Fact-checked by Grok 2 weeks ago

Ehrenfest theorem

The Ehrenfest theorem is a foundational result in that establishes a direct correspondence between the time evolution of expectation values of dynamical observables and the equations of . Formulated by Austrian theoretical physicist in 1927, it demonstrates that the average behavior of a quantum system, as captured by expectation values, follows classical trajectories under slowly varying potentials, thereby bridging the microscopic quantum realm with macroscopic . At its core, the theorem derives from the time-dependent and asserts two key relations for a particle in one dimension: the time derivative of the expectation value of \langle \hat{x} \rangle equals the expectation value of divided by , \frac{d}{dt} \langle \hat{x} \rangle = \frac{\langle \hat{p} \rangle}{m}, and the time derivative of the expectation value equals the negative expectation value of , \frac{d}{dt} \langle \hat{p} \rangle = -\left\langle \frac{\partial V}{\partial x} \right\rangle, where V(x) is the . These equations mirror Hamilton's canonical equations of , with quantum operators replacing classical variables, and highlight how quantum wave packets propagate in a manner approximating classical point particles when the de Broglie is small compared to variations in the potential. The significance of the Ehrenfest theorem lies in its role as an early demonstration of the correspondence principle, proposed by , which posits that should reduce to in the limit of large quantum numbers or macroscopic scales. It has been extended beyond non-relativistic cases to relativistic , including Dirac particles and systems with , underscoring its versatility in unifying quantum and classical descriptions across diverse physical contexts. However, the theorem's approximations break down for rapidly varying potentials or chaotic systems, where quantum effects like spreading wave packets lead to deviations from classical behavior over the so-called Ehrenfest time.

Introduction

Definition and Statement

The Ehrenfest theorem, named after the Austrian Paul , asserts that the expectation values of the position and momentum observables in evolve in time according to equations that mirror the classical equations of motion. Formulated in 1927, the theorem provides a bridge between quantum and classical descriptions by showing how averages over quantum states can reproduce classical dynamics on macroscopic scales or for localized wave packets. In its standard one-dimensional form, for a particle of m subject to a potential V(x), the theorem states: \frac{d}{dt} \langle \hat{x} \rangle = \frac{\langle \hat{p} \rangle}{m} \frac{d}{dt} \langle \hat{p} \rangle = -\left\langle \frac{d V}{d x} \right\rangle Here, \langle \cdot \rangle denotes the expectation value with respect to the system's \psi(x, t) at time t, \hat{x} is the , and \hat{p} = -i \hbar \frac{d}{dx} is the . These relations indicate that the " " moves with the " " ( over ), while the " " changes due to the force from the . More generally, the Ehrenfest theorem applies to any Hermitian operator \hat{A} in a Hamiltonian H = \frac{\hat{p}^2}{2m} + V(\hat{x}), yielding \frac{d}{dt} \langle \hat{A} \rangle = \left\langle \frac{\partial \hat{A}}{\partial t} \right\rangle + \frac{i}{\hbar} \langle [ \hat{H}, \hat{A} ] \rangle, where [ \cdot, \cdot ] is the . For time-independent \hat{A}, this reduces to the commutator term alone, which specializes to the position and momentum forms above when \hat{A} = \hat{x} or \hat{p}. This general expression underscores the theorem's role in deriving the dynamics of expectation values from the .

