Fact-checked by Grok 2 weeks ago

Quantum dynamics

Quantum dynamics is the branch of that investigates the of , governed by the time-dependent i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi, where \psi represents the wave function, \hbar is the reduced Planck's constant, and \hat{H} is the operator encoding the system's total energy. This equation describes how quantum states propagate unitarily in isolated systems, preserving probabilities and enabling predictions of phenomena such as wave packet spreading, tunneling, and . At its core, quantum dynamics builds on the foundational , including wave-particle duality and the probabilistic interpretation of the wave function, where |\psi|^2 yields the probability density for particle positions. The Hamiltonian typically decomposes into terms, such as \frac{\hat{p}^2}{2m} for \hat{p} and mass m, V(\hat{x}), allowing modeling of diverse systems from orbitals to molecular . For time-independent Hamiltonians, solutions involve energy eigenstates, with the full expressed as a superposition \psi(t) = \sum_n c_n e^{-i E_n t / \hbar} \phi_n, revealing oscillatory phases that underpin dynamic behaviors like Rabi oscillations in . The field employs multiple theoretical pictures to analyze evolution: the , where states evolve while operators remain fixed; the , where operators evolve as \hat{A}(t) = e^{i \hat{H} t / \hbar} \hat{A} e^{-i \hat{H} t / \hbar}, mirroring classical dynamics; and the , useful for perturbative treatments of time-varying perturbations. These frameworks highlight conservation laws, such as probability via unitarity (\hat{U}^\dagger \hat{U} = \hat{I}) and energy for time-independent cases, while the Heisenberg uncertainty principle \Delta x \Delta p \geq \hbar/2 imposes fundamental limits on trajectory predictability. Historically rooted in the 1920s development of by pioneers like , who formulated the equation in 1926, quantum dynamics has evolved to address open systems through master equations like the Lindblad form, accounting for dissipation and decoherence in realistic environments. In recognition of its importance, the proclaimed 2025 as the International Year of Quantum Science and Technology. Its applications span , where it explains spectral line broadening; condensed matter, modeling transport; and , simulating gate operations in quantum computers. Recent advances leverage numerical methods, such as trajectory-guided Gaussian wave packets, to simulate complex many-body dynamics intractable by classical means.

Fundamental Concepts

Time Evolution Operator

In quantum mechanics, the time evolution operator U(t, t_0) governs the deterministic change of a quantum state from an initial time t_0 to a later time t, such that the state vector transforms as |\psi(t)\rangle = U(t, t_0) |\psi(t_0)\rangle. This operator is unitary, satisfying U^\dagger(t, t_0) U(t, t_0) = I and U(t, t_0) U^\dagger(t, t_0) = I, which ensures that the evolution preserves the inner product between any two states and thus maintains the probabilistic interpretation of ./03%3A__Time-Evolution_Operator/3.01%3A_Time-Evolution_Operator) The time evolution operator derives directly from the time-dependent i \hbar \frac{\partial}{\partial t} |\psi(t)\rangle = H(t) |\psi(t)\rangle, where H(t) is the operator. For a time-independent Hamiltonian, the solution is the simple exponential U(t, t_0) = \exp\left( -\frac{i}{\hbar} H (t - t_0) \right), obtained by formal assuming the U(t_0, t_0) = I. For time-dependent Hamiltonians, the formal solution involves time-ordering to account for non-commuting s at different times: U(t, t_0) = \mathcal{T} \exp\left( -\frac{i}{\hbar} \int_{t_0}^t H(t') \, dt' \right), where \mathcal{T} is Dyson's time-ordering , which arranges all operator products in chronological order from left to right. This expression, introduced by in 1949, expands as a perturbative summing all possible time-ordered contractions of the interaction terms. Key properties of U(t, t_0) stem from the Hermiticity of the Hamiltonian. Unitarity guarantees that probabilities remain conserved, as the norm \langle \psi(t) | \psi(t) \rangle = 1 holds for all t if it holds at t_0. For time-independent Hamiltonians, the evolution exhibits time-reversal invariance: applying U(-t, 0) reverses the state to its initial form, since U(-t, 0) = U^\dagger(t, 0) = \exp\left( \frac{i}{\hbar} H t \right), reflecting the reversible nature of isolated quantum dynamics. The Hamiltonian serves as the infinitesimal generator of this evolution, with U(\delta t, t) \approx I - \frac{i}{\hbar} H \delta t for small time increments. A representative example is the , where H = \frac{p^2}{2m} with no potential. The operator is U(t, 0) = \exp\left( -\frac{i p^2 t}{2 m \hbar} \right), which in the position representation causes wave packets to spread diffusively over time due to the quadratic momentum dependence, illustrating without external forces. Another example is the one-dimensional with H = \frac{p^2}{2m} + \frac{1}{2} m \omega^2 x^2. Its , the position-space matrix element \langle x | U(t, 0) | x' \rangle, is given by K(x, t; x', 0) = \sqrt{\frac{m \omega}{2 \pi i \hbar \sin(\omega t)}} \exp\left[ \frac{i m \omega}{2 \hbar \sin(\omega t)} \left( (x^2 + x'^2) \cos(\omega t) - 2 x x' \right) \right], which periodically revives the initial every period T = 2\pi / \omega, demonstrating coherent oscillatory behavior.

