Fact-checked by Grok 2 weeks ago

Completely positive map

A completely positive map is a linear map \phi: \mathcal{A} \to \mathcal{B} between C*-algebras \mathcal{A} and \mathcal{B} such that the amplified map \phi^{(n)}: M_n(\mathcal{A}) \to M_n(\mathcal{B}), defined by applying \phi entrywise to n \times n matrices over \mathcal{A}, preserves positivity for every positive integer n, meaning \phi^{(n)}(X) \geq 0 whenever X \geq 0. This condition strengthens the notion of a positive map, which only requires \phi(A) \geq 0 for positive elements A \in \mathcal{A}, ensuring that completely positive maps maintain the positivity of operators even when embedded into larger matrix systems. Completely positive maps play a central role in operator algebra theory and quantum information science, where they characterize physically realizable transformations of quantum states. Introduced in the foundational work of W. Forrest Stinespring in 1955, these maps admit a dilation to *-homomorphisms via Stinespring's representation theorem, which states that for a completely positive map \phi: \mathcal{A} \to B(\mathcal{H}), there exists a Hilbert space \mathcal{K}, a *-representation \pi: \mathcal{A} \to B(\mathcal{K}), and an isometry V: \mathcal{H} \to \mathcal{K} such that \phi(a) = V^* \pi(a) V for all a \in \mathcal{A}. In finite-dimensional quantum mechanics, trace-preserving completely positive maps, known as quantum channels or CPTP maps, describe the evolution of density operators under open system dynamics and can be expressed in Kraus form as \Phi(\rho) = \sum_i K_i \rho K_i^\dagger, where \{K_i\} are Kraus operators satisfying \sum_i K_i^\dagger K_i = I. Key properties include complete boundedness, with the completely bounded norm equaling the for unital maps, and closure under and adjoints. Examples encompass *-homomorphisms, which are completely positive, and compressions like \phi(a) = P a P for a P, as well as conditional expectations onto subalgebras. In quantum information, completely positive maps underpin entanglement theory and distinguish valid quantum operations from non-physical ones, such as the partial transpose on entangled states, which is positive but not completely positive. Further developments, including Choi's matrix characterization and Arveson's extension theorem, extend their applicability to subalgebras and provide tools for classification and entropy inequalities in operator systems.

Fundamentals

Definition

In the theory of operator algebras, a C*-algebra is a complex Banach algebra equipped with an involution operation satisfying the C*-identity \|a^* a\| = \|a\|^2 for all elements a. The prototypical example is B(\mathcal{H}), the C*-algebra of all bounded linear operators on a complex Hilbert space \mathcal{H}, where the norm is the operator norm and the involution is the adjoint operation. A linear map \Phi between C*-algebras is called positive if it maps the cone of positive elements—self-adjoint elements with non-negative spectrum—to positive elements. Such maps preserve the order structure induced by the positive cone but may fail to do so when extended to matrix-valued inputs. A linear map \Phi: \mathcal{A} \to \mathcal{B} between C*-algebras \mathcal{A} and \mathcal{B} is completely positive if for every positive integer n \in \mathbb{N}, the amplified map \Phi^{(n)}: M_n(\mathcal{A}) \to M_n(\mathcal{B}), which applies \Phi entrywise to n \times n matrices over \mathcal{A}, is positive. Here, M_n(\mathcal{A}) denotes the C*-algebra of n \times n matrices with entries in \mathcal{A} (equivalently, \mathcal{A} \otimes_{\min} M_n(\mathbb{C})), and the tensor product is the minimal C*-tensor product. Positivity of the amplified map requires that it sends positive elements to positive elements, ensuring preservation of the positive cone even under tensoring with arbitrary finite-dimensional matrix systems. The concept of complete positivity strengthens the notion of positivity to capture maps that remain positivity-preserving in all finite-dimensional extensions, playing a foundational role in the representation theory of C*-algebras. It was introduced by W. F. Stinespring in 1955, in the context of studying linear functions on C*-algebras that arise from representations.

