations, J. Phys. A, Math. Theor. 41 (2008) 205302. [14] M.A. Nielsen, I.L. Chuang, Quantum Computation and Quantum Information, Cambridge University Press, Cambridge, 2000. [15] N. Boulant, T.F. Havel, M.A. Pravia, D.G. Cory, Robust method for estimating the Lindblad operators of a dissipative quantum process from measurements of the density operator at multiple time points, Phys. Rev. A 67 (2003) 042322. [16] J.M. Dominy, L.C. Venuti, A. Shabani, D.A. Lidar, Evolution prediction from to- mography, Quantum Inf. Process. 16 (2017) 78. [17] D. Chruściński, S. Pascazio, A brief history of the GKLS equation, Open Syst. Inf. Dyn. 24 (2017) 1740001. [18] C.D. Meyer, Matrix Analysis and Applied Linear Algebra, SIAM, Philadelphia, 2000. [19] J.E. Prussing, The principal minor test for semidefinite matrices, J. Guid. Control Dyn. 9 (1986) 121–122. [20] J. Kołodyński, J.B. Brask, M. Perarnau-Llobet, B. Bylicka, Adding dynamical gen- erators in quantum master equations, Phys. Rev. A 97 (2018) 062124. [21] D. Chruściński, A. Rivas, E. Størmer, Divisibility and information flow notions of quantum Markovianity for noninvertible dynamical maps, Phys. Rev. Lett. 121 (2018) 080407. [22] M.-D. Choi, Completely positive linear maps on complex matrices, Linear Alge- bra Appl. 10 (1975) 285–290. [23] T. Heinosaari, M. Ziman, The Mathematical Language of Quantum Theory: From Uncertainty to Entanglement, Cambridge University Press, New York, 2011. [24] D.E. Evans, J.T. Lewis, Dilations of Irreversible Evolutions in Algebraic Quantum Theory, Dublin Institute for Advanced Studies, Dublin, 1977. [25] E.C.G. Sudarshan, P.M. Mathews, J. Rau, Stochastic dynamics of quantum- mechanical systems, Phys. Rev. 121 (1961) 920–924. [26] V. Gorini, A. Kossakowski, E.C.G. Sudarshan, Completely positive dynamical semigroups of n-level systems, J. Math. Phys. 17 (1976) 821–825. [27] G. Lindblad, On the generators of quantum dynamical semigroups, Commun. Math. Phys. 48 (1976) 119–130. [28] M.J.W. Hall, J.D. Cresser, L. Li, E. Andersson, Canonical form of master equations and characterization of non-Markovianity, Phys. Rev. A 89 (2014) 042120. [29] M. Žnidarič, Geometry of local quantum dissipation and fundamental limits to local cooling, Phys. Rev. A 91 (2015) 052107. [30] M.B. Ruskai, S. Szarek, E. Werner, An analysis of completely-positive trace- preserving maps on M2 , Linear Algebra Appl. 347 (2002) 159. [31] D. Braun, O. Giraud, I. Nechita, C. Pellegrini, M. Žnidarič, A universal set of qubit quantum channels, J. Phys. A, Math. Theor. 47 (2014) 135302. [32] I. Bengtsson, K. Zyczkowski, Geometry of Quantum States: An Introduction to Quantum Entanglement, Cambridge University Press, New York, 2006. 134 Chapter C. Articles C.4 Article: Positivity and Complete positivity of differ- entiable quantum processes Physical Review A 96, 062127 (2017). Click to go to the webpage, Click to go to arXiv. PHYSICAL REVIEW A 96, 062127 (2017) Quantum non-Markovianity and localization David Davalos1 and Carlos Pineda1,2 1Instituto de Física, Universidad Nacional Autónoma de México, México D. F. 01000, México 2University of Vienna, Faculty of Physics, Boltzmanngasse 5, 1090 Wien, Austria (Received 22 August 2017; published 21 December 2017) We study the behavior of non-Markovianity with respect to the localization of the initial environmental state. The “amount” of non-Markovianity is measured using divisibility and distinguishability as indicators, employing several schemes to construct the measures. The system used is a qubit coupled to an environment modeled by an Ising spin chain kicked by ultrashort pulses of a magnetic field. In the integrable regime, non-Markovianity and localization do not have a simple relation, but as the chaotic regime is approached, simple relations emerge, which we explore in detail. We also study the non-Markovianity measures in the space of the parameters of the spin coherent states and point out that the pattern that appears is robust under the choice of the interaction Hamiltonian but does not have a classical-like phase-space structure. DOI: 10.1103/PhysRevA.96.062127 I. INTRODUCTION Open quantum systems were recognized as an important subfield of quantum mechanics early in their history [1], because understanding them allows one to explain ubiquitous phenomena, such as spontaneous decay [2]. Later, the Lindblad equation was proposed to describe the evolution of the reduced density matrix of a quantum system weakly coupled to a memoryless environment [3–5]. Environments that lie outside that approximation (Lindblad equation) have attracted the attention of the community in later years. This is, arguably, because we now have such delicate control of quantum systems that memory effects become experimentally relevant [6] and environment engineering is possible [7,8] to mitigate or even use such effects [6,9,10]. A whole community is now dedicated to the study of such systems, known as non-Markovian environments. Numerous efforts have been made to define non-Markovianity (NM) in a precise manner, to measure it, and to take advantage of it (see the previous review papers and Refs. [11,12]). Many systems have been studied under this program, both theoretically and experimentally [6]. Currently, there are many examples of non-Markovian environments that produce a variety of effects. However, not much is known regarding what the key properties that might boost the non-Markovianity of an environment are. Some properties, such as the structure of the phase space of the classical counterpart of the environment, have proven to be crucial; however, what happens when we do not find such a classical analog? In this paper, we focus on two questions. First, is the value of the several measures of non-Markovianity for long times only dependent on the effective dimension of the Hilbert space? Second, is there a hidden underlying classical structure in the environment that we can unveil with the help of these measures? To study these questions, we consider a qubit coupled to a kicked spin chain, which has integrable, mixed, and chaotic dynamical regimes [13,14], but, as far as we know, no semiclassical analog. The interaction between qubit and environment is set up so as to have dephasing, so all the deco- herence effects on the qubit are contained in a suitably defined fidelity of the environment. To quantify NM, we use two commonly used measures [15,16] and a third that was recently introduced and which has a direct relation with a physical task [11]. We find complex relations between NM and the localization of initial environmental states in the integrable and mixed regimes, which depend on the peculiarities of each NM measure. In fact, in Ref. [17] a relation between localization, induced by disordered, and a particular non-Markovianity measure was explored for an environment consisting of an array of cavities. In the case of the recently introduced measures [11], the effective dimension of the Hilbert space of the environmental states has an important role which leads to more complex behavior. In the chaotic regime, due to the ergodic properties of the Hamiltonian, the relation is simpler and almost homogeneous. Regarding the search for underlying classical structure, we focus our attention on the features that emerge in the space of the parameters of the initial states (spin coherent states) when the NM and the inverse participation ratio (IPR) [18] are calculated. We searched for the characteristic finely granulated fractal structure predicted by the Kolmogorov–Arnold–Moser (KAM) theorem but found only a coarse nonfractal one. The paper is organized as follows. In Sec. II, we give a brief introduction to the measures used for non-Markovianity and for localization of quantum states. In Sec. III, we present the general scheme of dephasing dynamics and the details of the dynamics. In Sec. IV, we present and discuss the results. We