Heisenberg Picture

In the Heisenberg picture of quantum mechanics, the state vectors of a quantum system are time-independent, while the operators representing physical observables evolve with time. This contrasts with the Schrödinger picture, where states evolve and operators are typically fixed. The picture was introduced by Werner Heisenberg in his foundational 1925 paper on matrix mechanics, providing a formulation that emphasizes observable quantities over wave functions. The transformation between the Schrödinger and Heisenberg pictures is achieved via the time evolution operator U(t), which is unitary. In the Schrödinger picture, operators A_S are time-independent (assuming no explicit time dependence), and states evolve as |\psi_S(t)\rangle = U(t) |\psi_S(0)\rangle. In the Heisenberg picture, states remain fixed at |\psi_H\rangle = |\psi_S(0)\rangle, and operators transform as A_H(t) = U^\dagger(t) A_S U(t), ensuring that expectation values match between pictures: \langle A \rangle_S(t) = \langle \psi_S(t) | A_S | \psi_S(t) \rangle = \langle \psi_H | A_H(t) | \psi_H \rangle. This equivalence preserves all physical predictions. The time evolution of Heisenberg operators follows the Heisenberg equation of motion. For an operator A_H(t) with possible explicit time dependence (i.e., A_S may depend on t), the equation is \frac{d A_H(t)}{dt} = \frac{i}{\hbar} [H, A_H(t)] + \frac{\partial A_H(t)}{\partial t}, where H is the Hamiltonian (assumed time-independent here), [ \cdot, \cdot ] denotes the commutator, and \hbar is the reduced Planck's constant. If A_S has no explicit time dependence, the partial derivative vanishes, simplifying to \frac{d A_H(t)}{dt} = \frac{i}{\hbar} [H, A_H(t)]. This form directly generalizes classical equations of motion, with the commutator [H, A_H]/i\hbar analogous to the Poisson bracket in Hamiltonian mechanics. The offers advantages in relating quantum dynamics to and computing values. The evolution mirrors classical trajectories more closely through the structure, facilitating the via the correspondence principle. It is particularly useful for calculating time-dependent values \langle A(t) \rangle = \langle \psi | A_H(t) | \psi \rangle, as states are static, simplifying functions and response applications. A concrete example illustrates this for a of m, with H = \frac{p^2}{2m} (no potential). The remains constant: p_H(t) = p_H(0), since [H, p_H] = 0. The evolves linearly: x_H(t) = x_H(0) + \frac{p_H(0)}{m} t, reflecting classical uniform motion. These satisfy the Heisenberg equation, as \frac{d x_H}{dt} = \frac{i}{\hbar} [H, x_H] = \frac{p_H}{m} and \frac{d p_H}{dt} = 0.