Distinction from Positive Maps

A linear map \Phi: \mathcal{B}(H) \to \mathcal{B}(K) between the algebras of bounded operators on Hilbert spaces H and K is called positive if it maps positive semidefinite operators to positive semidefinite operators, that is, if A \geq 0 implies \Phi(A) \geq 0. The key distinction between positive and completely positive maps lies in their under tensor extensions. Positivity requires only that the map preserves positivity when applied directly to operators on the original (corresponding to n=1 in the definition of complete positivity), whereas complete positivity demands preservation of positivity for all extensions \Phi \otimes \mathrm{id}_{M_n} acting on tensor products with arbitrary finite-dimensional matrix algebras M_n. This stronger condition for higher n is because it accounts for scenarios involving "entanglement-like" structures: when a subsystem interacts with an ancillary system, the input state may be entangled, and a merely positive map might fail to yield a valid positive output on the extended , whereas a completely positive map ensures physical validity regardless of such correlations. Choi's theorem provides a precise in the finite-dimensional case: a \Phi: M_m(\mathbb{C}) \to M_n(\mathbb{C}) between matrix algebras is completely positive if and only if its associated Choi matrix, defined as \sum_{i,j=1}^m |i\rangle\langle j| \otimes \Phi(|i\rangle\langle j|), is . This distinction has significant implications in quantum theory, as positive maps that are not completely positive exist—for instance, the transposition map on matrices—and such maps cannot represent physically realizable quantum operations, as they may violate positivity when subsystems are entangled with auxiliaries.

Characterizations

Kraus Operator Representation

The Kraus operator theorem provides a concrete operator-sum decomposition for completely positive maps. Specifically, every completely positive map \Phi: \mathcal{B}(\mathcal{H}) \to \mathcal{B}(\mathcal{K}), where \mathcal{H} and \mathcal{K} are complex Hilbert spaces, admits a representation of the form \Phi(A) = \sum_{k} V_k A V_k^*, with \{V_k\} a (possibly countable) family of bounded linear operators from \mathcal{H} to \mathcal{K}. This decomposition exists if and only if \Phi is completely positive, and the sum converges in the strong operator topology. For trace-preserving maps, the operators satisfy the completeness relation \sum_k V_k^* V_k = I_{\mathcal{H}}, ensuring \operatorname{Tr}(\Phi(A)) = \operatorname{Tr}(A) for all A \in \mathcal{B}(\mathcal{H}). Similarly, for unital maps, \sum_k V_k V_k^* = I_{\mathcal{K}}, preserving the identity operator. In the general non-unital, non-trace-preserving case, no such global relation holds, though the representation still characterizes the map's complete positivity. The Kraus representation, introduced by K. Kraus in 1971, is fundamental in quantum information theory. The theorem's proof begins from complete positivity and leverages the Choi-Jamiołkowski isomorphism to construct the Kraus operators explicitly. Consider the Choi operator associated with \Phi, obtained by applying the map to one half of a maximally entangled state on \mathcal{H} \otimes \mathcal{H}. Since \Phi is completely positive, this Choi operator is . Diagonalizing it as C_\Phi = \sum_m \lambda_m |\psi_m\rangle\langle\psi_m| with \lambda_m \geq 0, the Kraus operators V_m are derived by reshaping the eigenvectors |\psi_m\rangle \in \mathcal{K} \otimes \mathcal{H} into operators via : specifically, the matrix elements are given by \langle k | V_m | j \rangle = \sqrt{\lambda_m} \langle k j | \psi_m \rangle for bases \{|k\rangle\} in \mathcal{K} and \{|j\rangle\} in \mathcal{H}. This approach formalizes the superoperator form, where \operatorname{vec}(\Phi(A)) = \sum_k (V_k \otimes \overline{V_k}) \operatorname{vec}(A), confirming the operator-sum structure. The minimal number of Kraus operators required in any such representation equals the Choi rank of \Phi, defined as the rank of the Choi operator C_\Phi. This rank provides a measure of the map's "complexity" and is invariant under the choice of representation. Representations are not unique; any two sets \{V_k\} and \{W_l\} yielding the same \Phi are related by a unitary mixing, W_l = \sum_k u_{lk} V_k, where u is a unitary matrix (or more generally, an isometry for finite-rank cases). This equivalence underscores the theorem's utility in computational quantum information, where finite-dimensional approximations suffice for practical implementations.