finish by summarizing the results in Sec. V. II. TOOLS A. Identifying non-Markovianity Many measures of non-Markovianity have been proposed: The two most widespread are the BLP (introduced by Breuer, Laine and Piilo in [15]) and RHP (introduced by Rivas, Huelga and Plenio in [16]) measures. The first is based on the violation of the contraction property of Markovian systems, i.e., decreasing distinguishability between initial quantum states. The second is based on the violation of a well-known mathematical property of Markovian process, divisibility of the quantum map. Both criteria come from the classical theory of Markovian stochastic process. A whole new set of measures 2469-9926/2017/96(6)/062127(13) 062127-1 ©2017 American Physical Society DAVID DAVALOS AND CARLOS PINEDA PHYSICAL REVIEW A 96, 062127 (2017) CP ?CP Et′,0 Et,t′ Et,0 ̺(0) ̺(t′) ̺(t) FIG. 1. Illustration of the concept of CP divisibility. The process E is CP divisible if all existing intermediate maps E(t,t ′) are complete positive and trace preserving. have been proposed [10]. One of these [11], proposed by the authors of this paper, is based on quantifying the probability of successfully performing a certain task. It is hard to strictly verify if a stochastic system fulfills the classical definition of Markovianity [19], since it depends on the whole history of the stochastic process. An additional caveat for quantum systems is the fact that in order to observe intermediate states of the system, one would have to measure, thus collapsing the wave function and thus also the probability distributions. This leads, among other problems, to violation of Kolmogorov consistency conditions even for closed quantum systems [10]. One can, however, check the necessary conditions for Markovianity that can be easily interpreted from a physical point of view. For example, notice that a classical stochastic process (not necessarily Markovian) can be described by a time-dependent right stochastic matrix A(t) that maps the initial probability distribution p(t = 0) to A(t) p(0) = p(t). Matrices describing the intermediate process, say the map from time t ′ to t  t ′  0, described by At,t ′ ≡ At,0A −1 t ′,0, will also be right stochastic matrices for Markovian processes. We argue that the intermediate process is a valid one, and if At,t ′ is right stochastic for all t  t ′  0, the process is said to be divisible. This construction can extended to the quantum case, replacing the divisibility concept with the completely positive map (CP map), which characterizes a valid quantum channel. Given a quantum process Et,0, we shall say that it is CP divisible if the intermediate dynamics Et,t ′ ≡ Et,0E−1 t ′,0, t  t ′  0 (1) are CP maps. Figure 1 illustrates the general idea for divisi- bility and CP divisibility. A general property of a CP-divisible process is that given any Hermitian operator  the trace norm decreases under the action of the map [9] ||E()||1  ||||1, where || · ||1 is the trace norm. In particular, choosing  = 1/2(̺1 − ̺2) we have D(E(̺1),E(̺2))  D(̺1,̺2), (2) where D(̺1,̺2) = 1/2||̺1 − ̺2||1 is the trace distance. This property shows the contraction of the state space under a Markovian process. This in turn shows how two initial conditions are increasingly forgotten and are more difficult to distinguish as the trace norm is directly related with the two state discrimination problem. Some authors define Markovianity with this property: If there exists a pair of quantum states such that the last equation does not hold, in Ref. [15] the process is said to be non-Markovian. B. Quantifying non-Markovianity Two well-known measures of non-Markovianity can be constructed, based on violations of either Eqs. (1) or (2). In particular, the authors of both measures constructed them adding up the local contributions of the chosen criterion. For the case of the RHP measure (based on divisibility), the authors define g(t) = lim ǫ→0+ ||J [E(t+ǫ,t)]||1 − 1 ǫ , (3) where J [E(t+ǫ,t)] is the Jamiołkowski isomorphism [20] that relates quantum channels and density matrices. In particular, it takes CP maps to positive operators with unit trace. Thus, if E(t+ǫ,t) is a CP map, the eigenvalues of the J [E(t+ǫ,t)] will all be positive and add up to one. Otherwise, they will still add up to one, but with negative contributions. Thus, g(t) is greater than zero if at time t the dynamics are not divisible; otherwise, g(t) = 0. The measure proposed in Ref. [16] is obtained by integrating the contributions of the non-CP-divisible behavior throughout the entire evolution: NRHP[E] = ∫ ∞ 0 g(t)dt. (4) The brackets here indicate functional dependency. In a similar spirit, we can integrate the deviations from the contractive behavior, expected for Markovian evolution. Considering the derivative of the trace distance σ (t,̺1,2(0)) = dD(̺1(t),̺2(t)) dt . (5) According to Eq. (2), σ  0 for Markovian dynamics. We can integrate this deviation to obtain the measure proposed in Ref. [15], where a maximization over all states is taken. Thus, NBLP[E] = max ̺1,̺2 ∫ σ>0 σ (t,̺1(0),̺2(0))dt. (6) These two measures have some serious drawbacks. In particular, they are not continuous in the spaces of functions, and small fluctuations can change the value of the measure by an arbitrarily large amount. Notice that these issues arise always with a finite Hilbert-size environment and also in finite number statistics. One has the option to cut the integration interval to a finite time or smooth out the fluctuations by windowing the data. One can also consider other proposals [11] which not only remove that problem but also provide a physical interpretation for the number obtained. The proposals are Nmax K [t ] = max tf ,τtf [K(tf ) − K(τ )] (7) and N 〈·〉 K [t ] = max{0, max tf [K(tf ) − 〈K(τ )〉τ0 Ḟ (t) F (t) dt = ∑ i [ln (F (bi)) − ln (F (ai))], (16) with bi and ai the times of the ith maximum and minimum of F (t) respectively. For the computation of the BLP measure, the states that maximize Eq. (6) are those lying on the equator of the Bloch sphere in antipodal positions. The trace distance is the Loschmidt echo, D(̺1(t),̺2(t)) = F (t). From Eq. (6), the measure is NBLP[E] = ∫ Ḟ>0 dF (t) dt dt = ∑ i [F (bi) − F (ai)]. (17) This shows a direct relation with both revivals and fluctuations of the Loschmidt echo of the environmental dynamics. Finally, measures Nmax K [t ] and N 〈·〉 K [t ], as long as they are invariant with respect to unitary operations in the qubit, will depend only on F in the same way that the particular K chosen depends on F . B. The environment The system used as environment is the homogeneous Ising spin-1/2 chain kicked by short pulses of magnetic field. This system was proposed by Prosen to study the relation between ergodicity and fidelity [13,14]. The Hamiltonian reads Henv = N−1 ∑ i=0 σ z i σ z i+1 + δ̂(t) N−1 ∑ i=0 b⊥σ x i + b‖σ z i , (18) where δ̂(t) = ∑∞ n=−∞ δ(t − n) and σN ≡ σ0. The first term corresponds to a homogeneous Ising interaction strength; b⊥ and b‖ are the perpendicular and parallel components of the magnetic field with respect to the direction of the Ising interaction; finally, δ̂(t) is a train of Dirac δs with period 1. This system has three well-known dynamical regimes. For both b⊥ = 0 or b‖ = 0 the chain is integrable [13]. For b‖ = b⊥ ≈ √ 2, the dynamics is chaotic in the sense of random matrix theory [22]. It follows that the nearest neighbor spacing distribution P (s) of the quasienergies resembles the one of the circular orthogonal ensemble; see the appendix. The third regime is an intermediate one where there is level repulsion but the system is not fully chaotic. The Floquet operator is U = exp ( −i N−1 ∑ i=0 b⊥σ x i + b‖σ z i ) exp ( −i N−1 ∑ i=0 σ z i σ z i+1 ) , (19) and the evolution operator for longer times is simply U (n) = Un. This model has the advantage that it can be split in one- and two-qubit operations, as the terms in each of