Fact-checked by Grok 2 weeks ago

Density matrix

In quantum mechanics, the density matrix, also known as the density operator, is a Hermitian, positive semi-definite operator with unit trace that provides a complete statistical description of a quantum system's state, generalizing the pure state wavefunction to accommodate mixed states arising from incomplete knowledge or ensemble averaging. Introduced by in 1927 and independently by , it formalizes the probabilities of measurement outcomes without specifying the underlying pure states, making it essential for open quantum systems and statistical mechanics. For a pure state represented by a wavefunction |\psi\rangle, the density matrix is \hat{\rho} = |\psi\rangle\langle\psi|, while for a mixed state with probabilities p_i over orthonormal pure states |i\rangle, it is \hat{\rho} = \sum_i p_i |i\rangle\langle i|, ensuring that expectation values of observables \hat{A} are computed as \langle A \rangle = \mathrm{Tr}(\hat{\rho} \hat{A}). This formulation extends classical probability distributions to quantum theory by incorporating coherence and superposition effects, and it underpins applications in quantum information, decoherence, and thermodynamics. Key properties include idempotence for pure states (\hat{\rho}^2 = \hat{\rho}) and the von Neumann entropy S = -\mathrm{Tr}(\hat{\rho} \ln \hat{\rho}) as a measure of mixedness, which quantifies quantum uncertainty beyond classical limits.

Basic Concepts

Definition and motivation

The density matrix, also known as the density operator or statistical operator, was introduced by in his 1927 papers to establish a rigorous foundation for , specifically addressing the description of ensembles of prepared under conditions of incomplete information. This approach was motivated by the necessity to extend beyond individual pure states, enabling the treatment of statistical mixtures where multiple systems or repeated measurements yield probabilistic outcomes reflective of underlying quantum uncertainties. Independently, proposed a similar concept around the same time for handling such ensembles in . In , the wave function provides a complete description of a single pure state, capturing all properties deterministically (up to ). However, real-world scenarios often involve statistical ensembles, such as or partially decohered systems, where the state is a probabilistic superposition—or more precisely, a classical —of pure states, necessitating a more general to compute averages without specifying the exact realization. The density matrix resolves this limitation by representing the system's state as a single that encodes both quantum and classical probabilities. Formally, for an ensemble with probabilities p_i and corresponding orthonormal pure states |\psi_i\rangle, the density matrix \rho is defined as \rho = \sum_i p_i |\psi_i\rangle \langle \psi_i|, where each p_i \geq 0 and \sum_i p_i = 1. This operator is Hermitian (\rho^\dagger = \rho), as the adjoint of each projector |\psi_i\rangle \langle \psi_i| is itself, and positive semi-definite, with eigenvalues p_i \geq 0. Additionally, the normalization condition ensures \operatorname{Tr}(\rho) = 1, reflecting the total probability of the ensemble. The utility of this representation is evident in calculating values: for a Hermitian A, the average \langle A \rangle over the ensemble is \langle A \rangle = \sum_i p_i \langle \psi_i | A | \psi_i \rangle. This simplifies to the form \langle A \rangle = \operatorname{Tr}(\rho A), derived by inserting the definition of \rho and using the cyclic property of the : \sum_i p_i \langle \psi_i | A | \psi_i \rangle = \sum_i p_i \operatorname{Tr}(|\psi_i\rangle \langle \psi_i | A) = \operatorname{Tr}\left( \left( \sum_i p_i |\psi_i\rangle \langle \psi_i | \right) A \right) = \operatorname{Tr}(\rho A). This expression unifies the treatment of pure and mixed states, with pure states appearing as the special case \rho = |\psi\rangle \langle \psi|.

Pure and mixed states

In , a pure state is represented by a density matrix of the form \rho = |\psi\rangle\langle\psi|, where |\psi\rangle is a normalized in the , satisfying \langle\psi|\psi\rangle = 1. This form ensures that the density matrix is a rank-one , idempotent such that \rho^2 = \rho, and has trace unity, \operatorname{Tr}(\rho) = 1. In contrast, a mixed state corresponds to a density matrix \rho that cannot be expressed as |\psi\rangle\langle\psi| for any single |\psi\rangle, typically arising from an ensemble of pure states. The distinction between pure and mixed states is mathematically characterized by the purity \operatorname{Tr}(\rho^2). For pure states, \operatorname{Tr}(\rho^2) = 1, reflecting maximal quantum and no classical . For mixed states, $0 < \operatorname{Tr}(\rho^2) < 1, with the value quantifying the degree of mixing; lower purity indicates greater classical-like probabilistic . Since \rho is Hermitian and positive semi-definite, it admits a spectral decomposition \rho = \sum_i \lambda_i |i\rangle\langle i|, where \lambda_i \geq 0 are the eigenvalues summing to 1 (interpretable as probabilities) and |i\rangle are the orthonormal eigenvectors. In this basis, pure states have exactly one \lambda_i = 1 and the rest zero, while mixed states have multiple nonzero \lambda_i < 1. Pure states embody full quantum superposition and coherence, whereas mixed states often result from incomplete knowledge of the system, represented as an average over an ensemble of pure states weighted by classical probabilities. For instance, if an ensemble consists of pure states |\psi_k\rangle with probabilities p_k, then \rho = \sum_k p_k |\psi_k\rangle\langle\psi_k|. This ensemble interpretation underscores how mixed states incorporate both quantum indeterminacy and classical ignorance. Any two pure state vectors |\psi\rangle and |\phi\rangle representing the same density matrix \rho = |\psi\rangle\langle\psi| = |\phi\rangle\langle\phi| are related by a unitary transformation, |\phi\rangle = U |\psi\rangle, where U is a unitary operator with U^\dagger U = I. This equivalence highlights the non-uniqueness of the state vector representation for pure states, but the density matrix \rho remains invariant under such transformations.

