Wigner quasiprobability distribution
The Wigner quasiprobability distribution, often referred to as the Wigner function, is a real-valued phase-space representation of the quantum state of a system, defined as a function W(x, p) of position x and momentum p that generalizes classical probability distributions to quantum mechanics while allowing for negative values to capture non-classical behaviors such as interference.[1][2] Introduced by physicist Eugene Wigner in 1932, it was originally formulated to compute quantum corrections to classical thermodynamic equilibrium distributions, bridging wave mechanics with statistical mechanics by transforming the density operator into a phase-space integral.[1][3] Mathematically, for a pure state described by wave function \psi(x), the Wigner function is given by W(x, p) = \frac{1}{\pi \hbar} \int_{-\infty}^{\infty} \psi^*\left(x + \frac{y}{2}\right) \psi\left(x - \frac{y}{2}\right) e^{2 i p y / \hbar} \, dy, or more generally for a mixed state with density operator \rho, W(x, p) = \frac{1}{\pi \hbar} \int_{-\infty}^{\infty} \langle x + y | \rho | x - y \rangle e^{2 i p y / \hbar} \, dy, where \hbar is the reduced Planck's constant; this expression ensures the function is normalized such that \int W(x, p) \, dx \, dp = 1.[2][3] Key properties include its marginal distributions recovering the correct quantum probability densities for position and momentum—specifically, \int W(x, p) \, dp = |\psi(x)|^2 and \int W(x, p) \, dx = |\tilde{\psi}(p)|^2, where \tilde{\psi}(p) is the momentum-space wave function—while its negativity, which violates the non-negativity of classical probabilities, serves as a witness for quantum non-classicality.[2][3] Additionally, the Wigner function is invariant under unitary transformations corresponding to classical canonical transformations and satisfies a quantum Liouville equation for time evolution, facilitating simulations of quantum dynamics.[3] Beyond its foundational role in quantum statistical mechanics, the Wigner distribution has become a cornerstone in fields like quantum optics, where it reveals phase-space portraits of light fields including coherent, squeezed, and entangled states; solid-state physics, for modeling electron transport in nanostructures; and quantum information science, aiding in the analysis of quantum computing gates and error correction via negativity measures.[2][3] Its ability to highlight quantum-classical boundaries has also spurred developments in experimental tomography, where direct measurements reconstruct W(x, p) from homodyne detection or weak measurements, enabling visualizations of quantum superpositions in systems like harmonic oscillators and photons.[2] Despite challenges posed by negativity in interpretation, the Wigner function remains a powerful tool for intuition and computation in modern quantum theory.[3]Introduction
Definition
The Wigner quasiprobability distribution serves as a phase-space formulation of quantum mechanics, representing quantum states using continuous variables x for position and p for momentum, which act as analogs to the classical phase-space coordinates. This distribution enables the visualization and analysis of quantum states in a manner reminiscent of classical statistical mechanics, while incorporating inherently quantum features.[4] For a pure quantum state described by the wave function \psi(x), the Wigner distribution is defined as W(x,p) = \frac{1}{\pi \hbar} \int_{-\infty}^{\infty} \mathrm{d}y \, \psi^*(x+y) \psi(x-y) \exp\left(\frac{2i p y}{\hbar}\right). This expression originates from the Weyl ordering prescription, which symmetrizes products of non-commuting position and momentum operators to correspond to classical phase-space functions.[4] The definition extends naturally to mixed quantum states via the density operator \rho, yielding W(x,p) = \frac{1}{\pi \hbar} \int_{-\infty}^{\infty} \langle x + y | \rho | x - y \rangle e^{2 i p y / \hbar} \, \mathrm{d}y. This formulation using the density matrix elements preserves the connection to the Weyl transform and applies to thermal or ensemble-averaged states.[5] The Wigner distribution satisfies the normalization condition \iint W(x,p) \, \mathrm{d}x \, \mathrm{d}p = 1, ensuring it functions as a quasi-probability density over phase space, with units such that W(x,p) has dimensions of inverse action (inverse of \hbar).[4]Relation to Classical Mechanics
In classical statistical mechanics, the phase-space probability density \rho_{\mathrm{cl}}(x, p) describes the statistical distribution of a system's position x and momentum p, evolving according to Liouville's equation \partial \rho_{\mathrm{cl}} / \partial t = \{H, \rho_{\mathrm{cl}}\}_{\mathrm{PB}}, where \{ \cdot, \cdot \}_{\mathrm{PB}} denotes the Poisson bracket and H(x, p) is the classical Hamiltonian. This density is nonnegative and normalized such that its marginals recover the position and momentum probability densities: \int \rho_{\mathrm{cl}}(x, p) \, dp = |\psi(x)|^2 and \int \rho_{\mathrm{cl}}(x, p) \, dx = |\tilde{\psi}(p)|^2, where \psi(x) and \tilde{\psi}(p) are the classical analogs of wave functions. Such a formulation allows for a unified treatment of statistical ensembles in phase space, facilitating comparisons between deterministic trajectories and probabilistic descriptions.[1] The Wigner quasiprobability distribution arises as a quantum mechanical analog to this classical \rho_{\mathrm{cl}}(x, p), motivated by the need to extend phase-space methods to quantum systems where wave functions and operators replace classical variables. In his seminal 1932 work, Eugene Wigner sought to formulate quantum corrections to thermodynamic equilibrium distributions, aiming to represent quantum states in a phase-space picture that mirrors classical statistical mechanics while accounting for noncommutativity.