Fact-checked by Grok 2 weeks ago

Canonical quantization

Canonical quantization is a foundational method in quantum physics for deriving quantum theories from classical Hamiltonian mechanics by promoting classical phase space variables—such as position q and momentum p—to non-commuting operators \hat{q} and \hat{p} that satisfy the canonical commutation relation [\hat{q}, \hat{p}] = i \hbar, where \hbar is the reduced Planck's constant. This procedure replaces classical Poisson brackets \{f, g\} with quantum commutators [ \hat{f}, \hat{g} ] / i \hbar, ensuring that the quantum dynamics reduce to classical equations in the appropriate limit via the correspondence principle. Introduced by Paul Dirac in 1926, it provides a systematic way to construct the Hamiltonian operator and solve the Schrödinger equation for systems ranging from single particles to fields. The historical development of canonical quantization traces back to the early formulation of , where Dirac's 1926 paper "On the Theory of Quantum Mechanics" formalized the replacement of Poisson brackets with commutators to unify and wave mechanics. This approach built on Heisenberg's 1925 formulation and Schrödinger's 1926 , offering a that treats dynamical variables abstractly without specifying a representation. Dirac's method was independently discovered around the same time by and , but Dirac's version emphasized the transformation rules for general observables, laying the groundwork for modern . By 1927, Dirac extended these ideas to radiation theory, introducing early concepts of for bosonic fields. In practice, canonical quantization proceeds in steps: first, identify the classical Lagrangian or Hamiltonian and derive canonical conjugate pairs; second, promote these to operators on a Hilbert space with the required commutation relations; third, form the quantum Hamiltonian and impose boundary conditions or normal ordering to handle ambiguities; and finally, compute expectation values or solve for energy eigenstates. For relativistic systems, it is applied in quantum field theory (QFT), where fields \phi(\mathbf{x}, t) and their conjugate momenta \pi(\mathbf{x}, t) become operator-valued distributions satisfying equal-time commutators like [\phi(\mathbf{x}, t), \pi(\mathbf{y}, t)] = i \hbar \delta^3(\mathbf{x} - \mathbf{y}), leading to particle creation and annihilation operators. This yields second quantization, essential for describing multi-particle systems and interactions in quantum electrodynamics and the Standard Model. Challenges include operator ordering ambiguities and ultraviolet divergences, often addressed via renormalization. Canonical quantization remains a of , underpinning applications from atomic spectra to condensed matter phenomena like the , though alternative approaches like path integrals offer complementary insights for certain calculations. Its elegance lies in preserving the structure of while incorporating quantum non-commutativity, ensuring unitarity and probabilistic interpretations.

Introduction

Definition and Principles

Canonical quantization is a systematic procedure for transitioning from a classical Hamiltonian formulation of a physical system to its quantum mechanical counterpart, primarily by promoting classical observables to operators and replacing the structure of classical dynamics with the algebra of quantum operators. This method preserves the formal structure of the classical theory, including symmetries and , while incorporating such as and superposition. Introduced by , the approach ensures that the reduces to the classical one in the appropriate limit, aligning with . In , the is parameterized by such as position q and p, where the dynamics are governed by Hamilton's equations, expressible through brackets, for example, \{q, p\} = 1. Canonical quantization motivates a shift to a quantum description in , where observables are represented by operators acting on state vectors. The core correspondence rule posits that the of two quantum operators \hat{A} and \hat{B} corresponds to the bracket of their classical counterparts A and B, specifically [\hat{A}, \hat{B}] = i \hbar \{A, B\}, with \hbar being the reduced Planck's constant. This replacement ensures that the quantum Heisenberg equations of motion mimic the classical Hamilton's equations. A fundamental instance of this principle is the canonical commutation relation for position and momentum operators, [\hat{q}, \hat{p}] = i \hbar, which encapsulates the non-commutativity inherent to quantum measurements and leads to the Heisenberg uncertainty principle. This operator-based formulation underpins the Heisenberg picture of quantum mechanics, where states evolve unitarily and operators are time-dependent, while the equivalent Schrödinger picture employs wave functions with time-independent operators, both emerging naturally from the quantization procedure. The method's success in reproducing known quantum results for simple systems, like the harmonic oscillator, validates its foundational role in non-relativistic quantum theory.