Example: light polarization

A fundamental example of the density matrix arises in the description of light polarization, where photons serve as effective two-level quantum systems. The standard basis consists of horizontal and vertical polarization states, denoted as |H⟩ and |V⟩, respectively. In this basis, the density matrix ρ is a 2×2 Hermitian operator that captures both the polarization direction and any incoherence. Consider a pure state representing fully polarized light at 45° to the horizontal. This corresponds to the coherent superposition |+⟩ = \frac{1}{\sqrt{2}} (|H⟩ + |V⟩), with density matrix ρ = |+⟩⟨+|. Explicitly, \rho = \frac{1}{2} \begin{pmatrix} 1 & 1 \\ 1 & 1 \end{pmatrix}, which features equal diagonal elements and maximal off-diagonal coherence, reflecting the quantum superposition. In contrast, unpolarized light—a classical mixture—has density matrix ρ = \frac{1}{2} (|H⟩⟨H| + |V⟩⟨V|), or \rho = \frac{1}{2} \begin{pmatrix} 1 & 0 \\ 0 & 1 \end{pmatrix}, with equal eigenvalues of 1/2 and vanishing off-diagonals, indicating no phase coherence between basis states. To measure polarization, consider an observable given by the projector P_θ onto the state polarized at angle θ, P_θ = |θ⟩⟨θ| where |θ⟩ = cos θ |H⟩ + sin θ |V⟩. The probability of transmission through a polarizer at θ is Tr(ρ P_θ). For the pure state ρ = |+⟩⟨+|, this yields sin²(θ) or cos²(θ - π/4), depending on alignment, while for the unpolarized mixed state, it simplifies to 1/2 regardless of θ, averaging over random orientations as in classical . This highlights how the density matrix distinguishes quantum coherence from classical statistical mixtures. The two-dimensional nature of polarization allows visualization on the Poincaré sphere, analogous to the for qubits. Pure states like |+⟩ lie on the sphere's surface, corresponding to points with full polarization (degree of polarization P=1), while mixed states such as unpolarized light occupy the interior (P<1), with the maximally mixed state at the center. Physically, a transition from pure to mixed states occurs via decoherence, where environmental interactions—such as scattering or absorption—randomize the relative phase between |H⟩ and |V⟩ components, suppressing off-diagonal elements and eroding quantum coherence.

Formal Properties

Ensembles and purifications

The density matrix \rho offers a statistical interpretation for quantum systems described by an ensemble consisting of probabilities \{p_i\} and corresponding pure states |\psi_i\rangle, such that \rho = \sum_i p_i |\psi_i\rangle\langle\psi_i|, where \sum_i p_i = 1 and each p_i \geq 0. This formulation, introduced by von Neumann, allows \rho to encapsulate the average behavior of the ensemble without specifying the individual states, making it particularly useful for describing incomplete knowledge of a system's preparation. A key feature of this interpretation is the non-uniqueness of the ensemble for a given \rho: any mixed density matrix (with \operatorname{Tr}(\rho^2) < 1) admits infinitely many distinct decompositions into ensembles \{p_i, |\psi_i\rangle\} that yield the same \rho, as different sets of pure states can be combined with adjusted probabilities to produce identical mixtures. For instance, a mixed state diagonal in its eigenbasis, \rho = \sum_i \lambda_i |i\rangle\langle i| with $0 < \lambda_i < 1, can be expressed as a convex combination of pure states in various ways beyond the trivial eigenstate decomposition. In contrast, pure states (\operatorname{Tr}(\rho^2) = 1) have a unique trivial ensemble consisting of a single state with probability 1. To address the incompleteness inherent in mixed states, the concept of purification embeds \rho into a pure state of a larger composite system. Specifically, for a mixed density matrix \rho_A on subsystem A, there exists a pure state |\Psi\rangle_{AB} on the bipartite system A \otimes B (where B acts as an auxiliary or environmental system) such that \rho_A = \operatorname{Tr}_B(|\Psi\rangle\langle\Psi|_{AB}), with the dimension of B at least as large as that of A. This purification is constructed using the spectral decomposition \rho_A = \sum_i \lambda_i |i\rangle\langle i|_A (where \lambda_i > 0 are the nonzero eigenvalues), yielding |\Psi\rangle_{AB} = \sum_i \sqrt{\lambda_i} |i\rangle_A |i\rangle_B in the basis, ensuring the over B recovers \rho_A. Note that purifications are not unique, as different choices for the basis states on B (via unitary transformations) produce equivalent \rho_A but distinct global pure states. The operation, central to purification, is formally defined for a density operator \rho_{AB} on A \otimes B as \operatorname{Tr}_B(\rho_{AB}) = \sum_k \langle k|_B \rho_{AB} |k\rangle_B, where \{|k\rangle_B\} is an for the of B. This trace effectively "averages" over the in B, yielding the reduced description \rho_A while discarding about B. Purification highlights underlying quantum correlations, as the apparent classical mixture in \rho_A arises from between A and B in the global pure state |\Psi\rangle_{AB}, revealing how interactions with an can mask coherent quantum features in subsystems.

Key mathematical properties

The density matrix \rho, as a Hermitian on a , satisfies the normalization condition \operatorname{Tr}(\rho) = 1. This trace preservation ensures that probabilities derived from \rho sum to unity, and for any bounded A, the is given by \langle A \rangle = \operatorname{Tr}(\rho A). A defining feature of \rho is its positive semi-definiteness: for any normalized |\psi\rangle, the \langle \psi | \rho | \psi \rangle \geq 0. Consequently, the eigenvalues \lambda_i of \rho are real and lie in the interval [0, 1], with their sum equaling 1 due to the condition. By the for compact Hermitian operators, \rho admits a in an of eigenvectors: \rho = \sum_i \lambda_i |\phi_i\rangle \langle \phi_i|, where \lambda_i \geq 0 are the eigenvalues and |\phi_i\rangle form a complete basis. This diagonal form facilitates computations of traces, expectation values, and other operator functions. The collection of all density matrices forms a in the space of Hermitian operators. Specifically, if \rho_1 and \rho_2 are density matrices and $0 \leq p \leq 1, then the p \rho_1 + (1-p) \rho_2 is also a density matrix, inheriting the , positivity, and Hermiticity properties. Density matrices exhibit unitary invariance: for any U, the transformed operator U \rho U^\dagger remains a density matrix, as unitarity preserves the , Hermiticity, and positivity. Commutator relations involving \rho and observables A are central to , with [\rho, A] quantifying non-commutativity that underlies principles, such as bounds on simultaneous variances \Delta A \Delta B \geq \frac{1}{2} |\langle [A, B] \rangle|.

