Fact-checked by Grok 2 weeks ago

Quantum operation

A quantum operation, also known as a , is a completely positive trace-preserving (CPTP) \mathcal{C}: \mathcal{L}(\mathcal{H}_i) \to \mathcal{L}(\mathcal{H}_o) that transforms density operators between input and output Hilbert spaces, ensuring the positivity of quantum states and the preservation of total probability under interactions with an environment or . These operations extend the unitary dynamics of closed to open systems, capturing phenomena such as decoherence, , and entanglement with external , while maintaining physical consistency through complete positivity (valid even when tensored with an identity map on an auxiliary system) and trace preservation (normalizing the output state). Formally introduced in the context of general state changes by Kraus in 1971, quantum operations admit equivalent representations, including the Kraus operator form—where \mathcal{C}(\rho) = \sum_k K_k \rho K_k^\dagger with operators satisfying \sum_k K_k^\dagger K_k = I—and the Stinespring dilation, which embeds the map into a larger followed by a over an environmental system. In , quantum operations are foundational for modeling realistic quantum processes in computation, communication, and sensing, enabling analyses of error correction, channel capacities, and non-Markovian dynamics, as well as higher-order extensions like superchannels that process channels themselves to study causal structures and quantum networks.

Introduction and Background

Historical Development

The concept of quantum operations traces its roots to the foundational work in during the early 20th century, particularly John 's introduction of density operators in his 1932 monograph Mathematische Grundlagen der Quantenmechanik. There, von Neumann formalized the statistical description of quantum systems using density matrices to represent mixed states and measurements, laying the groundwork for describing non-unitary evolutions beyond pure state projections. This framework addressed the need to handle ensembles and interactions with measurement apparatus, though it initially focused on closed systems. Advancements in operator algebras during the mid-20th century further developed the mathematical structure for quantum evolutions. In 1955, W. Forrest Stinespring established the dilation theorem, which characterizes completely positive maps as arising from unitary representations on larger Hilbert spaces, providing a rigorous basis for positive-preserving transformations in C*-algebras. This result, originally motivated by , became essential for modeling dissipative processes in quantum systems. Building on this, Karl Kraus introduced Kraus operators in 1971 to describe general state changes due to external interventions, offering an operator-sum representation for operations on density operators in open quantum systems. The 1970s saw parallel developments that solidified the theory. Man-Duen Choi's 1975 work on completely positive linear maps provided a concrete characterization via the matrix, linking positivity conditions to physical realizability in finite-dimensional matrix algebras. Concurrently, the Gorini-Kossakowski-Sudarshan-Lindblad (GKSL) equation, formulated independently by V. Gorini, A. Kossakowski, and E.C.G. Sudarshan in 1976 and by G. Lindblad in the same year, defined the generators of continuous-time quantum dynamical semigroups, ensuring complete positivity and preservation for Markovian evolutions. These contributions shifted focus toward open systems, where environmental interactions lead to decoherence and . In the 1980s and 1990s, the rise of quantum information theory integrated these mathematical tools into practical frameworks for non-unitary evolutions, emphasizing information processing and entanglement. Key figures like William K. Wootters advanced this by exploring quantum channels in contexts such as teleportation and no-cloning theorems, formalizing how operations preserve or degrade quantum correlations in noisy environments. This period marked the transition from abstract operator theory to applications in quantum computing and communication, with the trace-preserving condition ensuring probabilistic interpretations remained intact.

Motivations in Quantum Information

Quantum operations, also known as quantum channels, arise from the fundamental limitations of unitary evolution in describing real quantum systems, which are invariably open and interact with their environments. In closed systems, time evolution is unitary and reversible, preserving coherence and information perfectly, but this idealization fails when the system couples to an external bath, leading to decoherence and irreversible information loss as quantum superpositions entangle with environmental degrees of freedom, resulting in mixed states. Such interactions are ubiquitous in physical implementations, making unitary models insufficient for capturing the non-unitary dynamics observed in practice. The necessity for quantum operations becomes evident in quantum information applications, where open system descriptions are essential for modeling realistic processes in , , and sensing. In , environmental noise degrades gate fidelities and coherence times, necessitating frameworks to simulate and mitigate error-prone operations that deviate from ideal unitaries. Similarly, in quantum protocols like , channels account for decoherence-induced errors to ensure secure information transfer, while in quantum sensing, they model effects that limit sensitivity in devices like atomic clocks or magnetometers. Quantum operations provide a unified —completely positive trace-preserving maps—to encapsulate diverse phenomena such as , , and projective measurements under a single paradigm, enabling the design of fault-tolerant schemes and error correction codes. Unlike classical maps, which preserve positivity of probability distributions through mere positivity, quantum operations must satisfy complete positivity to ensure the positivity of matrices for arbitrary input states, a requirement stemming from and the structure of composite systems. This stricter condition prevents unphysical negative eigenvalues that could arise in quantum evolutions, distinguishing quantum from classical . For instance, in protocols, channels model the noisy transmission of entangled states, quantifying fidelity loss due to environmental coupling and guiding improvements in experimental realizations. Early quantum information experiments, such as those involving error-prone gates at institutions like and NIST, highlighted the need for such models to analyze decoherence in rudimentary quantum processors and communication setups.

