Fact-checked by Grok 2 weeks ago

Quasiprobability distribution

A quasiprobability distribution is a phase-space representation of a quantum mechanical density operator that generalizes classical probability distributions by permitting negative, singular, or complex values, thereby capturing inherently quantum features such as interference and nonclassical correlations that cannot be described by positive probabilities alone. These distributions serve as powerful tools for visualizing and analyzing quantum states, bridging the gap between classical statistical mechanics and quantum theory, and are particularly valuable in identifying nonclassicality—states where the distribution fails to admit a positive classical interpretation. The foundational Wigner quasiprobability distribution, introduced by Eugene Wigner in 1932, provides a symmetric formulation in position-momentum phase space and integrates to yield expectation values of observables, though its negativity signals quantum delocalization. In quantum optics, quasiprobability distributions gained prominence with the Glauber-Sudarshan P representation, independently developed by Roy J. Glauber and E. C. G. Sudarshan in 1963 to describe coherent and thermal light fields via diagonal expansions in coherent states; negativity or singularities in the P function indicate nonclassical light, such as squeezed or entangled photon states. Complementing this, the Husimi Q function, proposed by Kôdi Husimi in 1940, offers a smoothed, always non-negative representation obtained by convolving the density matrix with a Gaussian coherent state, useful for bounding quantum uncertainties but at the cost of added noise. The Wigner function remains central due to its balance of properties, enabling direct comparisons to classical phase-space flows while revealing quantum corrections through its oscillatory behavior. Beyond optics, quasiprobability distributions extend to quantum information science for quantifying coherence and entanglement, quantum thermodynamics for defining work and heat via characteristic functions, and foundational studies of quantum measurement, where negative values underscore the impossibility of simultaneous sharp position-momentum measurements. Their experimental reconstruction, via quantum state tomography or interferometric techniques, has advanced with modern quantum technologies, confirming theoretical predictions of nonclassicality in systems like superconducting circuits and trapped ions.

Fundamentals

Definition and Motivation

In classical mechanics, phase space provides a framework for describing the probabilistic evolution of systems using positive probability distributions over position and momentum coordinates. Quantum mechanics, however, introduces phenomena such as superposition and interference that cannot be adequately represented by such classical distributions, necessitating alternative formulations to visualize quantum states in a phase-space-like manner. Quasiprobability distributions emerged as a solution to this challenge, originating in the 1930s with Eugene Wigner's introduction of a phase-space representation to investigate quantum corrections to classical statistical mechanics and thermodynamic equilibrium. This approach was later extended in the 1940s by Kôdi Husimi for density matrix properties and significantly developed in the 1960s by Roy J. Glauber and E. C. G. Sudarshan within quantum optics to describe coherent states and light fields. The primary motivation for these distributions is to bridge the gap between quantum and classical descriptions, enabling the computation of expectation values and dynamics in a familiar phase-space setting while accommodating inherently quantum features like nonlocality. Such negativity serves as an indicator of nonclassical behavior, distinguishing quantum systems from their classical counterparts. Characteristic functions often generate these distributions via Fourier transforms, providing a unified mathematical framework. At their core, quasiprobability distributions represent a quantum state given by the density operator ρ through a function W(α), where α is a complex variable parameterizing the phase space (often interpreted as a scaled position-momentum pair). These distributions satisfy the normalization condition ∫ W(α) d²α = Tr(ρ) = 1, ensuring they reproduce the total probability of the state, yet they may take negative or complex values, violating the positivity axiom of classical probabilities and thereby encoding quantum interference effects.