Dynamics and Evolution

Von Neumann equation

The time evolution of the density operator \rho(t) for an isolated quantum system is described by the von Neumann equation, which provides the quantum mechanical analogue to the classical Liouville equation and extends the to mixed states. This equation governs the reversible dynamics in closed systems, where the H dictates the unitary transformation of the state. To derive the von Neumann equation, consider first a pure state |\psi(t)\rangle evolving according to the time-dependent Schrödinger equation i \hbar \frac{d}{dt} |\psi(t)\rangle = H |\psi(t)\rangle. The corresponding density operator is \rho(t) = |\psi(t)\rangle \langle \psi(t)|. Differentiating \rho(t) with respect to time yields i \hbar \frac{d\rho}{dt} = i \hbar \left( \frac{d}{dt} |\psi\rangle \langle \psi| + |\psi\rangle \frac{d}{dt} \langle \psi| \right) = H |\psi\rangle \langle \psi| - |\psi\rangle \langle \psi| H = [H, \rho], where the commutator [H, \rho] = H \rho - \rho H. For a mixed state \rho(t) = \sum_i p_i |\psi_i(t)\rangle \langle \psi_i(t)|, with probabilities p_i and each |\psi_i(t)\rangle evolving unitarily under the same H, the derivation extends linearly, resulting in the same form i \hbar \frac{d\rho}{dt} = [H, \rho]. Equivalently, this is written as the Liouville-von Neumann equation \frac{d\rho}{dt} = -\frac{i}{\hbar} [H, \rho]. The solution to this equation preserves unitarity. For a time-independent Hamiltonian, the unitary operator is U(t) = e^{-i H t / \hbar}, and the density operator evolves as \rho(t) = U(t) \rho(0) U^\dagger(t). This ensures key conservation laws: the trace \operatorname{Tr}(\rho(t)) = 1 remains constant, as \frac{d}{dt} \operatorname{Tr}(\rho) = -\frac{i}{\hbar} \operatorname{Tr}([H, \rho]) = 0 due to the cyclic property of the trace. Similarly, the purity \operatorname{Tr}(\rho^2(t)) is invariant, reflecting the absence of decoherence in closed systems; its time derivative vanishes under the commutator form. A illustrative example is the precession of a two-level system, such as a particle in a constant \mathbf{B} = B \hat{z}. The is H = -\frac{\gamma B}{2} \sigma_z, where \gamma is the and \sigma_z is the Pauli-z . For an initial mixed state with density elements \rho_{11}(0), \rho_{22}(0) = 1 - \rho_{11}(0), and off-diagonal coherences \rho_{12}(0) = \rho_{21}^*(0), the equation yields diagonal elements that remain constant (\rho_{11}(t) = \rho_{11}(0)) while the off-diagonals precess as \rho_{12}(t) = \rho_{12}(0) e^{i \gamma B t}, demonstrating coherent Rabi-like oscillations without relaxation.

Open quantum systems

In realistic quantum systems, interactions with an uncontrollable environment render the closed-system unitary evolution inadequate for describing the dynamics; the effective evolution of the reduced density matrix \rho for the system is instead obtained by tracing out the environmental degrees of freedom, leading to non-unitary behavior that captures dissipation and decoherence. This approach is essential for modeling processes where the total Hilbert space is too large to track fully, focusing on the system's observable properties. For Markovian open quantum systems—those where environmental correlations decay rapidly—the dynamics are governed by the Lindblad master equation, which extends the von Neumann equation with dissipative terms: \frac{d\rho}{dt} = -\frac{i}{\hbar} [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's Hamiltonian and the L_k are Lindblad operators encoding the interaction with the environment. This form ensures complete positivity and trace preservation, maintaining the physical validity of \rho as a density matrix under the semigroup evolution. A sketch of the derivation begins with a purification of \rho in an enlarged including the , evolving unitarily under a total ; the reduced dynamics then emerge via the , assuming weak system- coupling such that the remains nearly unperturbed, combined with the Markov approximation to neglect memory effects in the bath correlations. Decoherence arises in these open systems as the induces rapid suppression of off-diagonal elements in \rho within a preferred "pointer basis," typically aligned with the interaction , thereby favoring diagonal, classical-like probability distributions over superpositions. This process explains the apparent emergence of classical behavior from without invoking measurement postulates. Representative examples of Lindblad operators include amplitude damping, which models energy dissipation like in a two-level , with L = \sqrt{\gamma} \sigma_- where \sigma_- is the lowering and \gamma the decay rate; this term drives the system toward the while suppressing coherences. Phase damping, or , captures pure loss of phase information without energy exchange, using L = \sqrt{\gamma} \frac{\sigma_z}{2}, which decays off-diagonals exponentially while leaving populations unchanged. As a non-Markovian alternative, the relaxes the Markov assumption, incorporating bath memory effects through time integrals of correlation functions, though it may violate complete positivity for strong couplings.

Measurement and Information

Expectation values and probabilities

In , the density matrix provides a unified framework for computing the probabilities of measurement outcomes, generalizing the to both pure and mixed states. For a projective corresponding to an with involving orthogonal projectors \{P_a\} satisfying \sum_a P_a = I, where I is the identity operator, the probability p(a) of obtaining outcome a for a in \rho is given by p(a) = \operatorname{Tr}(\rho P_a). This expression arises from the trace's invariance under cyclic permutations and the completeness of the projectors, ensuring \sum_a p(a) = 1. The expectation value of an A with non-degenerate eigenvalues \{a\} and corresponding projectors \{P_a\} is then \langle A \rangle = \sum_a a p(a) = \operatorname{Tr}(\rho A), which holds more generally for any Hermitian A without assuming degeneracy, as the trace formulation directly incorporates the operator's properties. This trace rule simplifies calculations for mixed states, where \rho encodes statistical mixtures, and aligns with the probabilistic interpretation of quantum measurements. Upon obtaining outcome a in a projective , the post-measurement collapses to \rho' = \frac{P_a \rho P_a}{p(a)}, assuming p(a) > 0, which normalizes the updated density matrix while projecting onto the eigenspace of P_a. This update rule, part of the measurement postulate, ensures the remains Hermitian and positive semi-definite with one, reflecting the irreversible nature of the . For more general measurements described by a positive-operator-valued measure (POVM) \{E_a\} with \sum_a E_a = I and each E_a positive semi-definite, the outcome probability generalizes to p(a) = \operatorname{Tr}(\rho E_a), accommodating non-projective effects like those in or approximate measurements. The corresponding post-measurement state is \rho' = \frac{\sqrt{E_a} \rho \sqrt{E_a}}{p(a)}, which preserves the POVM's positivity and completeness while updating the state consistently. This formulation, introduced in the operational approach to quantum probability, extends projective measurements to a broader class of physically realizable instruments. Simultaneous measurements of compatible observables are possible when their projectors commute, i.e., [P_a, P_b] = 0 for all a, b, allowing a and via the product of traces. Incompatibility, marked by non-commuting projectors, precludes such joint measurements, underscoring the density matrix's role in quantifying quantum correlations.