[1] This approach enables the computation of expectation values as phase-space integrals, similar to classical mechanics, but adapted to quantum density operators \hat{\rho}. Central to this analogy is the Weyl quantization procedure, which maps classical phase-space functions to quantum operators via symmetric (Weyl) ordering, ensuring that the Wigner distribution serves as the phase-space representation—or Weyl symbol—of \hat{\rho}. Wigner employed this symmetric ordering to define his distribution, allowing it to reproduce classical limits while incorporating quantum effects through a systematic expansion in powers of \hbar.[1] Consequently, the Wigner distribution shares the marginal properties of \rho_{\mathrm{cl}}(x, p), yielding the correct quantum position and momentum probabilities upon integration. A fundamental distinction from the classical \rho_{\mathrm{cl}}(x, p) is that the Wigner distribution can assume negative values, reflecting quantum interference phenomena absent in classical mechanics. These negativities, confined to regions on the scale of \hbar, highlight nonclassical correlations and prevent a direct probabilistic interpretation, yet they preserve the overall analogy by enabling quantum analogs of classical phase-space averages.[1]Mathematical Properties
General Properties
The Wigner quasiprobability distribution W(x, p) associated with a quantum state described by the density operator \rho is always real-valued. This reality follows directly from the Hermitian property of \rho, ensuring that the off-diagonal elements in the position representation conjugate appropriately in the integral definition of W(x, p).[6] A fundamental property is its normalization over phase space: \iint W(x, p) \, dx \, dp = 1. This reflects the trace of the density operator, \operatorname{Tr}(\rho) = 1, and positions W(x, p) as a valid quasiprobability measure despite potential negativities. The marginal distributions recover the standard quantum probability densities: integrating over momentum yields the position probability density, \int W(x, p) \, dp = |\psi(x)|^2 for a pure state with wave function \psi(x), while integrating over position gives the momentum probability density, \int W(x, p) \, dx = |\phi(p)|^2, where \phi(p) is the momentum-space wave function. These marginals highlight how W(x, p) bridges position and momentum representations without direct measurement of joint probabilities.[7][8] The purity of the state, \operatorname{Tr}(\rho^2), which quantifies mixedness (ranging from 1 for pure states to $1/d for a d-dimensional maximally mixed state), is expressed as \operatorname{Tr}(\rho^2) = 2\pi \hbar \iint W(x, p)^2 \, dx \, dp. This integral serves as a phase-space analogue of the Hilbert-Schmidt inner product, with the L^2 norm of W(x, p) directly probing quantum coherence. For pure states, the distribution is bounded, |W(x, p)| \leq 1/(\pi \hbar), with equality achieved for energy eigenstates of the harmonic oscillator, where the function reaches its maximum at specific phase-space loci corresponding to classical turning points.[9] The Wigner distribution exhibits covariance under Galilean transformations, preserving its form up to shifts and shears in phase space. Under a boost with velocity v, it transforms as W'(x, p) = W(x, p - m v), while spatial translations and rotations in the (x, p) plane induce corresponding phase-space displacements, reflecting the underlying symplectic structure. Additionally, the overlap between two states is captured by the phase-space inner product: \operatorname{Tr}(\rho_1 \rho_2) = 2\pi \hbar \iint W_1(x, p) W_2(x, p) \, dx \, dp, providing a direct link between trace operations in Hilbert space and integrals over quasiprobabilities. These transformation properties underpin the utility of W(x, p) in analyzing quantum dynamics and symmetries.[10]Positivity and Negativity
One distinctive feature of the Wigner quasiprobability distribution W(x,p) is its potential to take negative values, in contrast to classical phase-space probability distributions, which are inherently non-negative. This non-positivity arises because W(x,p) does not always qualify as a valid probability measure in the sense of Krein-Q spaces, where positive definiteness on test functions is required; instead, the negativity reflects quantum interference effects that have no classical analog.[11] A quantitative measure of this non-classicality is provided by the integrated negativity, defined as N = \frac{1}{2} \left( \iint |W(x,p)| \, dx \, dp - 1 \right), which captures the total "volume" of negative regions and directly quantifies the extent of quantum interference in the state. For classical states, N = 0, while quantum states exhibit N > 0, with the maximum value for pure states approaching $1/2 in the highly non-classical limit. This metric has been established as a reliable indicator of non-classical behavior, distinct from other quasiprobability representations like the Husimi Q-function, which remains non-negative.[11] Hudson's theorem provides a precise characterization of when the Wigner distribution is non-negative: for a pure quantum state, W(x,p) \geq 0 everywhere if and only if the state is Gaussian, such as a coherent state. This result underscores that positivity is a rare property, limited to minimum-uncertainty Gaussian wavefunctions, and highlights the ubiquity of negativity as a hallmark of quantum superpositions.