Formal Definition

Completely Positive Maps

A linear map \Phi: \mathcal{B}(\mathcal{H}) \to \mathcal{B}(\mathcal{K}) between the C^*-algebras of bounded operators on finite-dimensional Hilbert spaces \mathcal{H} and \mathcal{K} is completely positive if it preserves the order structure under arbitrary tensor extensions with the identity map. Specifically, for every integer n \geq 1, the extended map \Phi \otimes \id_n: \mathcal{B}(\mathcal{H} \otimes \mathbb{C}^n) \to \mathcal{B}(\mathcal{K} \otimes \mathbb{C}^n) maps operators to operators. This condition ensures that \Phi maintains the physical interpretability of quantum states when the system interacts with an arbitrary ancillary system, preventing non-physical negative probabilities in subsystems. The Choi-Jamiolkowski isomorphism establishes a one-to-one correspondence between completely positive maps and positive semidefinite operators, facilitating their analysis. For a map \Phi with \dim \mathcal{H} = d < \infty, the associated Choi operator is given by \chi_\Phi = \sum_{i,j=1}^d |i\rangle\langle j| \otimes \Phi(|i\rangle\langle j|), where \{|i\rangle\}_{i=1}^d is an orthonormal basis of \mathcal{H}. The map \Phi is completely positive if and only if \chi_\Phi \geq 0. Equivalently, in the quantum information context, the Choi state is \rho_\Phi = \frac{1}{d} (\id \otimes \Phi)(|\Omega\rangle\langle\Omega|), where |\Omega\rangle = \sum_{i=1}^d |i\rangle|i\rangle is the unnormalized maximally entangled state; complete positivity holds if and only if \rho_\Phi \geq 0 (positive semidefinite). Mere positivity of \Phi—requiring only that \Phi maps positive semidefinite operators on \mathcal{B}(\mathcal{H}) to those on \mathcal{B}(\mathcal{K})—is insufficient for quantum operations, as it fails to guarantee positivity for composite systems involving entanglement. To illustrate, consider the action on an entangled input: if \Phi is positive but not completely positive, then \Phi \otimes \id_n applied to a maximally entangled state on \mathcal{H} \otimes \mathbb{C}^n (for suitable n) yields an operator with negative eigenvalues, which would imply invalid reduced density matrices with negative probabilities in a physical setup. A concrete counterexample is the transpose map \Phi(A) = A^T, which is positive since it preserves eigenvalues and thus maps Hermitian positive semidefinite matrices to the same. However, for d=2, the Choi operator of the transpose is \chi_T = \begin{pmatrix} 1 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 1 \end{pmatrix}, which has eigenvalues $1, 1, 1, -1 and is thus not positive semidefinite, confirming that the transpose is not completely positive. This distinction underscores why completely positive maps are essential for describing legitimate quantum evolutions, as positive maps alone can violate coherence in entangled scenarios.

Trace-Preserving Condition

A quantum operation \Phi acting on the space of density operators is trace-preserving if \operatorname{Tr}[\Phi(\rho)] = \operatorname{Tr}[\rho] for all density operators \rho, which ensures that the total probability associated with the quantum state remains conserved at 1 throughout the evolution. This condition is essential for describing physically realizable deterministic processes in quantum mechanics, where the map \Phi models the dynamics without introducing or destroying probability mass. Mathematically, the trace-preserving condition is equivalent to the dual map \Phi^\dagger (defined via the Hilbert-Schmidt inner product \langle A, B \rangle = \operatorname{Tr}[A^\dagger B]) satisfying \Phi^\dagger(I) = I, where I is the identity operator, meaning the dual preserves the maximally mixed state in the Heisenberg picture. Alternatively, for an orthonormal basis \{|i\rangle\} of the Hilbert space, the condition holds if \sum_k \langle i | E_k^\dagger E_k | i \rangle = 1 for each i, where \{E_k\} are operators associated with the map, reflecting conservation on basis projections. Without the trace-preserving condition, a completely positive map may describe subnormalized states, as in quantum measurements where outcomes yield probabilities less than 1, or non-physical amplifiers that increase trace beyond 1, rendering them unsuitable for modeling closed-system evolutions or quantum channels. In contrast, completely positive trace-preserving (CPTP) maps, which build on complete positivity by adding this requirement, represent the full class of valid quantum operations. CPTP maps are trace-preserving by definition but not necessarily unital, where unitality requires \Phi(I) = I; the latter fixes the maximally mixed state but is absent in maps involving directed dissipation, such as the amplitude damping channel that models qubit relaxation to the ground state via Kraus operators E_0 = \begin{pmatrix} 1 & 0 \\ 0 & \sqrt{1-\gamma} \end{pmatrix} and E_1 = \begin{pmatrix} 0 & \sqrt{\gamma} \\ 0 & 0 \end{pmatrix} for damping probability \gamma. This distinction highlights that trace preservation ensures probabilistic integrity, while unitality implies symmetry in noise affecting coherent superpositions. In the Choi-Jamiołkowski representation, where the map \Phi is associated with the Choi operator J(\Phi) = (\mathrm{id} \otimes \Phi)(|\Phi^+\rangle\langle\Phi^+|) and |\Phi^+\rangle = \sum_i |i\rangle|i\rangle is the unnormalized maximally entangled state, trace preservation holds if and only if the partial trace over the second (output) system satisfies \operatorname{Tr}_B[J(\Phi)] = I_A, confirming the map's validity without ancillary computations.