Stinespring Dilation

Stinespring's dilation theorem provides a fundamental representation for completely positive maps in the context of operator algebras. Specifically, for a completely positive map \Phi: \mathcal{B}(H) \to \mathcal{B}(K) between the algebras of bounded operators on Hilbert spaces H and K, there exists a Hilbert space M, a *-representation \pi: \mathcal{B}(H) \to \mathcal{B}(M), and a bounded linear operator W: K \to M such that \Phi(A) = W^* \pi(A) W for all A \in \mathcal{B}(H). If \Phi is contractive (i.e., \|\Phi\| \leq 1), then W can be chosen to be an isometry. This representation embeds the map into a larger algebra via an isometric extension, highlighting its structural properties. The theorem, introduced by W. Forrest Stinespring in 1955, shows that completely positive maps are compressions of *-homomorphisms. The proof of the theorem relies on a construction analogous to the Gelfand-Naimark-Segal (GNS) representation for states, generalized to completely positive maps. One forms a pre-Hilbert space from the algebraic tensor product \mathcal{B}(H) \odot K with the inner product \langle A \xi, B \eta \rangle = \langle \Phi(A^* B) \eta, \xi \rangle, completes it to a Hilbert space M, and defines the representation \pi(A) (B \eta) = A B \eta on the dense subspace, extending by continuity. The operator W is defined by W \xi = 1 \cdot \xi (using the unit), ensuring the dilation formula holds. This dilation is minimal when the representation \pi is non-degenerate, meaning the subspace generated by \pi(\mathcal{B}(H)) acting on the range of W is dense in M. In quantum information theory, the Stinespring dilation corresponds to a purification of the map's action on density operators: applying \Phi to a state \rho is equivalent to evolving a purified state on the extended system H \otimes L' (where L' represents an environment) unitarily and then partially tracing out the environment, preserving the complete positivity and trace preservation when \Phi is a quantum channel. For non-contractive completely positive maps with \|\Phi\| = c > 1, the bounded operator W satisfies \|W\|^2 = c, without needing separate normalization; the general form accommodates the norm directly.

Choi-Jamiołkowski Isomorphism

The Choi–Jamiołkowski isomorphism establishes a bijective correspondence between linear maps \Phi: \mathcal{M}_d(\mathbb{C}) \to \mathcal{M}_d(\mathbb{C}) and bipartite operators on \mathbb{C}^d \otimes \mathbb{C}^d, enabling a matrix-level characterization of complete positivity. Specifically, the Choi matrix C_\Phi associated with \Phi is defined as C_\Phi = \sum_{i,j=1}^d |i\rangle\langle j| \otimes \Phi(|i\rangle\langle j|), where \{|i\rangle\}_{i=1}^d denotes an orthonormal basis for \mathbb{C}^d. The map \Phi is completely positive if and only if C_\Phi is positive semidefinite. This representation transforms the problem of verifying complete positivity into checking the positivity of a single operator, providing a practical tool for analysis in finite dimensions. The isomorphism, developed by M.-D. Choi in 1975 and independently by A. Jamiołkowski in 1972, is central to quantum information applications. An equivalent construction arises from vectorization techniques, which linearize matrices into vectors via \mathrm{vec}(X) = \sum_{k,l} X_{kl} |k\rangle \otimes |l\rangle for a matrix X = \sum_{k,l} X_{kl} |k\rangle\langle l|. For the maximally entangled state |\Omega\rangle = \frac{1}{\sqrt{d}} \sum_{i=1}^d |i\rangle \otimes |i\rangle, the relation \mathrm{vec}(\Phi(A)) = ( \mathrm{id} \otimes \Phi ) ( |\Omega\rangle\langle\Omega| ) \mathrm{vec}(A) holds, up to normalization conventions. Here, the Choi matrix corresponds to d ( \mathrm{id} \otimes \Phi ) ( |\Omega\rangle\langle\Omega| ), linking the isomorphism to entanglement in quantum information contexts. This vectorized form facilitates computational implementations and derivations of map properties. To recover the original map from the Choi matrix, one applies the partial trace over the first subsystem: \Phi(A) = \mathrm{Tr}_1 \left[ C_\Phi (I \otimes A^T) \right], where the transpose A^T ensures consistency with the standard column-major vectorization convention. This inversion formula confirms the isomorphism's invertibility, allowing direct reconstruction of \Phi from C_\Phi. While the isomorphism is primarily formulated for finite-dimensional matrix algebras \mathcal{M}_d(\mathbb{C}), extensions to general C*-algebras rely on approximate or state-based versions using the GNS construction to approximate the role of the Choi operator. These generalizations preserve the core idea of associating maps with positive elements but require careful handling of infinite-dimensional settings to maintain well-definedness.