[11] In multipartite systems, Wigner negativity can signal entanglement even when the reduced states of individual subsystems have non-negative Wigner functions, as the global distribution's negative regions arise from quantum correlations across parties. For instance, certain entangled non-Gaussian states maintain local positivity while exhibiting global negativity, enabling the use of integrated negativity as an entanglement witness beyond Gaussian criteria.[12] For pure states, the Wigner function is bounded below by W(x,p) \geq -1/(\pi \hbar), a pointwise limit that constrains the depth of negativity and implies an upper bound on the integrated negativity of N \leq 1/2. This bound is nearly saturated in states like odd Schrödinger cat states, where large coherent superpositions lead to pronounced negative interference fringes, achieving maximal non-classicality in the asymptotic limit of large displacement amplitude.[11]Examples
Simple Quantum States
The Wigner quasiprobability distribution for the ground state of the quantum harmonic oscillator is a positive, Gaussian function centered at the origin in phase space, illustrating a classical-like probability distribution without negativities. In standard notation, it takes the form W(x,p) = \frac{1}{\pi \hbar} \exp\left( - \frac{ m \omega x^2}{\hbar} - \frac{ p^2}{m \omega \hbar} \right), where m is the mass, \omega the angular frequency, x the position, and p the momentum.[13] This expression arises from the minimum-uncertainty wave function of the ground state and remains nonnegative everywhere, reflecting the absence of quantum interference effects in this simplest bound state.[14] For a position eigenstate |x_0\rangle, the Wigner function manifests as a vertical line in phase space at x = x_0, uniform along the momentum axis, highlighting complete delocalization in momentum. The explicit form in the continuous limit, derived from the density matrix, is W(x,p) = \frac{1}{2\pi \hbar} \delta(x - x_0), though the eigenstate's non-normalizability requires regularization via limiting procedures, such as narrow Gaussians, revealing a broad distribution in momentum without oscillations.[15] This uniformity demonstrates the quasiprobability's ability to capture the infinite momentum uncertainty inherent in a perfectly localized position state. By symmetry between position and momentum representations, the Wigner function for a momentum eigenstate |p_0\rangle features a horizontal line at p = p_0 uniform along the position axis. The analogous expression is W(x,p) = \frac{1}{2\pi \hbar} \delta(p - p_0), exhibiting similar non-normalizability, which underscores the dual delocalization in position for a definite momentum.[15] The Wigner functions for excited energy eigenstates of the harmonic oscillator, labeled by quantum number n \geq 1, incorporate Laguerre polynomials and display regions of negativity, signaling non-classical interference. The general form is W_n(x,p) = \frac{(-1)^n}{\pi \hbar} L_n\left( \frac{4 H_\mathrm{cl}}{\hbar \omega} \right) \exp\left( -\frac{2 H_\mathrm{cl}}{\hbar \omega} \right), where H_\mathrm{cl} = \frac{p^2}{2m} + \frac{1}{2} m \omega^2 x^2 is the classical Hamiltonian and L_n the Laguerre polynomials. For the first excited state (n=1), L_1(z) = 1 - z, yielding negative lobes symmetric around the origin, which quantify quantum revivals and tunneling-like behaviors absent in classical distributions.[14][13] For a free particle described by an initial Gaussian wave packet, the Wigner function evolves as a shearing and spreading Gaussian in phase space, remaining nonnegative and preserving the minimum uncertainty product over time. The explicit time-dependent expression is W(x,p,t) = \frac{1}{\pi \hbar} \exp\left( -\frac{(x - v t)^2}{2 \sigma_t^2} - \frac{(p - m v)^2 \sigma_t^2 2}{\hbar^2 (1 + (\hbar t / (m \sigma_0^2))^2)} \right), where \sigma_t^2 = \sigma_0^2 (1 + (\hbar t / (m \sigma_0^2))^2) is the time-dependent position variance, v the mean velocity, and \sigma_0 the initial width; this illustrates dispersion without introducing negativities, contrasting bound-state interferences.[16]Coherent and Squeezed States
The Wigner quasiprobability distribution for a coherent state |\alpha\rangle takes the form of a displaced Gaussian in phase space: W(x,p) = \frac{2}{\pi \hbar} \exp\left( -2 \left| \frac{x + i p}{\sqrt{2\hbar}} - \alpha \right|^2 \right), where \alpha is a complex parameter encoding the displacement. This distribution is always positive and peaks at the classical phase-space coordinates corresponding to the expectation values \langle x \rangle = \sqrt{2\hbar} \operatorname{Re}(\alpha) and \langle p \rangle = \sqrt{2\hbar} \operatorname{Im}(\alpha), with equal uncertainties in both quadratures matching the vacuum fluctuations. The squeezed vacuum state, obtained by applying the squeezing operator to the vacuum, has a Wigner function that is an elliptical Gaussian centered at the origin, featuring reduced variance in one quadrature (e.g., position) and correspondingly increased variance in the orthogonal quadrature (e.g., momentum) to satisfy the Heisenberg uncertainty principle. This anisotropy highlights the non-classical reduction of noise below the standard quantum limit in the squeezed direction, while the distribution remains strictly non-negative. A squeezed coherent state combines both effects, with its Wigner function describing a displaced elliptical Gaussian that exhibits shearing in phase space due to the combined action of displacement and squeezing operators. The elongation or compression along specific directions depends on the squeezing parameter and phase, allowing tailored control over quadrature uncertainties for enhanced