Kraus Representation

Theorem Statement

The Kraus representation theorem asserts that every completely positive trace-preserving (CPTP) map \Phi: \mathcal{B}(\mathcal{H}_\text{in}) \to \mathcal{B}(\mathcal{H}_\text{out}), where \mathcal{H}_\text{in} and \mathcal{H}_\text{out} are complex Hilbert spaces and \mathcal{B}(\mathcal{H}) denotes the algebra of bounded linear operators on \mathcal{H}, acting on density operators \rho can be expressed in the form \Phi(\rho) = \sum_k E_k \rho E_k^\dagger, where \{E_k\} is a set of bounded linear operators (known as Kraus operators) from \mathcal{H}_\text{in} to \mathcal{H}_\text{out} satisfying the completeness relation \sum_k E_k^\dagger E_k = I_\text{in}, with I_\text{in} the identity operator on \mathcal{H}_\text{in}. In the finite-dimensional case, where \dim(\mathcal{H}_\text{in}) = d < \infty and \dim(\mathcal{H}_\text{out}) = e < \infty, the theorem holds for maps between the spaces of trace-class operators on these Hilbert spaces, and there exists a Kraus representation with at most d e operators, as this bound corresponds to the maximum rank of the associated Choi matrix. This theorem was introduced by Karl Kraus in 1971, with a detailed treatment in his 1983 monograph, building on the earlier Stinespring dilation theorem from 1955, which provides a unitary representation of completely positive maps on larger auxiliary spaces. The theorem applies directly to trace-class operators in finite dimensions but extends to infinite-dimensional settings through approximations by finite-rank projections or Stinespring dilations on enlarged spaces. The minimal Kraus rank, defined as the smallest number of operators required in any such representation (equal to the rank of the Choi matrix), is at most d e and serves as a measure of the channel's structural complexity, influencing bounds on its information transmission capacities such as the quantum capacity.

Construction and Uniqueness

For the case \dim(\mathcal{H}_\text{in}) = \dim(\mathcal{H}_\text{out}) = d, the Kraus operators for a completely positive trace-preserving (CPTP) map \Phi can be constructed using its Choi matrix, which provides a systematic way to extract the operators from the map's representation. The Choi matrix J(\Phi) of \Phi acting on a d-dimensional Hilbert space is defined as J(\Phi) = \sum_{i,j=1}^d |i\rangle\langle j| \otimes \Phi(|i\rangle\langle j|), where \{|i\rangle\} is an orthonormal basis. To obtain the Kraus operators, perform the eigendecomposition of J(\Phi) = \sum_k \lambda_k |\psi_k\rangle\langle\psi_k|, where \lambda_k \geq 0 are the eigenvalues and |\psi_k\rangle are the corresponding eigenvectors. Each eigenvector |\psi_k\rangle is then reshaped (or "unvectorized") into a d \times d matrix V_k such that \mathrm{vec}(V_k) = |\psi_k\rangle, and the Kraus operators are given by E_k = \sqrt{\lambda_k} V_k for each k with \lambda_k > 0. This process ensures that \Phi(\rho) = \sum_k E_k \rho E_k^\dagger satisfies the CPTP conditions, as the positive eigenvalues and the structure of the Choi matrix guarantee complete positivity and trace preservation. In general, for \dim(\mathcal{H}_\text{out}) = e \neq d, the reshaping yields e \times d matrices V_k. The algorithmic steps for this construction are as follows: first, compute the Choi matrix J(\Phi) by applying \Phi to the basis elements |i\rangle\langle j|; second, diagonalize J(\Phi) to obtain the eigenvalues \{\lambda_k\} and eigenvectors \{|\psi_k\rangle\}; third, for each non-zero \lambda_k, reshape |\psi_k\rangle into the matrix V_k; and finally, scale by the square root to form E_k = \sqrt{\lambda_k} V_k. This method is efficient for numerical implementations and directly implements the Choi-Kraus theorem by yielding a valid Kraus representation from any CPTP map. An equivalent approach involves computing a square root of J(\Phi), say X such that J(\Phi) = X X^\dagger, and then taking the Kraus operators as the column matrices of X, though the eigendecomposition provides a canonical form. The Kraus representation is inherently non-unique: if \{E_k\} is a set of Kraus operators for \Phi, then another valid set is \{E'_l = \sum_k u_{lk} E_k\}, where U = (u_{lk}) is any satisfying the column \sum_l |u_{lk}|^2 = 1 for each k. This freedom arises because the operator-sum \sum_k E_k \rho E_k^\dagger remains invariant under such unitary mixtures, reflecting the multiplicity in the Stinespring underlying the representation. Different choices of Kraus operators can thus describe the same , with the unitary U parameterizing the of representations. A minimal Kraus representation, with the smallest number of operators, is achieved by selecting the Kraus equal to the of the of the Choi matrix, which is the number of positive eigenvalues in its eigendecomposition. This , known as the Choi , provides a lower bound on the number of Kraus operators needed and corresponds to the minimal of the in the Stinespring . Representations with more operators can always be reduced to this minimal form by absorbing redundancies via unitary transformations. As an illustrative example, consider the bit-flip channel, a simple noisy quantum operation that flips a qubit with probability p. Its Choi matrix can be computed explicitly, leading to Kraus operators E_0 = \sqrt{1-p} \, I and E_1 = \sqrt{p} \, \sigma_x, where I is the identity matrix and \sigma_x is the Pauli-X operator. This representation is minimal, with Kraus rank 2, and directly follows from the eigendecomposition of the Choi matrix parameterized by the error probability p. Non-uniqueness is evident, as one could mix these operators with a unitary matrix to obtain equivalent sets, such as phase-rotated versions that preserve the channel's action.