Mathematical Framework

Time-Dependent Schrödinger Equation

The time-dependent Schrödinger equation provides the foundational postulate for the time evolution of a quantum system's state in the Schrödinger picture of quantum mechanics. It is expressed as i \hbar \frac{\partial}{\partial t} |\psi(t)\rangle = \hat{H} |\psi(t)\rangle, where |\psi(t)\rangle denotes the time-dependent state vector in Hilbert space, \hat{H} is the Hamiltonian operator embodying the system's total energy, \hbar is the reduced Planck constant, and the equation holds for isolated quantum systems. This differential equation dictates how the quantum state changes deterministically under the influence of the Hamiltonian, enabling predictions of observables' expectation values over time. Erwin Schrödinger derived this equation in 1926 through his development of wave mechanics, motivated by analogies between classical wave equations and de Broglie's matter waves, initially postulating a real-valued form before adopting the complex version to match empirical spectra like the atom's. His work reconciled wave-particle duality, providing a linear, deterministic framework distinct from the probabilistic of Heisenberg and . The equation's introduction marked a pivotal unification in , earning Schrödinger the 1933 shared with Dirac. For systems with time-independent Hamiltonians, solutions often involve separating the wave function into spatial and temporal components, assuming \psi(\mathbf{r}, t) = \phi(\mathbf{r}) T(t), which yields T(t) \propto e^{-i E t / \hbar} and reduces the problem to the time-independent \hat{H} \phi(\mathbf{r}) = E \phi(\mathbf{r}) for energy eigenvalues E. These separable solutions correspond to states, where probability densities |\psi|^2 remain time-invariant despite phase . The equation's validity relies on boundary conditions ensuring the Hamiltonian is self-adjoint, such as square-integrability of \psi over the configuration space (vanishing at infinity for unbound systems) or periodicity for periodic potentials, which define the operator's domain. These conditions guarantee unitarity of the evolution, preserving the normalization \langle \psi(t) | \psi(t) \rangle = 1 if initially normalized, as the inner product remains constant due to the Hermitian nature of \hat{H}. To see this, differentiate the norm: \frac{d}{dt} \langle \psi | \psi \rangle = \frac{1}{i\hbar} \left( \langle \psi | \hat{H} | \psi \rangle - \langle \psi | \hat{H} | \psi \rangle^* \right) = 0, confirming probability conservation. While primarily formulated for pure states |\psi\rangle, the equation extends to mixed states via the density operator \hat{\rho} = \sum_i p_i |\psi_i\rangle \langle \psi_i |, evolving according to the analogous equation i \hbar \frac{\partial \hat{\rho}}{\partial t} = [\hat{H}, \hat{\rho}], though the focus here remains on pure-state dynamics as originally postulated.

Ehrenfest Theorem

The establishes a connection between and by showing that the of expectation values of position and momentum operators follows equations analogous to . Formulated by in 1927, it demonstrates how quantum systems can exhibit classical-like behavior in their average properties, particularly when wave packets remain localized. The theorem is derived from the time-dependent in the , where s evolve in time. For a general time-independent Hermitian A (with possible explicit time dependence \partial A / \partial t), the time of its is given by \frac{d}{dt} \langle A \rangle = \left\langle \frac{\partial A}{\partial t} \right\rangle + \frac{i}{\hbar} \langle [H, A] \rangle, where H is the , [H, A] = HA - AH is the , and \langle \cdot \rangle denotes the with respect to the . This expression arises from the unitary of the state and the definition of values, providing an exact quantum analog to the classical formulation for the rate of change of observables. Consider a standard non-relativistic Hamiltonian for a single particle, H = \frac{p^2}{2m} + V(x), where p is the momentum operator, m is the mass, and V(x) is the potential. Applying the general formula yields specific equations for position x and momentum p. The time derivative of the position expectation value is \frac{d}{dt} \langle x \rangle = \frac{\langle p \rangle}{m}, using the canonical commutation relation [x, p] = i\hbar. For momentum, \frac{d}{dt} \langle p \rangle = -\left\langle \frac{dV}{dx} \right\rangle, which follows from the commutator [V(x), p] = -i\hbar \, dV/dx. These relations mirror the classical definitions \dot{x} = p/m and \dot{p} = -dV/dx, showing that the "center of mass" of the quantum wave function accelerates according to Newton's second law. While the holds exactly for expectation values, it does not describe the evolution of individual quantum trajectories, as inherently lacks well-defined paths due to the . The classical correspondence is approximate and breaks down when the wave packet spreads significantly, such as over times scales where quantum becomes comparable to the system's size, limiting its validity to coherent or localized states. This bridging role highlights the theorem's importance in understanding the of classical regimes from quantum dynamics.