the exponentials commute with one another, and one can thus express the exponential as a multiplication of exponentials each with only one or two particles involved. In order to map local features of the non-Markovianity and have initially null correlations in any part of the complete system, we use the spin coherent states as initial states of the environment. They are invariant under permutations and can be regarded as a macroscopic state. Coherent states are defined as a coherent displacement of the fiducial state |J = j ; mz = j 〉: |ϑ,ϕ〉 = e−iϕSze−iϑSy |j ; j 〉 = D(j ) ϑ,ϕ|j ; j 〉, (20) where the total spin is given by j = N/2, D(j ) ϑ,ϕ is the rotation matrix in the subspace of spin j . These states form a complete basis in the symmetric subspace. In fact, one can parametrize these states in a Poincaré sphere, and rewrite |ϑ,ϕ〉 = ( cos ϑ 2 |0〉 + sin ϑ 2 eiϕ|1〉 )⊗N . (21) The environmental Hamiltonian is invariant under external rotations: The translation operator, which takes state ⊗i |ψi〉 to state ⊗i |ψi+1〉, commutes with Eq. (19). This symmetry foliates the Hilbert space in quasimomentum k subspaces [22]. As the translation symmetry leaves Eq. (20) invariant, such states live in the k = 0 subspaces, and as the evolution respects the symmetry, it will remain in such subspace. The calculation of the IPR is thus simply P−1 ϑ,ϕ = dimHk=0 ∑ i=1 ∣ ∣ 〈 φ (k=0) i ∣ ∣ϑ,ϕ 〉∣ ∣ 4 . (22) C. Interaction operator We shall study three kinds of couplings (local, global, and generic) and look for common trends and differences. Local and generic couplings will break the symmetry of the environment, whereas the global one is chosen to maintain it. We continue by presenting the local perturbations. As mentioned above, the interaction was chosen to induce a dephasing channel, for sake of simplicity. The operator V appearing in Eq. (13) can be seen as a perturbation operator of the environment dynamics [see Eq. (14)]. For the case of global perturbations, we probed altering either the magnetic field or the Ising interaction between neighbors, which correspond to choosing V as Vb ≡ δ1(t) N−1 ∑ i=0 σ x i , VJ ≡ N−1 ∑ i=0 σ z i σ z i+1. (23) Analogously, for the local interaction of the qubit with the environment, we chose the coupling as V0,1 ≡ σ z 0 σ z 1 , V0 ≡ δ1(t)σ x 0 , (24) where only two and one qubits of the environment, respec- tively, interact directly with the central qubit. Finally, to study the generic case, we consider the simplest choice, inspired in ergodicity arguments of quantum chaos [23]. We select V from one of the classical ensembles, namely the Gaussian unitary ensemble (GUE). We shall denote that case as VGUE, and it corresponds to a global and structureless perturbation. 062127-4 QUANTUM NON-MARKOVIANITY AND LOCALIZATION PHYSICAL REVIEW A 96, 062127 (2017) FIG. 2. BLP (black triangles, left axis) and RHP (blue squares, right axis) measures as a function of the IPR for initial coherent states of the environment, Eq. (20), distributed uniformly on the Poincaré sphere. Each column corresponds to a different kind of coupling of the qubit to the environment [see Eqs. (23) and (24)], whereas different rows correspond to different dynamical regimes of the environment. The parameters used for this and the rest of the figures are indicated at the beginning of Sec. IV. The results for the global and local field perturbation, Vb and V0, respectively, are very similar to their global and local Ising counterparts. IV. RESULTS The unitary dynamics in qubit plus environment [defined by Eqs. (13) and (18) and the interactions discussed in Sec. III C] induce a specific dephasing channel, Eq. (15), once the initial state of the environment is specified. In our case, such state is a coherent state, Eq. (20), specified by the parameters ϑ and ϕ. The environment, a spin chain, will be used in integrable, mixed, and chaotic regimes, varying b⊥ = 0.1, 1, and 1.4 respectively while fixing b‖ = 1.4. We use b⊥ = 0.1 instead of 0 for integrable dynamics, in order to avoid degeneracies in the spectrum and have a well-defined IPR. Corresponding spectral statistics are presented in the appendix. For all calculations, we chose the coupling parameter ǫ = 0.1. We performed numerical calculations of the measures of NM using time cutoffs of tcut = 10 000 and a mesh in coherent- state parameters (ϑ,ϕ) of ϑ = ϕ = 0.1; the two measures Eqs. (4) and (6) were slightly modified to accommodate to the intrinsic discrete time structure of Eq. (18). We also considered a time cutoff in the integrals of the measures, as the fluctuations caused by a finite-dimensional environment would send the aforementioned measures to infinity. The IPR of the initial environmental states were calculated with respect to the eigenbasis of U+ for simplicity. Since we are taking a small ǫ, the IPR does not vary considerably if instead of U+, we consider U− or a Floquet operator with an intermediate ǫ. We discuss first the relation of the different measures of NM with respect to the IPR. Next we study the dependence of these quantities with respect to the choice of the state of the environment; that is, we study the structure of the environment that can be seen, studying the decoherence of the qubit. The section is closed with some comments on the generality of the results when one varies the dimension of the environment and the total evolution time considered. A. Dependence of non-Markovianity on the state localization We study the behavior of NM, using NRHP and NBLP in Sec. IV A 1 and then using Nmax K and N 〈·〉 K in Sec. IV A 2, with K being D or G. In the first section, we focus in the cases which the coupling is via global and local nearest neighbor Ising interaction, VJ and V0,1, respectively, and a global VGUE operator. In the second section, we focus only on global VJ and VGUE. These interactions represent well what happens for the other cases for each study. 1. Using BLP and RHP measures In Fig. 2, we show, for different initial conditions of the environment and a coupling of the type VJ , the value of NM using BLP and RHP measures as a function of the IPR. In the integrable regime, the two measures have different behaviors; NBLP grows for increasing IPR until it reaches a maximum around P−1 ∼ 0.4, where it starts to decrease. NRHP has an approximate monotonic decreasing behavior, showing a change of slope around P−1 ∼ 0.4 and another close to P−1 ∼ 0.6. A local coupling, namely V0,1, yields similar results; however, the peak in the BLP measure is sharper and the decay of RHP measure is faster (Fig. 2 second column). The behavior of NBLP can be explained qualitatively by studying the fidelity which, for the dephasing case, is related to the distinguishability via the equation D(t) = |f (t)|2. In Fig. 3, FIG. 3. Typical behavior of the fidelities of the environment, Eq. (18), in the integrable regime with a global Ising perturbation VJ , for several coherent states, Eq. (20). We consider 10 and 16 qubits, shown in black and orange curves, respectively. The figure shows the fidelity for the state |ϑ = 2.8,ϕ = 4.8〉 (with IPR equal to 0.457 and 0.375 for 10 and 16 qubits, respectively), which is among the states that yield larger values for measures based on D(t) (dashed curves). Fidelities for the states that give low values of the BLP measure are the dotted and solid curves, obtained from the states |ϑ = 3.0,ϕ = 2.2〉 (IPR equal to 0.994, 0.984) and |ϑ = 1.5,ϕ = 3.5〉 (IPR equal to 0.046, 0.010), respectively, which are high and low localized states. 