Characteristic Functions

In quantum optics, the characteristic function provides a Fourier transform representation of the density operator ρ, facilitating the derivation of quasiprobability distributions in phase space. The symmetric characteristic function is defined as χ(ξ) = Tr[ρ exp(ξ a† - ξ* a)], where a and a† denote the annihilation and creation operators, respectively, satisfying the bosonic commutation relation [a, a†] = 1. This form arises from the symmetrically ordered expansion of the density operator and corresponds to the case of Weyl ordering. Variants of the characteristic function incorporate different operator orderings, such as normal or antinormal, to generate corresponding quasiprobabilities. The symmetric version can be expressed using the displacement operator as χ_sym(ξ) = Tr[ρ D(ξ)], where D(ξ) = exp(ξ a† - ξ* a) is the coherent displacement operator. In contrast, normally ordered (antinormally ordered) versions involve placing creation (annihilation) operators to the left (right) in the expansion, leading to distinct phase-space representations. A unified framework employs an ordering parameter s to parameterize these functions, yielding the s-ordered characteristic function χ(ξ; s) = Tr[ρ D(ξ)] exp\left( \frac{s}{2} |\xi|^2 \right), where s ∈ [-1, 1]. For s = 0, this reduces to the symmetric case associated with Weyl ordering; s = 1 corresponds to normal ordering, and s = -1 to antinormal ordering. Quasiprobability distributions are obtained from the s-ordered characteristic function via a Fourier transform: W(\alpha; s) = \frac{1}{\pi^2} \int d^2 \xi \, \chi(\xi; s) \exp(\alpha \xi^* - \alpha^* \xi), where α is a complex phase-space variable. This integral representation allows continuous interpolation between different quasiprobabilities, with the parameter s controlling the degree of smoothing in the distribution.

Specific Distributions

Wigner Function

The Wigner function represents the quasiprobability distribution corresponding to the s=0 ordering parameter within the family of s-ordered quasiprobability distributions introduced by Cahill and Glauber. Originally formulated by Eugene Wigner in 1932 as a phase-space analogue to classical Liouville distributions for quantum statistical mechanics, it enables the computation of expectation values of symmetrically ordered operators via phase-space integrals. The explicit expression for the Wigner function W(\alpha) in the complex phase-space variable \alpha (where \alpha = (q + i p)/\sqrt{2\hbar} in quadrature coordinates with \hbar = 1) is given by W(\alpha) = \frac{1}{\pi} \int d^2\beta \, \langle \alpha + \beta | \rho | \alpha - \beta \rangle \exp\left(2i \Im(\alpha^* \beta)\right), where \rho is the density operator and the integral runs over the complex plane of the displacement parameter \beta. Equivalently, it arises as the Fourier transform of the s=0 characteristic function: W(\alpha) = \frac{1}{\pi^2} \int d^2\beta \, \chi^{(0)}(\beta) \exp(\alpha \beta^* - \alpha^* \beta), with \chi^{(0)}(\beta) = \Tr[\rho \, D(\beta)], and D(\beta) = \exp(\beta a^\dagger - \beta^* a) the displacement operator for the bosonic mode annihilation operator a. Key properties of the Wigner function include its reality, W(\alpha) = W^*(\alpha), and normalization \int d^2\alpha \, W(\alpha) = 1, ensuring it acts as a properly normalized phase-space density despite deviations from positivity. For pure states |\psi\rangle, the marginals yield the standard quantum probability distributions: \int W(x, p) \, dx = |\psi(p)|^2 (momentum distribution) and \int W(x, p) \, dp = |\psi(x)|^2 (position distribution), where x and p are scaled quadratures. As a phase-space representation, the Wigner function visualizes quantum states with oscillatory interference patterns akin to classical densities but incorporating quantum corrections through negative regions, which arise from wavefunction overlaps and signal nonclassical behavior. While the signed integral over phase space equals 1 (or $1/(\pi \hbar) in dimensional form with \hbar restored), the allowance for negatives implies that the total positive volume exceeds this value, quantifying the extent of quantum interference. For computing the Wigner function of simple states, such as Fock or coherent states, one evaluates the defining integral or trace using the known density matrix elements \rho_{mn} = \langle m | \rho | n \rangle in the number basis, often leveraging generating functions involving associated Laguerre polynomials for analytic forms; for instance, the vacuum state yields W(\alpha) = (2/\pi) \exp(-2 |\alpha|^2).