Physical Interpretations and Properties

Unitary Equivalence

Two sets of Kraus operators \{E_k\} and \{F_l\} represent the same quantum operation \Phi if there exists a V such that F_l = \sum_k V_{lk} E_k for each l, where the summation is over the Kraus index k and the matrix V acts on the index space. This relation ensures that the operator-sum decomposition \Phi(\rho) = \sum_k E_k \rho E_k^\dagger = \sum_l F_l \rho F_l^\dagger remains unchanged, as the unitarity of V preserves the trace-preserving completely positive structure. The unitary equivalence stems from the underlying Stinespring dilation of \Phi, where the operation is realized as a unitary on the coupled to an , followed by a over the . The Kraus operators arise by expanding the isometry in an of the ; redefining this basis via a unitary on the induces the mixing captured by V, without altering the physical . This freedom implies that, except for unitary channels (which admit a Kraus up to a ), every quantum operation possesses infinitely many Kraus representations, parameterized by the on the environment dimension. Despite this non-uniqueness, key invariants persist across equivalent representations. The Kraus rank, the smallest number of operators required in any representation, remains fixed and equals the rank of the Choi matrix (or equivalently, the Choi-Jamiolkowski state) of \Phi. Another invariant is the entanglement fidelity F_e(\Phi) = \langle \Phi \rangle (\Phi \otimes \mathcal{I})(|\Omega\rangle\langle \Omega|), where |\Omega\rangle is a maximally entangled state and \langle \Phi \rangle = \operatorname{Tr}_B [\cdot] is the partial trace; this quantity, determined by the largest eigenvalue of the Choi state, is unchanged under unitary mixing of Kraus operators. Geometrically, the Kraus operators can be interpreted as components of a vector in the Hilbert-Schmidt space of operators on the system Hilbert space, with the completeness relation \sum_k \mathrm{vec}(E_k^\dagger E_k) = \mathrm{vec}(I) constraining the configuration. Unitary equivalence then acts as an orthogonal transformation in this higher-dimensional operator space, rotating the "frame" of the Kraus vectors while preserving the overall geometry of the channel, such as the shape of the image of the Bloch ball under \Phi. A representative example is the qubit depolarizing channel, which randomly applies one of the Pauli errors with equal probability or leaves the state unchanged. A canonical Kraus set is \{M_0 = \sqrt{1-p}\, I, \, M_1 = \sqrt{p/3}\, \sigma_x, \, M_2 = \sqrt{p/3}\, \sigma_y, \, M_3 = \sqrt{p/3}\, \sigma_z\}, where $0 \leq p \leq 1 parameterizes the noise strength and \{\sigma_i\} are the . Equivalent sets arise by applying a $3 \times 3 to mix \{M_1, M_2, M_3\}, reflecting the channel's SU(2) rotational invariance in the error subspace while fixing M_0; for instance, a in the Pauli basis yields a new set that implements the same isotropic decoherence. This Kraus rank of 4 is invariant, as the channel's Choi matrix has full rank for p > 0.

Physical Realizability

The Stinespring dilation theorem provides the foundational criterion for the physical realizability of quantum operations: a \Phi on the space of operators is completely positive and trace-preserving (CPTP) if and only if it can be realized by a unitary on the in with an ancillary , followed by a over the environment. This representation ensures that the map preserves the positivity and trace of operators, aligning with the requirements of for describing evolutions of open s. In explicit form, the dilation is given by \Phi(\rho) = \Tr_E \left[ U (\rho \otimes |0\rangle\langle 0|_E) U^\dagger \right], where U acts unitarily on the composite of the system and , and |0\rangle_E denotes a pure initial of the . The Kraus operators \{E_k\} associated with \Phi arise naturally from this as E_k = \langle k|_E U |0\rangle_E, with \{|k\rangle_E\} forming an for the . For a minimal exact , the 's dimension must be at least the Kraus of \Phi, defined as the smallest number of operators needed in any Kraus representation of the map. This framework implies that all CPTP maps are physically implementable in principle, as the dilation corresponds to a closed-system unitary that is always achievable under . In contrast, maps that are not completely positive can yield output operators with negative eigenvalues, leading to unphysical negative probabilities. However, fundamental bounds exist: for instance, the prohibits the existence of a CPTP map that perfectly clones an arbitrary unknown , though approximate partial cloners that are CPTP can be constructed.