Properties

Basic Properties

Completely positive maps exhibit several fundamental algebraic and analytic properties that arise directly from their definition. One key property is contractivity: for a unital completely positive map \Phi: \mathcal{A} \to \mathcal{B} between unital C*-algebras, the operator norm satisfies \|\Phi\| = 1. This follows from the preservation of positivity and the unital condition, ensuring that \|\Phi(a)\| \leq \|a\| for all a \in \mathcal{A} with \|a\| \leq 1. Unital completely positive maps, those satisfying \Phi(I) = I, preserve the unit and thus map the identity to itself. In the context of quantum information theory, completely positive trace-preserving (CPTP) maps, which are unital in the , necessarily preserve the trace: for any density operator \rho, \operatorname{tr}(\Phi(\rho)) = \operatorname{tr}(\rho). This trace-preservation is a direct consequence of the dual map being unital, ensuring that CPTP maps model valid quantum channels without probability loss. A related property is the monotonicity of norms, where for a completely positive map \Phi, the completely bounded norm equals the operator norm: \|\operatorname{id} \otimes \Phi\| = \|\Phi\|. This equality links directly to complete boundedness, as the supremum over matrix amplifications \sup_n \|\operatorname{id}_n \otimes \Phi\| = \|\Phi\| for unital cases, reflecting the map's stability under tensor products. Completely positive maps also satisfy a Schwarz-type inequality. Specifically, for a completely positive map \Phi: \mathcal{A} \to \mathcal{B(H)} and elements A \in \mathcal{A}, B \in \mathcal{B(H)} with B positive, the inner product satisfies |\langle \Phi(A), B \rangle| \leq \sqrt{\langle \Phi(A^* A), I \rangle} \sqrt{\langle B^* B, I \rangle}, where \langle X, Y \rangle = \operatorname{tr}(X^* Y) denotes the Hilbert-Schmidt inner product. This inequality extends the classical Cauchy-Schwarz bound and holds due to the complete positivity ensuring positivity in amplified systems. Finally, the composition of completely positive maps is completely positive: if \Phi: \mathcal{A} \to \mathcal{B} and \Psi: \mathcal{B} \to \mathcal{C} are completely positive, then \Psi \circ \Phi: \mathcal{A} \to \mathcal{C} is completely positive. This closure under composition makes completely positive maps a cone within the space of linear maps on operator algebras.

Complete Boundedness

A linear map \Phi: A \to B between C*-algebras is said to be completely bounded if the extended maps \mathrm{id}_n \otimes \Phi: M_n(A) \to M_n(B) are bounded for every n \in \mathbb{N}, where M_n denotes the algebra of n \times n complex matrices and \mathrm{id}_n is the identity map on M_n. The completely bounded norm is then given by \|\Phi\|_{\mathrm{cb}} = \sup_{n \in \mathbb{N}} \|\mathrm{id}_n \otimes \Phi\|. Every completely positive map is completely bounded, and a seminal result establishes that for any completely positive \Phi, the completely bounded norm equals the operator norm: \|\Phi\|_{\mathrm{cb}} = \|\Phi\|. This equality highlights the tight control complete positivity exerts over the behavior of the map under tensor extensions. In the 1980s, Paulsen proved a key characterization linking complete positivity directly to complete boundedness via decomposition: a map is completely positive if and only if it is completely bounded and decomposes into completely positive "parts," meaning it lacks a nontrivial negative component in its Wittstock decomposition. Specifically, by Wittstock's theorem, every completely bounded map \Phi admits a decomposition \Phi = \Phi_+ - \Phi_- into two completely positive maps \Phi_+ and \Phi_- with \|\Phi_+ + \Phi_-\|_{\mathrm{cb}} \leq \|\Phi\|_{\mathrm{cb}}, and \Phi is completely positive precisely when \Phi_- = 0. This decomposition underscores that complete positivity corresponds to the absence of "negative" contributions in the completely bounded framework. For positive maps, an equivalent condition is that \|\Phi\|_{\mathrm{cb}} = \|\Phi\|, as deviations indicate incomplete positivity. These results have significant applications in the study of operator bimodules, where Haagerup demonstrated that completely bounded bimodule maps between operator modules admit analogous decompositions into completely positive components, facilitating extensions and injectivity properties essential for operator space theory. In Haagerup's framework, such maps preserve the bimodule structure while maintaining complete boundedness, enabling the construction of Haagerup tensor products and resolutions in noncommutative geometry. Notably, complete boundedness strengthens boundedness: while every completely positive is bounded (\|\Phi\| \leq 1 for contractive cases), bounded maps need not be completely bounded. A is the transpose t: M_2(\mathbb{C}) \to M_2(\mathbb{C}), which is bounded with \|t\| = [1](/page/1) but has \|t\|_{\mathrm{cb}} = \sqrt{2} > [1](/page/1), illustrating how tensoring with matrices can amplify norms for non-completely positive maps.