precision in measurements. Schrödinger cat states, formed as superpositions of two or more coherent states (e.g., even cat |\alpha\rangle + |-\alpha\rangle or odd cat |\alpha\rangle - |-\alpha\rangle, normalized), display Wigner functions consisting of overlapping Gaussians with additional interference fringes that produce regions of negativity. These negative dips, particularly prominent at the phase-space origin for small separations between the coherent components, underscore the non-classical interference inherent in such macroscopic superpositions. In quantum optics, coherent and squeezed states serve as minimum-uncertainty Gaussian states, and their Wigner functions vividly illustrate the geometric transformations induced by the displacement operator D(\alpha) and squeezing operator S(\xi), providing an intuitive phase-space picture of laser light and noise-reduced fields. Per Hudson's theorem, these are precisely the pure states whose Wigner functions are non-negative everywhere.[17]Dynamics
Evolution Equation
The time evolution of the Wigner quasiprobability distribution W(\mathbf{x}, \mathbf{p}, t) is governed by the von Neumann equation for the density operator \hat{\rho}(t), transformed into phase space via the Weyl correspondence. This yields Moyal's equation, \frac{\partial W}{\partial t} = \{H_w, W\}_M, where H_w(\mathbf{x}, \mathbf{p}) is the Weyl symbol of the Hamiltonian \hat{H}, and \{\cdot, \cdot\}_M denotes the Moyal bracket.[3] The Moyal bracket is defined as \{A, B\}_M = \frac{2}{\hbar} A \sin\left(\frac{\hbar}{2} \Lambda\right) B, with the bidifferential operator \Lambda = \overleftarrow{\partial_{\mathbf{x}}} \overrightarrow{\partial_{\mathbf{p}}} - \overleftarrow{\partial_{\mathbf{p}}} \overrightarrow{\partial_{\mathbf{x}}}. Its series expansion in powers of \hbar is \{A, B\}_M = \sum_{n=0}^\infty (-1)^n \frac{\hbar^{2n}}{2^{2n} (2n+1)!} \left( A \overleftarrow{\partial_{\mathbf{x}}}^{2n+1} \overrightarrow{\partial_{\mathbf{p}}}^{2n+1} B - A \overleftarrow{\partial_{\mathbf{p}}}^{2n+1} \overrightarrow{\partial_{\mathbf{x}}}^{2n+1} B \right) + terms from mixed derivatives in the expansion of \Lambda^{2n+1}, where the leading term (n=0) is the classical Poisson bracket \{A, B\}_{\mathrm{PB}} and subsequent terms introduce quantum corrections involving higher-order derivatives.[3] For a general potential V(\mathbf{x}) in the Hamiltonian H_w = \frac{\mathbf{p}^2}{2m} + V(\mathbf{x}), the quantum corrections manifest as higher even-order terms in \hbar, such as -\frac{\hbar^2}{24} \partial_x^3 V \, \partial_p^3 W (in 1D) at order \hbar^2 in the semiclassical expansion. These terms vanish exactly when H_w is quadratic in \mathbf{x} and \mathbf{p}, reducing the evolution to the classical Liouville equation \frac{\partial W}{\partial t} = \{H_w, W\}_{\mathrm{PB}}.[3][18] A formal solution to the evolution equation can be expressed using the Moyal star product \star, defined as A \star B = A \exp\left(\frac{i\hbar}{2} (\overleftarrow{\partial_{\mathbf{x}}} \overrightarrow{\partial_{\mathbf{p}}} - \overleftarrow{\partial_{\mathbf{p}}} \overrightarrow{\partial_{\mathbf{x}}} )\right) B, via the star exponential W(\mathbf{x}, \mathbf{p}, t) = W(\mathbf{x}, \mathbf{p}, 0) \star \exp\left( - \frac{i t}{\hbar} H_w(\mathbf{x}, \mathbf{p}) \right). The explicit \hbar-dependence in the Moyal bracket and star product underscores the semiclassical nature of the formalism, with the classical limit recovered as \hbar \to 0.[19][3]Harmonic Oscillator Time Evolution
The Hamiltonian for the quantum harmonic oscillator is H = \frac{p^2}{2m} + \frac{1}{2} m \omega^2 x^2, where m is the mass, \omega is the angular frequency, x is the position, and p is the momentum. The Weyl symbol of this operator, H_w(x,p), coincides exactly with the classical Hamiltonian H_\mathrm{cl}(x,p) = H(x,p), as the Hamiltonian is quadratic in the phase-space variables. The time evolution of the Wigner quasiprobability distribution W(x,p,t) for this system follows the Moyal equation \partial W / \partial t = \{H_w, W\}_M, where \{\cdot, \cdot\}_M denotes the Moyal bracket. For quadratic Hamiltonians such as this one, all higher-order quantum corrections in the Moyal bracket vanish, reducing it precisely to the classical Poisson bracket: \{H_w, W\}_M = \{H_\mathrm{cl}, W\}_P. Consequently, the evolution equation simplifies to \partial W / \partial t = \{H_\mathrm{cl}, W\}_P, which is identical to the classical Liouville equation for phase-space densities.[3] This exact classical form of the evolution implies that the Wigner distribution flows along classical trajectories in phase space without quantum distortion. For an initial Gaussian Wigner function—such as that of a coherent state—the distribution undergoes rigid elliptical rotation around the origin at the classical frequency \omega, preserving its shape and width throughout the dynamics. This rotation corresponds directly to the periodic motion of the classical harmonic oscillator. The means of the distribution, \langle x \rangle and \langle p \rangle, evolve according to the Ehrenfest theorem, tracing out classical trajectories: \frac{d}{dt} \langle x \rangle = \frac{\langle p \rangle}{m} and \frac{d}{dt} \langle p \rangle = -m \omega^2 \langle x \rangle. Although the energy spectrum of the quantum harmonic oscillator remains discrete due to quantization, the phase-space flow of the Wigner function itself exhibits purely classical behavior, independent of \hbar.[3]Limits and Interpretations
Classical Limit