Canonical Commutation Relations

In canonical quantization, the core algebraic structure is provided by the canonical commutation relations, which replace the classical Poisson brackets with non-commuting operator relations. Paul Dirac proposed that for a system with Cartesian position coordinates q_i and conjugate momenta p_j, the corresponding quantum operators satisfy [q_i, p_j] = i \hbar \delta_{ij}, \quad [q_i, q_j] = 0, \quad [p_i, p_j] = 0, where \hbar = h / 2\pi is the reduced Planck's constant and \delta_{ij} is the Kronecker delta. These relations, originally formulated with h and later adjusted to \hbar in modern notation, can be verified in the Schrödinger representation using wave functions, where the momentum operator is p_j = \frac{h}{i} \frac{\partial}{\partial q_j}, such that the commutator applied to any function X yields [q_i, p_j] X = i h \delta_{ij} X. This verifies the algebraic postulate by the correspondence \{A, B\}_{\text{classical}} \to \frac{1}{i \hbar} [ \hat{A}, \hat{B} ]. For systems in q_r and velocities \dot{q}_r, the procedure begins with the classical L(q_r, \dot{q}_r), from which the momenta are defined via the p_r = \frac{\partial L}{\partial \dot{q}_r} and the as H = \sum_r p_r \dot{q}_r - L(q_r, p_r). The commutation relations are then imposed on the operator versions \hat{q}_r and \hat{p}_s in the same form [ \hat{q}_r, \hat{p}_s ] = i \hbar \delta_{rs}, ensuring the quantum \hat{H}( \hat{q}_r, \hat{p}_s ) generates the correct dynamics via the Heisenberg equations \frac{d \hat{q}_r}{dt} = \frac{i}{\hbar} [ \hat{H}, \hat{q}_r ] = \frac{\partial H}{\partial p_r} and similarly for \hat{p}_r. These relations underpin the Heisenberg uncertainty principle, a direct consequence of the non-commutativity. Specifically, for any state, the standard deviations satisfy \Delta q \, \Delta p \geq \frac{\hbar}{2}, with equality for Gaussian wave packets. This inequality was first derived rigorously by Earle Hesse Kennard from the commutation relations [q, p] = i \hbar, showing that the variance product \sigma_q^2 \sigma_p^2 \geq \left( \frac{\hbar}{2} \right)^2 follows from the Cauchy-Schwarz inequality applied to expectation values. The commutation relations admit multiple representations that realize them equivalently. In the position basis, states are described by wave functions \psi(q) = \langle q | \psi \rangle, where \hat{q} \psi(q) = q \psi(q) and \hat{p} \psi(q) = -i \hbar \frac{d}{dq} \psi(q), satisfying [ \hat{q}, \hat{p} ] \psi(q) = i \hbar \psi(q). In the momentum basis, wave functions are \phi(p) = \langle p | \psi \rangle, with \hat{p} \phi(p) = p \phi(p) and \hat{q} \phi(p) = i \hbar \frac{d}{dp} \phi(p), again fulfilling the commutator. These bases are connected by Fourier transform, \phi(p) = \frac{1}{\sqrt{2\pi \hbar}} \int_{-\infty}^{\infty} e^{-i p q / \hbar} \psi(q) \, dq. A illustrative example is the , with classical H = \frac{p^2}{2m} + \frac{1}{2} m \omega^2 q^2. Direct quantization replaces q \to \hat{q} and p \to \hat{p} under the commutation [ \hat{q}, \hat{p} ] = i \hbar, yielding the \hat{H} = \frac{\hat{p}^2}{2m} + \frac{1}{2} m \omega^2 \hat{q}^2. The eigenvalues are \hat{H} | n \rangle = \left( n + \frac{1}{2} \right) \hbar \omega | n \rangle for n = 0, 1, 2, \dots, obtained by ladder operators a = \sqrt{\frac{m \omega}{2 \hbar}} \left( \hat{q} + \frac{i}{m \omega} \hat{p} \right) and a^\dagger, satisfying [a, a^\dagger] = 1, with \hat{H} = \hbar \omega \left( a^\dagger a + \frac{1}{2} \right).

Historical Development

Early Formulations

The , developed in the 1910s and 1920s, introduced ad hoc quantization rules to reconcile with emerging quantum phenomena, particularly in atomic spectra. Bohr's 1913 model of the posited discrete orbits where the is quantized as m v r = n \frac{h}{2\pi}, with n an integer and h Planck's constant, to explain stable orbits and emission. This was extended by in 1916 through the Bohr-Sommerfeld quantization condition, which generalized quantization to multi-dimensional systems by requiring that the action integral over a closed classical path satisfy \oint p \, dq = n h, where p is and q the conjugate coordinate, allowing for elliptical orbits and in spectra. These rules successfully predicted energy levels for hydrogen-like atoms but relied on selective application to classical trajectories, lacking a systematic framework for general mechanical systems. In 1925, formulated as a radical departure from classical descriptions, replacing continuous trajectories with discrete arrays representing observables like and , whose elements corresponded to transition amplitudes between quantum states. This approach quantized classical frequencies ad hoc by promoting them to non-commuting operators satisfying relations derived from spectral data, such as the commutation rules for matrix products, enabling calculations of energy levels without reference to orbits. and soon formalized this in their 1925 paper, introducing symmetric ordering for products of non-commuting variables, exemplified by replacing the classical q p with the quantum \frac{1}{2} (q p + p q) to ensure Hermiticity and consistency in the . Erwin Schrödinger introduced wave mechanics in 1926, proposing that quantum systems are described by wave functions satisfying a analogous to classical wave , with the time-independent -\frac{\hbar^2}{2m} \nabla^2 \psi + V \psi = E \psi yielding quantized energies as eigenvalues. In a subsequent paper that year, Schrödinger proved the mathematical equivalence of wave mechanics to Heisenberg's by showing that solutions to the wave equation reproduce the matrix eigenvalues and transformation rules, unifying the two formulations under a common structure. Despite this equivalence, early quantization rules exhibited limitations, notably ambiguities in ordering non-commuting operators when mapping classical Hamiltonians to quantum ones, as different orderings could yield distinct spectra, highlighting the need for unambiguous prescriptions. These initial efforts, while groundbreaking, suffered from inconsistencies in handling operator products, such as the ambiguity in symmetrizing terms like q p, which Born and Jordan addressed partially but not generally, foreshadowing more systematic approaches. The ad hoc nature of applying quantization to specific systems, without preserving classical Poisson brackets universally, restricted applicability to complex interactions.

