Fact-checked by Grok 2 weeks ago

Classical limit

The classical limit in physics denotes the regime where the predictions of converge to those of , ensuring theoretical consistency across scales. This is formally approached by taking the reduced Planck's \hbar to zero, or equivalently, by considering systems with large quantum numbers (such as high-energy states) or macroscopic dimensions where quantum fluctuations become negligible compared to classical behavior. Central to this concept is Niels Bohr's , introduced in 1913, which posits that must asymptotically reproduce classical results in the of large quantum numbers, thereby bridging the with the emerging framework of wave and . Mathematically, the classical limit is demonstrated through the (1927), which states that the of values for \langle \hat{x} \rangle and \langle \hat{p} \rangle in obeys the same differential equations as the x and p in classical Hamilton's equations, provided the wave function remains sufficiently localized (e.g., as a narrow Gaussian packet). This theorem highlights how quantum averages mimic classical trajectories under conditions where uncertainties in and are small relative to the system's scale, aligning with the correspondence principle. In contemporary interpretations, the classical limit also involves decoherence, a process where quantum superpositions are suppressed through interactions with an environment, leading to the emergence of classical-like probabilities and definite outcomes without invoking wave function collapse. Decoherence explains why macroscopic objects exhibit classical behavior despite underlying quantum descriptions, as environmental entanglement rapidly localizes the system's state in position space. These mechanisms collectively ensure that quantum theory encompasses classical physics as a valid approximation, resolving foundational tensions while preserving the predictive power of both paradigms.

Fundamentals

Definition and Scope

The classical limit in refers to the regime where the theory reduces to through the formal procedure of taking the limit as the reduced Planck's constant ħ approaches zero, rendering quantum effects such as and negligible compared to classical predictions. In this limit, quantum observables converge to functions on classical , and the quantum dynamics approach the deterministic flow of . This reduction is not always straightforward, as the does not universally yield Newton's without additional assumptions, but it establishes the foundational compatibility between the two theories. The scope of the classical limit is primarily confined to non-relativistic quantum mechanics, where it applies to systems described by the , excluding relativistic extensions or quantum field theories unless specifically adapted. Within this domain, the limit manifests when phase space volumes associated with the system's motion substantially exceed the quantum scale set by ħ, allowing classical trajectories to dominate. Action-angle variables provide a useful framework for this scope, as they transform the into a form dependent only on the action variables, facilitating the identification of classical periodic orbits in the limit. Physically, the classical limit corresponds to macroscopic systems or high-energy quantum states where the Heisenberg uncertainty principle yields spreads in position and momentum (Δx Δp ≈ ħ) that are insignificant relative to the overall scale of the system, effectively suppressing quantum fluctuations. For instance, in everyday macroscopic objects, the de Broglie wavelength is minuscule, aligning quantum expectations with classical intuition. Mathematically, this boundary is marked by conditions such as large quantum numbers n ≫ 1, under which quantized energy levels and orbits closely approximate their classical counterparts. The Ehrenfest theorem exemplifies this alignment by demonstrating that the time evolution of expectation values follows classical equations when wave packets remain localized.