Examples

Classical and Simple Cases

The map, defined by \Phi(A) = A for any A, is a example of a completely positive map. It preserves the positivity of operators and, when tensored with the map on any auxiliary space, the composition \mathrm{id}_n \otimes \Phi remains positive for all n, confirming its complete positivity. This map corresponds to a trivial Kraus representation with a single operator, the itself. In the qubit case (d=2), the depolarizing channel provides a simple non-trivial example of a completely positive trace-preserving map, given by \Phi(\rho) = (1-p)\rho + p \frac{\mathrm{Tr}(\rho)}{2} I for $0 \leq p \leq 1, where \rho is a density operator and I is the $2 \times 2 identity. This map mixes the input state with the maximally mixed state and admits a Kraus operator representation \Phi(\rho) = \sum_{k=0}^3 A_k \rho A_k^\dagger, with A_0 = \sqrt{1 - \frac{3p}{4}}\, I, A_1 = \sqrt{\frac{p}{4}}\, \sigma_x, A_2 = \sqrt{\frac{p}{4}}\, \sigma_y, and A_3 = \sqrt{\frac{p}{4}}\, \sigma_z, where \sigma_x, \sigma_y, \sigma_z are the Pauli matrices. The completeness relation \sum_k A_k^\dagger A_k = I ensures trace preservation, and the form guarantees complete positivity. When p=0, it reduces to the identity map; when p=1, it is the completely depolarizing map outputting the maximally mixed state regardless of input. The transposition map T(A) = A^T illustrates a positive map that fails complete positivity. It maps positive semidefinite matrices to positive semidefinite matrices, as the eigenvalues of A and A^T coincide. However, it is not completely positive, as demonstrated by the counterexample where \mathrm{id}_2 \otimes T applied to the maximally entangled state projector |\Phi^+\rangle\langle\Phi^+| (with |\Phi^+\rangle = \frac{1}{\sqrt{2}}(|00\rangle + |11\rangle)) yields a matrix with a negative eigenvalue of -\frac{1}{2}. This partial transpose operation on the Bell state projector produces the operator \frac{1}{2} \begin{pmatrix} 1 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 1 \end{pmatrix}, which is not positive semidefinite. In classical settings, stochastic maps arise naturally as completely positive maps restricted to diagonal density operators (probability distributions). A stochastic matrix F = (f_{s,t}) with f_{s,t} \geq 0 and \sum_t f_{s,t} = 1 defines a map on distributions p \mapsto q via q_t = \sum_s p_s f_{s,t}, which extends to a completely positive map on the full matrix algebra by acting diagonally. Completely positive trace-preserving maps thus serve as the quantum analogue of these classical stochastic maps, preserving the classical structure on diagonal inputs. For doubly stochastic matrices, where both rows and columns sum to 1, the corresponding maps preserve the uniform distribution and form a convex set whose extreme points are permutation matrices, analogous to the classical Birkhoff-von Neumann theorem but in the context of completely positive maps.