Dirac's Contributions

Paul Dirac played a central role in transforming the ad hoc ideas of early into a systematic procedure known as canonical quantization. In his 1925 paper "The Fundamental Equations of ," Dirac proposed replacing classical variables q with multiplication s by q and classical momentum variables p with differential s -i\hbar \frac{d}{dq} in the representation, thereby bridging classical with a quantum . This substitution allowed the to emerge naturally from the classical Hamilton-Jacobi equation, providing a concrete method to quantize single-particle systems. Building on this, Dirac introduced a general in the same paper, mapping classical s to quantum s via the [A, B] = i\hbar \{A, B\}_{\text{PB}}, where \{A, B\}_{\text{PB}} denotes the and [A, B] = AB - BA the . This analogy ensured that the algebraic structure of was preserved in the quantum domain, with fundamental relations like [q, p] = i\hbar directly following from \{q, p\}_{\text{PB}} = 1. Dirac's approach synthesized elements of , offering a Hamiltonian-based quantization scheme independent of specific representations. Dirac's 1930 monograph provided a rigorous formalization of canonical quantization, extending the method to q_r and conjugate momenta p_r satisfying [q_r, p_s] = i\hbar \delta_{rs}, [q_r, q_s] = [p_r, p_s] = 0. In this framework, the Hamiltonian operator H(q_r, p_r) governs through i\hbar \frac{\partial \psi}{\partial t} = H \psi, applicable to systems in arbitrary coordinates. For systems with constraints, Dirac outlined procedures in by incorporating them into the structure before quantization, ensuring consistency via modified commutators that respect the constraints on the . This handling of constraints, such as those from or boundaries, laid groundwork for quantizing complex mechanical systems. Dirac's quantization rules profoundly influenced later developments, including Richard Feynman's , which originated from Dirac's 1933 suggestion to express quantum amplitudes via classical action extrema in .

First Quantization Methods

Single-Particle Mechanics

In canonical quantization of a single non-relativistic particle, the classical for a , H = \frac{p^2}{2m}, is promoted to the \hat{H} = \frac{\hat{p}^2}{2m}, where \hat{p} = -i[\hbar](/page/H-bar) \frac{d}{dx} in the position representation. The time-independent \hat{H} \psi = E \psi yields plane wave solutions \psi(x) = e^{i p x / [\hbar](/page/H-bar)}, which are momentum eigenstates with continuous spectrum E = \frac{p^2}{2m}. These solutions satisfy the [\hat{x}, \hat{p}] = i[\hbar](/page/H-bar) and describe delocalized particle states without boundaries. For particles in central potentials V(r), the Hamiltonian becomes \hat{H} = \frac{\hat{p}^2}{2m} + V(r), and the orbital angular momentum operator \hat{\mathbf{L}} = \mathbf{r} \times \hat{\mathbf{p}} is introduced via canonical quantization. Its components satisfy the commutation relations [\hat{L}_x, \hat{L}_y] = i\hbar \hat{L}_z and cyclic permutations, ensuring \hat{\mathbf{L}} generates rotations and commutes with \hat{H}, conserving angular momentum. In spherical coordinates, the wave function separates as \psi(\mathbf{r}) = R(r) Y_{l m}(\theta, \phi), where Y_{l m} are spherical harmonics as eigenfunctions of \hat{L}^2 and \hat{L}_z, with eigenvalues l(l+1)\hbar^2 and m\hbar. A key example is the , where V(r) = -\frac{e^2}{r} (in ) and the reduced mass \mu approximates the . The radial arises from separating variables in \hat{H} \psi = E \psi: -\frac{\hbar^2}{2\mu} \frac{1}{r^2} \frac{d}{dr} \left( r^2 \frac{dR}{dr} \right) + \left[ V(r) + \frac{\hbar^2 l(l+1)}{2\mu r^2} \right] R = E R, yielding bound states for E < 0 labeled by n = 1, 2, \dots, with energies E_n = -\frac{\mu e^4}{2 \hbar^2 n^2} independent of l and m, reflecting degeneracy. The solutions involve associated for R(r), ensuring normalizability. To ensure operators correspond to observables, they must be Hermitian, requiring symmetrization of non-commuting products from classical expressions. For instance, the classical x p becomes the quantum operator \frac{1}{2} (\hat{x} \hat{p} + \hat{p} \hat{x}), which satisfies \left( \frac{1}{2} (\hat{x} \hat{p} + \hat{p} \hat{x}) \right)^\dagger = \frac{1}{2} (\hat{x} \hat{p} + \hat{p} \hat{x}) despite [\hat{x}, \hat{p}] = i\hbar. This rule applies generally to maintain real eigenvalues and probabilities. An illustrative calculation is the one-dimensional with \hat{H} = \frac{\hat{p}^2}{2m} + \frac{1}{2} m \omega^2 \hat{q}^2. Introducing ladder operators \hat{a} = \sqrt{\frac{m \omega}{2 \hbar}} \left( \hat{q} + \frac{i \hat{p}}{m \omega} \right), \quad \hat{a}^\dagger = \sqrt{\frac{m \omega}{2 \hbar}} \left( \hat{q} - \frac{i \hat{p}}{m \omega} \right), satisfying [\hat{a}, \hat{a}^\dagger] = 1, the Hamiltonian simplifies to \hat{H} = \hbar \omega \left( \hat{a}^\dagger \hat{a} + \frac{1}{2} \right). The energy eigenvalues are then E_n = \hbar \omega \left( n + \frac{1}{2} \right) for n = 0, 1, 2, \dots, with \hat{a}^\dagger raising and \hat{a} lowering states by \hbar \omega. The ground state satisfies \hat{a} |0\rangle = 0, yielding zero-point energy \frac{1}{2} \hbar \omega.