Historical Development

The origins of the classical limit concept emerged in the late 19th and early 20th centuries amid efforts to reconcile with experimental anomalies in . In December 1900, proposed quantization for oscillators in to resolve the , introducing Planck's constant ħ as a fundamental scale; this quantization implied that classical continuous distributions would be recovered in the low-frequency (long-wavelength) or high-temperature limit where quantum effects become negligible. Building on this, Albert Einstein's 1905 paper on the posited light quanta (photons) with E = hν, explaining discrete absorption while noting that in the limit of many quanta or low frequencies, wave-like reemerges. These precursors highlighted the need for quantum theories to transition smoothly to classical descriptions under conditions of macroscopic scales or large quantum numbers. Niels Bohr's 1913 atomic model advanced this transition by postulating stationary states with quantized angular momentum L = nħ (n integer) for electrons in hydrogen-like atoms, preventing classical radiative collapse while predicting spectral lines. In the high-n limit, corresponding to large orbits, the model yielded radiation frequencies and intensities matching classical electrodynamics, as the quantized orbits approximate continuous classical trajectories and the selection rules align with dipole radiation. This intuitive bridge, later formalized as the correspondence principle in Bohr's 1918–1923 works, guided subsequent developments by demanding quantum predictions coincide with classical ones for n → ∞. The mid-1920s marked the birth of , explicitly incorporating classical limits. Werner Heisenberg's 1925 matrix mechanics, developed with and , replaced classical and with non-commuting arrays, deriving transition amplitudes from spectral data; the formalism ensured compliance with the correspondence principle, where matrix elements for large quantum numbers reproduce classical Fourier components of periodic motion. Independently, Erwin Schrödinger's 1926 wave mechanics posited a ψ satisfying iħ ∂ψ/∂t = Ĥψ, derived via the optical-mechanical analogy to Hamilton-Jacobi theory; in the short-wavelength limit (λ → 0 or ħ → 0), solutions reduce to classical rays, recovering deterministic trajectories. The equivalence of these approaches, proven in 1926, solidified as a theory whose classical limit is not ad hoc but intrinsic. Post-1930s formalizations provided mathematical rigor to the ħ → 0 limit. Paul Dirac's 1930 framed observables as operators on , showing that commutators [Â, B̂] = iħ Ĉ̂ become classical brackets {A, B} = C in the ħ → 0 scaling, thus deriving classical dynamics from quantum algebra. John von Neumann's 1932 extended this by axiomatizing quantum states and observables in von Neumann algebras, where the classical limit corresponds to commutative subalgebras mimicking phase-space functions, ensuring statistical predictions align with classical probabilities for macroscopic systems. These works established the classical limit as a rigorous asymptotic regime in . From the onward, modern perspectives shifted emphasis from strict ħ → 0 to environmentally induced emergence of classicality via decoherence. Zurek's seminal 1981 paper on pointer states and subsequent works (e.g., 1991 review) demonstrated that interactions with an rapidly suppress quantum superpositions, selecting preferred states (e.g., position eigenstates) that behave classically due to entanglement and tracing out environmental , resolving the without invoking collapse. This framework, refined through einselection (environment-induced superselection) in Zurek's 2003 synthesis, explains classical stability in open quantum systems at finite ħ, influencing fields like up to 2025. By 2025, decoherence remains central to understanding classical emergence in noisy quantum devices, bridging foundational theory with practical quantum technologies.

Quantum Mechanical Formulation

Correspondence Principle

The correspondence principle, introduced by Niels Bohr in 1913, posits that in the limit of large quantum numbers, the quantum description of physical processes must asymptotically approach the corresponding classical description, particularly for radiation emission and absorption. In his paper "On the Quantum Theory of Line-Spectra," Bohr emphasized the analogy between quantum transitions and classical radiation theory, stating that the frequency of light emitted in a quantum jump between stationary states should match the frequency of the classical harmonic motion for large quantum numbers. This principle served as a foundational heuristic, ensuring consistency between the emerging quantum theory and well-established classical electromagnetism without relying on a complete dynamical framework. A key aspect of Bohr's formulation involves the matching of quantum amplitudes to the components of classical motion. As elaborated in his 1920 work "On the Series Spectra of the Elements," for large s, the relative probability of a quantum corresponds directly to the of the relevant in the classical orbital motion. This was applied to atomic spectra, where the frequency \Delta E / \hbar between high-lying states aligns with the classical oscillation frequency of the electron in Rydberg states, providing a bridge between discrete quantum energy levels and continuous classical orbits. For instance, in the , the lines converge to the classical frequency as the principal n \to \infty, illustrating how quantum predictions recover classical patterns in the high-energy limit. The principle played a crucial role in the , guiding the development of quantization rules for systems like the without full derivations from first principles. Bohr used it to infer selection rules and intensities for spectral lines, extending the 1913 model to more complex atoms by assuming quantum behavior mimics classical in the appropriate regime. However, it is not a rigorous but rather a consistency check, relying on intuitive analogies rather than formal proofs. Its limitations become evident in modern contexts, such as , where the principle's original formulation inadequately addresses field quantization and coherent states, requiring generalizations beyond large quantum numbers for bound systems.