Quantum Information Examples

In quantum information theory, completely positive maps play a crucial role in modeling realistic noise processes that affect quantum systems, such as decoherence and dissipation. One prominent example is the amplitude damping channel, which describes the energy relaxation of a qubit due to spontaneous emission or coupling to a thermal bath. This channel is represented by the Kraus operators A_0 = \begin{pmatrix} 1 & 0 \\ 0 & \sqrt{1 - \gamma} \end{pmatrix} and A_1 = \begin{pmatrix} 0 & \sqrt{\gamma} \\ 0 & 0 \end{pmatrix}, where \gamma \in [0, 1] parameterizes the damping strength, corresponding to the probability of decay from the excited state |1\rangle to the ground state |0\rangle. The action of the channel on a density operator \rho is \mathcal{E}(\rho) = A_0 \rho A_0^\dagger + A_1 \rho A_1^\dagger, which preserves the trace and leads to a contraction of the Bloch vector along the z-axis while damping off-diagonal coherences. Another key example is the bit-phase flip channel, a Pauli channel that models combined bit and phase errors in quantum bits, such as those arising from magnetic field fluctuations or imperfect control pulses. It applies the identity operator with probability $1 - p and the bit-phase flip operator Y = iXZ (where X and Z are Pauli matrices) with probability p, and is given by the Kraus operators K_0 = \sqrt{1 - p} \, I and K_1 = \sqrt{p} \, Y. The channel map is \mathcal{N}(\rho) = (1 - p) \rho + p Y \rho Y^\dagger, resulting in decoherence in the x and z directions of the Bloch sphere, with the contraction factor $1 - 2p for the x and z components (y component preserved). Entanglement-breaking maps represent a special class of completely positive trace-preserving maps that destroy all quantum entanglement when applied to one part of a bipartite system. A map \Phi is entanglement-breaking if (\mathrm{id} \otimes \Phi)(\rho_{AB}) is a separable state for any entangled input \rho_{AB}. These maps are characterized by having Kraus operators of rank at most one, i.e., each K_i = |\psi_i\rangle \langle \phi_i| for some vectors |\psi_i\rangle and |\phi_i\rangle, allowing representation as a "measure-and-prepare" process: the input is measured in some basis, and a classical mixture of output states is prepared based on the outcome. This property makes them useful for modeling classical communication embedded in quantum protocols, as they sever quantum correlations without preserving coherence. An illustrative case highlighting the necessity of complete positivity is the attempt to construct a universal NOT gate, which would transpose an unknown qubit state in the computational basis, mapping |\psi\rangle = \alpha |0\rangle + \beta |1\rangle to \alpha^* |1\rangle + \beta^* |0\rangle. Such a map is positive but not completely positive, as its Choi matrix has negative eigenvalues, violating the requirement for physical realizability on entangled systems. Consequently, no completely positive trace-preserving implementation exists for a deterministic universal NOT gate, though approximate versions can be achieved via probabilistic or suboptimal channels.