Many-Particle Systems

In first quantization, systems of fixed particle number N are described by wave functions \Psi(\mathbf{x}_1, \mathbf{x}_2, \dots, \mathbf{x}_N) in the $3N-dimensional configuration space, where \mathbf{x}_j = (\mathbf{r}_j, s_j) includes position and spin coordinates. For distinguishable particles, the Hilbert space is the tensor product of single-particle spaces. However, for identical particles, the wave function must respect exchange symmetry: symmetric under particle permutations for bosons (integer spin) and antisymmetric for fermions (half-integer spin). This symmetrization postulate enforces quantum statistics and, for fermions, the Pauli exclusion principle, which limits occupation numbers to 0 or 1 per state. The for non-relativistic identical particles is \hat{H} = \sum_{j=1}^N \hat{h}(\mathbf{x}_j) + \frac{1}{2} \sum_{j \neq k} \hat{V}(\mathbf{x}_j, \mathbf{x}_k), where \hat{h} is the single-particle and \hat{V} the two-body interaction, both symmetric under exchanges. The allowed states are projections onto the symmetric (bosons) or antisymmetric (fermions) subspaces. For bosons, the wave function is a permanent: \Psi(\mathbf{x}_1, \dots, \mathbf{x}_N) = \frac{1}{\sqrt{N! \prod n_i!}} \sum_P P \prod_{j=1}^N \phi_{\lambda_j}(\mathbf{x}_j), where P permutes particles and n_i are occupation numbers. For fermions, it is a Slater determinant: \Psi(\mathbf{x}_1, \dots, \mathbf{x}_N) = \frac{1}{\sqrt{N!}} \det \left[ \phi_{\lambda_j}(\mathbf{x}_i) \right], ensuring antisymmetry. These symmetries underpin physical phenomena, such as the stability of matter via the Pauli principle, which prevents electrons from collapsing into the lowest orbital in multi-electron atoms. A representative example is the of the (N=2 electrons, fermions). The spatial is symmetric, \phi_{1s}(\mathbf{r}_1) \phi_{1s}(\mathbf{r}_2), multiplied by the antisymmetric spin singlet \frac{1}{\sqrt{2}} [\alpha(s_1) \beta(s_2) - \beta(s_1) \alpha(s_2)]. Equivalently, using spin-orbitals, it is the \Psi(1,2) = \frac{1}{\sqrt{2}} \left[ u_{1s\uparrow}(1) u_{1s\downarrow}(2) - u_{1s\uparrow}(2) u_{1s\downarrow}(1) \right], where u_{nlm s}(\mathbf{x}) = \phi_{nlm}(\mathbf{r}) \chi_s(s), with hydrogen-like 1s orbitals \phi_{1s}. This form incorporates exchange effects and approximate correlation.

Challenges in Quantization

Bracket Transformations

In canonical quantization, the classical between q and momentum p, defined as \{q, p\}_{\text{PB}} = 1, is mapped to the quantum [q, p]/i\hbar = 1, where \hbar is the reduced Planck's constant. This correspondence, first systematically proposed by Dirac, extends to general observables by replacing the \{f, g\}_{\text{PB}} with the [ \hat{f}, \hat{g} ] / i\hbar, preserving the algebraic structure of in the quantum regime. However, this mapping introduces ambiguities when quantizing products of non-commuting operators, such as qp versus pq, since [q, p] = i\hbar implies qp \neq pq. Different ordering choices, like (creation operators before annihilation) or antinormal ordering, can lead to distinct quantum Hamiltonians, affecting physical predictions unless resolved by a consistent . For instance, the classical Hamiltonian H = qp for a might be quantized as the non-Hermitian qp, but to ensure a real-valued energy spectrum, the symmetric Hermitian form \hat{H} = (qp + pq)/2 is often adopted, symmetrizing the product to mitigate ordering artifacts. Weyl ordering addresses these ambiguities by defining a unique quantization map for functions f(q, p) on via an , associating each classical function to a Hermitian through the Weyl . The Weyl-quantized \hat{f}_W is given by \hat{f}_W = \frac{1}{(2\pi \hbar)^n} \int d^n q' d^n p' \, f\left(q + \frac{q'}{2}, p - \frac{p'}{2}\right) e^{i (p' \hat{q} - q' \hat{p})/\hbar}, where n is the number of dimensions, effectively averaging over all orderings to produce a symmetric, unitary-preserving quantization. This approach, originally formulated by Weyl, avoids operator-ordering issues by embedding the quantization in phase-space integrals, ensuring consistency for polynomials and facilitating the transition to deformation quantization frameworks. The Moyal bracket provides a non-commutative phase-space analog to the Poisson bracket, capturing quantum corrections through a star-product structure. Defined as [[f, g]]_M = \frac{2}{\hbar} f \sin\left( \frac{\hbar}{2} (\overleftarrow{\partial_q} \overrightarrow{\partial_p} - \overleftarrow{\partial_p} \overrightarrow{\partial_q}) \right) g, it expands in powers of \hbar, with the leading term recovering the classical Poisson bracket and higher terms encoding quantum non-commutativity. Introduced by Moyal, this bracket enables a fully quantum description of dynamics on phase space without explicit operator algebra, resolving ambiguities by incorporating the Weyl-ordered star product into the bracket formalism. Dirac's simple replacement rule serves as a foundational precursor, but bracket transformations like the Moyal form offer a more refined, ambiguity-free extension for complex systems.