Ehrenfest Theorem

The Ehrenfest theorem establishes that the time evolution of expectation values of position and momentum operators in quantum mechanics follows equations analogous to the classical Hamilton's equations of motion. Specifically, for a particle in a potential V(x) with Hamiltonian H = \frac{p^2}{2m} + V(x) in the Schrödinger picture, the theorem states \frac{d}{dt} \langle x \rangle = \frac{\langle p \rangle}{m}, \quad \frac{d}{dt} \langle p \rangle = -\left\langle \frac{dV}{dx} \right\rangle. This result was first derived by Paul Ehrenfest in 1927. The derivation begins with the general formula for the time derivative of the expectation value of a time-independent operator A: \frac{d}{dt} \langle A \rangle = \frac{i}{\hbar} \langle [H, A] \rangle, which follows from the Schrödinger equation i\hbar \frac{\partial}{\partial t} |\psi\rangle = H |\psi\rangle and its adjoint, applied to \langle A \rangle = \langle \psi | A | \psi \rangle. For A = x, the commutator is [H, x] = -\frac{i\hbar}{m} p using the canonical commutation relation [x, p] = i\hbar and [V(x), x] = 0, yielding \frac{d}{dt} \langle x \rangle = \frac{\langle p \rangle}{m}. For A = p, [H, p] = i\hbar \frac{dV}{dx} since [ \frac{p^2}{2m}, p ] = 0, giving \frac{d}{dt} \langle p \rangle = -\left\langle \frac{dV}{dx} \right\rangle. In the classical limit, these equations imply that expectation values \langle x \rangle and \langle p \rangle trace classical trajectories when the quantum wave packet is sufficiently localized, such that the uncertainty \Delta x and \Delta p are much smaller than the and scales of the potential (e.g., \Delta x \ll \lambda where \lambda is the scale over which V(x) varies significantly). Under this , \left\langle \frac{dV}{dx} \right\rangle \approx \frac{[dV](/page/DV)}{dx} \big|_{\langle x \rangle}, reducing the to the classical m \ddot{x} = -\frac{dV}{dx}. The holds exactly for quadratic potentials, such as the V(x) = \frac{1}{2} m \omega^2 x^2, where \left\langle \frac{dV}{dx} \right\rangle = m \omega^2 \langle x \rangle for any state, so quantum expectation values oscillate precisely as classical trajectories. It is approximate for other potentials due to wave packet spreading, which violates the localization assumption over long times. For a (V=0), the theorem gives constant \langle p \rangle and linear \langle x \rangle(t) = \langle x \rangle(0) + \frac{\langle p \rangle}{m} t, matching classical motion, though the position variance grows as \sigma_x^2(t) = \sigma_x^2(0) + \left( \frac{\hbar t}{2 m \sigma_x(0)} \right)^2.