The semiclassical expansion of the Wigner quasiprobability distribution reveals its convergence to classical phase-space descriptions. In this regime, the quantum evolution equation for the Wigner function, involving the Moyal bracket {H, W}_M, reduces to the classical Liouville equation governed by the Poisson bracket {H, W}P as \hbar \to 0, where H is the Hamiltonian.[20] The resulting classical distribution \rho{cl} is strictly positive and becomes sharply peaked along classical trajectories, reflecting the localization of quantum states onto deterministic paths in phase space.[21] This expansion highlights the Wigner function's utility in bridging quantum and classical dynamics, particularly for integrable systems where the distribution aligns with invariant tori.[22] In the strict classical limit \hbar \to 0, the Wigner function loses all negativity, transforming into a genuine probability density that reproduces classical marginal distributions for position and momentum.[23] For bound systems, such as energy eigenstates, the distribution crystallizes into delta functions confined to the classical energy shells, which are the f-dimensional tori defined by the conserved energy in phase space for systems with f degrees of freedom.[22] This localization ensures that quantum delocalization effects, manifested as oscillations or spreading in finite-\hbar distributions, vanish entirely. From the perspective of Bohmian mechanics, the Wigner function admits an interpretation as |R|^2 times a classical density derived from the guiding wave function \psi = R e^{iS/\hbar}, where R is the amplitude and S the phase; in the classical limit, corrections from the quantum potential Q = -\frac{\hbar^2}{2m} \frac{\nabla^2 R}{R} become negligible, yielding pure classical trajectories.[24] This view underscores the emergence of deterministic Bohmian motion from the quasiprobability framework as quantum fluctuations subside.[25] In many-particle systems, the thermodynamic limit effectively drives \hbar_{eff} \to 0 through scaling with particle number N, causing the Wigner function to average over quantum correlations and converge to the classical ensemble distribution in phase space.[26] States exhibiting non-negative Wigner functions in this limit, such as coherent states, serve as criteria for classicality; Hudson's theorem establishes that, for pure states in continuous-variable systems, positivity of the Wigner function holds if and only if the state is Gaussian, aligning precisely with these classical analogs.[27]Role in Quantum Interpretations
The Wigner quasiprobability distribution plays a central role in Weyl quantization, where it serves as a phase-space representation that resolves ambiguities in operator ordering by employing the symmetric Weyl ordering rule. In this framework, classical functions on phase space are mapped to quantum operators via the Weyl transform, and the inverse Wigner transform reconstructs the quasiprobability distribution from the density operator, providing a geometric interpretation of quantum states on the symplectic phase space. This approach underpins geometric quantization, which extends the quantization procedure to general symplectic manifolds by prequantizing the phase space and incorporating polarization to select a Hilbert space of sections, with the Wigner function facilitating the correspondence between classical and quantum observables.[28][29] In the context of hidden variable theories, the negativity of the Wigner distribution demonstrates incompatibility with non-contextual hidden variable models, as established by the equivalence between Wigner negativity and quantum contextuality. Specifically, a state possessing negative regions in its Wigner function cannot be described by a non-contextual hidden variable theory, in line with the Bell-Kochen-Specker theorem, because such negativity reflects the impossibility of assigning definite values to all observables independently of measurement context. However, contextual hidden variable models remain compatible, as they allow for context-dependent value assignments that accommodate the quasiprobability's signed nature without requiring non-negativity.[30] Within Bohmian mechanics, the Wigner distribution relates to the guiding equation for particle trajectories through its polar decomposition, where the phase of the Wigner function determines the probability current analogous to the Bohmian velocity field. The quantum potential, derived from the amplitude and phase of the wave function, manifests in the phase-space dynamics captured by the Wigner formalism, influencing trajectory spreading and non-classical motion. Negativity in the Wigner function arises from interference terms between wave components, which in Bohmian terms correspond to the non-local effects of the quantum potential that prevent classical-like trajectories in superposition states.[24] In the many-worlds interpretation, the Wigner distribution provides a phase-space measure of decoherence, illustrating how quantum superpositions evolve into branching structures of effectively classical distributions. Decoherence suppresses interference fringes in the Wigner function, leading to a proliferation of positive, localized quasiprobability peaks that correspond to the divergent world branches, each resembling a classical probability distribution in the emergent multiverse. This phase-space perspective highlights how the universal wave function's evolution naturally selects preferred bases without collapse, with the Wigner's negativity quantifying the residual quantum coherence prior to full branching. Objective collapse models, such as continuous spontaneous localization (CSL) theories, incorporate the Wigner distribution to model the suppression of negativity through stochastic localization processes that drive quantum states toward classicality. In these models, spontaneous collapses localize the wave function in position space, which manifests in phase space as a damping of oscillatory interference terms in the Wigner function, reducing its negative volumes and promoting positive, peaked distributions akin to classical probabilities. This mechanism resolves the quantum measurement problem by dynamically suppressing superpositions, with the rate of negativity reduction tied to the model's localization strength, ensuring consistency with observed macroscopic realism.[31]Applications