Historical Background

The Ehrenfest theorem emerged during the formative years of , building on Paul Ehrenfest's prior contributions to the . Born in 1880 in , Ehrenfest studied under and later became a professor at the University of in 1912, where he played a pivotal role in elucidating foundational concepts in and . In 1916, he introduced the concept of adiabatic invariants, which demonstrated that certain quantities, such as the action integral in , remain unchanged under slow variations of parameters; this principle provided a bridge between classical and quantized systems, influencing Niels Bohr's atomic model by justifying the quantization of orbital . The theorem's development coincided with the rapid evolution of in the mid-1920s. Following Werner Heisenberg's in 1925 and Erwin Schrödinger's wave mechanics in 1926, physicists grappled with reconciling the probabilistic nature of with classical determinism. Ehrenfest, known for his clarity in pedagogical expositions, sought to address this by exploring how quantum expectation values could mimic classical trajectories, aligning with Bohr's that should reduce to in the limit of large quantum numbers. His work was motivated by Schrödinger's 1926 paper on the continuous transition from micro- to macro-mechanics, which highlighted the need for such a correspondence in dynamics. In 1927, Ehrenfest published a concise two-page paper in Zeitschrift für Physik, titled "Bemerkung über die angenäherte Gültigkeit der klassischen Mechanik innerhalb der Quantenmechanik" (Remark on the approximate validity of classical mechanics within quantum mechanics). Therein, he derived equations showing that the time derivatives of the expectation values of position and momentum operators follow Hamilton's classical equations of motion, provided the wave function remains sufficiently localized. This result, now known as the Ehrenfest theorem, was presented as a direct consequence of the Schrödinger equation, emphasizing its approximate nature for systems where quantum uncertainties are small compared to the scales of motion. The theorem's publication marked a significant milestone in ' acceptance, as it provided mathematical reassurance that the new theory encompassed as a limiting case, alleviating concerns among skeptics like . It influenced subsequent developments, including generalizations to and semiclassical approximations, and remains a cornerstone for understanding quantum-to-classical transitions. Ehrenfest died in 1933.

Mathematical Derivation

In the Schrödinger Picture

In the , the time evolution of the is governed by the time-dependent , i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi, where \psi(\mathbf{r}, t) is the wave function, \hat{H} is the operator, and \hbar is the reduced Planck's constant. operators, such as \hat{\mathbf{x}} and \hat{\mathbf{p}}, are time-independent, while values \langle \hat{A} \rangle = \int \psi^* \hat{A} \psi \, dV evolve according to the theorem. The general form of the Ehrenfest theorem for the time of the value of an \hat{A} (with possible explicit time dependence) is derived by differentiating \langle \hat{A} \rangle with respect to time and substituting the for \partial \psi / \partial t and its : \frac{d}{dt} \langle \hat{A} \rangle = \left\langle \frac{\partial \hat{A}}{\partial t} \right\rangle + \frac{i}{\hbar} \langle [\hat{H}, \hat{A}] \rangle, where [\hat{H}, \hat{A}] = \hat{H} \hat{A} - \hat{A} \hat{H} is the . For operators without explicit time dependence, such as \hat{\mathbf{x}} and \hat{\mathbf{p}} in the standard formulation, the first term vanishes, simplifying to \frac{d}{dt} \langle \hat{A} \rangle = \frac{i}{\hbar} \langle [\hat{H}, \hat{A}] \rangle. This expression bridges quantum values to classical dynamics by relating the evolution to the structure. Consider a single particle in three dimensions with Hamiltonian \hat{H} = \frac{\hat{\mathbf{p}}^2}{2m} + V(\hat{\mathbf{x}}), where m is the and V is the operator depending only on position. For the \hat{\mathbf{x}}, the relevant is [\hat{H}, \hat{x}_i] = \left[ \frac{\hat{p}^2}{2m}, \hat{x}_i \right], since V(\hat{\mathbf{x}}) commutes with \hat{x}_i. Using the [\hat{x}_i, \hat{p}_j] = i \hbar \delta_{ij}, it follows that [\hat{p}^2, \hat{x}_i] = -2 i \hbar \hat{p}_i, so [\hat{H}, \hat{x}_i] = -\frac{i \hbar}{m} \hat{p}_i. Thus, \frac{d}{dt} \langle \hat{x}_i \rangle = \frac{\langle \hat{p}_i \rangle}{m}. This holds for each component i = x, y, z, mirroring the classical velocity \mathbf{v} = \mathbf{p}/m. For the momentum operator \hat{\mathbf{p}}, the commutator is [\hat{H}, \hat{p}_i] = [V(\hat{\mathbf{x}}), \hat{p}_i], as the kinetic term commutes with \hat{p}_i. The position-momentum commutation yields [V(\hat{\mathbf{x}}), \hat{p}_i] = i \hbar \frac{\partial V}{\partial x_i}, so \frac{d}{dt} \langle \hat{p}_i \rangle = -\left\langle \frac{\partial V}{\partial x_i} \right\rangle. Combining these, the second derivative of position satisfies m \frac{d^2}{dt^2} \langle \mathbf{x} \rangle = -\langle \nabla V(\mathbf{x}) \rangle, analogous to Newton's second law \mathbf{F} = m \mathbf{a} with the force as the expectation value of the potential gradient. This derivation, originally presented by in 1927, underscores the theorem's role in connecting quantum and for expectation values.