Husimi Q-Function

The Husimi Q-function, also known as the Q-representation, is a quasiprobability distribution defined for a quantum state described by the density operator \rho as Q(\alpha) = \frac{1}{\pi} \langle \alpha | \rho | \alpha \rangle, where \alpha is a complex number labeling the coherent state |\alpha\rangle and \langle \alpha | is its bra counterpart. This formulation arises directly from the overlap of the density operator with the projector onto the coherent state, providing a phase-space representation that smooths quantum features through the inherent uncertainty of these states. The Q-function possesses several key properties that distinguish it from other quasiprobability distributions. It is always non-negative, Q(\alpha) \geq 0, because the density operator \rho is positive semidefinite, ensuring \langle \alpha | \rho | \alpha \rangle \geq 0 for all coherent states. Additionally, it is normalized such that \int Q(\alpha) \, d^2\alpha = 1, reflecting the trace condition \operatorname{Tr}(\rho) = 1 and the overcomplete basis property of coherent states with the measure d^2\alpha / \pi. However, the Q-function is inherently broader than classical probability distributions due to the vacuum fluctuations encoded in the coherent states, which introduce an irreducible Gaussian smoothing; this results in a lower bound on position and momentum uncertainties that exceeds the Heisenberg limit by a factor related to the vacuum noise variance of $1/2. Within the general framework of s-parameterized quasiprobability distributions, the Husimi Q-function corresponds to the case of antidiagonal or anti-normal ordering, characterized by the parameter s = -1. This ordering arises from the normally ordered expansion of the density operator using creation and annihilation operators, where anti-normal ordering reverses the sequence, leading to a representation that prioritizes the smoothing effect over sharp marginal distributions. The always-positive nature of the Q-function offers significant advantages for practical applications in quantum optics. It serves as a bona fide probability distribution, facilitating intuitive visualization of quantum states in phase space without the interpretive challenges posed by negative regions in other distributions. Furthermore, its direct measurability through techniques like balanced homodyne detection makes it particularly valuable for quantum state tomography, where reconstructing the full distribution from experimental data enables reliable inference of state properties.

Glauber-Sudarshan P-Function

The Glauber–Sudarshan P-function, also known as the P-representation, provides a phase-space formulation of the density operator in quantum optics by expanding it in the overcomplete basis of coherent states. Introduced independently by Roy J. Glauber and E. C. G. Sudarshan in 1963, it was developed to describe statistical properties of light beams using coherent state expansions, bridging quantum mechanical and semiclassical descriptions of optical fields. The P-function P(\alpha) is defined such that the density operator \rho of a quantum state is given by \rho = \int d^2 \alpha \, P(\alpha) \, |\alpha \rangle \langle \alpha |, where \alpha is a complex number labeling the coherent states |\alpha \rangle, and the integral extends over the entire complex plane. This representation corresponds to the s = 1 case in the general family of s-ordered quasiprobability distributions. For classical optical states, such as coherent or thermal states, P(\alpha) is non-negative and properly normalized, \int d^2 \alpha \, P(\alpha) = 1, allowing an interpretation as a classical probability density for the field's complex amplitude. However, for nonclassical states, P(\alpha) exhibits quantum features, including negative regions or singularities more irregular than delta functions, violating the conditions of a valid classical probability distribution. In particular, for pure Fock states |n \rangle with n \geq 1, the P-function becomes highly singular, often expressed in terms of derivatives of delta functions, necessitating regularization techniques like truncation or smoothing for numerical evaluation. This ill-behaved nature arises because Fock states cannot be represented as mixtures of coherent states without negative weights. The P-function is linked to the photon number distribution through its moments, where the probability of detecting n photons is obtained via p(n) = \int d^2 \alpha \, P(\alpha) e^{-|\alpha|^2} \frac{|\alpha|^{2n}}{n!}, highlighting its utility in computing normally ordered expectation values. The P-function relates to the Husimi Q-function through deconvolution, expressed as a Gaussian convolution: Q(\alpha) = \frac{1}{\pi} \int d^2 \beta \, P(\beta) \, \exp(-|\alpha - \beta|^2), which smooths the potentially singular P into a positive, bounded distribution. This inverse relationship underscores the P-function's role as an "unsmoothed" representation, amplifying nonclassical effects but complicating direct measurement.