Derivations and Approximations

Time Evolution of Expectation Values

In the , the of operators is given by \hat{x}(t) = U^\dagger(t) \hat{x} U(t) and \hat{p}(t) = U^\dagger(t) \hat{p} U(t), where U(t) = e^{-i \hat{H} t / \hbar} is the time-evolution operator and \hat{H} is the . The values \langle \hat{x}(t) \rangle = \langle \psi | \hat{x}(t) | \psi \rangle and \langle \hat{p}(t) \rangle = \langle \psi | \hat{p}(t) | \psi \rangle, taken with respect to a fixed state |\psi\rangle, then satisfy the Ehrenfest equations \frac{d}{dt} \langle \hat{x} \rangle = \frac{\langle \hat{p} \rangle}{m} and \frac{d}{dt} \langle \hat{p} \rangle = -\left\langle \frac{\partial V}{\partial x} \right\rangle, which parallel the classical provided the average force approximates the force evaluated at the average position. For Hamiltonians quadratic in position and momentum, such as the harmonic oscillator \hat{H} = \frac{\hat{p}^2}{2m} + \frac{1}{2} m \omega^2 \hat{x}^2, the solutions for \langle \hat{x}(t) \rangle and \langle \hat{p}(t) \rangle coincide exactly with classical trajectories, with no quantum corrections arising over time. This exact correspondence exemplifies the classical limit in time-dependent scenarios. A key illustration is the dynamics of coherent states in the harmonic oscillator, which are Gaussian wavepackets minimizing the uncertainty product and exhibiting classical-like behavior. These states, originally proposed by Schrödinger and formalized by Glauber, evolve under the Hamiltonian as |\alpha(t)\rangle = e^{-i \omega t / 2} |\alpha(0) e^{-i \omega t}\rangle, preserving their Gaussian form without spreading. The expectation values follow \langle \hat{x}(t) \rangle = \sqrt{\frac{2\hbar}{m\omega}} |\alpha(0)| \cos(\omega t + \phi), \quad \langle \hat{p}(t) \rangle = -\sqrt{2 m \hbar \omega} |\alpha(0)| \sin(\omega t + \phi), where \phi is the initial phase, tracing a classical elliptical orbit in phase space with amplitude determined by |\alpha(0)|. This motion mirrors the solution to Hamilton's equations for a classical oscillator with the same initial energy. Deviations from classical evolution occur in nonlinear (anharmonic) potentials, where the Ehrenfest equations hold but the nonlinearity causes \left\langle \frac{\partial V}{\partial x} \right\rangle \neq \frac{\partial V}{\partial \langle x \rangle} as the wavepacket samples varying forces across its width. This mismatch leads to gradual divergence of quantum expectation values from classical ones, accompanied by wavepacket spreading that amplifies the effect. Numerical studies of Gaussian wavepacket in anharmonic potentials, such as the V(x) = D_e (1 - e^{-\alpha (x - x_e)})^2, demonstrate this breakdown. Initial localized Gaussians follow classical trajectories for the center of mass over times comparable to the classical period, but emerges due to the position-dependent , with the packet width increasing as \Delta x(t) \approx \Delta x(0) \sqrt{1 + (\hbar t / m (\Delta x(0))^2)^2} modulated by . In thawed Gaussian approximations applied to such systems, agreement persists until a spreading timescale t \sim \omega^{-1} (E_{cl} / \hbar \omega)^{1/2}, after which quantum delocalization dominates and classical predictability is lost. The of expectation values also underpins the quantum , obtained by applying Ehrenfest dynamics to the \frac{1}{2} (\hat{x} \hat{p} + \hat{p} \hat{x}), yielding \frac{d}{dt} \left\langle \frac{1}{2} (\hat{x} \hat{p} + \hat{p} \hat{x}) \right\rangle = 2 \langle \hat{T} \rangle + \left\langle x \frac{\partial V}{\partial x} \right\rangle. For bound states or long-time averages where the left side vanishes, this reduces to $2 \langle \hat{T} \rangle = \left\langle x \frac{\partial V}{\partial x} \right\rangle. In the classical limit for potentials with V = \frac{1}{2} m \omega^2 x^2, it simplifies to \langle \hat{T} \rangle = \langle V \rangle, aligning with classical equipartition and time-averaged energies for periodic motion. This connection illustrates how recover classical statistical relations for bound systems under conditions.