Applications in Quantum Dynamics

Open System Evolution

In open quantum systems, the time evolution of the density operator \rho(t) is described by quantum operations that account for interactions with an uncontrollable , leading to and decoherence. Unlike closed systems governed by unitary evolution, open system dynamics are modeled by completely positive trace-preserving (CPTP) maps that form a dynamical under the Markovian . This framework ensures that the evolution preserves the positivity of the density operator and its , capturing irreversible processes while maintaining quantum where possible. The standard continuous-time description is given by the Gorini-Kossakowski-Sudarshan-Lindblad (GKSL) master equation, which generates a CPTP : \frac{d\rho}{dt} = -i [H, \rho] + \sum_k \left( L_k \rho L_k^\dagger - \frac{1}{2} \{ L_k^\dagger L_k, \rho \} \right), where H is the system and the L_k are Lindblad operators encoding environmental effects. This form was independently derived by Gorini, Kossakowski, Sudarshan, and Lindblad, establishing the most general Markovian dynamics consistent with . The dissipative term ensures complete positivity, preventing non-physical negative probabilities, and the anticommutator maintains trace preservation. The GKSL equation arises under the Markovian approximation, which assumes weak system-environment coupling and a memoryless bath, allowing a time-local that neglects correlations building up over time. This approximation is valid when the environment relaxation time is much shorter than the system's evolution timescale, leading to of coherences and approach to . For discrete-time evolution, the continuous dynamics can be approximated by composing successive CPTP maps \Phi_n \circ \cdots \circ \Phi_1, where each \Phi_j represents an infinitesimal time step, often expressed in Kraus form for computational purposes. This stepwise approach is useful in numerical simulations or when analyzing stroboscopic dynamics in periodically driven systems. A representative example is the damped , modeling photon loss in a , with H = \omega a^\dagger a and Lindblad operator L = \sqrt{\gamma} a at zero temperature, where \gamma is the decay rate. The master equation yields exponential damping of the mean number \langle a^\dagger a \rangle(t) = \langle a^\dagger a \rangle(0) e^{-\gamma t} and decoherence of off-diagonal elements in the number basis at rate \gamma/2. Quantum operations in open systems quantify to the , with Lindblad operators determining decoherence rates that suppress superpositions and drive the toward classical mixtures. The decoherence timescale is inversely proportional to the strength of the L_k, as seen in the decay of coherences \rho_{ij}(t) \approx \rho_{ij}(0) e^{-\Gamma t} for rate \Gamma derived from the dissipator. Additionally, these dynamics produce , with the rate \dot{S} = -\operatorname{Tr} [\mathcal{L}(\rho) \ln \rho] (where \mathcal{L} is the Liouvillian) being non-negative and quantifying irreversibility, vanishing only at steady states. This measure, introduced by Spohn, links thermodynamic dissipation to the semigroup's approach to .

Quantum Channels

In quantum information theory, a is defined as a completely positive trace-preserving (CPTP) map \Phi: \mathcal{B}(\mathcal{H}_A) \to \mathcal{B}(\mathcal{H}_B) that models the transmission of quantum states from an input \mathcal{H}_A to an output \mathcal{H}_B, accounting for noise and decoherence inherent in physical processes. This formalism captures the most general form of quantum evolution without measurement, ensuring that probabilities remain normalized and positivity is preserved even under tensorization with identity maps. A prototypical example is the random unitary channel, where \Phi(\rho) = \sum_i p_i U_i \rho U_i^\dagger, with \{p_i\} a probability distribution over unitary operators \{U_i\}, representing stochastic unitary dynamics such as those induced by fluctuating control fields. Standard noise models classify common quantum channels based on their physical origins. The amplitude damping channel describes energy relaxation in a qubit system coupled to a thermal bath, with Kraus operators E_0 = \begin{pmatrix} 1 & 0 \\ 0 & \sqrt{1-\gamma} \end{pmatrix} and E_1 = \begin{pmatrix} 0 & \sqrt{\gamma} \\ 0 & 0 \end{pmatrix}, where \gamma is the damping probability; it maps the toward the while preserving the . The phase damping channel models pure , where off-diagonal elements decay due to , with Kraus operators E_0 = \begin{pmatrix} 1 & 0 \\ 0 & \sqrt{1-\lambda} \end{pmatrix} and E_1 = \begin{pmatrix} 0 & 0 \\ 0 & \sqrt{\lambda} \end{pmatrix}, \lambda being the dephasing strength; it leaves populations unchanged but erodes coherences. The depolarizing channel represents full , transforming any input state \rho to \Phi(\rho) = p \rho + (1-p) \frac{I}{d} (for d), effectively replacing the state with the maximally mixed one with probability $1-p; this arises in isotropic environments. Channel capacities quantify the information transmission limits of quantum channels. The classical capacity, given by the Holevo quantity \chi(\Phi) = \max_{\{p_i, \rho_i\}} \sum_i p_i S(\Phi(\rho_i)) - S(\Phi(\sum_i p_i \rho_i)), where S denotes the and the maximum is over ensembles of input states, represents the supremum of reliable classical bit transmission rates without entanglement. The entanglement-assisted quantum capacity, which allows pre-shared entanglement between sender and receiver to enhance transfer, is Q_E(\Phi) = \frac{1}{2} \max_\rho I(A;B)_\sigma, where I(A;B)_\sigma = S(\rho_A) + S(\rho_B) - S(\sigma_{AB}), \sigma = (id \otimes \Phi)(|\psi\rangle\langle\psi|_{AR}) for a purification |\psi\rangle of input \rho_A, and the maximum is over input states \rho; this equals half the entanglement-assisted classical capacity. Quantum can be composed to model complex systems. The serial concatenation \Phi \circ \Psi applies channel \Psi first, followed by \Phi, describing sequential noisy processes like multi-stage transmission. Parallel composition via the \Phi \otimes \Psi acts independently on separate systems, enabling multi-party or multi-mode communication protocols. In , bosonic Gaussian channels provide a key example, acting on infinite-dimensional Hilbert spaces of bosonic modes and preserving Gaussian states under affine transformations of quadrature operators; they encompass lossy, additive noise, and squeezing channels, with capacities computable via Williamson and playing a central role in optical quantum communication.