062127-5 DAVID DAVALOS AND CARLOS PINEDA PHYSICAL REVIEW A 96, 062127 (2017) we show its evolution in the integrable regime, for three initial conditions and two different environment sizes. For high and low values of localization, oscillations of D(t) are constrained around high and low values of asymptotic fidelity, respectively. Therefore, the relatively low values of non-Markovianity belong to the high and low values of localization. There are also states with high IPR that lead to distinguishabilities that oscillate with large amplitude but at a low frequency; those states have low asymptotic fidelity. The maximum value of NM is achieved at ∼0.4, where fidelity can oscillate with a large amplitude. One can understand the behavior of NRHP with similar arguments [see Eq. (16)] but this time taking into account the role of the logarithm. For high localized states, the typical values of the minimums and maximums of D(t) are very close to one or with lower frequency, yielding very small values of the logarithm and thus low values of the RHP measure. As the IPR decreases, the minimums in D(t) diminishes faster than the maximums, and one reaches quickly the regime in which − ln (F (ai))∼O(1), causing an increasing of the measure until IPR ∼ 0.4. For small values of the localization, the typical minimum is very close to zero, for which the logarithm is large, in absolute value. One can approximate NRHP ≈ ∑ i ln (F (bi)) + n ln (F (ã−1)), where n is the number of minimums included in the interval of the computation of the measure and F (ã) is its typical value. The value of the measure is now seen to be directly related with the localization, giving again a monotonic behavior with different slope. For both measures, low localized states tend to cluster. These states are localized in the equator of the Poincaré sphere (see Fig. 9). This explains the two leaflike structures connected by a stem in the integrable regime. In the mixed and chaotic regimes, fidelities begin with a fast decay, after which they fluctuate around the inverse of the effective dimension of the state (Fig. 4). Since the asymptotic fidelity is inversely proportional to the effective dimension of Hilbert space, the scale of the NM is lower in these regimes with respect to the integrable. The IPR is also small due to FIG. 4. Typical behavior of the fidelities of the environment in the chaotic (blue curves) and mixed (black thick curves) regimes for 10 (solid curves) and 16 (dashed curves) qubits, with the coupling V = VJ ; the initial state is a coherent state characterized by |ϑ = 0.7,ϕ = 0.8〉; see Eq. (20). A fast decay and fluctuations around a value determined by the effective dimension of the Hilbert spaces explains the values of the different measures of non-Markovianity. FIG. 5. RHP and BLP measures of the spin chain using a random coupling, chosen from the GUE, in the integrable regime (main panel) and the chaotic regime (inset). We observe a monotonic decreasing behavior for both measures in all regimes, with a short growth for BLP measure in the integrable regime. ergodic properties of the Hamiltonian. In the mixed regime, the slope of the data using VJ and V0,1 is positive for BLP measure, while for RHP it is clearly decreasing for both perturbations, mimicking the integrable cases. Thus, both measures behave differently also in the mixed regime. In the chaotic regime, we expect full ergodic properties, and consequently, a reasoning similar to that of the mixed case will follow, however, with smaller IPR. Indeed, all initial conditions cluster around a smaller region but a slope, consistent with the mixed cases, is observed. Finally, we show the results when a random potential provides the coupling in Eq. (13); namely, when we take V = VGUE. The dependence of NM on the IPR is shown in Fig. 5 for both the integrable and the chaotic cases. Its behavior is qualitatively similar to the one observed for the other couplings, when comparing among integrable cases, mixed and chaotic. However, there are some quantitative differences. For example, the BLP measure still has an initial growth but is very short compared with the case of V = VJ . The same arguments as before can be stated to explain the general features of the behavior. 2. Using measure schemes Nmax K and N 〈·〉 K In the previous section, we considered measures BLP and RHP, which are based on the nonmonotonicity of distinguisha- bility, as measured by D(t), and of divisibility, as measured by G(t) = ∫ t 0 g(τ )dτ . In this section, we use measures based on the same quantities, but use Eqs. (7) and (8) to obtain a quantity that can be directly related to a physical process [11] and contrast its behavior with measures BLP and RHP. For the integrable case, we observe that there are two dis- tinct behaviors, for both measures Nmax K and N 〈·〉 K , regardless of whether they are based on D or G(t). In Fig. 6, we show the results for the case in which the coupling is VJ . These two different behaviors are associated with the two hemispheres of the Poincaré sphere, and its details can be understood by studying the evolution of fidelity. In particular, for Nmax D , one of the branches displays a maximum (IPR ∼ 0.4), then it decays linearly. The other branch, corresponding to the 062127-6 QUANTUM NON-MARKOVIANITY AND LOCALIZATION PHYSICAL REVIEW A 96, 062127 (2017) FIG. 6. Measures N 〈·〉 K (blue) and Nmax K (black) with D(t) [left column] and G(t) [right column], for the spin chain using global Ising perturbation VJ , as a function of the initial IPR of the environment; see Fig. 2. The initial states of the environment are coherent states uniformly chosen from the northern or southern hemispheres of the Poincaré sphere and indicated by the hollow and filled markers, respectively. In the integrable regime (and in the mixed for Nmax D ), we see two different behaviors, coming from the two hemispheres of the Poincaré sphere. The results for local Ising interaction, V0,1, are very similar to the presented here. Results for global and local field perturbations, Vb and V0 respectively, presented only the behavior plotted by filled markers. southern hemisphere (π/2 < ϑ  π ), has a slight increase with IPR. The behavior of N 〈·〉 D is similar; however, it is scaled down, and instead of a slight increase, the southern hemisphere displays a small increase with IPR. A quantitatively similar behavior is seen when we base our measures in G(t), with the bending point being again at IPR ∼ 0.4, for Nmax G . N 〈·〉 G is also a scaled down and slightly deformed version of N 〈·〉 D . For low localized states, the explanation of the aforementioned behavior is similar to the one given for BLP and RHP measures. Since the size of the fluctuations of the fidelity depend on the effective dimension of the state, N 〈·〉 K and Nmax K increase as we take more localized initial environmental states. For highly localized states in the integrable regime, there are two families of states. One, with asymptotic fidelity greater than 1/2 and whose fidelity has a high frequency but small amplitude, and other with asymptotic fidelity smaller than 1/2 but with a fidelity that has smaller frequency and a larger oscillation amplitude. Since the schemes under discussion depend mainly in the amplitude of the oscillations, they are critically sensitive to the asymptotic fidelity of the environmental states. This FIG. 7. Typical behavior of the fidelities in the integrable regime for V = VGUE. In solid black we plot the fidelity of the state |ϑ = 3.2,ϕ = 1.1〉 as a representative state of highly localized states, and |ϑ = 2.2,ϕ = 2.4〉 in dashed gray as a representative of low localized states. High localized states lead to a low frequency of occurrence of pairs of local minima and maxima, while for localized states such frequency is increased. This explains the different behaviors among BLP and Nmax ,〈·〉 D . feature is a significant difference between the newly proposed schemes [11] and the more often used BLP and RHP. In the mixed and chaotic regimes, the behavior of the measures is monotonically increasing. Since all coherent states have a small IPR, the same arguments given before for low localized states in the integrable regime hold to explain such monotonicity. For the chaotic regime, the measures also tend FIG. 8. Relation between N 〈·〉 K and Nmax K with IPR, using a global random perturbation. We consider the two measures, based on both D(t) and G(t) and a spin chain of eight spins for an ensemble of 40 matrices. Measures based on G are almost constant in all regimes. For Nmax ,〈·〉 D in the integrable regime, we observe different behaviors for each hemisphere of the Poincaré sphere. 