Semiclassical Methods

Semiclassical methods in offer systematic approximations that connect quantum descriptions to classical dynamics, particularly in regimes where the reduced ħ is small relative to the characteristic scales of the . These techniques expand quantum observables, such as energy levels and wavefunctions, in powers of ħ, with the leading order recovering classical results and higher orders incorporating quantum corrections like and tunneling. They are grounded in the , where quantum values align with classical trajectories in the ħ → 0 limit, and are widely used for analytically intractable problems in atomic, molecular, and . A cornerstone of semiclassical methods is the Wentzel–Kramers–Brillouin (WKB) approximation, developed independently in 1926, which addresses the one-dimensional time-independent for potentials varying slowly compared to the de Broglie wavelength. The wavefunction is approximated as ψ(x) ≈ [p(x)]^{-1/2} \exp\left( \pm \frac{i}{\hbar} \int^x p(x') , dx' \right), where p(x) = \sqrt{2m(E - V(x))} is the classical , leading to the |∇S|^2 = 2m(E - V) for the action S in higher dimensions. For bound states, the WKB quantization condition ∫_{turning points} p(x) , dx = \left(n + \frac{1}{2}\right) \pi \hbar yields accurate energy levels, improving upon the older Bohr–Sommerfeld by including a Maslov index correction for turning points. This method excels in predicting tunneling probabilities through barriers, with transmission coefficients T ≈ \exp\left( - \frac{2}{\hbar} \int |p(x)| , dx \right), and extends to multidimensional integrable systems via the Einstein–Brillouin–Keller (EBK) scheme, where action-angle variables satisfy similar quantization ∫ p_i , dq_i = \left(n_i + \frac{\mu_i}{4}\right) h, with μ_i as Maslov indices. In non-integrable and chaotic systems, semiclassical methods shift focus to periodic orbits, as captured by Gutzwiller's trace formula from 1967, which expresses the quantum density of states d(E) as a smooth Weyl term plus oscillatory contributions from classical periodic orbits: d(E) ≈ \frac{1}{2\pi \hbar} \sum_\gamma A_\gamma \exp\left( i \frac{S_\gamma}{\hbar} - \frac{i}{2} \mu_\gamma \pi \right) \cos(\theta_\gamma), where S_\gamma is the action, A_\gamma the stability amplitude, μ_\gamma the Maslov index, and θ_\gamma a phase. This formula links quantum spectral fluctuations to classical chaos, explaining level repulsion and statistics akin to random matrix theory in the semiclassical limit. Complementing this, Feynman's path integral formulation provides a semiclassical propagator via the van Vleck–Pauli–Morette determinant, where the kernel K(x_f, t; x_i, 0) ≈ \sum_{classical paths} \left| \det \frac{\partial^2 S}{\partial x_f \partial x_i} \right|^{1/2} \exp\left( \frac{i}{\hbar} S - \frac{i \pi}{2} \nu \right), with ν as the index of the Hessian, enabling approximations for time evolution and scattering amplitudes. These approaches highlight how quantum interference arises from classical paths, with applications in quantum chaos and molecular dynamics.

Broader Contexts

Relativistic Extensions

In , the classical limit is explored through the behavior of expectation values and semiclassical approximations for wave equations like the Dirac and Klein-Gordon equations, where the limit \hbar \to 0 or c \to \infty recovers classical trajectories while accounting for relativistic effects such as . This extension builds on the non-relativistic by deriving analogous relations for relativistic cases, ensuring that quantum averages align with classical under appropriate conditions. For the describing particles, the classical limit \hbar \to 0 is obtained via the Foldy-Wouthuysen transformation, which decouples components and yields the for a classical with intrinsic . In this representation, the transformation facilitates a semiclassical approximation where quantum fluctuations diminish, leading to classical trajectories. In the Klein-Gordon case for spin-0 particles, the Ehrenfest theorem applied to the equation shows that expectation values of position and momentum satisfy relativistic equations of motion in the classical limit. This limit emerges when \hbar \to 0, suppressing probability current ambiguities and negative energy issues inherent to the Klein-Gordon equation, thus restoring a probabilistic interpretation consistent with classical scalar field propagation. The Zitterbewegung, or trembling motion, in the Dirac equation arises from interference between positive and negative energy states, manifesting as rapid oscillations around the classical trajectory with amplitude on the order of the Compton wavelength \hbar / mc. In the classical limit \hbar \to 0, these oscillations disappear, as the spin-related terms proportional to \hbar^2 vanish, leaving a smooth classical path; alternatively, in the high-energy limit where particle energy greatly exceeds mc^2, the relative amplitude of the trembling diminishes, averaging to the expected classical relativistic trajectory. To recover non-relativistic from these relativistic formulations, one imposes either c \to \infty (reducing to the for the ) or \hbar \to 0 while keeping velocities subluminal, effectively decoupling relativistic corrections and yielding Newtonian dynamics. Recent numerical simulations of Dirac fermions in , modeled via semiclassical approximations like the frozen Gaussian method, demonstrate this emergence: under strain-induced curved , wave packet dynamics in the \hbar \to 0 regime exhibit classical trajectories along pseudospin directions, validating the relativistic classical limit in condensed-matter analogs without invoking field . These computations, performed on tight-binding Hamiltonians, show reduced quantum spreading and alignment with Lorentz-like forces in effective fields, bridging high-energy with in 2020s experiments.