Applications

In Quantum Information Theory

In quantum information theory, completely positive trace-preserving (CPTP) maps, also known as quantum channels, provide the fundamental mathematical framework for describing the evolution of open quantum systems, where interactions with an environment lead to decoherence and dissipation. These maps ensure that the positivity of density operators is preserved under dynamics, even when tensoring with auxiliary systems, making them suitable for modeling realistic quantum processes. The Gorini-Kossakowski-Sudarshan-Lindblad (GKSL) master equation generates such CPTP maps for Markovian dynamics, representing the most general form of time evolution for density operators under weak coupling to a large environment. This equation was independently derived by Gorini, Kossakowski, and Sudarshan, as well as by Lindblad, establishing that the infinitesimal generator of the semigroup of CPTP maps must take a specific Lindblad form to maintain complete positivity and trace preservation. The no-cloning theorem, a cornerstone of quantum information, demonstrates that no CPTP map exists that can perfectly copy an arbitrary unknown quantum state onto a blank state, preventing the replication of quantum information in a way that would undermine principles like quantum key distribution. This result arises because any attempt at universal cloning via a CPTP map would violate linearity or positivity when applied to superpositions, as shown in proofs extending the original argument to general quantum operations. The theorem implies that quantum states cannot be faithfully duplicated without introducing errors, a property that distinguishes quantum from classical information and enables secure protocols. In quantum error correction, CPTP maps model the noise processes afflicting , such as bit-flip or phase-flip errors, allowing for the of codes that protect logical qubits encoded in larger physical systems. A quantum error-correcting code can correct a set of errors represented by Kraus operators from a CPTP map if the code subspace satisfies the Knill-Laflamme conditions, which ensure that the errors act orthogonally on codewords without mixing information between correctable errors. These conditions, formulated in the operator-sum representation inherent to completely positive maps, guarantee the existence of a recovery CPTP map that restores the original state with high fidelity. This framework has enabled fault-tolerant quantum computing by mitigating decoherence through redundancy. Post-2000 developments have highlighted the role of completely positive maps in proving quantum computational advantage, particularly by modeling noisy quantum devices in complexity-theoretic separations between classical and quantum resources. For instance, in interactive proof systems, CPTP maps underpin quantum proofs for classical theorems, demonstrating exponential separations where quantum strategies outperform classical ones under general noise models. Additionally, resource theories of quantum non-Markovianity treat deviations from CP-divisibility as a valuable resource, contrasting with the standard Markovian CPTP channels. In these theories, non-Markovian processes—where intermediate maps are not completely positive—enable tasks like enhanced state preparation or correlation preservation that Markovian channels cannot achieve, with measures quantifying non-Markovianity based on information backflow. Seminal work established such measures using trace distance, framing non-Markovianity within broader resource-theoretic bounds on quantum correlations. High-impact reviews have integrated these into unified resource theories, emphasizing how non-Markovian CP maps (or their absence of divisibility) provide advantages in open-system control.

In Operator Algebras

In the structural theory of C*-algebras, completely positive maps play a pivotal role in K-theory and index theory by inducing order-preserving maps on the K_0 groups. Specifically, for C*-algebras A and B, a completely positive map φ: A → B defines a group homomorphism φ_*: K_0(A) → K_0(B) that preserves the positive cone, ensuring that positive elements in K_0(A) map to positive elements in K_0(B). This property is essential for studying extensions and ideals, as it allows the preservation of order structures in non-commutative topology. Arveson's extension provides a foundational for extending completely positive maps from subalgebras to the entire . The states that if E is an operator subsystem of a unital A and φ: E → B(H) is a unital completely positive map into the bounded operators on a Hilbert space H, then φ extends to a unital completely positive map on A. This result, which relies on the complete boundedness of such maps, facilitates the analysis of representations and dilations in operator algebra theory. In non-commutative , completely positive maps serve as generators for semigroup actions on C*-algebras, modeling time evolutions in . These maps form the basis for completely positive , which are strongly continuous one-parameter of completely positive contractions preserving the algebraic structure. Such capture irreversible and are dilated to unitary representations via the Stinespring , enabling the study of and in non-commutative settings. Modern developments since the 2010s have extended these ideas to approximate completely positive maps, which are crucial for the classification of nuclear C*-algebras through dimension theory. Nuclear C*-algebras are characterized by the completely positive approximation property, where the identity map is approximated by completely positive maps factoring through finite-dimensional algebras. The nuclear dimension, defined via the minimal number of such approximations needed to cover the algebra, provides a regularity condition that, combined with K-theoretic invariants, classifies simple, separable, nuclear C*-algebras absorbing the infinite Cuntz algebra. This framework has advanced the Elliott classification program by bounding dimensions for classes like Z-stable algebras.