No-Go Theorems

In canonical quantization, no-go theorems reveal fundamental limitations on constructing a consistent mapping from classical algebras to quantum algebras that preserves the structure for all observables. These results, emerging in the mid-20th century, demonstrated that Dirac's proposed —replacing brackets {f, g} with commutators (1/iħ)[Q(f), Q(g)]—cannot be realized exactly for general classical theories beyond simple cases. Groenewold's theorem, established in , states that there exists no linear mapping Q from the of classical observables (smooth functions on ) to quantum operators that preserves all brackets exactly, except for those up to quadratic in the coordinates and momenta. Specifically, while the mapping succeeds for the fundamental commutator [Q(q), Q(p)] = iħ (corresponding to {q, p} = 1), it fails for higher- terms. For example, the classical observable q²p² can be expressed as {q³, p³}/9 or as {q p², q² p}/3. Under the assumed mapping, the first yields the quantum operator \hat{q}^2 \hat{p}^2 - 2 i \hbar \hat{q} \hat{p} + (2/3) (i \hbar)^2, while the second gives \hat{q}^2 \hat{p}^2 - 2 i \hbar \hat{q} \hat{p} + (1/3) (i \hbar)^2, differing at \hbar^2 and thus contradicting the that Q({f, g}) = (1/i \hbar) [Q(f), Q(g)]. This introduces discrepancies that grow with the degree of the polynomials involved. Building on this, Van Hove's from 1951 extends the obstruction to systems with an excess of classical invariants, showing that no unitary of the commutation relations exists for phase spaces where the number of independent constants of motion exceeds the dimension of the configuration space. In such cases, the attempt to quantize the full fails because the quantum representation cannot accommodate all the required symmetries without violating irreducibility or unitarity. The implications of these theorems are profound: exact canonical quantization is only feasible for systems with finitely many and low-order observables, such as the , where quadratic approximations suffice; for more complex theories, like those with nonlinear interactions or infinite , approximations or alternative schemes are necessary. Historically, these no-go results, arising shortly after Dirac's influential work on quantization rules, underscored the ambiguities in bracket transformations and spurred the development of alternative approaches, including Weyl and Born-Jordan quantization methods that prioritize specific orderings over full preservation.

Quantization Conditions

In canonical quantization, Dirac formulated foundational axioms to ensure a consistent transition from classical to quantum mechanics. The correspondence principle requires that the quantum commutator of operators corresponding to classical observables f and g approximates the classical Poisson bracket in the limit \hbar \to 0, specifically [\hat{f}, \hat{g}] / i\hbar \to \{f, g\}. Irreducibility of the representation demands that the Hilbert space representation of the canonical commutation relations has no nontrivial invariant subspaces under the action of the full set of operators, guaranteeing a unique physical description up to unitary equivalence. Unitarity stipulates that the time evolution operator and transformations induced by symplectomorphisms are unitary on the Hilbert space, preserving probabilities and inner products. Modern axiomatic frameworks extend these ideas to more general quantization maps Q_\hbar: C^\infty(M) \to \Psi^{-\infty}(M) from smooth functions on the to pseudodifferential operators, incorporating stricter conditions for well-defined classical limits. Continuity in the \hbar \to 0 limit requires that \|Q_\hbar(f)\| \to \|f\|_\infty uniformly for bounded functions, ensuring quantum observables converge to classical ones. Preservation of traces mandates that for trace-class operators, \operatorname{Tr}(Q_\hbar(f)) = \int_M f \, d\mu (up to normalization), linking quantum expectation values to classical phase space integrals. Hermiticity preservation ensures that if f is real-valued, then Q_\hbar(f) is , maintaining the reality of observables. These axioms address ambiguities in non-strict quantizations, such as operator ordering, by enforcing algebraic and topological consistency. Berezin's conditions formalize quantization via coherent states on quantizable Kähler manifolds (M, \omega), where a holomorphic L with -i\omega / \hbar admits global sections forming a . The quantization map Q_\hbar(f) is defined as the Toeplitz operator Q_\hbar(f) \psi = P_\hbar (f \psi), with P_\hbar the orthogonal projection onto holomorphic sections of L^\otimes n (where \hbar = 1/n). Key conditions include linearity, positivity (mapping positive functions to positive operators), and the asymptotic property that the contravariant symbol \tilde{f}_\hbar(x) = \langle e_x^{(n)}, Q_\hbar(f) e_x^{(n)} \rangle / \|e_x^{(n)}\|^2 satisfies \tilde{f}_\hbar \to f as \hbar \to 0, or in suitable norms. Additionally, the map preserves the via complex conjugation and satisfies the correspondence [Q_\hbar(f), Q_\hbar(g)] / i\hbar \to \{f, g\}. These ensure a covariant quantization scheme suitable for symmetric spaces. These axiomatic requirements play a crucial role in distinguishing valid quantization schemes, particularly in resolving ordering ambiguities like Weyl (symmetric) versus normal (coherent state) ordering. For instance, Weyl ordering satisfies strict continuity and trace preservation on flat phase spaces, while normal ordering aligns better with Berezin's coherent state conditions on curved manifolds, as it reproduces expectation values in overcomplete bases. No-go theorems motivate these axioms by highlighting impossibilities in exact mappings, but constructive conditions like trace preservation select schemes that approximate classical statistics. A representative example is the trace formula for the partition function: under trace-preserving quantization, \operatorname{Tr}(e^{-\beta \hat{H}}) \sim (2\pi \hbar)^{-n} \int e^{-\beta H(q,p)} \, dq \, dp as \hbar \to 0, ensuring thermodynamic consistency between quantum and classical regimes.