Deformations in Other Theories

In quantum field theory, the classical limit is approached by taking the Planck constant ħ to zero, which effectively reduces the quantum theory to a classical field theory where fields obey deterministic equations of motion derived from the action principle. For scalar fields, this limit transforms the quantized Klein-Gordon field into a classical scalar field satisfying the Klein-Gordon equation (\square + m^2)\phi = 0, where \square = \partial^\mu \partial_\mu is the d'Alembertian operator; in the massless case (m=0), this recovers the classical wave equation \square \phi = 0. Similarly, in gauge theories, the classical limit of non-Abelian Yang-Mills quantum field theory yields classical Yang-Mills theory, whose Abelian specialization (e.g., U(1) gauge group) corresponds to classical electromagnetism governed by Maxwell's equations. Additionally, in relativistic contexts, taking the speed of light c \to \infty within these classical field theories further yields non-relativistic approximations, such as the Schrödinger-Poisson system from scalar field theories. The provides a clear framework for understanding this classical limit in QFT. The quantum transition amplitude is given by a functional over configurations, \int \mathcal{D}\phi \, e^{iS[\phi]/\hbar}, where S[\phi] is the classical . In the \hbar \to 0, the (or stationary phase) dominates the , concentrating contributions around the classical configurations \phi_\mathrm{cl} that extremize the , i.e., solutions to the Euler-Lagrange equations \delta S / \delta \phi = 0. Fluctuations around these saddle points yield semiclassical corrections, but the leading term reproduces the classical equations exactly, underscoring how quantum interference suppresses non-classical paths. This approach extends the single-particle to infinite in theories. Deformations of standard quantum theories often incorporate the classical limit as a recovery mechanism in low-energy regimes. A prominent example is , a non-commutative arising from the deformation of the Poincaré with deformation $1/\kappa (related to the Planck scale). In this framework, the \kappa-Poincaré modifies commutation relations among coordinates and momenta, but in the low-energy limit ($1/\kappa \to 0), a nonlinear change of generators maps the deformed structure back to the undeformed Poincaré , restoring classical spacetime symmetries and Lorentz invariance. This ensures compatibility with at observable energies while allowing effects at high scales. Recent advancements in (LQG), a non-perturbative quantization of , similarly demonstrate a classical limit through semiclassical approximations. For instance, formulations of LQG recover the in the limit of large spin networks or via saddle-point evaluations of the partition function, with studies from the onward refining this limit to include couplings and cosmological models. Despite these successes, challenges arise in achieving a strict classical limit in certain QFTs, particularly due to divergences in massless theories. Loop contributions in the \hbar \to 0 expansion do not always vanish, as higher-order diagrams can persist or diverge at low momenta, complicating the reduction to purely classical equations; for example, in , IR divergences from soft emissions require regularization (e.g., via finite volume or mass terms) before taking the limit, preventing a naive recovery of without additional infrared cutoffs. These issues highlight the non-trivial interplay between quantum fluctuations and classical in field theories with unbounded .