Relation to Classical Mechanics

Correspondence Principle

The correspondence principle, first articulated by Niels Bohr in 1913, asserts that for systems with large quantum numbers, the frequencies of quantum transitions coincide with the classical frequencies of the corresponding periodic motions. Specifically, in the limit of high excitation, the quantum radiation frequencies \nu_{n' \to n''} for transitions between nearby stationary states asymptotically approach the classical Fourier components \omega_\tau = \tau \omega, where \tau represents integer multiples of the fundamental frequency \omega. This matching ensures that quantum predictions align with classical electrodynamics for observable phenomena in the macroscopic regime, establishing a rational continuity between the old quantum theory and classical physics without introducing arbitrary postulates. A formal of emerges in the as \hbar \to 0, where the quantum [x, p] = i\hbar reduces to the classical \{x, p\} = 1, thereby mapping the non-commutative algebra of quantum operators onto the commutative structure of classical phase-space functions. An illustrative case is the , where the quantized energy levels E_n = -13.6 \, \text{eV}/n^2 for large principal quantum numbers n yield transition energies \Delta E \approx (dE/dn) \Delta n that correspond to the classical orbital frequency of Keplerian motion, with the approximation improving as n increases. The exemplifies this correspondence by showing that the expectation values of dynamical variables evolve according to classical equations in the appropriate limit. Through such mechanisms, the principle upholds the consistency of quantum dynamics with , validating as a fundamental extension rather than a contradictory replacement.