062127-7 DAVID DAVALOS AND CARLOS PINEDA PHYSICAL REVIEW A 96, 062127 (2017) FIG. 9. The different columns correspond to density plots of several measures of non-Markovianity and the IPR, for a chain with 10 qubits using the homogeneous perturbation VJ . For Nmax D some smaller structures appear as we go into the chaotic regime. We can also observe a relation in the integrable and mixed regimes between IPR and all non-Markovianity measures. The results for local Ising interaction, V0,1, are very similar, just with more extended depressions. The results for the global and local field perturbations, Vb and V0, respectively, are very similar to their global and local Ising counterparts. to homogenize; this is expected given that the initial states have similar effective dimension, as they appear random in the eigenbasis of the Floquet operator. For a random coupling to the environment, measures Nmax ,〈·〉 D have a monotonic behavior with respect to IPR. However, in contrast to the behavior of the BLP measure, non- Markovianity increases with the inverse participation ratio. This surprising change can be explained when noticing that the BLP measure depends on the number of pairs of minima and maxima that appear in the fidelity in a given interval, while Nmax ,〈·〉 D depend only on the amplitude of the fluctuations of F (t). As we take more localized initial environmental states, the size of the fluctuations is increased as the pairs of minima and maxima appear less frequently (shown in Fig. 7), which explains the aforementioned effect. The behavior in the mixed regime, which is also monotonic increasing, has the same explanation. In the chaotic regime, the values of NM also tend to homogenize, having the same explanation as the one given for V = VJ for this regime. Now using G(t) as indicator, all measure schemes in all regimes yield almost constant NM with respect to the IPR (right panels of Fig. 8). This behavior is expected for the chaotic regime; what remains to be explained is its emergence in the integrable and mixed regimes. To do this, we can find an upper bound for the change of Nmax G in the whole interval of localization; we shall call this NG . From Eq. (7), Nmax G = ln (F (tf )) − ln (F (τ )), where tf and τ are the maximum and the minimum attained to the maximization required by the definition. Now, since the logarithm is a monotonic function, the measure Nmax D is attained to the same times, allowing us to write Nmax G = ln (Nmax D + F (τ )) − ln (F (τ )) ≈ ln (Nmax D ) + F (τ )/Nmax D − ln (F (τ )). Therefore, the total change is NG =  ln (Nmax D ) + (F (τ )/Nmax D ) −  ln (F (τ )). The last term can be ignored since F (τ ) is typically very similar for any value of localization. The second term is negative since Nmax D changes faster than F (τ ) and its absolute value is smaller than the first term which is positive. Therefore, Nmax G is upper bounded by  ln (Nmax D ) and its numerical values for the integrable and mixed regime are 0.4 and 0.08 respectively. There is a similar explanation for N 〈·〉 G using typical values of the average instead of the minima. We finish this section by summarizing the results and commenting on practical consequences of the relations we found between non-Markovianity and IPR. The integrable regime shows the richest behavior when we use a structured coupling to the environment. In our case, we observed a wide variety which includes up to two different behaviors for the two hemispheres of the Poincaré sphere. In general, the different measures behave differently and depend on the details of the fidelity. However, the IPR determines coarsely the value of the non-Markovianity. As mentioned in Sec. II B, measure N 〈·〉 D is directly related to the task of storing information safely; we can see that to perform such a task with a high probability of success, we need an environment in the integrable regime, a structured interaction, and states with 062127-8 QUANTUM NON-MARKOVIANITY AND LOCALIZATION PHYSICAL REVIEW A 96, 062127 (2017) intermediate localization. When the environment is in the chaotic regime, the behavior is not so rich, as the coherent states are quite delocalized, and the non-Markovianity seems to be self averaging. In the mixed regime of the environment, we have an intermediate behavior. B. Underlying structure In Ref. [24], the authors show that non-Markovianity, via long time fluctuations of fidelity, is able to resolve complex phase space structures of the environment using initial coherent states. In particular, the fractal nature of the phase space is clearly visible in the mixed regime. We investigated the spin chain in a similar way, using spin coherent states as initial environmental states, studying now the measures of NM and the IPR as functions of the parameters of the spin coherent states. Our goal is to study the visible structures and how they change during the transition from integrability to chaos. In the integrable regime (top of Fig. 9), the values of the NM measures mimic the behavior of the IPR close to the equator of the Poincaré sphere (ϑ = π/2); close to the poles, the situation is different. The equator of the Poincaré sphere corresponds to low localized states, and this in turn leads to local minimums for all examined measures of NM. When moving toward the poles, which are very localized states, one finds a local maximum, and then in the vicinity of the pole, a local minimum, for all cases except for Nmax D near the north pole. The difference arises from the different asymptotic fidelities of the chosen high localized states. This picture deepens the understanding of the behavior already seen in Figs. 2 and 6. For the mixed regime, the features on the NM measures are mainly governed by the IPR. High localization leads to local maximums in the measure Nmax D and local minimums for the RHP measure. For the BLP measure, there is also an interesting feature. The local maximum of IPR, located around ϑ ≈ ϕ ≈ 2.5, leads to a local minimum on the NM which is partially surrounded by a maximum. This behavior is actually similar to the one at the poles in the integrable regime. In the chaotic regime, the relation of the measures with the localization practically vanishes. Regarding the transition from integrability to chaos, using the BLP and RHP measures, there is not a notable change in the size of the structures as it does for environments with a classical analog [24,25]. This might be due to the absence of such structures, or, that simply due to the relative size of the coherent states in this system, they are not able to resolve small structures. More quantitatively, the fluctuations of the spin co- herent states in the Poincaré sphere (chosen to have radius one) scale as ∼N−1 [26], i.e., as [ln2 (dimH)]−1, while for coherent states in the torus fluctuations scale as (dimH)−1 [27]. The situation is different when usingNmax D . In the transition to chaos, a finer structure emerges. Although such features do not appear classical, in the sense of the appearance and breaking of KAM tori, it is clear that there is a finer granularity than is typically expected in this transition; these structures are robust with respect to changes in parameters and times of FIG. 10. Density plots of the NM measures and the IPR for the chain with eight qubits, using random potentials V = VGUE averaged over 40 matrices. Figures show an emerging fine structure in the transition from integrability to chaos in N 〈·〉 G . The structures observed in the NM measures are correlated (or