In Quantum Optics and Mechanics
In quantum optics, the Wigner quasiprobability distribution provides a phase-space representation that reveals the structure of photon number states, such as Fock states, as concentric rings centered at the origin, with the negativity in the central region for odd photon numbers highlighting their non-classical nature.[32] This visualization contrasts with classical probability distributions, allowing direct identification of quantum features like the oscillatory behavior in the photon-number distribution derived from marginal integrals of the Wigner function.[33] Quadrature measurements, essential for probing these states, are performed using balanced homodyne detection, where a local oscillator interferes with the signal field to measure the field quadratures at various phases, enabling reconstruction of the full Wigner function through pattern functions or Radon transforms.[32] Wigner tomography, pioneered in the 1990s, reconstructs quantum states from these homodyne projections, providing a complete characterization of optical fields with high fidelity, as demonstrated in experiments with attenuated laser light and parametric down-conversion sources.[34] The Wigner distribution is instrumental in identifying non-classical properties of light, such as squeezing and antibunching. For squeezed states, the Wigner function exhibits elongated elliptical contours along one quadrature axis, reflecting reduced uncertainty below the vacuum level while remaining non-negative, thus distinguishing Gaussian non-classicality without violating the positivity criterion.[33] Antibunching, indicative of sub-Poissonian photon statistics, manifests in the Wigner function through irregular contours and potential negativities near the origin, correlating with second-order coherence functions measured via intensity correlations, as verified in homodyne tomography of single-photon sources.[35] These features have been crucial for validating non-classical light generation in cavity quantum electrodynamics and nonlinear optical processes from the 1980s onward.[36] In foundational quantum mechanics, the Wigner function facilitates phase-space formulations of path integrals, where the propagator is expressed as an integral over classical trajectories weighted by quantum phases, capturing interference effects beyond the classical Liouville equation.[37] This approach validates the WKB approximation through semiclassical limits of Wigner propagation, where higher-order quantum corrections to the Moyal bracket align with WKB wave functions for bound states, as shown in analyses of harmonic oscillators and anharmonic potentials during the 1990s and 2000s.[38] For interference phenomena, the Wigner function visualizes quantum superpositions in phase space, such as in the double-slit experiment, where negative regions emerge between interfering paths, quantifying the volume of non-classical interference inaccessible to classical rays.[39] From the 1930s revival of phase-space methods to the 2000s, the Wigner distribution has played a key role in quantum state engineering in optics, enabling the design and verification of superpositions like Schrödinger cat states through nonlinear interactions and homodyne readout, with negativities confirming their fragile quantum coherence against decoherence. These applications, building on coherent states as displaced vacuum Gaussians, underscore the Wigner's utility in bridging harmonic oscillator dynamics with engineered non-Gaussian resources.[33]Outside Quantum Mechanics
The Wigner quasiprobability distribution, independently reinvented by Jean Ville in 1948, found early application in signal analysis as a quadratic representation of the local time-frequency energy of signals, predating its broader recognition in quantum contexts.[40] This formulation, known as the Wigner-Ville distribution, treats the distribution as a phase-space tool for non-stationary signals, where time replaces position and frequency replaces momentum.[41] In signal processing, the Wigner-Ville distribution serves as a high-resolution time-frequency representation for analyzing non-stationary signals, such as chirps in radar systems, where it reveals instantaneous frequency variations by concentrating energy along the signal's trajectory in the time-frequency plane.[42] However, it introduces cross-terms that appear as oscillatory artifacts between multiple signal components, analogous to interference patterns, which can complicate interpretation but are mitigated through smoothing techniques like the pseudo Wigner-Ville distribution. These properties make it valuable for applications in radar signal detection and biomedical signal analysis, where precise localization of transient events is essential. In classical optics, the Wigner function describes the phase-space distribution of paraxial light beams, providing a ray-optical interpretation of wave propagation without invoking quantum mechanics.