In the Heisenberg Picture

In the of , the state vectors are time-independent, while the operators evolve according to the Heisenberg equation of motion. For any \hat{A} that does not explicitly depend on time, the is given by \frac{d \hat{A}}{dt} = \frac{1}{i \hbar} [\hat{A}, \hat{H}], where \hat{H} is the and [\cdot, \cdot] denotes the . This formulation provides a natural framework for deriving the Ehrenfest theorem, as it directly yields the for the \hat{\mathbf{x}} and \hat{\mathbf{p}} operators in a system with \hat{H} = \frac{\hat{\mathbf{p}}^2}{2m} + V(\hat{\mathbf{x}}). Consider the one-dimensional case for clarity, with \hat{H} = \frac{\hat{p}^2}{2m} + V(\hat{x}). The time derivative of the is \frac{d \hat{x}}{dt} = \frac{1}{i \hbar} [\hat{x}, \hat{H}] = \frac{1}{i \hbar} \left[ \hat{x}, \frac{\hat{p}^2}{2m} \right], since [\hat{x}, V(\hat{x})] = 0. Using the commutation relation [\hat{x}, \hat{p}] = i \hbar, it follows that [\hat{x}, \hat{p}^2] = 2 i \hbar \hat{p}, so \frac{d \hat{x}}{dt} = \frac{\hat{p}}{m}. Similarly, for the , \frac{d \hat{p}}{dt} = \frac{1}{i \hbar} [\hat{p}, \hat{H}] = \frac{1}{i \hbar} [\hat{p}, V(\hat{x})]. The [\hat{p}, V(\hat{x})] = - i \hbar \frac{d V}{d \hat{x}}, leading to \frac{d \hat{p}}{dt} = -\frac{d V}{d \hat{x}}. These operator equations mirror the classical Hamilton's equations. Taking values with respect to a fixed state |\psi\rangle, the Ehrenfest theorem emerges: \frac{d \langle \hat{x} \rangle}{dt} = \frac{\langle \hat{p} \rangle}{m}, \quad \frac{d \langle \hat{p} \rangle}{dt} = -\left\langle \frac{d V}{d \hat{x}} \right\rangle, where \langle \cdot \rangle = \langle \psi | \cdot | \psi \rangle. Combining these yields Newton's second law in terms of values: m \frac{d^2 \langle \hat{x} \rangle}{dt^2} = -\left\langle \frac{d V}{d \hat{x}} \right\rangle. In three dimensions, this generalizes to m \frac{d^2 \langle \mathbf{x} \rangle}{dt^2} = -\langle \nabla V(\mathbf{x}) \rangle. This result holds regardless of the wave function's spread, though for broad wave packets, the force term \langle \nabla V \rangle may deviate from the classical \nabla V(\langle \mathbf{x} \rangle ). The theorem underscores the correspondence principle, as the expectation values follow classical trajectories for systems where quantum uncertainties are small compared to the potential's scale. In the Heisenberg picture, this evolution is operator-driven, facilitating connections to classical phase-space dynamics.