Properties and Interpretations

Ordering and Smoothing Parameters

The s-ordered quasiprobability distributions form a continuous family parameterized by a real number s, unifying various phase-space representations of quantum states. The general expression for the s-ordered distribution W(\alpha; s) associated with a density operator \rho is given by W(\alpha; s) = \frac{1}{\pi} \int d^2\lambda \, \exp\left(-\frac{2|\lambda|^2}{s}\right) \langle \alpha + \lambda | \rho | \alpha - \lambda \rangle, valid for s > 0, or equivalently through the s-ordered characteristic function \chi(\xi; s) = \operatorname{Tr}[\rho D(\xi)] \exp\left(\frac{s}{2} |\xi|^2 \right), where D(\xi) is the displacement operator, yielding W(\alpha; s) = \frac{1}{\pi^2} \int d^2\xi \, \chi(\xi; s) \exp(\alpha \xi^* - \alpha^* \xi). These forms, introduced by Cahill and Glauber, allow interpolation between different distributions while preserving normalization \int d^2\alpha \, W(\alpha; s) = 1. The parameter s tunes the degree of smoothing and the associated operator ordering in the expansion of \rho. Specifically, s = 1 corresponds to the normal-ordered Glauber-Sudarshan P-function, s = 0 to the symmetrically ordered Wigner function, and s = -1 to the anti-normally ordered Husimi Q-function; values of s outside [-1, 1] extend the family but may not yield valid representations for all states. As s decreases from 1 to -1, the distributions become progressively smoother due to implicit Gaussian broadening, with negativity or singularities in W(\alpha; s) for nonclassical states increasing in prominence for higher s (less smoothing) and diminishing toward positivity for lower s. This smoothing manifests as Gaussian convolutions between distributions; for instance, the Q-function is the Wigner function convolved with a unit-variance Gaussian, Q(\alpha) = \int d^2\beta \, W(\beta) \frac{2}{\pi} \exp(-2|\alpha - \beta|^2). The s-parameter directly links to operator ordering conventions in quantum optics. Normal ordering (s=1) places creation operators to the left of annihilation operators, anti-normal ordering (s=-1) reverses this, and symmetric (Weyl) ordering (s=0) balances them, facilitating evaluations of expectation values like \operatorname{Tr}[\rho \{a^\dagger, a\}^s]. Related parameterizations, such as the t-ordered form with t = s, appear in extensions for computing time-ordered correlation functions, maintaining the same interpretive framework. Specific distributions like the P-, Wigner, and Q-functions emerge as special cases at these canonical values of s.

Indicators of Non-Classicality

Quasiprobability distributions serve as powerful tools for detecting non-classical features in quantum states, as their deviations from classical probability distributions—such as negativity or failure to remain non-negative—signal behaviors impossible in classical optics. In particular, the Wigner function exhibits negative regions for states like Fock states or squeezed vacuum, quantifying quantum interference that has no classical analog. The volume of these negative regions provides a measure of non-classicality, with larger volumes indicating stronger quantum signatures; for instance, the integrated negativity serves as a resource quantifier in quantum information tasks. A foundational criterion for classicality stems from the Glauber-Sudarshan P-function: a quantum state is classical if and only if its P-representation is non-negative everywhere, as positive P-functions correspond to mixtures of coherent states describable by classical stochastic fields. States with non-positive (negative or singular in a non-probabilistic sense) P-functions, such as single-photon or higher Fock states, thus exhibit non-classicality by violating this positivity condition. This criterion underscores how quasiprobabilities extend beyond classical limits, enabling identification of quantum correlations. Mandel's Q parameter further probes non-classicality through photon-number statistics, defined as Q = \frac{\langle (\Delta n)^2 \rangle - \langle n \rangle}{\langle n \rangle}, where n is the photon number; negative Q values indicate sub-Poissonian statistics, a hallmark of non-classical light without classical counterparts, as seen in resonance fluorescence or single-photon sources. This parameter relates to the variance in the P-function representation, where deviations from Poissonian behavior (Q=0 for coherent states) highlight quantum squeezing or antibunching. For higher-order non-classicality, the volume of negative regions in the Wigner function offers a quantitative criterion beyond mere presence of negativity, capturing the extent of quantum interference and entanglement in multi-mode systems. Similarly, the Husimi Q-function, while always non-negative, can violate classical moment bounds—such as inequalities on higher-order correlations—that hold for classical states, providing indirect evidence of non-classicality through phase-space inequalities. These measures extend to resource quantification, where negativity volumes bound the utility in quantum protocols. Post-2000 developments have integrated these indicators into practical applications, such as certifying non-classical states in quantum computing via phase-space inequalities on reconstructed quasiprobabilities, enabling resource verification without full tomography. In optomechanics, Wigner negativity volumes assess mechanical quantum states, confirming non-classical motion in hybrid systems and advancing continuous-variable quantum technologies. Ongoing research refines these tools for scalable quantum networks.