anticorrelated) with the IPR. 062127-9 DAVID DAVALOS AND CARLOS PINEDA PHYSICAL REVIEW A 96, 062127 (2017) FIG. 11. Relation of the IPR and the averaged square root of the asymptotic fidelity, F(t), for 10 and 16 qubits. The figure shows the splitting in the NM vs IPR relation, explaining both the peculiar behavior of the results shown in Fig. 6 and the spreading of the non-Markovianity measures, for a fixed IPR, as the dimension is increased. integration. We consider this one of the central results of this work. Let us now comment on the results using V = VRMT, shown in Fig. 10. In the integrable regime, measures Nmax D and N 〈·〉 G completely mimic the behavior of the IPR, while the BLP measure is anticorrelated with the IPR. Such behavior is a consequence of the way fidelity contributes to the different measures. Recall that the BLP measure depends mainly on the frequency with which the pairs of minima and maxima occur in D(t), while schemes Nmax K and N 〈·〉 K depend mainly on the amplitude. In the mixed and chaotic regimes, the situation is similar: The IPR is correlated with Nmax D and anticorrelated with BLP. However, for N 〈·〉 G the landscapes appear to have almost no correlation with IPR. It is also important to underline that the measures N 〈·〉 G and Nmax D show a nonfractal structure in the transition to chaos, as in the results using VJ . Results using RHP measure are very similar to the ones for BLP; the ones for N 〈·〉 D resemble the ones for Nmax D , and the results using Nmax G reveal only a random landscape for all regimes. C. Generality of the results This section is devoted to a discussion the validity of the main results presented above for a larger number of qubits and for different cutoff times. We first discuss three key features, namely (i) the decreasing behavior of the BLP and RHP measures for high localized states (shown in Figs. 2 and 5); (ii) the same property for measures Nmax K and N 〈·〉 K , but only for the hemisphere which contains the states with low asymptotic fidelity (Fig. 6); and (iii) the peculiar behavior of measures based on D(t) (also shown in Fig. 6), which exhibits a clear change on the slope as localization is increased. Let us now comment how these observations behave as the dimension of the environment is increased, and for sake of brevity only for measures Nmax K (shown in Fig. 12). The results show that the patterns are preserved; however, as the dimension increases the data becomes diffused, i.e., for each value of IPR there is a wider FIG. 12. Relation of the NM with the IPR and with the averaged asymptotic fidelity F(t), for 10 and 16 qubits using V = VJ (compare with Fig. 6). The figures show the data spreading of the relation NM versus IPR when the dimension is increased (upper row). Such feature is not present in the relation of NM versus F(t) (lower row). We have the same situation. 062127-10 QUANTUM NON-MARKOVIANITY AND LOCALIZATION PHYSICAL REVIEW A 96, 062127 (2017) FIG. 13. Density plots of the measures of NM and of the IPR for the chain with 16 qubits and V = VJ . As in the case of the chain with 10 qubits, the fine structures in Nmax D is present. The local maximums in the IPR also dictate where are the dominant structures of local maximums or minimums in the NM. range of NM. This is due to the relation between asymptotic fidelity F(t) and IPR (shown in top panel of Fig. 11), which is linear (for each hemisphere of the Poincaré sphere) but spreads out for a larger number of qubits. It is interesting that this observation also reveals the origin of the above-mentioned splitting of the relation between localization and non-Markovianity, due to the different values of asymptotic fidelities of high localized states. Therefore, by plotting the relation of NM versus F(t) (shown in the bottom panel of Fig. 12), it can be seen that the splitting and the spreading of the data are removed, revealing that the relation of NM is simpler as a function of the effective dimension of the Hilbert space of the initial states. Next, we shall study the emergent structures in the computed measures for the system with a higher dimension (we used a spin chain with 16 qubits). It yields basically the same behavior as for the 10 qubits case (shown in Fig. 13), but there is an emergence of smaller, finer features in the landscapes of measure Nmax D (N 〈·〉 G ), which has basically an identical landscape. We conclude that such fine structures become smaller as the dimension is increased. A general characteristic of the landscapes, especially in the integrable and mixed regimes, is that the local maximums in the IPR determines the most visible structures in the NM. They appear as local maximums or minimums depending on the chosen measure and/or in the asymptotic averaged fidelity of the coherent states of the region. We finalize this section by discussing the validity of our observations for other cutoff times. In Fig. 14, we show the values of all the measures treated in this paper for the integrable case and for one state of the environment, as a function of the cutoff time. Measures BLP and RHP are normalized by tcut to avoid their trivial linear dependence. The figure shows that all measures saturate quickly to its asymptotic value, except Nmax G , which saturates more slowly than others but more quickly with respect to the system size. We discussed only FIG. 14. Measures of NM as a function of the cutoff time, using the coherent state |ϑ = 2.8,ϕ = 4.8〉 for 10 (solid lines) and 16 (dashed lines) qubits in the integrable regime. BLP and RHP measures are normalized by tcut to remove their linear dependence on tcut; it is clear from this plot that without such normalization they grow mainly linearly with time. The figure shows that the cutoff time used throughout the paper is appropriate to understand the results for asymptotic times. 062127-11 DAVID DAVALOS AND CARLOS PINEDA PHYSICAL REVIEW A 96, 062127 (2017) the results for one state in one regime since the exploration for other cases gives very similar results. V. CONCLUSIONS We performed numerical calculations of the non- Markovianity of a qubit coupled to an environment modeled by a unitary kicked spin chain in a coherent state. Several dynamical regimes of the chain, couplings between qubit and environment, and measures of non-Markovianity were considered. Additionally, the inverse participation ratio of the environment (with respect to the coupled environment) was calculated. We explored the relation of NM versus IPR and showed that the schemes Nmax K and N 〈·〉 K proposed in Ref. [11] have important and potentially useful differences with respect to the more common measures BLP and RHP. We showed that that the first mentioned schemes reveal the asymptotic fidelity of the environmental state, leading to two clearly different behaviors of the measures in function of the IPR. Regarding the validity of the former results, we showed that the relations between non-Markovianity and localization for larger environments remain the same. However, self averaging was not observed. A central result of the paper is the identification of a maximum of the NM for intermediately localized environmental states, when using distinguishability as indicator. Such a scenario could be used to protect classical information more efficiently [11]. In the second part of the work, we presented a study of the NM and the IPR as functions of the parameters of the Poincaré sphere in which the initial coherent environmental states live. We concluded that there are structures mainly depicted by the IPR in all dynamical regimes; these are robust under the election of the interaction Hamiltonian and the dimension of the environment. We have shown that although such structures are not classical-like (in the sense that they do not present KAM behavior), they become finer in the transition to chaos when using measures Nmax D and N 〈·〉 D . Such features