Connection to Classical Physics

Relation to Hamilton's Equations

The Ehrenfest theorem establishes a direct correspondence between the time evolution of expectation values in quantum mechanics and the classical Hamilton's equations of motion. In classical , for a with H = \frac{p^2}{2m} + V(x), the are given by \frac{dx}{dt} = \frac{\partial H}{\partial p} = \frac{p}{m}, \quad \frac{dp}{dt} = -\frac{\partial H}{\partial x} = -\frac{\partial V}{\partial x}. These describe the deterministic trajectories of x and p. In , the Ehrenfest theorem shows that the expectation values \langle x \rangle and \langle p \rangle obey analogous equations derived from the or the . Specifically, \frac{d}{dt} \langle x \rangle = \frac{\langle p \rangle}{m}, \quad \frac{d}{dt} \langle p \rangle = -\left\langle \frac{\partial V}{\partial x} \right\rangle, where the averages are taken over the . These relations emerge from the structure of quantum operators, particularly [x, p] = i\hbar, and the time-evolution operator i\hbar \frac{d\hat{A}}{dt} = [\hat{A}, \hat{H}] for any \hat{A}. Thus, the center of a quantum follows classical trajectories under the averaged potential, illustrating the correspondence principle. This parallelism holds more generally for multidimensional systems or arbitrary Hamiltonians quadratic in momenta, where the expectation values evolve along canonical trajectories in phase space. For nonlinear potentials, the quantum force \left\langle \frac{\partial V}{\partial x} \right\rangle approximates the classical \frac{\partial V}{\partial \langle x \rangle} when the wave function is sufficiently localized, such as in the semiclassical limit \hbar \to 0. The theorem's Hamiltonian structure can be formalized using Lie-Poisson brackets on the Ehrenfest group, a semidirect product of phase space and unitary transformations, ensuring conservation laws like energy preservation for time-independent Hamiltonians. Originally formulated by Paul Ehrenfest in 1927, this connection underscores how quantum mechanics recovers classical dynamics for expectation values, bridging the two frameworks without resolving quantum uncertainties like wave packet spreading.

Limitations and the Semiclassical Approximation

While the Ehrenfest theorem provides an exact description of the time evolution of expectation values in quantum mechanics, its utility in establishing a correspondence with classical mechanics relies on approximations that replace quantum operators with their expectation values, such as \langle V(\hat{q}) \rangle \approx V(\langle \hat{q} \rangle). This semiclassical approximation holds when the quantum wave function remains localized relative to the scale over which the potential V(q) varies significantly, ensuring that quantum fluctuations are negligible. However, this assumption breaks down as the wave packet spreads due to dispersion, leading to deviations where the expectation value of the potential no longer matches the potential evaluated at the mean position. For instance, in a free particle or anharmonic potential, the wave packet width \Delta q(t) \sim \sqrt{\Delta q(0)^2 + \left( \frac{\Delta p(0) t}{m} \right)^2} for a Gaussian wave packet, causing the approximation to fail on timescales longer than the initial localization time. In chaotic systems, the limitations become particularly pronounced due to the exponential sensitivity to initial conditions, quantified by the Lyapunov exponent \lambda. The Ehrenfest time \tau_E, which marks the timescale over which the semiclassical approximation remains valid, scales as \tau_E \sim \lambda^{-1} \ln(S / \hbar), where S is a classical action scale. Beyond \tau_E, quantum wave packets stretch along unstable manifolds and spread rapidly, violating the localization required for the classical correspondence and leading to a breakdown in the predictability of expectation values. This effect is especially relevant in quantum chaos studies, where even macroscopic systems can exhibit deviations from classical trajectories on surprisingly short timescales, such as seconds or minutes for certain macroscopic systems. The semiclassical approximation addresses these limitations by incorporating higher-order quantum corrections beyond the mean-field level of the Ehrenfest theorem. Approaches like the Weyl symbol expansion or initial-value representations (e.g., Heller's semiclassical ) expand the around classical trajectories, including terms proportional to \hbar to account for evolution and . For example, in the frozen Gaussian approximation, the wave function is propagated as a superposition of Gaussians centered on classical paths, extending the validity of Ehrenfest-like dynamics up to the Ehrenfest time while capturing spreading effects. These methods are particularly effective in transitional regimes, such as near-classical limits for coherent states, but still fail in strongly quantum domains like tunneling or high-dimensional chaos where full quantum simulations are required.