Dynamics and Applications

Time Evolution

The time evolution of quasiprobability distributions in open quantum systems follows from the Lindblad master equation, \dot{\rho} = -i[H, \rho] + \sum_k \left( L_k \rho L_k^\dagger - \frac{1}{2} \{L_k^\dagger L_k, \rho\} \right), where H is the system Hamiltonian and the L_k represent dissipative interactions with the environment. To derive the dynamics, the equation is transformed using the s-ordered characteristic function \chi_s(\xi, t) = \Tr\left[\rho(t) \exp(\xi a^\dagger - \xi^* a)\right] \exp\left(\frac{s}{2} |\xi|^2\right), yielding a linear evolution equation for \chi_s that depends on the forms of H and the L_k. The quasiprobability distribution W_s(\alpha, t) is recovered via the Fourier transform W_s(\alpha, t) = \frac{1}{\pi^2} \int d^2\xi \, \chi_s(\xi, t) \exp\left(\alpha \xi^* - \alpha^* \xi - \frac{s}{2} |\xi|^2\right). For the Wigner function (s=0), the evolution takes the form of a partial differential equation combining classical-like advection from the Hamiltonian with quantum jumps from dissipation: \frac{\partial W}{\partial t} = \{H_\mathrm{cl}, W\}_P + \sum_k \left[ \text{drift and diffusion terms from } L_k \right], where \{ \cdot, \cdot \}_P denotes the Poisson bracket with the classical Hamiltonian H_\mathrm{cl}. In the case of a damped harmonic oscillator at zero temperature (H = \omega (a^\dagger a + 1/2), L = \sqrt{\gamma} a), the equation in complex phase-space variables \alpha is \frac{\partial W}{\partial t} = -(\gamma/2 + i\omega) \left( \alpha \frac{\partial W}{\partial \alpha} + \alpha^* \frac{\partial W}{\partial \alpha^*} \right) + \frac{\gamma}{2} \frac{\partial^2 W}{\partial \alpha \partial \alpha^*}, capturing both the oscillatory motion and damping toward the vacuum state. For general s-parameterized distributions, the evolution equations are Fokker--Planck-like, with s-dependent drift vector A^{(s)}(\alpha) and diffusion matrix D^{(s)}(\alpha): \frac{\partial W_s}{\partial t} = -\sum_i \frac{\partial}{\partial \alpha_i} \left[ A_i^{(s)}(\alpha) W_s \right] + \frac{1}{2} \sum_{i,j} \frac{\partial^2}{\partial \alpha_i \partial \alpha_j^*} \left[ D_{ij}^{(s)}(\alpha) W_s \right]. The diffusion terms scale with (1 - s), leading to increased smoothing for s < 0 (e.g., Husimi Q-function) and potential higher-order derivatives for intermediate s, which are often truncated to yield valid stochastic interpretations. These dynamics preserve the normalization \int W_s(\alpha, t) \, d^2\alpha = 1 for all s and t, reflecting \Tr[\rho(t)] = 1. However, the integrated negativity of the Wigner function, quantifying nonclassical features, can dynamically increase under certain non-Markovian or coherent processes or decrease due to decoherence in open systems.