Second Quantization for Fields

Operator Formalism

In the operator formalism of , the transition from discrete particle descriptions to continuous relativistic fields involves promoting the classical field φ(x) to an operator-valued distribution φ̂(x) that acts on a of quantum states, enabling the description of variable particle numbers and field excitations. This approach builds on the procedure by treating fields as fundamental dynamical variables, with the conjugate π̂(x) defined as π̂(x) = ∂ℒ/∂(∂_t φ), where ℒ is the density. The core of the formalism lies in imposing canonical commutation relations at equal times to preserve the Poisson bracket structure of classical field theory in the quantum regime: [\hat{\phi}(t, \mathbf{x}), \hat{\pi}(t, \mathbf{y})] = i \hbar \delta^3(\mathbf{x} - \mathbf{y}), with the remaining equal-time commutators vanishing: [\hat{\phi}(t, \mathbf{x}), \hat{\phi}(t, \mathbf{y})] = 0, \quad [\hat{\pi}(t, \mathbf{x}), \hat{\pi}(t, \mathbf{y})] = 0. These relations are extended to all spacetime points via the field equations of motion, ensuring consistency with relativity. For the Klein-Gordon field, the Hamiltonian operator takes the form \hat{H} = \int d^3\mathbf{x} \left[ \frac{1}{2} \hat{\pi}^2 + \frac{1}{2} (\nabla \hat{\phi})^2 + \frac{1}{2} m^2 \hat{\phi}^2 \right], which generates time evolution and expresses the total energy in terms of field operators. The Hilbert space is constructed as a Fock space, consisting of the vacuum state |0⟩ annihilated by all destruction operators and multi-particle states generated by applying creation operators derived from the mode expansion of the field φ̂(x). This structure accommodates arbitrary numbers of particles and antiparticles, resolving issues with fixed-particle interpretations in relativistic quantum mechanics. To uphold causality, the commutator must satisfy [φ̂(x), φ̂(y)] = 0 for spacelike separations (x - y)^2 < 0, preventing superluminal influences and ensuring locality in the theory.

Scalar Field Quantization

The canonical quantization of a real provides a foundational example of in , treating the field as an infinite collection of harmonic oscillators. This approach resolves issues with single-particle relativistic wave equations, such as negative probabilities and negative energies, by interpreting the field itself as the fundamental entity composed of particles and antiparticles. The procedure begins with the classical theory of a free real satisfying the Klein-Gordon equation, which arises from the Lagrangian density \mathcal{L} = \frac{1}{2} \partial_\mu \phi \partial^\mu \phi - \frac{1}{2} m^2 \phi^2, where \phi(x) is the scalar field, m is the particle mass, and the summation convention is used with the Minkowski metric \eta^{\mu\nu} = \operatorname{diag}(1, -1, -1, -1). This form ensures Lorentz invariance and yields the equation of motion (\square + m^2) \phi = 0 via the Euler-Lagrange equations. To quantize, the field \phi and its conjugate momentum \pi = \partial_0 \phi are promoted to operators satisfying the equal-time commutation relations [\phi(\mathbf{x}, t), \pi(\mathbf{y}, t)] = i \hbar \delta^3(\mathbf{x} - \mathbf{y}), with other commutators vanishing. The Hamiltonian is then H = \int d^3x \left( \frac{1}{2} \pi^2 + \frac{1}{2} (\nabla \phi)^2 + \frac{1}{2} m^2 \phi^2 \right), which resembles that of infinitely many harmonic oscillators. This operator formalism was systematically applied to the scalar field in the seminal work establishing field quantization for relativistic scalar theories. The field operator is expanded in momentum modes as a Fourier integral over plane waves on Minkowski space: \phi(x) = \int \frac{d^3 k}{(2\pi)^3} \frac{1}{\sqrt{2 \omega_{\mathbf{k}}}} \left( a_{\mathbf{k}} e^{-i k \cdot x} + a^\dagger_{\mathbf{k}} e^{i k \cdot x} \right), where k^\mu = (\omega_{\mathbf{k}}, \mathbf{k}) with \omega_{\mathbf{k}} = \sqrt{\mathbf{k}^2 + m^2} on the mass shell, and the a_{\mathbf{k}}, a^\dagger_{\mathbf{k}} are annihilation and creation operators, respectively. Imposing the canonical commutation relations leads to the algebra [a_{\mathbf{k}}, a^\dagger_{\mathbf{k}'}] = (2\pi)^3 \delta^3(\mathbf{k} - \mathbf{k}'), with all other commutators zero. This mode decomposition reveals the field as a superposition of positive- and negative-frequency components, corresponding to particles and antiparticles. The particle interpretation emerges from the construction, where the vacuum state |0\rangle is defined by a_{\mathbf{k}} |0\rangle = 0 for all \mathbf{k}, annihilating no particles. The state a^\dagger_{\mathbf{k}} |0\rangle represents a single particle with \mathbf{k} and \omega_{\mathbf{k}}, while multi-particle states are built by applying multiple creation operators, such as \frac{(a^\dagger_{\mathbf{k}})^n}{\sqrt{n!}} |0\rangle for n identical bosons. The total number operator N = \int \frac{d^3 k}{(2\pi)^3} a^\dagger_{\mathbf{k}} a_{\mathbf{k}} counts particles, and the is \langle 0 | H | 0 \rangle = \frac{1}{2} \int \frac{d^3 k}{(2\pi)^3} \omega_{\mathbf{k}}, which is infinite but regularized in interactions. This framework interprets excitations of the as spin-0 bosons, resolving the one-particle relativistic issues by allowing creation and annihilation processes. A striking observable consequence of the quantized scalar field's vacuum fluctuations is the , where boundary conditions imposed by two parallel conducting plates modify the allowed modes, leading to a finite, negative shift and an attractive force between the plates proportional to -\frac{\pi^2 \hbar c}{240 a^4} per unit area, with plate separation a. Although originally derived for the , the effect analogously arises for scalar fields with appropriate boundary conditions, demonstrating the physical reality of zero-point fluctuations.