Quantum Measurements

Projective Measurements

Projective measurements, also known as measurements, represent an ideal class of quantum measurements where the measurement apparatus interacts with the system in a way that projects the onto one of a set of orthogonal subspaces. These measurements are characterized by a complete set of orthogonal projectors \{P_m\}, satisfying \sum_m P_m = I and P_m P_n = \delta_{mn} P_m, where I is the identity operator and \delta_{mn} is the . Upon obtaining outcome m, the post-measurement state of the system is given by \rho' = \frac{P_m \rho P_m}{\operatorname{Tr}[P_m \rho]}, where \rho is the pre-measurement density operator, assuming \operatorname{Tr}[P_m \rho] > 0. This projection collapses the state onto the eigenspace corresponding to P_m, enforcing between different outcomes. In the framework of quantum operations, a projective can be modeled as a completely positive trace-preserving (CPTP) map when the classical measurement outcome is discarded. Specifically, the overall evolution, including the outcome register, forms a CPTP map on the enlarged , but tracing over the outcome yields the decohering \Phi(\rho) = \sum_m P_m \rho P_m. Here, the Kraus operators are the projectors themselves, K_m = P_m, satisfying the \sum_m K_m^\dagger K_m = \sum_m P_m = I. The probability of outcome m is p_m = \operatorname{Tr}[P_m \rho], which integrates the into the operational description. A canonical example is the of along the z-axis, where the projectors are P_+ = |0\rangle\langle 0| and P_- = |1\rangle\langle 1|, with |0\rangle and |1\rangle denoting the eigenstates of \sigma_z. For an initial state \rho = |\psi\rangle\langle\psi| with |\psi\rangle = \alpha |0\rangle + \beta |1\rangle, the probabilities are p_+ = |\alpha|^2 and p_- = |\beta|^2, and the channel becomes \Phi(\rho) = p_+ |0\rangle\langle 0| + p_- |1\rangle\langle 1|, fully diagonalizing \rho in the measurement basis. Projective operations inherently lead to the loss of quantum coherence, as they eliminate off-diagonal elements of \rho in the basis defined by the projectors, effectively destroying superpositions aligned with the measured . This decoherence arises because each P_m \rho P_m term projects onto a , and the over outcomes averages without preserving between subspaces.

Generalized Measurements

Generalized measurements in , unlike the projective measurements discussed previously, are formalized using positive operator-valued measures (), which allow for a wider range of physically realizable observation processes. A is defined as a collection of positive semi-definite operators \{E_m\}_{m=1}^M on the \mathcal{H}, satisfying the normalization condition \sum_{m=1}^M E_m = I, where I is the identity operator. For a \rho, the probability of obtaining outcome m is given by p_m = \operatorname{Tr}(E_m \rho). The corresponding post-measurement state, assuming the standard Lüders rule for the collapse, is \rho_m = \frac{\sqrt{E_m} \rho \sqrt{E_m}}{p_m}, which preserves the trace and positivity of the density operator. As quantum operations, generalized measurements are more completely described by quantum instruments, which specify the evolution of the state conditional on each outcome. An instrument \{\Phi_m\} consists of completely positive trace-non-increasing maps \Phi_m: \mathcal{B}(\mathcal{H}) \to \mathcal{B}(\mathcal{H}) such that \sum_m \Phi_m(\rho) = \rho for any input state \rho, with the POVM elements given by E_m = \sum_j K_{m j}^\dagger K_{m j}, where \{\Phi_m(\rho) = \sum_j K_{m j} \rho K_{m j}^\dagger\} is the Kraus representation for each branch. This framework captures the full dynamics, including any decoherence or information gain beyond mere probability assignment. Naimark's dilation theorem establishes that any on \mathcal{H} can be realized as a projective on an enlarged \mathcal{H} \otimes \mathcal{K}, via an V: \mathcal{H} \to \mathcal{H} \otimes \mathcal{K} such that the projectors \Pi_m = V E_m V^\dagger form a resolution of the identity on the extended space. This embedding demonstrates the physical implementability of POVMs using standard projective techniques, often by coupling the system to an ancillary system and performing a joint projection. A representative example is the trine POVM for phase estimation on a , which optimally discriminates three symmetric states on the equator of the . The elements are E_m = \frac{2}{3} |\psi_m\rangle\langle\psi_m| for m = 0,1,2, where |\psi_m\rangle = \frac{1}{\sqrt{2}} (|0\rangle + e^{i 2\pi m /3} |1\rangle), and the sum \sum_m E_m = I holds due to the equiangular overlap structure that cancels off-diagonal contributions. One key advantage of POVMs over projective measurements is their ability to implement weak measurements, where the interaction extracts partial information with minimal disturbance to the , enabling repeated or sequential observations without full collapse.