Von Neumann entropy

The provides a fundamental measure of the or mixedness inherent in a described by a density matrix \rho. It is defined as S(\rho) = -\operatorname{Tr}(\rho \log_2 \rho), where the logarithm is taken with base 2 to express the in units of bits, and the is over the . This definition extends the classical Shannon entropy to , quantifying the information content or statistical disorder of \rho. When \rho is diagonalized in its eigenbasis with eigenvalues \{\lambda_i\}, the simplifies to S(\rho) = -\sum_i \lambda_i \log_2 \lambda_i, where the sum runs over the non-zero eigenvalues, as \rho is positive semidefinite with trace 1. Several key properties characterize the von Neumann entropy. It is non-negative, S(\rho) \geq 0, with equality if and only if \rho represents a pure state (i.e., \rho = |\psi\rangle\langle\psi| for some vector |\psi\rangle). The entropy is concave, meaning that for any density matrices \rho_1, \rho_2 and $0 \leq p \leq 1, S(p \rho_1 + (1-p) \rho_2) \geq p S(\rho_1) + (1-p) S(\rho_2), which reflects the averaging effect on mixedness under convex combinations. Additionally, S(\rho) is invariant under unitary transformations: S(U \rho U^\dagger) = S(\rho) for any unitary operator U, as the trace operation preserves this structure. For product states, the entropy exhibits additivity: S(\rho_A \otimes \rho_B) = S(\rho_A) + S(\rho_B), allowing independent systems to contribute separately to the total entropy. A related quantity is the quantum relative entropy, or Umegaki relative entropy, defined as S(\rho \parallel \sigma) = \operatorname{Tr}(\rho \log_2 \rho - \rho \log_2 \sigma) = -S(\rho) - \operatorname{Tr}(\rho \log_2 \sigma), for density matrices \rho and \sigma where the support of \rho is contained in that of \sigma (otherwise defined as +\infty). This measure is non-negative, S(\rho \parallel \sigma) \geq 0, with equality \rho = \sigma, and it quantifies the distinguishability or between two quantum states, playing a central role in quantum hypothesis testing and resource theories. The von Neumann entropy also satisfies subadditivity for composite systems: if \rho_{AB} is the density matrix of a bipartite system and \rho_A = \operatorname{Tr}_B(\rho_{AB}) is the reduced state on subsystem A, then S(\rho_{AB}) \leq S(\rho_A) + S(\rho_B). This inequality bounds the total entropy by the sum of marginal entropies, implying that correlations (such as entanglement) cannot increase the entropy beyond this limit. In quantum statistical mechanics, the von Neumann entropy corresponds to the thermodynamic entropy, appearing in the Helmholtz free energy F = \langle H \rangle - T S(\rho) for a system with Hamiltonian H at temperature T, where it governs equilibrium properties and phase transitions.

Representations and Analogies

Wigner function

The Wigner function provides a quasi-probability distribution in phase space that represents the density matrix \rho, offering a bridge between quantum and classical descriptions of quantum systems. It is defined via the Weyl transform as W(q,p) = \frac{1}{\pi \hbar} \int_{-\infty}^{\infty} \langle q + y | \rho | q - y \rangle e^{2 i p y / \hbar} \, dy, where q and p are position and momentum coordinates, respectively, and \hbar is the reduced Planck's constant. This integral transform maps the operator \rho to a function on phase space, preserving key quantum features while allowing for phase-space analysis akin to classical statistical mechanics. A fundamental property of the Wigner function is that its marginal distributions recover the quantum probability densities: integrating over momentum yields the position probability \int W(q,p) \, dp = |\psi(q)|^2 for a pure state |\psi\rangle, and integrating over position gives the momentum probability \int W(q,p) \, dq = |\tilde{\psi}(p)|^2, where \tilde{\psi} is the Fourier transform of \psi. Unlike classical probability distributions, W(q,p) can take negative values, which signal non-classical quantum interference effects and prevent a direct probabilistic interpretation. For pure states, the expression simplifies to W(q,p) = \frac{1}{\pi \hbar} \int_{-\infty}^{\infty} \psi^*(q+y) \psi(q-y) e^{2 i p y / \hbar} \, dy, highlighting its origin as a Fourier transform of the off-diagonal elements of the density matrix in the position basis. To address the negativity issue, alternative representations such as the Husimi and the Glauber-Sudarshan P-function serve as smoothed or deconvolved counterparts to the Wigner function. The , obtained by convolving the Wigner function with a Gaussian of width set by the vacuum state uncertainty, is always non-negative and acts as an overlap with coherent states, providing a true at the cost of added smoothing. In contrast, the P-function represents the density matrix as an integral over coherent states with a distribution P that can be highly singular or negative for non-classical states, offering less smoothing but greater sensitivity to quantum features. For Gaussian states, the Wigner function takes a particularly simple form and remains positive definite. Coherent states, which are minimum-uncertainty Gaussian wave packets displaced in phase space, yield a Gaussian Wigner function centered at the classical phase-space point, with no negative regions, illustrating a direct quantum-classical correspondence for these states. This positivity aligns with Hudson's theorem, which states that a pure state's Wigner function is non-negative everywhere if and only if the state is Gaussian. In , quantum of the density matrix, governed by the , translates to the Wigner function via the product and . The [f, g]_M = f \star g - g \star f (where \star is the Moyal product incorporating \sin terms from the ) replaces the classical , enabling the description of as a deformed classical flow in .