[43] Beam evolution through first-order optical systems, characterized by ABCD ray-transfer matrices, transforms the Wigner function via a linear shearing operation in phase space, preserving its marginals for position and momentum distributions.[44] This analogy to quantum propagation enables efficient simulation of beam propagation, diffraction, and partial coherence in lens systems and free space.[45] In statistical mechanics, the Wigner function facilitates semiclassical approximations for many-body systems by representing the quantum density operator in phase space, allowing interpolation between classical Liouville dynamics and quantum corrections. For equilibrium states, it yields the classical Boltzmann distribution in the high-temperature limit while incorporating ħ-dependent terms for quantum effects in dense gases or solids.[22] This approach is particularly useful in deriving transport coefficients and thermodynamic properties via phase-space integrals.[46] Extensions to acoustics and general wave phenomena generalize the Wigner distribution to arbitrary phase spaces for describing wave packet propagation, such as in sound fields or elastic waves, where it captures energy density in position-wavenumber coordinates.[47] In acoustic signal processing, it analyzes transient pressure waves, revealing dispersion and interference in non-homogeneous media, much like its role in optics.[48] This framework unifies treatments of classical waves across domains by emphasizing conserved phase-space volumes under linear transformations.[49]Modern Uses in Quantum Computing
In quantum information theory, the negativity of the Wigner quasiprobability distribution serves as a key indicator of nonclassical resources essential for achieving quantum computational advantage. Specifically, states exhibiting Wigner negativity enable operations beyond Clifford circuits, such as in magic state distillation protocols, where non-stabilizer states are purified to support universal quantum computation.[50] A quantitative measure of this resource is provided by mana, defined as the logarithm of the sum of the absolute values of the components of the discrete Wigner function (its l1-norm), which bounds the efficiency of magic state transformations and distillation thresholds.[51] This negativity is not only necessary for quantum speedup but also certifiable through fidelity-based witnesses with Fock states, requiring fewer measurements than full tomography.[50] In continuous-variable quantum computing, the Wigner distribution facilitates the encoding of discrete-variable qubits into phase space via the Gottesman-Kitaev-Preskill (GKP) code, where logical states manifest as periodic lattices in the quadrature plane.[52] For Gaussian states—prevalent in photonic implementations—the Wigner function enables efficient simulation of linear optical gates, as these operations correspond to affine transformations in phase space that preserve the Gaussian form. This representation aids in verifying gate fidelities and scaling hybrid discrete-continuous architectures, with negativity quantifying the "magic" needed for non-Gaussian extensions to universality.[52] Quantum machine learning leverages phase-space formulations of the Wigner distribution to enhance kernel methods, where data is embedded into bosonic Hilbert spaces and kernel matrices are estimated via quantum feature maps.[53] In hybrid quantum-classical algorithms, Wigner negativity ensures that these kernels capture nonclassical correlations intractable for classical simulation, providing a pathway to advantage in tasks like classification on datasets such as MNIST.[53] For instance, linear-optical networks with postselected non-Gaussian measurements amplify this negativity, improving kernel expressivity while tensor network contractions validate the phase-space dynamics.[53] For error correction in continuous-variable systems, the Wigner function helps detect non-Gaussian errors, which manifest as negativity or irregular features and deviate from Gaussian noise in bosonic codes. These errors, often arising from photon loss or nonlinear interactions, can be identified through phase-space reconstruction.[54] Tensor network methods, such as the time-dependent variational principle on matrix product states, propagate approximate Wigner functions via the truncated Wigner approximation, accurately simulating far-from-equilibrium dynamics in two-dimensional spin models up to intermediate times.[55] These techniques bridge limitations in simulating quantum dynamics by concentrating computational resources on entangled regions, facilitating benchmarks for quantum advantage in non-Gaussian resource theories.[55]Measurement and Computation
Characterization Methods
The Wigner quasiprobability distribution can be characterized experimentally through quantum tomography, which involves reconstructing the full phase-space function from a series of quadrature measurements. In quantum optics, balanced homodyne detection serves as the primary setup, where the signal field is mixed with a strong local oscillator on a 50/50 beam splitter, and the resulting photocurrents from balanced photodetectors yield the quadrature distribution for various phases of the local oscillator. By collecting these marginal distributions over a complete range of phases (0 to 2π), the Wigner function is obtained via the inverse Radon transform, providing a complete tomographic reconstruction of the quantum state.