Examples

Harmonic Oscillator

The Ehrenfest theorem finds a particularly exact correspondence with classical mechanics in the case of the quantum harmonic oscillator, where the potential is quadratic. For the Hamiltonian H = \frac{p^2}{2m} + \frac{1}{2} m \omega^2 x^2, the theorem yields the time evolution equations for expectation values: \frac{d}{dt} \langle x \rangle = \frac{\langle p \rangle}{m}, \quad \frac{d}{dt} \langle p \rangle = - m \omega^2 \langle x \rangle. These are identical to the classical Hamilton's equations of motion, and they hold for any initial quantum state |\psi\rangle, without approximations. Solving these equations gives the explicit classical-like trajectories: \langle x(t) \rangle = \langle x(0) \rangle \cos(\omega t) + \frac{\langle p(0) \rangle}{m \omega} \sin(\omega t), \langle p(t) \rangle = \langle p(0) \rangle \cos(\omega t) - m \omega \langle x(0) \rangle \sin(\omega t). This exact agreement arises because the F = - \frac{\partial V}{\partial x} = - m \omega^2 x is linear in , so the quantum \langle F \rangle = - m \omega^2 \langle x \rangle matches the classical evaluated at \langle x \rangle. In contrast, for nonlinear potentials, higher moments introduce quantum corrections. Coherent states provide a quantum analog to classical motion, minimizing the spread in position and momentum while satisfying the Ehrenfest equations. These states |\alpha\rangle, eigenstates of the annihilation operator a |\alpha\rangle = \alpha |\alpha\rangle, evolve as |\alpha(t)\rangle = | \alpha(0) e^{-i \omega t} \rangle, preserving their Gaussian wavepacket form and yielding: \langle x(t) \rangle = \sqrt{\frac{\hbar}{m \omega}} \left( \alpha(0) e^{-i \omega t} + \alpha^*(0) e^{i \omega t} \right) / \sqrt{2}, with minimum uncertainty \Delta x \Delta p = \hbar / 2. This makes coherent states useful for modeling classical-like quantum behavior in systems like lasers or trapped ions.

Particle in a General Potential

The Ehrenfest theorem provides a framework for understanding the time evolution of expectation values for a quantum particle subject to a general potential V(x) in one dimension. The Hamiltonian for such a system is given by H = \frac{p^2}{2m} + V(x), where p is the , m is the particle mass, and V(x) is an arbitrary . The theorem implies that the expectation values \langle x \rangle and \langle p \rangle satisfy equations analogous to the classical . Specifically, \frac{d}{dt} \langle x \rangle = \frac{\langle p \rangle}{m}, which follows from the [H, x] = i\hbar p / m, using the general form \frac{d}{dt} \langle A \rangle = \frac{i}{\hbar} \langle [H, A] \rangle for a time-independent A. Similarly, \frac{d}{dt} \langle p \rangle = -\left\langle \frac{dV}{dx} \right\rangle, derived from the [H, p] = -i\hbar dV/dx, representing the quantum analog of the classical force F = -dV/dx. These relations hold exactly for any wave function and potential, demonstrating how quantum expectation values can mimic classical trajectories under certain conditions. In practice, the second approximates the classical form \frac{d}{dt} \langle p \rangle \approx -\frac{dV}{d \langle x \rangle} when the potential varies slowly compared to the spatial extent of the wave packet, i.e., when the in \Delta x is small relative to the scale over which V(x) changes significantly. This semiclassical is valid for coherent states or narrow Gaussian wave packets propagating in potentials, where the is nearly across the packet. For example, in a linear potential V(x) = -F x (corresponding to a constant F), the equations reduce exactly to acceleration \frac{d^2}{dt^2} \langle x \rangle = F/m, with no needed since dV/dx is constant. However, for highly nonlinear potentials or delocalized wave functions, quantum effects like spreading or tunneling cause deviations, limiting the classical . This general case extends the insights from specific potentials, such as the , by highlighting the theorem's broad applicability while underscoring its limitations in capturing full , such as wave packet dispersion. The results affirm the correspondence principle, where recovers classical behavior for macroscopic systems or short times before quantum spreading dominates.