Classical limits and analogies

The density matrix in finds its classical analog in the Liouville density f(q, p, t), a over coordinates q and momenta p, which governs the statistical description of classical ensembles. This function evolves deterministically according to the Liouville , \frac{\partial f}{\partial t} = \{ H, f \}_{\text{Poisson}}, where \{ \cdot, \cdot \}_{\text{Poisson}} denotes the with respect to the classical Hamiltonian H(q, p). This preserves the phase-space volume and ensures incompressibility of the flow, mirroring the unitary of pure quantum states but in a commutative framework. In the semiclassical limit as \hbar \to 0, the Wigner function, a phase-space representation of the density matrix, approaches the classical Liouville density for systems with large actions, where quantum oscillations average out. The further bridges this gap by demonstrating that the expectation values \langle q \rangle and \langle p \rangle of the density matrix follow classical trajectories derived from Hamilton's equations, \frac{d \langle q \rangle}{dt} = \frac{\partial H}{\partial p} and \frac{d \langle p \rangle}{dt} = -\frac{\partial H}{\partial q}, provided the wave packet remains localized on scales much larger than \hbar. This correspondence holds particularly well for short times up to the Ehrenfest time, beyond which quantum spreading disrupts the classical mimicry. Quantum features of the density matrix, such as negativities in the Wigner function, signify nonclassical and vanish in the macroscopic due to the suppression of quantum coherence over large action scales. These negativities, absent in classical distributions, disappear as \hbar becomes negligible relative to system size, allowing the Wigner function to become a bona fide positive probability density akin to f(q, p, t). Decoherence plays a crucial role in enforcing this classical appearance, as interactions with an rapidly suppress off-diagonal elements of the density matrix in the basis, rendering it diagonal and interpretable as a classical over positions. This process, driven by entanglement with environmental , selects preferred states (pointer states) that mimic classical trajectories, effectively erasing quantum superpositions without invoking collapse. The von Neumann equation, i \hbar \frac{d \rho}{dt} = [H, \rho], reduces to the Liouville equation in the classical limit, with the commutator [H, \rho] becoming i \hbar \{ H, \rho \}_{\text{Poisson}} up to higher-order terms in \hbar, thus recovering incompressible classical flow. However, this limit has limitations: quantum systems exhibit revivals, where wave packets reform periodically due to discrete energy levels, contrasting with the irreversible diffusion in classical periodic motion. Additionally, quantum chaos suppresses sensitivity to initial conditions compared to classical chaos, as level repulsion stabilizes spectra and prevents exponential divergences beyond the Ehrenfest time.

Advanced Topics

C*-algebraic formulation

In the C*-algebraic formulation of quantum mechanics, the observables are elements of a unital \mathcal{A}, and physical states are described by positive linear functionals \phi: \mathcal{A} \to \mathbb{C} satisfying \phi(I) = 1, where I denotes the multiplicative identity. This abstract setting generalizes the approach to infinite-dimensional s and non-separable systems, avoiding direct reliance on a fixed . The density matrix \rho, in cases where \mathcal{A} = B(\mathcal{H}) for a \mathcal{H}, corresponds precisely to the functional \phi(A) = \Tr(\rho A) for all A \in \mathcal{A}, preserving the trace's cyclic property and ensuring \phi is completely positive. The key mathematical properties of trace and positivity underpin this correspondence, as \rho must be self-adjoint, positive semi-definite, and trace-normalized to yield a valid state functional. The Gelfand-Naimark-Segal (GNS) construction provides a canonical way to realize any such state \phi concretely: it builds a pre-Hilbert space from the left ideal \mathcal{A}_\phi = \{ a \in \mathcal{A} \mid \phi(a^* a) = 0 \}, completing it to a \mathcal{H}_\phi with inner product \langle a | b \rangle = \phi(b^* a), and defines a representation \pi_\phi: \mathcal{A} \to B(\mathcal{H}_\phi) by \pi_\phi(a) \xi_b = \xi_{a b}, where \xi_a are vectors in the space. The state \phi then becomes the vector state \phi(a) = \langle \Omega_\phi | \pi_\phi(a) \Omega_\phi \rangle, with \Omega_\phi = \xi_I as the cyclic and separating vector, thus embedding the abstract algebra into operators on a tailored to the state. This construction is faithful if \phi is faithful, and it extends naturally to mixed states in . Applying the GNS representation to the algebra of observables yields a von Neumann algebra as the double commutant \pi(\mathcal{A})'', which captures the weak closure and includes all bounded operators relevant to the state. In finite-dimensional quantum mechanics, these von Neumann algebras are Type I factors, isomorphic to the full operator algebra B(\mathcal{H}) for \dim \mathcal{H} < \infty, where the center is trivial and projections correspond to direct summands of \mathcal{H}. This classification extends the formulation to infinite factors, accommodating systems like quantum fields where Type II or III structures arise, but Type I remains foundational for standard quantum mechanical models. Thermal equilibrium states in this framework are characterized by the Kubo-Martin-Schwinger (KMS) condition, which specifies a state \phi on a C*-dynamical system (\mathcal{A}, \mathbb{R}, \alpha) generated by a Hamiltonian evolution \alpha_t(A) = e^{i t H} A e^{-i t H}. A state \phi is KMS at inverse temperature \beta > 0 if the function f_{AB}(z) = \phi(\alpha_{i z}(A) B) is analytic in the strip $0 < \Im z < \beta, continuous on the boundary, and satisfies the boundary condition \phi(A \alpha_\beta(B)) = \phi(B A), where \alpha_\beta is the . The canonical example is the Gibbs state \rho = e^{-\beta H}/Z with partition function Z = \Tr(e^{-\beta H}), which uniquely satisfies the KMS condition for gapped systems and models equilibrium in infinite-volume limits. Tomita-Takesaki theory further refines the treatment of states on , associating to a faithful normal state \phi on a M \subset B(\mathcal{H}) with cyclic and separating \Omega (from the GNS ) a modular \Delta = S^* S, where S is the closure of the S a \Omega = a^* \Omega for a \in M. This is positive self-adjoint, generating the modular automorphism group \sigma_t^\phi(A) = \Delta^{i t} A \Delta^{-i t}, which preserves the state and encodes time-reversal symmetries intrinsic to \phi. The theory connects to via the relative modular \Delta_{\psi|\phi} between states \phi, \psi, enabling the Araki-Uhlmann relative S(\phi || \psi) = -\langle \Omega_\phi | \log \Delta_{\psi|\phi} \Omega_\phi \rangle, a measure of distinguishability that generalizes and satisfies monotonicity under channels. In relativistic , the C*-ic formulation manifests through algebraic quantum field theory (AQFT), where observables are assigned to open sets O in as a net of local algebras \mathcal{R}(O), satisfying isotony \mathcal{R}(O_1) \subset \mathcal{R}(O_2) for O_1 \subset O_2, microcausality [\mathcal{R}(O_1), \mathcal{R}(O_2)] = 0 for spacelike separated O_1, O_2, and under Poincaré transformations. States are defined on the quasi-local \mathcal{A} = \bigcup_K \prod_{O_i \in K} \mathcal{R}(O_i) over compact covers K, as positive linear functionals extending to local algebras, with the Reeh-Schlieder theorem ensuring density of local operators in the GNS . This structure accommodates infinite , Haag duality for type III factors in local algebras, and states via KMS-like conditions on thermal representations.