Operator Correspondences

The Wigner-Weyl transform establishes a correspondence between classical phase-space functions and quantum operators, providing a foundational mapping for quasiprobability distributions. Originally proposed by Weyl in 1927 as a quantization rule and extended by Wigner in 1932 to define a phase-space representation of quantum states, the transform quantizes a classical symbol f(x, p) into an operator \hat{O} via integration over phase space. In the context of quantum optics, using complex phase-space variables \alpha, this takes the form \hat{O} = \int d^2\alpha \, f(\alpha) \Delta(\alpha), where \Delta(\alpha) is the Stratonovich-Weyl kernel. For the symmetric case corresponding to the Wigner function (s=0), the kernel is given by \Delta(\alpha) = \frac{2}{\pi} D(2\alpha) (-1)^{a^\dagger a} D^\dagger(2\alpha), with D(\alpha) the displacement operator D(\alpha) = \exp(\alpha a^\dagger - \alpha^* a), ensuring the mapping preserves traces and expectation values. This kernel facilitates the inversion, yielding the Wigner quasiprobability W(\alpha) from the density operator \hat{\rho} as the Weyl symbol. The transform's unitarity in the trace inner product guarantees a one-to-one correspondence between bounded operators and suitable phase-space functions. Generalizations to s-parameterized quasiprobabilities, introduced by Cahill and Glauber in 1969, extend this correspondence to different operator orderings, where s controls the smoothing: s=1 for normal ordering (Glauber-Sudarshan P-function), s=0 for Weyl (symmetric, Wigner), and s=-1 for antinormal (Husimi Q-function). The generalized kernel incorporates a Gaussian convolution factor \exp\left(\frac{s}{2} |\alpha - \beta|^2\right), linking the quasiprobability W_s(\alpha) to the s-ordered characteristic function. This framework connects to *-product quantization, where the Moyal product defines non-commutative multiplication in phase space, and the Moyal bracket [f, g]_M = \frac{f \star g - g \star f}{i\hbar} (with \star the bidirectional arrow product) captures quantum commutators as Poisson-like brackets in the semiclassical limit. Applications of these correspondences include quantizing classical Hamiltonians by specifying operator ordering, resolving ambiguities in products like \hat{x}\hat{p} through Weyl symmetrization (\hat{x}\hat{p} + \hat{p}\hat{x})/2, which aligns with the correspondence principle for large quantum numbers n \gg 1 or small \hbar. In multimode systems, the transform extends to entangled states via tensor products of single-mode kernels, enabling phase-space representations of multipartite correlations, such as in continuous-variable quantum information protocols. These mappings underpin time evolution in phase space but focus here on static quantization structures.

Quantum Harmonic Oscillator Examples

The quantum harmonic oscillator provides a fundamental testing ground for quasiprobability distributions, as its states and dynamics reveal key features of quantum nonclassicality and decoherence. Coherent states, Fock states, and damped evolutions illustrate how these distributions capture both classical-like and quantum behaviors, with the Wigner function often highlighting negativities absent in true probability distributions. For a coherent state |\alpha_0\rangle, the Wigner function is a Gaussian centered at \alpha_0 in phase space, given by W(\alpha) = \frac{2}{\pi} \exp\left(-2|\alpha - \alpha_0|^2\right), which remains nonnegative and resembles a classical probability distribution for the oscillator's position and momentum. The Husimi Q-function for this state is always positive, forming a smoothed Gaussian Q(\alpha) = \frac{1}{\pi} \exp\left(-|\alpha - \alpha_0|^2\right), while the Glauber-Sudarshan P-function is delta-like, P(\alpha) = \delta^{(2)}(\alpha - \alpha_0), reflecting the state's perfect classical correspondence. These features underscore the coherent state's role as the quantum analog of a classical driven oscillator. In contrast, Fock states |n\rangle (energy eigenstates) exhibit nonclassical traits through the Wigner function, which incorporates Laguerre polynomials and displays oscillatory patterns for n \geq 1, leading to regions of negativity that signal quantum interference. The explicit form is W(\alpha) = \frac{2 (-1)^n}{\pi} \exp\left(-2|\alpha|^2\right) L_n\left(4|\alpha|^2\right), where L_n denotes the n-th Laguerre polynomial; for n=0, it reduces to the vacuum Gaussian, but higher n introduce negative lobes, quantifying nonclassicality via the volume of negative regions. The Q-function for Fock states remains nonnegative, Q(\alpha) = \frac{\exp(-|\alpha|^2) |\langle \alpha | n \rangle|^2}{\pi}, with |\langle \alpha | n \rangle|^2 = e^{-|\alpha|^2} \frac{|\alpha|^{2n}}{n!}, resembling a Poissonian intensity distribution smoothed by the coherent state overlap. The P-function for |n\rangle is highly singular, involving derivatives of deltas, further emphasizing the state's departure from classical optics. Under damping, as modeled by a Lindblad master equation for the oscillator coupled to a thermal bath, quasiprobabilities evolve to reveal decoherence. Starting from an initial coherent state, the Wigner function undergoes diffusion, with its Gaussian width increasing due to noise terms in the Fokker-Planck equation, while the center spirals inward following classical damped trajectories. Over time, this spreads the distribution toward a thermal state, characterized by a broader, centered Gaussian at steady state, marking the loss of initial coherence. For Fock-like initial states, the oscillatory negativities in the Wigner function decay exponentially under zero-temperature damping, with the negativity measure \eta_W = \int \left( |W| - W \right)^+ d^2\alpha vanishing at a timescale t^* \sim \gamma^{-1} \ln\left( \frac{2\bar{n}+2}{2\bar{n}+1} \right), where \gamma is the damping rate and \bar{n} the bath occupancy, transitioning the system to classical statistics. This qualitative behavior—negativity suppression via broadening and thermalization—highlights damping's role in erasing quantum signatures.