Semiclassical Dynamics

Semiclassical dynamics encompasses approximation techniques that link quantum wave functions to classical trajectories, particularly in the limit where the de Broglie wavelength is small compared to the scale of potential variations. These methods rely on the classical action S, obtained from the Hamilton-Jacobi equation, to construct approximate solutions to the Schrödinger equation. By expanding the wave function in powers of \hbar, semiclassical approximations capture quantum effects like interference and tunneling while retaining classical structure. The Wentzel-Kramers-Brillouin ( is a cornerstone of semiclassical dynamics, providing an asymptotic solution for the one-dimensional in slowly varying potentials. For the time-independent case, the -\frac{\hbar^2}{2m} \frac{d^2\psi}{dx^2} + V(x)\psi = E\psi is solved by assuming \psi(x) \approx A(x) \exp\left( \frac{i}{\hbar} S(x) \right), where S(x) is the classical . Substituting this and expanding in powers of \hbar yields the (S'(x))^2 = p^2(x) = 2m(E - V(x)) at leading order, with S(x) = \int^x p(x') dx', and the transport equation A'(x)/A(x) = -\frac{1}{2} p'(x)/p(x) at next order, giving A(x) \propto 1/\sqrt{p(x)}. Thus, the approximate wave function is \psi(x) \approx \frac{C}{\sqrt{p(x)}} \exp\left( \frac{i}{\hbar} \int^x p(x') dx' \right) in classically allowed regions where p(x) > 0. This form ensures unit current and matches the classical probability density $1/|v(x)| \propto 1/|p(x)|. The approximation was independently developed by Wentzel, Kramers, and Brillouin in 1926. Near classical turning points where p(x) = 0, the WKB approximation breaks down due to rapid variations. A uniform approximation connects the oscillatory solutions on either side using Airy functions: in the vicinity of a turning point at x_0, define the variable \xi = \left( \frac{2m}{\hbar^2 |V'(x_0)|} \right)^{1/3} (x - x_0), and the wave function becomes \psi(x) \approx \frac{C}{\sqrt{|p(x)|}} \mathrm{Ai}(\xi), where Ai is the Airy function. This matches the exponential decay in the forbidden region \psi(x) \approx \frac{C}{\sqrt{|p(x)|}} \exp\left( -\frac{1}{\hbar} \int_{x_0}^x |p(x')| dx' \right) for x > x_0 if V(x) > E. For bound states, the quantization condition arises from matching phases across turning points: \int_{x_1}^{x_2} p(x) dx = \left( n + \frac{1}{2} \right) \pi \hbar, n = 0,1,2,\dots, accurately predicting energy levels for potentials like the harmonic oscillator or hydrogen atom in the large-n limit. For time-dependent problems, the extends to the time-dependent i\hbar \partial_t \psi = \hat{H}(t) \psi by assuming \psi(x,t) \approx A(x,t) \exp\left( \frac{i}{\hbar} S(x,t) \right). The leading-order Hamilton-Jacobi equation becomes \partial_t S + \frac{1}{2m} ( \partial_x S )^2 + V(x,t) = 0, solved along classical trajectories, with the amplitude satisfying a \partial_t A^2 + \partial_x (A^2 v) = 0, where v = \partial_x S / m. This yields wave packets that propagate along classical paths, useful for short-time dynamics in adiabatic or slowly varying fields. Turning points in time-dependent cases are handled analogously with local Airy matching, though caustics require additional uniform treatments. In multidimensional systems, the semiclassical time-evolution K(\mathbf{x},t; \mathbf{x}_0,0) connects initial and final wave functions via \psi(\mathbf{x},t) = \int K(\mathbf{x},t; \mathbf{x}_0,0) \psi(\mathbf{x}_0,0) d\mathbf{x}_0. The Van Vleck approximation gives K(\mathbf{x},t; \mathbf{x}_0,0) \approx \left( \frac{1}{2\pi i \hbar} \right)^{n/2} \sqrt{ \left| \det \left( -\frac{\partial^2 S}{\partial \mathbf{x} \partial \mathbf{x}_0} \right) \right| } \exp\left( \frac{i}{\hbar} S(\mathbf{x},t; \mathbf{x}_0,0) - \frac{i\pi}{2} \nu \right), where n is the , S is the classical from the initial to final point, the determinant is the Van Vleck prefactor ensuring normalization, and \nu is the Maslov index counting caustics encountered. This form generalizes the one-dimensional WKB and is exact for quadratic Hamiltonians like the or . The Van Vleck determinant vanishes at caustics, signaling the need for uniform approximations like Airy or integrals. Applications of these methods include calculating tunneling probabilities through barriers, where the is approximately T \approx \exp\left( -\frac{2}{\hbar} \int_{x_1}^{x_2} |p(x)| dx \right), with turning points x_1, x_2 defined by E = V(x); this semiclassical estimate agrees well with exact results for parabolic or Eckart barriers, capturing the exponential suppression central to and processes. For spectra, the WKB quantization condition yields energy levels that approach exact values as \hbar \to 0 or for high-lying states, as demonstrated in anharmonic oscillators where deviations are O(\hbar^2). In the 1960s, Gutzwiller extended these ideas to incorporate periodic orbits in the trace formula for the , d(E) \approx \bar{d}(E) + \sum_{\mathrm{po}} A_{\mathrm{po}} \cos(S_{\mathrm{po}}/ \hbar - \mu_{\mathrm{po}} \pi /2), where the sum is over primitive orbits, providing a semiclassical basis for spectral statistics in integrable systems.

Advanced Developments

Open Quantum Systems

Open quantum systems refer to quantum systems that interact with an external , or , leading to non-unitary and the of irreversible processes. These interactions cause the system to exchange energy, information, or particles with the , resulting in phenomena such as and decoherence that cannot be captured by the unitary dynamics of isolated s. The theoretical framework employs the reduced density operator \rho for the , obtained by tracing out the from the total . The dynamics of open quantum systems under Markovian conditions—where the system's evolution lacks memory of past states—are governed by the Gorini-Kossakowski-Sudarshan-Lindblad (GKSL) , formulated independently in 1976. This is expressed as \frac{d\rho}{dt} = -\frac{i}{\hbar} [H, \rho] + \sum_k \left( L_k \rho L_k^\dagger - \frac{1}{2} \{ L_k^\dagger L_k, \rho \} \right), where H is the effective of the system, and the L_k are Lindblad operators encoding the dissipative mechanisms. The GKSL form guarantees that the evolution is completely positive and trace-preserving, preserving the physical interpretability of \rho as a over states. To derive the GKSL equation from a microscopic system-bath Hamiltonian, the Born-Markov approximation is invoked, assuming weak system-bath coupling (such that the total state factorizes approximately as \rho_{SB} \approx \rho_S \otimes \rho_B) and rapid relaxation of the bath (enabling a Markovian, time-local description without correlations over long times). These assumptions hold when the bath correlation time is much shorter than the system's evolution timescale, common in scenarios like dilute gases or optical cavities interacting with vacuum fluctuations. A prominent effect in open quantum systems is decoherence, the progressive loss of quantum superpositions due to entanglement with the , which suppresses the off-diagonal elements of \rho in the basis of pointer states stable under . This process explains the transition from quantum interference to classical probabilities without invoking collapse. Illustrative models include the spin-boson model, which captures a two-level system linearly coupled to a bath, elucidating dissipative dynamics like quantum tunneling under Ohmic or sub-Ohmic . In , master equations in Lindblad form describe phenomena such as cavity damping or atomic , where the bath is the modes, leading to photon loss rates proportional to the coupling strength. Recent advances as of 2025 include developments in many-body open quantum systems, where dissipative and coherent dynamics interplay in platforms like quantum simulators, enabling studies of nonequilibrium phases. Additionally, robust control techniques have been proposed to mitigate imperfections and stabilize desired nonequilibrium steady states in driven-dissipative systems.