Fermionic and Gauge Fields

The canonical quantization of fermionic fields extends the second quantization formalism to half-integer spin particles, requiring anticommutation relations to incorporate the Pauli exclusion principle and ensure consistency with relativistic causality. The Dirac field, describing electrons and other spin-1/2 fermions, serves as the prototypical example. The classical Dirac field is a four-component spinor satisfying the Dirac equation, and upon quantization, the field operator \psi(x) and its adjoint \psi^\dagger(x) obey the equal-time canonical anticommutation relations \{\psi_\alpha(\mathbf{x},t), \psi^\dagger_\beta(\mathbf{y},t)\} = \hbar \delta_{\alpha\beta} \delta^3(\mathbf{x}-\mathbf{y}), while all other anticommutators vanish. These relations arise from the need to treat fermions as indistinguishable particles obeying Fermi-Dirac statistics, preventing unphysical states with multiple occupancy of the same mode. The free Dirac field Lagrangian density is given by \mathcal{L} = i \bar{\psi} \gamma^\mu \partial_\mu \psi - m \bar{\psi} \psi, where \bar{\psi} = \psi^\dagger \gamma^0, m is the fermion mass, and \gamma^\mu are the Dirac matrices. This Lagrangian leads to the Dirac equation through the Euler-Lagrange equations and provides the starting point for deriving the Hamiltonian in the canonical formalism. The field is expanded in a mode decomposition analogous to the scalar case, but with fermionic creation b^\dagger_s(\mathbf{k}) and annihilation b_s(\mathbf{k}) operators for particles, and similar operators d^\dagger_s(\mathbf{k}), d_s(\mathbf{k}) for antiparticles, satisfying \{b_s(\mathbf{k}), b^\dagger_{s'}(\mathbf{k}'\}\}=\delta_{ss'}(2\pi)^3\delta^3(\mathbf{k}-\mathbf{k}') and likewise for the d operators, with vanishing mixed anticommutators. The vacuum is defined as the state annihilated by all b_s(\mathbf{k}) and d_s(\mathbf{k}), and the Hamiltonian for the free field becomes H = \int \frac{d^3k}{(2\pi)^3} \, \sum_s \left[ \sqrt{|\mathbf{k}|^2 + m^2} \left( b^\dagger_s(\mathbf{k}) b_s(\mathbf{k}) + d^\dagger_s(\mathbf{k}) d_s(\mathbf{k}) \right) \right] + E_0, where E_0 accounts for zero-point contributions regularized appropriately. The spin-statistics theorem underpins this quantization scheme, dictating that integer-spin fields (bosons) must satisfy commutation relations while half-integer-spin fields (fermions) require anticommutation relations to maintain Lorentz invariance and positive energy spectra in relativistic . Violations would lead to negative-probability interpretations or acausal propagation, as demonstrated in the theorem's proof. For gauge fields, such as the , canonical quantization encounters additional challenges due to the redundancy from invariance, which introduces and unphysical . The A_\mu(x) is subject to the Lorenz condition \partial^\mu A_\mu = 0 in covariant approaches, but in the canonical formalism, the Coulomb \nabla \cdot \mathbf{A} = 0 is preferred, as it manifestly separates transverse (physical) modes from longitudinal ones and sets the A_0 to satisfy as a . This eliminates two unphysical per mode, leaving two polarization states for the . However, residual freedom persists under time-independent transformations \mathbf{A} \to \mathbf{A} + \nabla \Lambda(\mathbf{x}) with \nabla^2 \Lambda = 0, which must be quotiented out in the Hilbert space via the Gupta-Bleuler condition or subsidiary conditions to ensure physical states are gauge-invariant. The field is quantized by imposing commutation relations on the transverse components, expanding \mathbf{A}(\mathbf{x},t) in Fourier modes with annihilation a_{\lambda}(\mathbf{k}) and creation a^\dagger_{\lambda}(\mathbf{k}) operators for helicities \lambda = 1,2, satisfying [a_{\lambda}(\mathbf{k}), a^\dagger_{\lambda'}(\mathbf{k}')] = \delta_{\lambda\lambda'} (2\pi)^3 \delta^3(\mathbf{k}-\mathbf{k}'). The free-field Hamiltonian is then H = \int \frac{d^3k}{(2\pi)^3} \, \sum_\lambda |\mathbf{k}| a^\dagger_{\lambda}(\mathbf{k}) a_{\lambda}(\mathbf{k}) + E_0. In (QED), the interaction between the Dirac field and the electromagnetic field is incorporated via in the , leading to a with the interaction term H_I = -e \int d^3x \, :\bar{\psi}(\mathbf{x},t) \gamma^\mu \psi(\mathbf{x},t): A_\mu(\mathbf{x},t), where : \dots :\ ) subtracts [vacuum](/page/Vacuum) divergences and \(e is the . In the gauge, the \mu=0 component handles instantaneous Coulomb interactions, while spatial components mediate transverse exchange, ensuring gauge invariance and unitarity in . This structure reproduces the Feynman rules for QED scattering processes.