References

  1. [1]
    [PDF] arXiv:1907.12427v2 [quant-ph] 10 Feb 2020
    Feb 10, 2020 · First, in Section 2, we present a more comprehensive review that demonstrates how quasiprobability distributions became a standard tool for ...
  2. [2]
  3. [3]
  4. [4]
    Optical experiment to test negative probability in context of quantum ...
    Dec 13, 2019 · Known as a quasiprobability distribution, they are regarded as a useful tool that provides significant insight into the underlying fundamentals ...
  5. [5]
    Quasiprobabilities in quantum thermodynamics and many-body ...
    In this tutorial, we present the definition, interpretation and properties of some of the main quasiprobabilities that can describe the statistics of ...
  6. [6]
  7. [7]
    Coherent and Incoherent States of the Radiation Field | Phys. Rev.
    Coherent and Incoherent States of the Radiation Field. Roy J. Glauber. Lyman Laboratory of Physics, Harvard University, Cambridge, Massachusetts. PDF Share. X ...
  8. [8]
    Equivalence of Semiclassical and Quantum Mechanical ...
    Equivalence of Semiclassical and Quantum Mechanical Descriptions of Statistical Light Beams. E. C. G. Sudarshan. Department of Physics and Astronomy ...
  9. [9]
    Experimental determination of a nonclassical Glauber-Sudarshan - P
    Aug 25, 2008 · In the zero-temperature limit, the SPATS includes the special case of the one-photon Fock state with the highly singular P function given in Eq.
  10. [10]
    Negativity of Quasiprobability Distributions as a Measure of ...
    Mar 19, 2020 · We demonstrate that the negative volume of any s-parametrized quasiprobability, including the Glauber-Sudashan P function, can be consistently defined.
  11. [11]
    Density Operators and Quasiprobability Distributions | Phys. Rev.
    The paper studies expanding density operators to simplify quantum expectation values, discussing weight functions, Wigner distribution, and a function related ...
  12. [12]
    Negativity of the Wigner function as an indicator of nonclassicality
    Jun 2, 2004 · Abstract: A measure of nonclassicality of quantum states based on the volume of the negative part of the Wigner function is proposed.
  13. [13]
    Observation of Sub-Poissonian Photon Statistics | Phys. Rev. Lett.
    Aug 1, 1983 · Observation of Sub-Poissonian Photon Statistics. R. Short and L. Mandel. Department of Physics and Astronomy, University of Rochester ...Missing: paper | Show results with:paper
  14. [14]
    Experimental Certification of Nonclassicality via Phase-Space ...
    Jan 15, 2021 · For the implementation of the nonclassicality inequality conditions, we reconstruct the Wigner and Husimi Q functions of the states via quantum- ...Missing: bounds | Show results with:bounds<|control11|><|separator|>
  15. [15]
    A short introduction to the Lindblad master equation | AIP Advances
    Feb 4, 2020 · The Lindblad (or Gorini-Kossakowski-Sudarshan-Lindblad) master equation plays a key role as it is the most general generator of Markovian dynamics in quantum ...
  16. [16]
    Wigner distribution function for the harmonic oscillator with ... - arXiv
    Oct 12, 1993 · We deduce the density matrix of the damped harmonic oscillator from the solution of the Fokker-Planck equation for the coherent state representation.
  17. [17]
  18. [18]