Limitations and Extensions

Non-Completely Positive Maps

Non-completely positive maps, also known as positive but not completely positive maps, are linear transformations on the space of operators that preserve the positivity of matrices—mapping operators to operators—but fail to do so when extended to composite systems via tensoring with the identity map on an ancillary . This distinguishes them from completely positive maps, which maintain positivity under such extensions and are essential for describing physical quantum evolutions. The primary issue with non-completely positive maps arises when applied to entangled states: the extended map (id ⊗ Φ) can produce operators with negative eigenvalues, leading to non-physical outcomes such as negative probabilities in quantum measurements. For instance, if an input state is entangled across the system and ancilla, the output may no longer represent a valid operator, violating the requirement that physical operations yield interpretable probabilities. Prominent examples include the transposition map in dimensions d \geq 3, defined by \Phi(\rho) = \rho^T, which preserves positivity because the transpose shares the same eigenvalues as the original but is not completely positive, as demonstrated by its action on the Choi matrix corresponding to a maximally entangled state. Another example is the reduction map applied to bipartite states via (\mathrm{id}_A \otimes \Gamma_B)(\rho_{AB}), where \Gamma(X) = \frac{\Tr_B(X)}{d_B} I_{d_B} - X, which is positive on separable states but produces negative eigenvalues for certain entangled states. In higher dimensions, the generalized reduction map \Phi(X) = \frac{\operatorname{Tr}(X)}{d} I_d - X similarly exhibits positivity without complete positivity and serves to detect entanglement through the negativity of its output. These maps cannot model physical quantum processes, as they undermine the consistency of in the presence of entanglement, but they find valuable application as entanglement witnesses: a bipartite state \rho is entangled if (\mathrm{id} \otimes \Phi)(\rho) has negative eigenvalues for some positive map \Phi that is not completely positive. The partial transpose map, \Phi(\rho_{AB}) = \rho^{T_B}_{AB}, exemplifies this utility, where negativity in the output spectrum indicates entanglement in systems with d \geq 3. Historically, early formulations of open quantum system dynamics in quantum optics during the 1960s often relied on positive maps without ensuring complete positivity, resulting in models that predicted unphysical behaviors for correlated systems; this confusion was resolved with the establishment of the complete positivity criterion in seminal works from the early 1970s.

Connections to Quantum Error Correction

Quantum operations play a central role in quantum error correction by modeling the noisy processes that degrade quantum information and enabling the design of recovery procedures that restore it. Error models in quantum systems are typically represented as completely positive trace-preserving (CPTP) maps, known as quantum channels, which capture the evolution of density operators under decoherence and noise. For instance, Pauli errors, such as bit-flip (X) and phase-flip (Z) operations, can be described as probabilistic mixtures forming CPTP maps; the bit-phase flip channel applies X with probability p, Z with probability p, both with probability p, and identity with probability $1-3p, ensuring the map is trace-preserving and completely positive. Correction of these errors is achieved through a recovery R, also a CPTP map, designed such that the R \circ \Phi approximates the identity map on the of encoded logical states, where \Phi denotes the error . This recovery process involves syndrome extraction to identify the error without disturbing the encoded information, followed by a corrective tailored to the detected . The efficacy of such corrections relies on the structure of quantum error-correcting codes, which encode logical qubits into a larger physical to protect against errors. A foundational result for correctability is provided by the Knill-Laflamme conditions, which specify that a can correct a set of errors \{E_a\} if, for logical states |i_L\rangle and |j_L\rangle in the code subspace, the inner products satisfy \langle i_L | E_a^\dagger E_b | j_L \rangle = \delta_{ij} c_{ab} for some c_{ab}, ensuring errors act uniformly across the code without mixing logical states. These conditions guarantee the existence of a recovery map that perfectly corrects the errors on the code subspace. An illustrative example is the 9-qubit Shor code, which encodes one logical into nine physical qubits and corrects any single-qubit Pauli error through a two-stage process: first correcting bit-flip errors using three-qubit repetition codes, then phase-flip errors via measurements implemented as quantum operations on ancillary qubits. The measurements project onto error subspaces, allowing the application of corrective Pauli operators to restore the logical state. In fault-tolerant quantum computing, the composition of multiple noisy quantum operations can still approximate ideal fault-free computation if the underlying error rate per operation remains below a threshold value, estimated in the 1990s to be around $10^{-2} or lower depending on the noise model and gate set. This underpins scalable quantum computation by showing that correction can suppress accumulation exponentially with distance.