Advanced Topics and Applications

Deriving the Schrödinger Equation

The Ehrenfest theorem establishes that, under the time-dependent Schrödinger equation, the expectation values of position and momentum operators evolve according to equations resembling Newton's second law and the classical definition of velocity. Specifically, for a particle of mass m in a potential V(\mathbf{r}), \frac{d}{dt} \langle \hat{\mathbf{r}} \rangle = \frac{\langle \hat{\mathbf{p}} \rangle}{m}, \quad \frac{d}{dt} \langle \hat{\mathbf{p}} \rangle = -\left\langle \nabla V(\hat{\mathbf{r}}) \right\rangle. This correspondence validates ' consistency with in the limit of well-localized wave packets. To derive the in reverse, one starts from the postulate that quantum expectation values must satisfy these classical-like equations for arbitrary states, not merely narrow wave packets, and works backward to infer the required form of the operator \hat{H} and the of the wave function \psi. Assume the commutation relations [\hat{x}, \hat{p}] = i\hbar and that the time derivative of an expectation value follows the general Heisenberg-like form \frac{d}{dt} \langle \hat{A} \rangle = \left\langle \frac{\partial \hat{A}}{\partial t} \right\rangle + \frac{i}{\hbar} \left\langle [\hat{H}, \hat{A}] \right\rangle, where the second term arises from the unitary time evolution generated by \hat{H}. For time-independent operators like \hat{x} and \hat{p}, this simplifies to \frac{d}{dt} \langle \hat{A} \rangle = \frac{i}{\hbar} \left\langle [\hat{H}, \hat{A}] \right\rangle. To match the Ehrenfest equations, \hat{H} must be of the form \hat{H} = \frac{\hat{p}^2}{2m} + V(\hat{x}), as this ensures the commutators yield the classical Poisson bracket structure scaled by i\hbar. Consider the momentum evolution: \frac{d}{dt} \langle \hat{p} \rangle = \frac{i}{\hbar} \left\langle [\hat{H}, \hat{p}] \right\rangle = -\left\langle \frac{\partial V}{\partial x} \right\rangle. Compute [\hat{H}, \hat{p}] = \left[\frac{\hat{p}^2}{2m}, \hat{p}\right] + [V(\hat{x}), \hat{p}]. The kinetic term vanishes since \hat{p}^2 commutes with \hat{p}, leaving [V(\hat{x}), \hat{p}]. For a function f(\hat{x}), the commutator is [f(\hat{x}), \hat{p}] = i\hbar \frac{df}{dx}, derived from the action on a test function: f(\hat{x}) \hat{p} \psi - \hat{p} f(\hat{x}) \psi = -i\hbar f \frac{d\psi}{dx} + i\hbar \frac{df}{dx} \psi + i\hbar f \frac{d\psi}{dx} = i\hbar \frac{df}{dx} \psi. Thus, [V(\hat{x}), \hat{p}] = i\hbar \frac{\partial V}{\partial x}, and \frac{i}{\hbar} \left\langle i\hbar \frac{\partial V}{\partial x} \right\rangle = - \left\langle \frac{\partial V}{\partial x} \right\rangle, confirming the force term. A similar calculation for position yields \frac{d}{dt} \langle \hat{x} \rangle = \frac{\langle \hat{p} \rangle}{m}, with [\hat{H}, \hat{x}] = -\frac{i\hbar}{m} \hat{p}. This requirement holds for all wave functions \psi only if the operator equations match exactly, justifying the drop from expectation values to operator identities via completeness of the function space (e.g., using sharply peaked test functions approximating Dirac deltas). The imaginary unit i in the evolution equation i\hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi emerges from the sign convention needed for forward time evolution consistent with classical dynamics; a real coefficient would reverse the flow. formalized this reverse inference in , arguing that the Schrödinger equation's structure is necessitated by the demand for classical in expectation values, bridging the quantum-to-classical without assuming the equation a priori.