Applications in quantum information

In quantum information theory, density matrices provide a complete description of mixed quantum states, enabling the modeling of realistic scenarios involving decoherence and noise. One key application is the representation of , which describe the evolution of a quantum system under environmental interactions. A quantum channel transforms an input density operator \rho to an output \rho' = \sum_i K_i \rho K_i^\dagger, where the Kraus operators \{K_i\} satisfy the completeness relation \sum_i K_i^\dagger K_i = I to ensure trace preservation. This formalism, rooted in the work of Kraus, allows for the simulation of noise processes in quantum devices. For instance, the depolarizing channel, which randomly replaces the state with the maximally mixed state with probability p, has Kraus operators including the identity and scaled by \sqrt{p/3}, modeling symmetric decoherence in noisy quantum circuits. Density matrices are essential for detecting entanglement in mixed states, where pure-state methods fail. Entanglement witnesses are Hermitian operators W such that \operatorname{Tr}(W \sigma) \geq 0 for all separable states \sigma, but \operatorname{Tr}(W \rho) < 0 for some entangled \rho. These witnesses can be optimized to detect specific forms of entanglement, as shown in approaches that maximize violation under physical constraints. For bipartite systems, the partial transpose criterion serves as a : a state is entangled if its partial transpose has negative eigenvalues, providing a necessary and sufficient test for $2 \times 2 and $2 \times 3 systems. Quantifiers like further measure entanglement degree; for two qubits, it is C(\rho) = \max(0, \sqrt{\lambda_1} - \sum_{i=2}^4 \sqrt{\lambda_i}), where \lambda_i are eigenvalues of \rho (\sigma_y \otimes \sigma_y) \rho^* (\sigma_y \otimes \sigma_y), enabling quantification in experimental mixed states. In quantum error correction, density matrices facilitate the diagnosis and mitigation of errors in encoded . Stabilizer codes define a codespace as the +1 eigenspace of a stabilizer group generated by commuting Pauli operators, with syndrome measurements projecting the density matrix onto error subspaces without disturbing the logical state. For the [[5,1,3]] code, stabilizers such as X_1 Z_2 Z_3 X_4 I_5 and Z_1 X_2 I_3 X_4 Z_5 (with cyclic permutations) allow syndrome extraction via ancillary qubits, recovering the original \rho up to a correctable Pauli error with in fault-tolerant implementations. This approach extends to mixed states, where the density operator evolves under error channels, and correction restores coherence, crucial for scalable . Quantum optics leverages density matrices for continuous-variable systems, where states are described in infinite-dimensional Hilbert spaces of bosonic modes. Beam splitters act as two-mode transformations on density operators, mixing fields via unitary U_{BS} = \exp[i \theta (a^\dagger b + a b^\dagger)], enabling entanglement generation from independent squeezed states. Squeezing, reducing uncertainty in one quadrature below the vacuum level, is modeled by Gaussian density matrices with covariance matrices exhibiting negative eigenvalues under partial transpose, confirming entanglement for applications like quantum teleportation of optical fields. State tomography reconstructs the full density matrix from repeated measurements, essential for verifying quantum operations. By performing projective measurements in multiple bases, expectation values \langle A_k \rangle = \operatorname{Tr}(\rho A_k) yield \rho = \sum_k \langle A_k \rangle \chi_k via a basis expansion, with fidelity F(\rho, \sigma) = \left[ \operatorname{Tr} \sqrt{\sqrt{\rho} \sigma \sqrt{\rho}} \right]^2 quantifying reconstruction accuracy against an ideal \sigma. For d-dimensional systems, d^2 - 1 independent measurements suffice, achieving fidelities above 99% in photonic implementations with to reduce shots. Recent advances integrate matrices into and sensing protocols. In variational quantum eigensolvers adapted for mixed states, the is variationally optimized to minimize energy under noisy channels, as in the variational quantum state eigensolver, which finds dominant eigenvectors of non-Hermitian Hamiltonians with applications to open-system dynamics, achieving near-ground-state fidelities in molecular simulations. For nitrogen-vacancy () centers in , via fast pulse sequences reconstructs states for nanoscale magnetometry, enabling sub-angstrom resolution sensing of biomolecular fields with coherence times extended to milliseconds under dynamical decoupling. These developments, up to 2025, underscore ' role in bridging theory and experiment in fault-tolerant quantum technologies.

Historical Development

Origins and key contributors

The concept of the density matrix originated in the mid-1920s amid the rapid development of , as physicists sought mathematical tools to handle statistical descriptions of quantum systems beyond pure states. played a pivotal role in its introduction through his 1927 trilogy of papers published in Nachrichten von der Gesellschaft der Wissenschaften zu , Mathematisch-Physikalische Klasse, where he developed a rigorous -based framework for . In the second paper of this series, titled "Wahrscheinlichkeitstheoretischer Aufbau der Quantenmechanik," von Neumann first presented the density , denoted as a statistical operator U (later ρ), as a means to represent statistical ensembles in the of open systems. This innovation allowed for the description of systems interacting with environments, where the state is a rather than a single wavefunction, addressing limitations in earlier wave and matrix formulations. Independently, in 1946, introduced the density matrix in his work on nuclear induction, applying it to describe spin ensembles in . Von Neumann's work was heavily influenced by Werner Heisenberg's from 1925, which emphasized non-commuting operators, and Paul Dirac's emerging bra-ket notation, which provided a compact Dirac delta-based approach to quantum states starting in his 1927-1928 publications. The primary motivation was to formulate thermodynamic ensembles purely within , avoiding classical probability distributions that had been used to interpret quantum statistics up to that point. By defining ρ as a weighted sum of operators over possible pure states, von Neumann enabled calculations of expectation values and probabilities for ensembles without invoking external classical assumptions. Independently, introduced a similar in his paper on the problem in wave mechanics, applying it to describe the reduced dynamics of an excited atom interacting with an . Although focused on radiative rather than broad ensembles, Landau's approach used an analogous to over environmental , prefiguring the density 's role in open quantum systems. Landau's ideas appeared in the context of calculating paramagnetic susceptibility in quantum gases, where statistical averaging over degenerate states required such a tool to capture irreversible processes without full environmental resolution. Von Neumann further refined and formalized the density matrix in his seminal 1932 monograph, , where ρ is defined as ρ = Σ p_i |ψ_i⟩⟨ψ_i| for an ensemble with probabilities p_i and pure states |ψ_i⟩. This text solidified the concept's place in , proving its hermiticity, unit , and idempotency for pure states, and establishing its to classical phase-space distributions in the commutative .