References

  1. [1]
  2. [2]
  3. [3]
  4. [4]
    Lecture Notes on the Theory of Open Quantum Systems - arXiv
    Feb 3, 2019 · This is a self-contained set of lecture notes covering various aspects of the theory of open quantum system, at a level appropriate for a one-semester graduate ...
  5. [5]
  6. [6]
    [PDF] quantum-computation-and-quantum-information-nielsen-chuang.pdf
    This comprehensive textbook describes such remarkable effects as fast quantum algorithms, quantum teleportation, quantum cryptography, and quantum error- ...
  7. [7]
    [PDF] February 9, 2018 2.1 Overview 2.2 Three Definitions of Quantum ...
    Form 2 is called the Kraus. Decomposition, where Ve are the Kraus Operators of the channel and V is still an isometry. Page 3. Lecture 2: February 9, 2018. 2-3.
  8. [8]
    [PDF] Lecture Notes for Ph219/CS219: Quantum Information Chapter 3
    amplitude-damping channel these two times are related and comparable: T2 ... Heisenberg-picture time evolution is unital rather than trace preserving; the ...
  9. [9]
    States, Effects, and Operations - SpringerLink
    Free delivery 14-day returnsBook Title · States, Effects, and Operations ; Book Subtitle · Fundamental Notions of Quantum Theory ; Editors · Karl Kraus, A. Böhm, J. D. Dollard, W. H. Wootters.
  10. [10]
    [PDF] 22.51 Course Notes, Chapter 8: Open Quantum Systems
    The Kraus representation theorem reconciles these two description, by stating that they are equivalent.
  11. [11]
    [PDF] Quantum Channels, Kraus Operators, POVMs
    The term “completely positive trace preserving map” is often used in the literature. ◦ This is approximately the same terminology used in QCQI where, however, b ...
  12. [12]
    [PDF] Chapter 4: Quantum channels
    There is an abstract, mathematically minded approach to the question, introducing notions of complete positivity. Contrasting this, one can think of putting ...
  13. [13]
    [PDF] arXiv:quant-ph/0605009v1 30 Apr 2006
    Stinespring's dilation theorem is the basic structure theorem for quantum channels: it states that any quantum channel arises from a unitary evolution on a ...
  14. [14]
    [PDF] Approximating quantum channels by completely positive maps with ...
    The Kraus rank of a quantum channel can thus legitimately be seen as a measure of its “complexity”: it quantifies the minimal amount of ancillary resources ...
  15. [15]
    [PDF] arXiv:quant-ph/0104027v1 5 Apr 2001
    The minimal Stinespring representation is thus unique up to a unitary transformation. c. Kraus operators: In most of the current litera- ture the Stinespring ...
  16. [16]
    [PDF] arXiv:quant-ph/0202124v2 22 Jan 2003
    A unique Kraus representation can be obtained by for example enforcing the Kraus operators to be orthogonal, as these would correspond to the unique ...
  17. [17]
    Quantum resource theories | Rev. Mod. Phys.
    Apr 4, 2019 · As described in Sec. II , the Stinespring dilation theorem ensures that every CPTP map on system A can be implemented by applying a unitary ...
  18. [18]
    Completely positive dynamical semigroups of N‐level systems
    May 1, 1976 · We establish the general form of the generator of a completely positive dynamical semigroup of an N‐level quantum system.
  19. [19]
    On the generators of quantum dynamical semigroups
    The notion of a quantum dynamical semigroup is defined using the concept of a completely positive map. An explicit form of a bounded generator of such a se.
  20. [20]
    Density matrix for the damped harmonic oscillator ... - AIP Publishing
    Sep 1, 1993 · The time evolution of the density matrix of the damped harmonic oscillator is studied within the Lindblad theory for open quantum systems.
  21. [21]
    Entropy production for quantum dynamical semigroups
    May 1, 1978 · We prove that the entropy production is convex and positive and that the entropy production is a measure of dissipativity of the semigroup.
  22. [22]
    [PDF] Entanglement cost of generalised measurements - Rinton Press
    The trine measurement is defined as a POVM on , where. = 2 3. ¼ = 1 3 for = (4). The three directions are equally spaced around a great circle on the Bloch ...
  23. [23]
    [PDF] Who's afraid of not completely positive maps? - UT Physics
    Positive as well as not positive maps are good candidates for describing open quantum evolution. ... For instance, for quantum information processors one.
  24. [24]
    Scheme for reducing decoherence in quantum computer memory
    The scheme reduces decoherence in quantum memory using a quantum analog of error-correcting codes, assuming decoherence acts independently on each bit.
  25. [25]
    [quant-ph/9604034] A Theory of Quantum Error-Correcting Codes
    Apr 26, 1996 · We develop a general theory of quantum error correction based on encoding states into larger Hilbert spaces subject to known interactions.