References

  1. [1]
    What is the limit ℏ → of quantum theory? | American Journal of Physics
    classical mechanics may be regarded as the limiting case of quantum mechanics when ℏ tends to zero.” In quantum mechanics, a single particle in an external ...INTRODUCTION · IV. THE DETERMINISTIC... · VI. THE COMBINED LIMIT OF...<|control11|><|separator|>
  2. [2]
    Bohr's Correspondence Principle
    Oct 14, 2010 · ... correspondence principle as some sort of asymptotic relation holding in the classical limit. There is a remarkable continuity in Bohr's ...
  3. [3]
    Ehrenfest Theorem - Richard Fitzpatrick
    This scheme guarantees that quantum mechanics yields the correct classical equations of motion in the classical limit. Whenever an ambiguity arises because ...
  4. [4]
    [PDF] 131 Chapter 7 Uncertainty, and the Classical Limit Previously, we ...
    ... states that in the classical limit the predictions of quantum mechanics should closely approximate those of classical physics. It was a powerful ...
  5. [5]
    The Role of Decoherence in Quantum Mechanics
    Apr 21, 2020 · The theory of decoherence is precisely the study of such situations. It is is relevant (or is claimed to be relevant) to a variety of questions ...
  6. [6]
    [PDF] Decoherence and the Classical Limit
    Classical mechanics deals with objects which have a precise location and move in a deterministic way as a function of time. By contrast, quantum mechanics ...
  7. [7]
    classical limit in nLab
    Oct 21, 2022 · For a given quantum system, the limit ħ → 0 \hbar \to 0 (in as far as that can be made sense of), is called its classical limit.
  8. [8]
    [quant-ph/9504016] The classical limit of quantum theory - arXiv
    Apr 24, 1995 · The classical limit of quantum theory is defined as where quantum observables converge to a function on phase space, and quantum dynamics ...Missing: volume | Show results with:volume
  9. [9]
    [1201.0150] What is the limit $\hbar \to 0$ of quantum theory ? - arXiv
    Dec 30, 2011 · Abstract:An analysis is made of the relation between quantum theory and classical mechanics, in the context of the limit \hbar \to 0.
  10. [10]
    Action-angle variables in quantum mechanics - AIP Publishing
    In this work it is shown how the quantum expressions may be reformulated in terms of the canonical set of action-angle variables, and specific examples of the.
  11. [11]
    Classical Limit - an overview | ScienceDirect Topics
    The classical limit refers to the scenario where quantum behavior transitions to classical behavior, particularly as the quantum phase space cell size ...
  12. [12]
    [PDF] QUANTUM MECHANICS
    the passage from quantum mechanics to the limit of classical mechanics occurs similarly to the passage from wave optics to geometrical optics. Let us recall ...
  13. [13]
    [PDF] THE ORIGINS OF THE QUANTUM THEORY
    The wheel made its final turn when Einstein connected the Bohr atom back to blackbody radiation. ... In the “classical limit,” for large orbits with high quantum ...
  14. [14]
    [1307.2132] The 1905 Relativity Paper and the "Light Quantum" - arXiv
    Jul 8, 2013 · Title:The 1905 Relativity Paper and the "Light Quantum" ; Subjects: History and Philosophy of Physics (physics.hist-ph) ; Cite as: arXiv:1307.2132 ...
  15. [15]
    Celebrating the Birth Centenary of Quantum Mechanics: A Historical ...
    May 5, 2025 · In 1913, in a series of three papers, Bohr proposed an atomic model that resolved the contradictions. While Einstein extended Planck's quantum ...
  16. [16]
    Bohr's Correspondence Principle
    Oct 14, 2010 · More precisely, Bohr's selection rule states that the transition from a stationary state n′ to another stationary state n″ is allowed if and ...
  17. [17]
    The 1925 revolution of matrix mechanics and how to celebrate it in ...
    Jan 1, 2025 · In 1925, Heisenberg, Born, and Jordan developed matrix mechanics as a strategy to solve quantum-mechanical problems.
  18. [18]
    Erwin Schrödinger's Path to Wave Mechanics
    In 1926, Schrödinger invoked Hamilton's optical-mechanical analogy that related ray optics and particle mechanics: His invention of wave mechanics.
  19. [19]
    What is the limit ℏ → 0 \hbar \to 0 of quantum theory - ResearchGate