References

  1. [1]
    [PDF] 9. Completely Positive Maps This section gives a very quick ...
    We say it is n-positive if (n) is positive, and we say that it is completely positive (c.p. or cp) if it is n-positive for all n. 1. A completely positive map ...
  2. [2]
    [PDF] Completely Positive Maps - Paul Skoufranis
    Examples of positive linear functionals are abundant in mathematics. ... defined completely positive map. ϕ : S → Qn≥1 A such that ϕ(A)=(ϕn(A))n≥1 ...
  3. [3]
    6.2: Positive and Completely Positive Maps - Physics LibreTexts
    Dec 8, 2021 · ... k = I is a completely positive map, and moreover, that any completely positive map can be expressed in this form. Exercises. Show that ...
  4. [4]
    Completely Positive Maps and Entropy Inequalities - Project Euclid
    The definition of a completely positive map has an interesting physical interpretation. Regard Φ as an operation [8, 9] on the space of normal states, i.e. ...
  5. [5]
    [PDF] COMPLETE POSITIVITY IN OPERATOR ALGEBRAS - UCSD Math
    Clearly (˜φ)n = ˜φn, so we are done. The following proposition states that a completely positive map must be com- pletely bounded. In its proof we need the ...
  6. [6]
    [PDF] On the difference between positivity and complete positivity of maps ...
    In contrast, positive maps are guaranteed to transform quantum states to other valid quantum states only when the input states are not entangled to other ...
  7. [7]
    [1207.6519] Dissipative and Non-dissipative Single-Qubit Channels
    Jul 27, 2012 · A canonical derivation of the Kraus representation of the former, via the Choi isomorphism is presented for the general case of a system's ...
  8. [8]
    Positive Functions on C * -Algebras - jstor
    say that Iu is completely positive if ,i(") is positive for each positive integer n. Received by the editors March 29, 1954. 211. Page 2. 212 W. F. STINESPRING ...
  9. [9]
    [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 ...Missing: original | Show results with:original
  10. [10]
    Linear transformations which preserve trace and positive ...
    Linear transformations which preserve trace and positive semidefiniteness of operators. Author links open overlay panel. A. Jamiołkowski.
  11. [11]
    Completely positive linear maps on complex matrices - ScienceDirect
    A linear map Φ from n to m is completely positive iff it admits an expression Φ(A)=Σ i V ∗ i AV i where V i are n×m matrices.Missing: theorem | Show results with:theorem
  12. [12]
    [2411.14800] Fixed Points of Completely Positive Trace-Preserving ...
    Nov 22, 2024 · Completely positive trace-preserving maps S, also known as quantum channels, arise in quantum physics as a description of how the density ...
  13. [13]
    a schwarz inequality for positive linear maps - Project Euclid
    A SCHWARZ INEQUALITY FOR POSITIVE LINEAR MAPS. ON C*-ALGEBRAS. BY. M-D Coz. 1. Introduction. Dvis [5] has derived Schwrz inequality for completely positive ...
  14. [14]
    Completely Bounded Maps and Operator Algebras
    In this book, first published in 2003, the reader is provided with a tour of the principal results and ideas in the theories of completely positive maps, ...
  15. [15]
    [PDF] Injectivity and decomposition of completely bounded maps
    Hence the Choi-Effros lifting theorem for completely positive maps [4] fails for completely bounded maps, even if A is abelian. If dim(A) < ~ , T has of ...<|control11|><|separator|>
  16. [16]
    [PDF] An Analysis of Completely-Positive Trace-Preserving Maps on M2
    One would like to find a simple way to characterize completely positive maps in terms of their action on the algebra B(H1) of the subsystem. The purpose of this ...
  17. [17]
    [PDF] arXiv:quant-ph/0410233v2 27 Sep 2005
    Sep 27, 2005 · The theorem provides an operational interpretation for trace- preserving completely positive maps, which are the natural quantum analogue of ...
  18. [18]
    On Birkhoff's theorem for doubly stochastic completely positive maps ...
    A study is made of the extreme points of the convex set of doubly stochastic completely positive maps of the matrix algebra n.
  19. [19]
    [PDF] Lecture Notes for Ph219/CS219: Quantum Information Chapter 3
    Kraus operators {Ma} is inverted by channel E2 with Kraus operators. {Nµ} ... amplitude-damping channel these two times are related and comparable: T2 ...
  20. [20]
    [quant-ph/0302031] General Entanglement Breaking Channels - arXiv
    Feb 4, 2003 · The only extreme points of the set of completely positive trace preserving maps which are also entanglement breaking are those known as ...
  21. [21]
    [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.Missing: original | Show results with:original
  22. [22]
    [math/0108102] Covering Dimension for Nuclear C*-Algebras II - arXiv
    Aug 15, 2001 · The completely positive rank is an analogue of topological covering dimension, defined for nuclear C*-algebras via completely positive approximations.