remain stable with respect to the cutoff time, indicating that they are not random fluctuations, and become finer as the dimension increases. ACKNOWLEDGMENTS We acknowledge the support by CONACyT and DGAPA- IN-111015, as well useful discussions with Heinz-Peter Breuer, Diego Wisniacki, and Thomas Gorin. APPENDIX: DYNAMICAL REGIMES The spin chain has well-known dynamical regimes in the sense of random matrix theory. The analysis of the spectra (the eigenphases of the Floquet operator) has been done for the chaotic regime and for 16 qubits in Ref. [22]. In this appendix, we present a brief analysis for the integrable regime for 12 qubits and for completeness also for the chaotic and mixed regimes, following the aforementioned work. In order to show the correspondence of the eigenphases of the Floquet operator with the results of random matrix theory, we have to identify the subspaces corresponding to FIG. 15. The figure shows the nearest neighbor spacing distri- butions P (s) of the spin chain with 12 qubits for two values of the control parameter. In the main figure, the dotted blue curve shows the P (s) for the chaotic regime, and the dashed black shows that for the integrable regime. The solid black curve shows the P (s) for the Poissonian orthogonal ensemble and the solid blue shows it for the circular orthogonal ensemble. In the inset, the dashed curve shows the P (s) for the mixed regime and the solid curve shows that for the Brody distribution with b = 0.77; see Ref. [28]. There is good agreement on all regimes with the predictions of random matrix theory. the good quantum numbers of the system. We then compute the distribution of the distance among the nearest neighbor eigenphases [named P (s)] in each symmetry sector. The homogeneous spin chain has a symmetry under translation of spins; i.e., the Hamiltonian remains invariant if we take the spin i to i + 1. Thus we will use the eigenspectra corresponding to the eigenspaces of the translation operator T for the analysis of P (s). The symmetry operator acts in the computational basis |α0, . . . ,αN−1〉 (αj ∈ {0,1}), as T |α0, . . . ,αN−1〉 = |αN−1,α0, . . . ,αN−2〉. Since T N = I, its eigenvalues are sim- ply exp (2πik/N ) with k an integer between 0 and N − 1. Therefore, the Hilbert space is foliated into N subspaces H = ⊕k∈Z/N Hk . The chain also has a reflection symmetry given the symmetry operator R, which transforms R|α0, . . . ,αN−1〉 = |αN−1, . . . ,α0〉. This symmetry commutes with the T in the subspace identified by k = 0, and for even N , also in k = N/2; for simplicity these subspaces are removed from the calculation. Figure 15 shows the averaged nearest neighbor spacing distribution over the relevant subspaces, and the ansatz corresponding to the different dynamical regimes [28]. For the integrable regime, we plot the Poisson distribution e−s ; for the chaotic, we plot the Wigner surmise; finally, for the mixed regime, we present the Brody distribution [29], Pq (s) = (q + 1)sqŴ ( q + 2 q + 1 )q+1 e −sq+1Ŵ ( q+2 q+1 )q+1 . The Brody parameter is denoted by q and takes the ansatz from the integrable case (q = 0) to the Gaussian orthogonal ensemble (q = 1), fitting smoothly with the nearest spacing distribution of the chain in the transition to chaos. 062127-12 QUANTUM NON-MARKOVIANITY AND LOCALIZATION PHYSICAL REVIEW A 96, 062127 (2017) [1] J. von Neumann, Wahrscheinlichkeitstheoretischer aufbau der quantenmechanik, Nachr. Ges. Wiss. Goettingen 1927, 245 (1927). [2] P. A. M. Dirac, The quantum theory of the emission and absorption of radiation, Proc. R. Soc. London, Ser. A 114, 243 (1927). [3] G. Lindblad, On the generators of quantum dynamical semi- groups, Commun. Math. Phys. 48, 119 (1976). [4] A. Kossakowski, On quantum statistical mechanics of non-hamiltonian systems, Rep. Math. Phys. 3, 247 (1972). [5] V. Gorini, A. Kossakowski, and E. C. G. Sudarshan, Completely positive dynamical semigroups of n-level systems, J. Math. Phys. 17, 821 (1976). [6] I. de Vega and D. Alonso, Dynamics of non-Markovian open quantum systems, Rev. Mod. Phys. 89, 015001 (2017). [7] F. Verstraete, M. M. Wolf, and I. J. Cirac, Quantum computation and quantum-state engineering driven by dissipation, Nat. Phys. 5, 633 (2009). [8] J. Nokkala, F. Galve, R. Zambrini, S. Maniscalco, and J. Piilo, Complex quantum networks as structured envi- ronments: Engineering and probing, Sci. Rep. 6, 26861 (2016). [9] Á. Rivas, S. F. Huelga, and M. B. Plenio, Quantum non- Markovianity: Characterization, quantification, and detection, Rep. Prog. Phys. 77, 094001 (2014). [10] H.-P. Breuer, E.-M. Laine, J. Piilo, and B. Vacchini, Colloquium: Non-Markovian dynamics in open quantum systems, Rev. Mod. Phys. 88, 021002 (2016). [11] C. Pineda, T. Gorin, D. Davalos, D. A. Wisniacki, and I. García- Mata, Measuring and using non-Markovianity, Phys. Rev. A 93, 022117 (2016). [12] P. M. Poggi, F. C. Lombardo, and D. A. Wisniacki, Driving- induced amplification of non-Markovianity in open quantum systems evolution, EPL 118, 20005 (2017). [13] T. Prosen, A new class of completely integrable quantum spin chains, J. Phys. A 31, L397 (1998). [14] T. Prosen, Exact time-correlation functions of quantum Ising chain in a kicking transversal magnetic field: Spectral analysis of the adjoint propagator in Heisenberg picture, Prog. Theor. Phys. Suppl. 139, 191 (2000). [15] H.-P. Breuer, E.-M. Laine, and J. Piilo, Measure for the Degree of Non-Markovian Behavior of Quantum Processes in Open Systems, Phys. Rev. Lett. 103, 210401 (2009). [16] Á. Rivas, S. Huelga, and M. Plenio, Entanglement and Non- Markovianity of Quantum Evolutions, Phys. Rev. Lett. 105, 050403 (2010). [17] S. Lorenzo, F. Lombardo, F. Ciccarello, and G. M. Palma, Quantum non-Markovianity induced by Anderson localization, Sci. Rep. 7, 42729(EP) (2017). [18] L. Benet, T. H. Seligman, and H. A. Weidenmüller, Quantum Signatures of Classical Chaos: Sensitivity of Wave Functions to Perturbations, Phys. Rev. Lett. 71, 529 (1993). [19] G. P. Basharin, A. N. Langville, and V. A. Naumov, The life and work of A. A. Markov, Linear Algebra Appl. 386, 3 (2004). [20] I. Bengtsson and K. Życzkowski, Geometry of Quantum States: An Introduction to Quantum Entanglement (Cambridge Univer- sity Press, Cambridge, UK, 2006). [21] A. Goussev, R. A. Jalabert, H. M. Pastawski, and D. Ariel Wisniacki, Loschmidt echo, Scholarpedia 7, 11687 (2012). [22] C. Pineda and T. Prosen, Universal and nonuniversal level statistics in a chaotic quantum spin chain, Phys. Rev. E 76, 061127 (2007). [23] F. Haake, Quantum Signatures of Chaos (Springer-Verlag, New York, 2006). [24] I. García-Mata, C. Pineda, and D. A. Wisniacki, Quantum non- Markovian behavior at the chaos border, J. Phys. A 47, 115301 (2014). [25] M. Žnidarič, C. Pineda, and I. García-Mata, Non-Markovian Behavior of Small and Large Complex Quantum Systems, Phys. Rev. Lett. 107, 080404 (2011). [26] A. B. Klimov and S. M. Chumakov, A Group-Theoretical Ap- proach to Quantum Optics: Models of Atom-Field Interactions (Wiley-VCH, New York, 2009). [27] I. García-Mata, C. Pineda, and D. Wisniacki, Non-Markovian quantum dynamics and classical chaos, Phys. Rev. A 86, 022114 (2012). [28] T. A. Brody, J. Flores, J. B. French, P. A. Mello, A. Pandey, and S. S. M. Wong, Random-matrix physics: Spectrum and strength fluctuations, Rev. Mod. Phys. 53, 385 (1981). [29] T. A. Brody, A statistical measure for the repulsion of energy levels, Lett. Nuovo Cimento 7, 482 (1973). 