Ehrenfest Time and Modern Applications

The Ehrenfest time, denoted t_E, represents the characteristic timescale over which the values of quantum observables, as governed by the Ehrenfest theorem, closely follow their classical counterparts before significant deviations arise due to quantum effects such as spreading. In regular (non-) systems, this time can be relatively long, but in systems, it is limited by the to initial conditions, scaling logarithmically with the inverse of Planck's constant as t_E \sim \frac{1}{\lambda} \ln \left( \frac{S_0}{\hbar} \right), where \lambda is the classical and S_0 is an initial action scale. Beyond t_E, quantum superpositions extend across the system's , breaking the local classical approximation inherent in the Ehrenfest theorem. This concept has found significant application in the study of , where the Ehrenfest time demarcates the regime of semiclassical validity. For instance, out-of-time-ordered correlators (OTOCs), which quantify information scrambling and chaos in , exhibit exponential growth up to approximately t_E, after which quantum interference effects dominate and lead to saturation. This behavior has been pivotal in bridging with physics and , providing a diagnostic for chaotic versus integrable dynamics in complex systems. Recent analyses extend the Ehrenfest theorem's utility beyond t_E by incorporating environmental decoherence, which suppresses spreading and restores approximate classical correspondence in open . As of 2025, further generalizations have been developed for open quantum systems and relativistic particles. In and , Ehrenfest dynamics—derived from the theorem—enable mixed quantum-classical simulations of non-adiabatic processes, such as electron-nuclear coupling in photochemical reactions. Here, electronic states are treated quantum mechanically while nuclei evolve classically under the mean-field potential from quantum expectation values, offering computational efficiency for large systems despite limitations near t_E in chaotic regimes. Applications include modeling ultrafast charge transfer in and , where the method captures real-time dynamics without full quantum treatment of all . Extensions like damped Ehrenfest approaches mitigate energy non-conservation issues, enhancing accuracy for condensed-phase simulations. Emerging uses in quantum simulation and leverage the Ehrenfest framework to approximate dynamics in noisy intermediate-scale quantum (NISQ) devices. For example, combined with Ehrenfest propagation simulates electron dynamics in warm dense matter, relevant to . In quantum chaos studies on quantum simulators, the Ehrenfest time informs the design of circuits to probe , aiding mitigation strategies by identifying timescales where classical-like behavior holds. These applications underscore the theorem's enduring role in scaling quantum models to practical computations, particularly where full quantum evolution is intractable.