Quantum Chaos

Quantum chaos refers to the quantum mechanical analogs of classical chaotic dynamics, where quantum systems whose classical counterparts exhibit sensitivity to initial conditions display distinct statistical behaviors in their spectra and wavefunctions, despite the absence of true exponential instability due to unitary time evolution. Unlike classical chaos, which is characterized by positive Lyapunov exponents and mixing, quantum chaos manifests through universal patterns predicted by random matrix theory (RMT), such as level repulsion in energy spectra, where the probability of small spacings between adjacent eigenvalues vanishes linearly, reflecting the orthogonality of eigenstates. This phenomenon was conjectured by Bohigas, Giannoni, and Schmit in 1984 to hold for time-independent quantum systems with chaotic classical limits, linking their spectral statistics to those of Gaussian ensembles in RMT. Another key signature is the scarring of quantum wavefunctions, where eigenstates concentrate enhanced probability density along unstable periodic orbits of the classical system, rather than spreading uniformly. This effect, first identified by Heller in 1984, arises from the interference of semiclassical contributions and provides a direct quantum imprint of classical instability, contrasting with the delocalized, random-wave-like nature of eigenstates in fully ergodic regimes. Berry's conjecture from the 1980s further delineates this distinction: in integrable , high-lying eigenstates localize on classical invariant tori following Einstein-Brillouin-Keller quantization, whereas in chaotic cases, they resemble superpositions of random plane waves, leading to ergodic filling of . These signatures can be probed using semiclassical methods like the trace formula, which relates spectral densities to classical periodic orbits. Paradigmatic models for studying include the quantized baker's map and the quantum kicked rotor. The baker's map, a two-dimensional area-preserving transformation that stretches and folds the unit square, was quantized by Balazs and Voros in 1989 using a Weyl representation, revealing RMT-like spectral statistics and wavefunction scarring for finite-dimensional Hilbert spaces. Similarly, the quantum kicked rotor, introduced by Casati, Chirikov, Ford, and Izrailev in 1979 as a periodically driven rotor , exhibits dynamical localization—suppression of classical diffusion due to quantum interference—while displaying chaotic signatures in its Floquet eigenstates for strong kicking strengths. These models highlight the suppression of classical chaos in the quantum domain through spectra and recurrent dynamics. To quantify to perturbations, measures such as quantal and the Loschmidt are employed, capturing the overlap between states evolved under unperturbed and slightly perturbed Hamiltonians. decays Gaussian-like at short times due to local differences but transitions to in regimes, reflecting akin to classical Lyapunov , as analyzed by Jalabert and Pastawski in 2001. The Loschmidt , defined as the squared of this overlap, revives weakly in systems due to pseudorandom , providing a -based diagnostic for quantum without requiring full . These tools underscore the reversible yet statistically irreversible nature of quantum dynamics in settings. Recent experimental progress as of 2025 includes a 2024 demonstration confirming quantum scarring patterns in a controlled , providing direct evidence of these phenomena. Furthermore, in 2025, Google Quantum AI achieved a 13,000-fold in simulating complex physics systems using a 65-qubit , advancing the study of quantum dynamics beyond classical supercomputers.