062127-13 Bibliography [AHFB15] Christian Arenz, Robin Hillier, Martin Fraas, and Daniel Burgarth. Distinguishing decoherence from alternative quantum theories by dynamical decoupling. Physical Review A - Atomic, Molecular, and Optical Physics, 92(2):22102, 2015. [AKM14] Markus Aspelmeyer, Tobias J. Kippenberg, and Florian Marquardt. Cavity optomechanics. Rev. Mod. Phys., 86:1391–1452, Dec 2014. [ARHP14] Ángel Rivas, Susana F Huelga, and Martin B Plenio. Quantum non- markovianity: characterization, quantification and detection. Rep. Prog. Phys., 77(9):094001, 2014. [BP07] H.P. Breuer and F. Petruccione. The Theory of Open Quantum Sys- tems. OUP Oxford, 2007. [BvL05] Samuel L. Braunstein and Peter van Loock. Quantum information with continuous variables. Rev. Mod. Phys., 77:513–577, Jun 2005. [CDG19] Gustavo Montes Cabrera, David Davalos, and Thomas Gorin. Pos- itivity and complete positivity of differentiable quantum processes. Physics Letters A, 383(23):2719–2728, 2019. [Cho75] Man-Duen Choi. Completely positive linear maps on complex ma- trices. Linear Algebra and its Applications, 10(3):285 – 290, 1975. [CLP07] N J Cerf, G Leuchs, and E S Polzik. Quantum Information with Continuous Variables of Atoms and Light. Imperial College Press, 2007. [CTZ08] Hilary A Carteret, Daniel R Terno, and Karol Zyczkowski. Dynam- ics beyond completely positive maps: Some properties and applica- tions. Physical Review A - Atomic, Molecular, and Optical Physics, 77(4), 2008. 149 150 BIBLIOGRAPHY [Cul66] W. J Culver. On the Existence and Uniqueness of the Real Log- arithm of a Matrix. Proceedings of the American Mathematical Society, 17(5):1146–1151, 1966. [Den89] L. V. Denisov. Infinitely Divisible Markov Mappings in Quantum Probability Theory. Theory Prob. Appl., 33(2):392–395, 1989. [EL77] D. E. Evans and J. T. Lewis. Dilations of Irreversible Evolutions in Algebraic Quantum Theory, volume 24 of Communications of the Dublin Institute for Advanced Studies: Theoretical physics. Dublin Institute for Advanced Studies, 1977. [EW07] J. Eisert and M. M. Wolf. Gaussian Quantum Channels. In Quantum Information with Continuous Variables of Atoms and Light, pages 23–42. Imperial College Press, feb 2007. [Exn85] Pavel Exner. Open Quantum Systems and Feynman Integrals, vol- ume 36. Springer Netherlands, Dordrecht, 1985. [FPMZ17] S. N. Filippov, J. Piilo, S. Maniscalco, and M. Ziman. Divisibility of quantum dynamical maps and collision models. Phys. Rev. A, 96(3):032111, 2017. [GKL13] A. D. Greentree, J. Koch, and J. Larson. Fifty years of Jaynes–Cummings physics. J. Phy. B, 46(22):220201, 2013. [Gor76] V. Gorini. Completely positive dynamical semigroups of N-level systems. J. Math. Phys., 17(5):821, 1976. [GSI88] Hermann Grabert, Peter Schramm, and Gert-Ludwig Ingold. Quan- tum Brownian motion: The functional integral approach. Physics Reports, 168(3):115–207, oct 1988. [GTW09] Alexei Gilchrist, Daniel R. Terno, and Christopher Wood. Vector- ization of quantum operations and its use. arXiv, page 12, 2009. [GVAW+03] Frédéric Grosshans, Gilles Van Assche, Jérôme Wenger, Rosa Brouri, Nicolas J. Cerf, and Philippe Grangier. Quantum key distri- bution using gaussian-modulated coherent states. Nature, 421:238, Jan 2003. [HHHH09] R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki. Quantum entanglement. Rev. Mod. Phys., 81(2):865–942, 2009. [Hol01] Alexander S Holevo. Statistical Structure of Quantum Theory, vol- ume 67 of Lecture Notes in Physics Monographs. Springer Berlin Heidelberg, Berlin, Heidelberg, 2001. BIBLIOGRAPHY 151 [Hol07] A S Holevo. One-mode quantum Gaussian channels: Structure and quantum capacity. Problems of Information Transmission, 43(1):1– 11, mar 2007. [Hol08] A S Holevo. Entanglement-breaking channels in infinite dimen- sions. Problems of Information Transmission, 44(3):171–184, 2008. [HR06] S. Haroche and J.-M. Raimond. Exploring the Quantum: Atoms, Cavities, and Photons. Oxford University Press, USA, 2006. [HSP10] Klemens Hammerer, Anders S. Sørensen, and Eugene S. Polzik. Quantum interface between light and atomic ensembles. Rev. Mod. Phys., 82:1041–1093, Apr 2010. [HZ12] T. Heinosaari and M. Ziman. The Mathematical Language of Quan- tum Theory: From Uncertainty to Entanglement. Cambridge Uni- versity Press, 2012. [JC63] E. T. Jaynes and F. W. Cummings. Comparison of quantum and semiclassical radiation theories with application to the beam maser. Proc. IEEE, 51:89, 1963. [KBDW83] Karl Kraus, A. Böhm, J. D. Dollard, and W. H. Wootters. States Ef- fects Operators, volume 190 of Lecture Notes in Physics. Springer Berlin Heidelberg, Berlin, Heidelberg, 1983. [KC09] A. B. Klimov and S. M. Chumakov. A Group-Theoretical Approach to Quantum Optics: Models of Atom-Field Interactions. Wiley- VCH, 2009. [KG97] Robert Karrlein and Hermann Grabert. Exact time evolution and master equations for the damped harmonic oscillator. Physical Re- view E, 55(1):153–164, 1997. [Kos72a] A. Kossakowski. On necessary and sufficient conditions for a gen- erator of a quantum dynamical semigroup. Bull. Acad. Pol. Sci., 20(12):1021, 1972. [Kos72b] A. Kossakowski. On quantum statistical mechanics of non- hamiltonian systems. Rep. Math. Phys., 3(4):247 – 274, 1972. [LB99] Seth Lloyd and Samuel L. Braunstein. Quantum computation over continuous variables. Phys. Rev. Lett., 82:1784–1787, Feb 1999. [Lin76] G. Lindblad. On the generators of quantum dynamical semigroups. Comm. Math. Phys., 48(2):119–130, 1976. 152 BIBLIOGRAPHY [Lin00] Göran Lindblad. Cloning the quantum oscillator. Journal of Physics A: Mathematical and General, 33(28):5059–5076, 2000. [LRW+18] Ludovico Lami, Bartosz Regula, Xin Wang, Rosanna Nichols, An- dreas Winter, and Gerardo Adesso. Gaussian quantum resource the- ories. Phys. Rev. A, 98:022335, Aug 2018. [MP12] Esteban A Martinez and Juan Pablo Paz. Supplementary material for the paper ” Dynamics and thermodynamics for linear quantum open systems. Phys. Rev. Lett., 2012. [NC11] Michael A. Nielsen and Isaac L. Chuang. Quantum Computation and Quantum Information: 10th Anniversary Edition. Cambridge University Press, New York, NY, USA, 10th edition, 2011. [PGD+16] C. Pineda, T. Gorin, D. Davalos, D. A. Wisniacki, and I. Garcı́a- Mata. Measuring and using non-Markovianity. Phys. Rev. A, 93:022117, 2016. [Red65] A. G. Redfield. The Theory of Relaxation Processes. In Advances in Magnetic and Optical Resonance, volume 1, pages 1–32. Academic Press, jan 1965. [RFZB12] T. Rybár, S. N. Filippov, M. Ziman, and V. Bužek. Simula- tion of indivisible qubit channels in collision models. J. Phys. B, 45(15):154006, 2012. [RH12] Ángel Rivas and Susana F. Huelga. Open Quantum Systems. SpringerBriefs in Physics. Springer Berlin Heidelberg, Berlin, Hei- delberg, 2012. [RPZ18] Ł. Rudnicki, Z. Puchała, and K. Zyczkowski. Gauge invariant infor- mation concerning quantum channels. Quantum, 2:60, April 2018. [RSW02] M. B. Ruskai, S. Szarek, and E. Werner. An analysis of completely- positive trace-preserving maps on M2. Lin. Alg. Appl., 347(1):159 – 187, 2002. [Sti06] W. Forrest Stinespring. Positive Functions on C∗-Algebras. Pro- ceedings of the American Mathematical Society, 6(2):211, feb 2006. [Tun85] Wu-Ki. Tung. Group theory in physics. World Scientific, 1985. [VDD01] F. Verstraete, J. Dehaene, and B. DeMoor. Local filtering operations on two qubits. Phys. Rev. A, 64(1):010101, 2001. [VSL+11] B. Vacchini, A. Smirne, E.-M. Laine, J. Piilo, and H.-P. Breuer. BIBLIOGRAPHY 153 Markovianity and non-markovianity in quantum and classical sys- tems. New J. Phys., 13(9):093004, 2011. [VV02] F. Verstraete and H. Verschelde. On quantum channels. Unpub- lished, 2002. [WC08] M. M. Wolf and J. I. Cirac. Dividing quantum channels. Comm. Math. Phys., 279(1):147–168, 2008. [WECC08] M. M. Wolf, J. Eisert, T. S. Cubitt, and J. I. Cirac. Assessing non- Markovian quantum dynamics. Phys. Rev. Lett., 101(15):150402, 2008. [Wol11] Mm Wolf. Quantum channels & operations: Guided tour. Lecture notes available at http://www-m5. ma. tum. . . . , 2011. [WPGP+12] Christian Weedbrook, Stefano Pirandola, Raúl Garcı́a-Patrón, Nico- las J. Cerf, Timothy C. Ralph, Jeffrey H. Shapiro, and Seth Lloyd. Gaussian quantum information. Reviews of Modern Physics, 84(2):621–669, may 2012. [ZB05] M. Ziman and V. Bužek. Concurrence versus purity: Influence of local channels on Bell states of two qubits. Phys. Rev. A, 72(5):052325, 2005.