References

  1. [1]
    [1907.02354] Ehrenfest's Theorem Revisited - arXiv
    Jul 4, 2019 · Ehrenfest's theorem (1927) is the first one which shows that classical physics can emerge from quantum physics as a kind of approximation.Missing: primary | Show results with:primary
  2. [2]
    Ehrenfest's Theorem - Richard Fitzpatrick
    Ehrenfest's Theorem. A simple way to calculate the expectation value of momentum is to evaluate the time derivative of $\langle x\rangle$ , and then multiply ...
  3. [3]
    Bemerkung über die angenäherte Gültigkeit der klassischen ...
    Zeitschrift für Physik; Article. Bemerkung über die angenäherte Gültigkeit ... Cite this article. Ehrenfest, P. Bemerkung über die angenäherte Gültigkeit ...
  4. [4]
  5. [5]
    [PDF] Ehrenfest's theorem
    We show that the rate of change of the average value of the momentum is the average value of the force, the negative gradient of the potential.
  6. [6]
    Ehrenfest's adiabatic theory and the old quantum theory, 1916–1918
    Sep 4, 2008 · I show that until 1918 the adiabatic hypothesis did not play an important role in the development of quantum theory. Article PDF. Download to ...
  7. [7]
    [PDF] Ehrenfest's adiabatic hypothesis in Bohr's quantum theory - arXiv
    Ehrenfest, Paul (1913b). “A mechanical theorem of Boltzmann and its relation to the theory of energy quanta.” Proceedings of the Amsterdam Academy 16, 591–597.
  8. [8]
    None
    ### Historical Background on Ehrenfest’s Theorem
  9. [9]
    [PDF] Quantum Theory I, Lecture 7 Notes - MIT OpenCourseWare
    This result is called Ehrenfest's theorem. It is important, when using the second equation, to remember to differentiate the potential first before taking ...
  10. [10]
    Ehrenfest Theorem - Richard Fitzpatrick
    ... Schrödinger picture. By contrast, the operator equation (3.46) only holds if $ {\bf x}$ and $ {\bf p}$ are understood to be Heisenberg dynamical variables ...
  11. [11]
    Ehrenfest Theorem - Richard Fitzpatrick
    This scheme guarantees that quantum mechanics yields the correct classical equations of motion in the classical limit.
  12. [12]
  13. [13]
    [PDF] Hamiltonian approach to Ehrenfest expectation values and ... - arXiv
    May 17, 2016 · Then, the Hamiltonian structure of Ehren- fest's theorem is shown to be Lie-Poisson for a semidirect-product Lie group, named the Ehrenfest.
  14. [14]
    [PDF] Chapter 6 Ehrenfest's Theorem
    Paul Ehrenfest originated a different answer. Erhenfest said replace the dynamical variables of classical mechanics with the expectation values of quantum.<|control11|><|separator|>
  15. [15]
    Ehrenfest, P. (1927) Zeitschrift für Physik, 45, 455. - References
    The main objective of this study is to show that decoherence is an inherent characteristic of quantum states caused by the quantum uncertainty relation.
  16. [16]
    [2206.07531] Bouncing Wave Packets, Ehrenfest Theorem, and ...
    Jun 15, 2022 · We illustrate these results for bouncing wave packets which first spread, then shrink, and return to their original form after a certain revival time.
  17. [17]
    [2305.18644] Classical and semi-classical limits in phase space
    May 29, 2023 · Abstract:A semiclassical approximation is derived by using a ... Ehrenfest's theorem, the resulting approximation is a linear first ...
  18. [18]
    How do wave packets spread? Time evolution on Ehrenfest ... - arXiv
    Dec 17, 2011 · A wave packet propagated along a hyperbolic trajectory becomes a Lagrangian state associated with the unstable manifold at the Ehrenfest time.Missing: theorem | Show results with:theorem
  19. [19]
    [PDF] 9. Harmonic Oscillator - MIT OpenCourseWare
    ... Ehrenfest theorem. Coherent states achieve this result. For this reasons, these states are also called quasi-classical. The coherent state was defined by ...
  20. [20]
    None
    ### Summary of Ehrenfest Theorem Application to Coherent States in Harmonic Oscillator
  21. [21]
  22. [22]
    [2306.13717] Ehrenfest's theorem beyond the Ehrenfest time - arXiv
    Jun 23, 2023 · Title:Ehrenfest's theorem beyond the Ehrenfest time ... Abstract:In closed quantum systems, wavepackets can spread exponentially in time due to ...Missing: mechanics | Show results with:mechanics
  23. [23]
    Ehrenfest dynamics with quantum mechanical nuclei - ScienceDirect
    We evaluate the performance of the Ehrenfest method for molecular dynamics simulation in cases where some atomic nuclei are designated as classical and others ...Missing: modern | Show results with:modern
  24. [24]
    A Quantum Ehrenfest (QuEh) Protocol for Reaction Path Following ...
    Jun 25, 2024 · An algorithm is described for quantum dynamics where an Ehrenfest potential is combined with fully quantum nuclear motion (Quantum-Ehrenfest, ...
  25. [25]
    Real-Time CASSCF (Ehrenfest) Modeling of Electron Dynamics in ...
    Nov 27, 2024 · We present a strategy for the modeling of charge carrier dynamics in organic semiconductors using conventional quantum chemistry methods, ...
  26. [26]
    The damped Ehrenfest (D-Eh) method: Application to non-adiabatic ...
    Mar 15, 2019 · In other work we have shown that [23] the Ehrenfest method can be derived as an approximation to full quantum dynamics methods.
  27. [27]
    Quantum simulation of exact electron dynamics can be ... - Nature
    Jul 10, 2023 · Stopping of deuterium in warm dense deuterium from Ehrenfest time-dependent density functional theory. Contrib. Plasma Phys. 56, 459–466 ...