References

  1. [1]
    [PDF] Quantum Dynamics - Arizona Math
    Aug 12, 1992 · The invention of quantum mechanics was the greatest revolution of modern science. It explained the properties of everyday matter, ...
  2. [2]
    [PDF] Quantum Mechanics: Time Dependence
    May 20, 2025 · Quantum dynamics is central to the understanding how quantum systems evolve in time and provides a direct link between a system's energy ...
  3. [3]
    [PDF] Lecture 4: Quantum Dynamics and Schroedinger Equation
    Nov 18, 2020 · Lecture 4: Quantum Dynamics and Schroedinger Equation. November 18, 2020. 1 Time Evolution of Quantum System. In the previous section we ...
  4. [4]
    Quantum Dynamics - an overview | ScienceDirect Topics
    Quantum dynamics (QD) refers to the study of systems using trajectory-guided basis sets of Gaussian wave packets or coherent states to solve the ...
  5. [5]
    [PDF] Time evolution operator
    We define the time evolution operator U(t, t0):. |α, t0; ti = U(t, t0)|α, t0i, which must satisfy physically relevant conditions.
  6. [6]
    [PDF] Unitary time evolution
    The restriction of time evolution to unitary operators means that certain kinds of evolution are impossible. One impossible task is quantum cloning. Suppose we ...
  7. [7]
    [PDF] Time Reversal - University of California, Berkeley
    Time reversal is a space-time symmetry, implemented by antiunitary transformations. In classical mechanics, it's x(-t). In quantum mechanics, it's ψr(x,t) = ψ∗ ...
  8. [8]
    [PDF] Time Evolution in Quantum Mechanics
    Time Evolution of the Density Operator. Equation (14), the Schrödinger equation, gives the time evolution of the state vector in quantum mechanics. As we ...
  9. [9]
    An Elementary Derivation of the Harmonic Oscillator Propagator
    Feb 14, 2004 · The harmonic oscillator propagator is found straightforwardly from the free particle propagator, within the imaginary-time Feynman path ...Missing: mechanics | Show results with:mechanics
  10. [10]
  11. [11]
    Heisenberg Equation of Motion
    Equation (244) shows how the dynamical variables of the system evolve in the Heisenberg picture. It is denoted the Heisenberg equation of motion.
  12. [12]
    Bemerkung über die angenäherte Gültigkeit der klassischen ...
    Download PDF ... Cite this article. Ehrenfest, P. Bemerkung über die angenäherte Gültigkeit der klassischen Mechanik innerhalb der Quantenmechanik.
  13. [13]
    [0907.1877] On the Ehrenfest theorem of quantum mechanics - arXiv
    Jul 10, 2009 · We give a mathematically rigorous derivation of Ehrenfest's equations for the evolution of position and momentum expectation values.
  14. [14]
    Bohr’s Correspondence Principle (Stanford Encyclopedia of Philosophy)
    ### Summary of Bohr's Correspondence Principle (1923)
  15. [15]
    Volume 3: The Correspondence Principle (1918–1923)
    Niels Bohr Collected Works, Volume 3: The Correspondence Principle (1918–1923), Edited by Pages iii-xii, 3-705 (1976), Download full volume.
  16. [16]
    15.2: Poisson bracket Representation of Hamiltonian Mechanics
    Jun 28, 2021 · In classical mechanics there is a formal correspondence between the Poisson bracket and the commutator. This can be shown by deriving the ...Missing: original paper
  17. [17]
    [PDF] Physics 228, Lecture 11 Monday, February 28, 2005 Bohr Model
    Feb 28, 2005 · Bohr's correspondence principle says that for large excitation levels, the quantum mechanical results should agree with classical expectations.
  18. [18]
  19. [19]
    The Correspondence Principle in the Statistical Interpretation of ...
    The Correspondence Principle in the Statistical Interpretation of Quantum Mechanics. J. H. Van VleckAuthors Info & Affiliations. February 15, 1928.Missing: propagator | Show results with:propagator
  20. [20]
  21. [21]
  22. [22]
    Environment-Independent Decoherence Rate in Classically Chaotic ...
    Mar 19, 2001 · This is done by analytical evaluation of the Loschmidt echo (i.e., the revival of a localized density excitation upon reversal of its time ...