Evolution of the concept

In the 1950s, the density matrix formalism began to intersect with , particularly through the work of Edwin T. Jaynes, who established a direct analogy between the of a density matrix and Shannon's measure of in classical . This connection highlighted the density matrix's role in quantifying incomplete knowledge of , bridging and . During the 1960s and 1970s, advancements in phase-space representations extended the utility of density matrices, with developments in the Wigner function enabling quasiprobability descriptions of quantum states that facilitated semiclassical approximations. Concurrently, the study of open quantum systems saw pivotal progress through the Gorini–Kossakowski–Sudarshan–Lindblad (GKSL) equation, derived independently by Vittorio Gorini, Andrzej Kossakowski, George Sudarshan, and Goran Lindblad, which provided a Markovian for the of density matrices under dissipative dynamics while preserving complete positivity. These formulations, building on earlier analyses like Sudarshan et al.'s 1961 work, became foundational for modeling environmental interactions in . The 1980s marked the rise of , where the density matrix proved essential for quantifying channel capacities, as exemplified by Alexander Holevo's 1973 bound—revisited and applied in this era—which limits the classical transmissible through quantum channels to the von Neumann entropy of the ensemble's density matrix. A key milestone was Charles Bennett's 1982 demonstration of reversible computation, which, when extended to quantum contexts, underscored the density matrix's importance in analyzing thermodynamic costs and preservation in unitary evolutions. In the 1990s and 2000s, Wojciech Zurek's decoherence theory utilized density matrices to explain the quantum-to-classical transition, showing how environmental entanglement rapidly diagonalizes the reduced density matrix in preferred bases, suppressing superpositions. This framework gained prominence in , as detailed in Michael Nielsen and Isaac Chuang's 2000 textbook, which integrated density matrices for describing mixed states, error correction, and algorithmic efficiency in noisy environments. From the 2010s to 2025, density matrices found extensions in quantum thermodynamics, where fluctuation theorems were generalized to open systems, relating work and distributions to the evolution of nonequilibrium density matrices, as in studies of quantum processes beyond classical Jarzynski equality analogs. Applications emerged in , with density matrices employed as kernels or features in quantum-enhanced models for and state reconstruction, leveraging their ability to encode correlations. In topological quantum matter, density matrices revealed mixed-state topological invariants, enabling characterization of phases robust to dissipation. Experimental milestones included high- multi-qubit density matrix , achieving over 95% fidelity for nine-qubit states with reduced measurements in photonic and superconducting platforms by the mid-2020s.