    An analysis is made of the relation between quantum theory and classical mechanics, in the context of the limit ℏ → 0 \hbar \to 0 .
  20. [20]
    Decoherence, einselection, and the quantum origins of the classical
    May 22, 2003 · The manner in which states of some quantum systems become effectively classical is of great significance for the foundations of quantum ...Missing: emergence | Show results with:emergence
  21. [21]
    [PDF] 1918 on the quantum theory of line–spectra. - UB
    The quantum theory of line–spectra rests upon the following fundamental assumptions: • I. That an atomic system can, and can only, exist permanently in a.
  22. [22]
    Bohr's Correspondence Principle
    Oct 14, 2010 · Bohr postulates that when the electron is in one of these stationary states, its motions can be adequately described by means of classical ...Missing: Kepler | Show results with:Kepler
  23. [23]
    [PDF] The Theory of Spectra and Atomic Constitution. - Project Gutenberg
    1920. Page 42. on the series spectra of the elements. 31 ber of electrons revolving about it. Although the nucleus is assumed to be very small in proportion ...
  24. [24]
    [PDF] The many faces of the Bohr atom - arXiv
    Thus, in this first discussion of the correspondence principle for frequencies he was in effect using Rydberg states as an illustration. Bohr's expectation ...<|control11|><|separator|>
  25. [25]
    [PDF] Ehrenfest's theorem - Neo-classical physics
    Remark on the approximate validity of classical mechanics within quantum mechanics. By P. Ehrenfest in Leiden, Holland. Translated by D. H. Delphenich. ( ...
  26. [26]
    [PDF] Chapter 6 Ehrenfest's Theorem
    Erhenfest said replace the dynamical variables of classical mechanics with the expectation values of quantum mechanics and you get the same relations.
  27. [27]
    None
    ### Summary of Ehrenfest's Theorem from the Document
  28. [28]
    [PDF] Semiclassical approximations in wave mechanics
    Three classes of approximation method are commonly employed in quantum mechanics. Perturbation techniques produce series expansions for quantities of.Missing: seminal | Show results with:seminal
  29. [29]
    Classical Limit of Relativistic Quantum Mechanical Equations in the ...
    Feb 8, 2013 · Using the Foldy-Wouthuysen representation allows the problem of finding a classical limit of relativistic quantum mechanical equations to be reduced.
  30. [30]
    A semiclassical description of relativistic spin without the use of ...
    In particular, the position variable experiences Zitterbewegung in noninteracting theory. The classical equations for the spin tensor are the same as those of ...
  31. [31]
    Relativistic laser-plasma interactions in the quantum regime
    Apr 21, 2011 · It shall be noted that the Zitterbewegung disappears if the ... However, in the classical limit ℏ → 0 , we instead have the plasma ...
  32. [32]
    [PDF] Frozen Gaussian Approximation for the Dirac Equation in Curved ...
    In this paper, we derive the frozen Gaussian approximation (FGA) for com- puting the solution to the Dirac equation in curved space in the semi-classical regime ...
  33. [33]
    (PDF) Semi-classical limit of the Dirac equation in curved space and ...
    Aug 6, 2025 · The semi-classical regime of static Dirac matter is derived from the Dirac equation in curved space-time. The leading- and next-to-leading-order ...
  34. [34]
    Limit in Quantum Field Theory | Phys. Rev.
    It is shown that as a rule the contributions of loop diagrams do not vanish in the limit ℏ → 0 . Remarks are made on the relevance of the ℏ → 0 limit in ...
  35. [35]
    [PDF] 2. Free Fields
    relativistic free theory is the classical Klein-Gordon (KG) equation for a real scalar field (x, t),. @µ@µ. + m2. = 0. (2.4). To exhibit the coordinates in ...
  36. [36]
    [PDF] Path Integral Methods and Applications - arXiv
    Secondly, the classical limit of quantum mechanics can be understood in a particularly clean way via path integrals. It is in quantum field theory, both ...
  37. [37]
    None
    Summary of each segment:
  38. [38]
    [PDF] A Short Review of Loop Quantum Gravity - arXiv
    Apr 9, 2021 · Formally, in the limit ~ → 0, one can approximate the path-integral (10) by perturbations around a saddle-point: A classical spacetime with ...