[56][32] Direct sampling methods offer an alternative for probing specific points in the Wigner phase space without full reconstruction. The eight-port homodyne detector, which combines the signal and a local oscillator using four 50:50 beam splitters and four photodetectors, enables simultaneous measurement of both quadratures, allowing direct access to sampled values of the Wigner function W(x, p) through photon-counting statistics or unbalanced homodyning. This approach is particularly useful for verifying nonclassical features at discrete phase-space locations, such as negative regions, by projecting the state onto pattern overlaps derived from Fock states.[57][32] For efficient state estimation, pattern functions provide a filtered projection technique that reconstructs the density matrix elements—and thus the Wigner function—from homodyne data. These functions, which are essentially Radon transforms of the density operator weighted by phase-space filters, allow direct averaging over measured quadrature outcomes to yield the quasiprobability without iterative processing, improving accuracy in low-photon-number regimes. This method extends to noisy environments by incorporating detection inefficiency into the filter design.[58] Despite these advances, characterization methods face significant limitations, including high sensitivity to detection noise, which can smear out negative values in the Wigner function and obscure nonclassicality. Additionally, many techniques rely on weak measurements to minimize state disturbance, requiring low interaction strengths that amplify statistical uncertainties and necessitate large ensembles of identically prepared states for reliable reconstruction.[32]Numerical Computation
Numerical computation of the Wigner quasiprobability distribution typically involves approximating the continuous phase-space integral on finite grids or expanding the density operator in suitable bases, enabling efficient evaluation for quantum states represented in discrete Hilbert spaces. For systems with finite-dimensional Hilbert space of dimension d, the discrete Weyl transform provides a framework to compute the Wigner function on a d \times d phase-space grid by mapping the density matrix to phase-space coordinates via translation and modulation operators. This approach leverages the fast Fourier transform (FFT) to perform the required symplectic Fourier transforms, achieving a computational complexity of O(d^2 \log d) for the core operations on the grid, making it suitable for moderate-sized quantum systems.[59] Pseudospectral methods offer an alternative by expanding the wave function or density operator in orthogonal bases tailored to the phase space, such as discrete Hermite functions for systems near the harmonic oscillator or Gabor functions for localized representations. In the Hermite basis, the Wigner function coefficients are obtained by solving a system of linear partial differential equations derived from the Weyl correspondence, allowing direct computation without explicit integration and with spectral convergence for smooth states. Gabor bases, which form overcomplete frames in phase space, facilitate efficient expansions for non-stationary states by minimizing the support in both position and momentum, reducing the number of terms needed for accurate representation compared to standard Fourier methods. These methods scale favorably for one-dimensional systems, with complexity depending on the basis truncation level, typically O(M^2) where M is the number of basis elements.[60] Another route computes the Wigner function through the ambiguity function, which is the Fourier transform of the characteristic function \chi(\xi, \eta) = \operatorname{Tr}[\rho \exp(i \xi \hat{X} + i \eta \hat{P})], where \rho is the density operator and \hat{X}, \hat{P} are position and momentum operators. The characteristic function is evaluated by expressing the displacement operator in the number basis or using matrix exponentiation, followed by a 2D inverse symplectic Fourier transform via FFT to yield the Wigner function on a discrete grid. This method is particularly efficient for Gaussian states, where \chi has a closed form, but extends to general states with numerical trace evaluation, maintaining O(d^2 \log d) complexity for d-dimensional systems.[61] The oscillatory nature of the Wigner integral often leads to numerical challenges in capturing regions of negativity, which arise from interference terms. Adaptive quadrature techniques, such as those based on Levin's method or Gaussian quadrature with dynamic node adjustment, are employed to integrate the highly oscillatory kernel \exp(2 i p y / \hbar) in the position representation, ensuring accurate resolution of negative lobes without excessive sampling. These approaches adapt the integration step size to the local oscillation frequency, improving convergence for states with significant quantum features like superpositions, and have been shown effective in computing Wigner functions for non-Gaussian states generated by Kerr nonlinearities.[62] Implementations in software libraries facilitate practical computations for quantum simulations. The Quantum Toolbox in Python (QuTiP) provides thequtip.wigner function, which computes the Wigner distribution using an FFT-based discrete Weyl approach for both pure states and density matrices, supporting Hilbert spaces up to dimension ~1000 before memory limits dominate, with visualization options for phase-space plots. Similarly, MATLAB toolboxes such as the Quantum Optics Toolbox and dedicated Wigner function scripts enable analogous computations via matrix representations, scaling to similar system sizes for time-dependent simulations of open quantum systems. For larger systems, hybrid methods combining these with sparse representations mitigate scaling issues, though full density matrix approaches remain O(d^2) in storage.[63][64]