References

  1. [1]
    [PDF] Physics 7230: Statistical Mechanics Lecture set 5: Density Matrix
    It is a generalizaiton of convention wavefunction formulation of quantum mechanics, and in equilibrium allows computation of the partition func- tion and ...Missing: authoritative sources
  2. [2]
    [PDF] Density Matrices - CMU Quantum Theory Group
    Density matrices are employed in quantum mechanics to give a partial description of a quantum system, one from which certain details have been omitted.
  3. [3]
    [PDF] The Quantum Density Matrix and Its Many Uses - arXiv
    Aug 16, 2023 · Abstract. The quantum density matrix generalises the classical concept of proba- bility distribution to quantum theory.
  4. [4]
    Density Matrix Theory and Applications | SpringerLink
    In stockThe density matrix was introduced by J. von Neumann in 1927 to describe statistical concepts in quantum mechanics. The main virtue of the density matrix is its ...
  5. [5]
    Mathematical foundations of quantum mechanics : Von Neumann ...
    Jul 2, 2019 · Mathematical foundations of quantum mechanics. by: Von Neumann, John, 1903-1957. Publication date: 1955. Topics: Matrix mechanics. Publisher ...
  6. [6]
    None
    Nothing is retrieved...<|separator|>
  7. [7]
    [PDF] Quantum Computation and Quantum Information
    Nielsen, Michel A., and Chuang, Isaac L. Quantum Computation and ... The density operator was introduced as a means of describing ensembles of quantum.
  8. [8]
    [PDF] Chapter 8 Density Matrix Formalism
    Note also that, if given a density matrix, it is easy to determine if it is a mixed or pure state by looking at Tr(⇢2: Tr(⇢2) = Xk h k| Xj pj | jih j| Xi.
  9. [9]
    [PDF] Course Notes: The Density Operator - Quantum and Atom Optics
    The density operator represents the state of a quantum system in a more general way than the state vector, and equivalently represents an observer's state of ...Missing: authoritative sources
  10. [10]
    Derivation from Bloch Equation to von Neumann Equation to ...
    Jun 6, 2022 · Here, we mathematically derive the space-independent von Neumann equation for electron spin from the classical Bloch equation.
  11. [11]
    A short introduction to the Lindblad master equation | AIP Advances
    Feb 4, 2020 · The concepts of reduced density matrix and partial trace are essential in the study of open quantum systems. If we want to calculate the ...Missing: seminal | Show results with:seminal
  12. [12]
    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.
  13. [13]
    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.<|separator|>
  14. [14]
    [1110.2122] A simple derivation of the Lindblad equation - arXiv
    Oct 10, 2011 · We start with the Born-Markov equation and, tracing out the bath degrees of freedom, we obtain an equation in the Lindblad form. The ...
  15. [15]
    None
    Nothing is retrieved...<|separator|>
  16. [16]
    [PDF] Lecture Notes for Ph219/CS219: Quantum Information Chapter 3
    Our next example is the dephasing channel, also called the phase-damping channel. ... The operators La are called Lindblad operators or quantum jump ...
  17. [17]
    [2402.06354] Taming the Bloch-Redfield equation: Recovering an ...
    Feb 9, 2024 · Abstract:Master equations play a pivotal role in investigating open quantum systems. In particular, the Bloch-Redfield equation stands out ...
  18. [18]
    An operational approach to quantum probability
    A mathmatical framework for the process of making repeated measurements on continuous observables in a statistical system.
  19. [19]
  20. [20]
    [math-ph/0102013] Entropy, von Neumann and the von ... - arXiv
    Feb 9, 2001 · This paper is an introduction to the von Neumann entropy in a historic approach. Von Neumann's gedanken experiment is repeated, which led him to the formula of ...
  21. [21]
    Proof of the strong subadditivity of quantum‐mechanical entropy
    Dec 1, 1973 · M. B. Ruskai, “A Generalization of the Entropy Using Traces on von Neumann Algebras,” preprint. O. Lanford III, in Statistical Mechanics and ...
  22. [22]
    Density matrix formulation of dynamical systems | Phys. Rev. E
    Nov 14, 2022 · We show that a classical density matrix is similar to the phase-space metric and evolves in time according to generalizations of Liouville's theorem and ...
  23. [23]
    Classical correspondence beyond the Ehrenfest time for open ...
    Dec 9, 2024 · The quantum and classical evolutions may be compared via the Wigner-Weyl representation. Due to decoherence, they are conjectured to match ...
  24. [24]
    Wigner Function for Harmonic Oscillator and The Classical Limit
    Apr 14, 2021 · In this paper, the classical limit of the Wigner function is shown using the quantum harmonic oscillator as an example.
  25. [25]
    Quantum revivals versus classical periodicity in the infinite square well
    Jan 1, 2001 · Given this qualitative difference, how can the classical motion possibly be the limit of the quantal time development? The resolution of this ...
  26. [26]
    [PDF] Algebraic Formulation of Quantum Theory
    Apr 17, 2012 · Extremal states on a. C⇤-algebra are called pure states. Non-pure states are called mixed states. Example: Density matrix. Consider the C ...
  27. [27]
    [PDF] On states of quantum theory - arXiv
    Jul 25, 2022 · A generalized quantum state is a state on a C∗−algebra that generalizes the notion of a density matrix in quantum mechanics. A density matrix ...
  28. [28]
    Gelfand-Naimark-Segal construction in nLab
    Sep 26, 2025 · The Gelfand-Naimark-Segal construction (“GNS construction”) represents a state on a star-algebra over the complex numbers.
  29. [29]
    [PDF] GNS and all that: a rough guide to algebras and states - LSE
    Apr 2, 2021 · This document begins with general motivations for the algebraic approach to quantum theory. Then it expounds the basics of this approach's ...
  30. [30]
    Notes on the type classification of von Neumann algebras - arXiv
    Feb 3, 2023 · These notes provide an explanation of the type classification of von Neumann algebras, which has made many appearances in recent work on entanglement in ...
  31. [31]
    [PDF] an introduction to von neumann algebras - NSF PAR
    Any factor M of type I is ∗-isomorphic to B(K), for some Hilbert space K. Proof. Let p ∈ M be a nonzero abelian projection. Then pMp is both abelian and a ...
  32. [32]
    [PDF] Equilibrium: KMS states
    A pair (A,тt) is a C*-dynamical system if A is a C*-algebra with an identity and R 3 t 7! тt is a strongly continuous group of *-automorphisms of A, namely.
  33. [33]
    [PDF] KMS states of C*-dynamical systems - Zero-Dimensional Symmetry
    Mar 1, 2019 · This is the KMS condition; it characterizes equilibrium for finite systems, and is the definition of equilibrium state at inverse temperature β ...
  34. [34]
    [PDF] Tomita-Takesaki Modular Theory vs. Quantum Information ... - arXiv
    Jul 14, 2016 · This idea is implicit in proofs based on Araki's relative modular operator. 6.3 Operator extension of strong subadditivity of entropy. Following ...
  35. [35]
    [PDF] Algebraic quantum field theory - Universität Hamburg
    Algebraic quantum field theory involves quantum systems with many degrees of freedom, and the algebra of observables is an associative involutive complex  ...
  36. [36]
    [1501.00234] Local state and sector theory in local quantum physics
    Jan 1, 2015 · We define a new concept of local states in the framework of algebraic quantum field theory (AQFT). Local states are a natural generalization of ...
  37. [37]
    [PDF] Algebraic Quantum Field Theory - PhilSci-Archive
    Feb 14, 2006 · Algebraic quantum field theory provides a general, mathematically precise description of the structure of quantum field theories.
  38. [38]
    Quantum channels and memory effects | Rev. Mod. Phys.
    Dec 10, 2014 · Any physical process can be represented as a quantum channel mapping an initial state to a final state.
  39. [39]
    Optimization of entanglement witnesses | Phys. Rev. A
    Oct 16, 2000 · Optimization of entanglement witnesses. M. Lewenstein1, B. Kraus2, J. I. Cirac2, and P. Horodecki3. 1Institute for Theoretical Physics ...
  40. [40]
    Variational quantum state eigensolver | npj Quantum Information
    Sep 21, 2022 · Here, we address the case when the matrix is a density matrix ρ. We introduce the variational quantum state eigensolver (VQSE), which is ...
  41. [41]
    Fast Quantum State Tomography in the Nitrogen Vacancy Center of ...
    To recover the full density matrix, we use a set of unitary operations that transform the density operator elements into the directly measurable observable. We ...
  42. [42]
    [PDF] Information Theory and Statistical Mechanics
    Secondly, the paradox regarding the entropy of com- bined systems, which was resolved only by adoption of the generic instead of the specific definition of ...Missing: historical | Show results with:historical
  43. [43]
    Wigner quasiprobability distribution - Wikipedia
    For instance, the Wigner distribution can and normally does take on negative values for states which have no classical model—and is a convenient indicator of ...Relation to classical mechanics · Mathematical properties · Positivity of the function
  44. [44]
    [PDF] A Brief History of the GKLS Equation - arXiv
    Nov 4, 2017 · The articles written by Vittorio Gorini, Andrzej Kossakowski and George Sudarshan (GKS) [2], and Göran Lindblad [3] belong to the list of the ...
  45. [45]
    [PDF] Lecture 18 -- Quantum Information Theory and Holevo's Bound
    Nov 10, 2015 · In this lecture, we will motivate and formally state Holevo's Theorem [Hol73], which does a good job of answering our question. We begin by ...Missing: 1980s Bennett reversible
  46. [46]
    Decoherence, einselection, and the quantum origins of the classical
    May 22, 2003 · Decoherence, caused by environment interaction, leads to einselection, a quantum process that enforces classicality by selective loss of ...
  47. [47]
    Fluctuation theorems for quantum processes | Phys. Rev. E
    Sep 30, 2013 · We present fluctuation theorems and moment generating function equalities for generalized thermodynamic observables and quantum dynamics.
  48. [48]
    Probing the Topology of Density Matrices | Phys. Rev. X
    Feb 28, 2018 · In this work, we explore a distinct notion of parallel transport for mixed states within the comparatively simple class of symmetry-protected ...
  49. [49]
    Experimental Sample-Efficient Quantum State Tomography via ...
    Our results show that PQST reduces measurement cost, achieving fidelities of 98.68% and 95.07% after measuring 75 and 99 observables for six-qubit and nine- ...Missing: 2020s | Show results with:2020s