Formal Mathematical Approaches

Deformation Quantization

Deformation quantization provides a framework for canonical quantization by deforming the algebra of classical observables—smooth functions on a Poisson manifold—into a non-commutative algebra via a star product parameterized by the deformation constant \nu = \hbar, where \hbar is the reduced Planck's constant. This approach, formalized in the late 1970s, treats quantum mechanics as a continuous deformation of classical mechanics, avoiding the direct construction of operators on Hilbert space and instead working in phase space. Unlike traditional operator-based methods, it globally resolves ambiguities in operator ordering by embedding them into the structure of the deformed product, ensuring consistency with the canonical commutation relations in the semiclassical limit. The star product \star is a bilinear, associative operation on the of functions C^\infty(M)[[\nu]], defined as f \star g = \sum_{r=0}^\infty \nu^r C_r(f, g), where C_0(f, g) = f g and the first-order term satisfies C_1(f, g) - C_1(g, f) = i \{f, g\}, with \{f, g\} denoting the . A canonical example is the on the \mathbb{R}^{2n} with the standard symplectic structure, given by (f \star_M g)(z) = \left. \exp\left(\frac{i\hbar}{2} P^{rs} \frac{\partial}{\partial x^r} \frac{\partial}{\partial y^s}\right) f(x) g(y) \right|_{x=y=z}, where P^{rs} is the Poisson bivector and z = (q, p). This product deforms the pointwise multiplication while preserving associativity and the unit function. The Weyl symbol map establishes a between quantum operators \hat{O} and classical symbols \tilde{O}(q, p) via a Fourier-type transform, \hat{O} = \frac{1}{(2\pi \hbar)^n} \int \tilde{O}(\xi, \eta) \exp\left(\frac{i}{\hbar} (\xi \cdot \hat{q} + \eta \cdot \hat{p})\right) d\xi d\eta, ensuring that the operator composition maps to the star product: \widehat{f \star g} = \hat{f} \hat{g}. This map, originally proposed by Weyl, connects phase-space functions to self-adjoint operators on L^2(\mathbb{R}^n), facilitating a phase-space formulation of quantum mechanics. A key property is the star commutator [f, g]_\star = f \star g - g \star f = i\hbar \{f, g\} + O(\hbar^3), which reproduces the canonical quantization rule [\hat{f}, \hat{g}] = i\hbar \widehat{\{f, g\}} to leading order in \hbar, while higher terms account for quantum corrections without ad hoc ordering prescriptions. This equivalence holds for small \hbar, aligning deformation quantization with standard canonical methods, but extends them globally by treating all observables on equal footing in . In applications, it proves useful for semiclassical approximations, where the WKB method emerges as an of the star product, and for coherent states, where the Berezin or star products on Kähler manifolds yield representations akin to the Bargmann-Fock space, minimizing uncertainty in . These tools have been applied to study and integrable systems, providing insights into the transition from classical to .

Geometric Quantization

Geometric quantization is a mathematical framework for implementing canonical quantization rules on a classical modeled as a (M, \omega), where \omega is the symplectic form derived from the structure of the phase space. Developed independently by and Jean-Marie Souriau, this approach constructs a of quantum states and operator representations of classical observables through geometric constructions on associated bundles over M. Unlike algebraic methods, it emphasizes the to ensure the quantum theory respects the classical structure, particularly for systems with symmetries such as coadjoint orbits of Lie groups. The first step in is prequantization, which associates to (M, \omega) a L \to M (the prequantum ) equipped with a \nabla whose form satisfies F_\nabla = \frac{i}{\hbar} \omega. This condition ensures that parallel transport around closed loops encodes the symplectic area via the , providing a geometric realization of the commutation relations at the level of U(1)-gerbes or bundles when \frac{[\omega]}{2\pi \hbar} \in H^2(M; \mathbb{Z}). The Kostant-Souriau construction guarantees the existence of such a bundle and for any prequantum condition, with sections of L forming a pre-Hilbert where classical functions f \in C^\infty(M) act via Kostant-Souriau operators \hat{f} = -i\hbar \nabla_{X_f} + f, where X_f is the of f. To obtain a true of square-integrable wave functions, prequantization is refined by choosing a polarization \mathcal{P}, a maximally subbundle of the complexified T_\mathbb{C}M that is integrable and positive with respect to \omega. This selects a "half-space" of directions in which the wave functions are holomorphic, effectively reducing the infinite-dimensional prequantum space to a finite-dimensional quantum for compact M. For Kähler manifolds, the vertical polarization—spanned by anti-holomorphic vector fields—is commonly used, aligning with representations. However, the metaplectic representation introduces a half-form correction: quantum states are taken as sections of the bundle \sqrt{L} \otimes (K^{1/2})^*, where K is the , to ensure unitarity under the Bochner-Kostant kernel or half-density quantization, accounting for the of the volume form along the polarization leaves. In the semiclassical limit, recovers the Bohr-Sommerfeld quantization condition as a on Lagrangian submanifolds (tori or orbits) within the polarized space: the of \omega / (2\pi \hbar) over such a must be an for it to support non-vanishing quantum states, ensuring the symplectic flux through Bohr-Sommerfeld leaves matches the quantized action. This manifests as a on the preimages of classical periodic orbits under the moment map. A canonical example is the quantization of the 2-sphere S^2, modeling the phase space of a particle's angular momentum under SU(2) symmetry, where M = S^2 is a coadjoint orbit with symplectic form \omega = j \sin\theta \, d\theta \wedge d\phi (in units where \hbar = 1) for total angular momentum quantum number j. The prequantum line bundle is the Hopf bundle S^1 \hookrightarrow S^3 \to S^2 with Chern number -2j, and the Kähler polarization (vertical) yields the Hilbert space of holomorphic sections as the space of spherical harmonics of degree $2j, decomposing into the irreducible representation of dimension $2j + 1 with highest weight j. This recovers the discrete spectrum of angular momentum operators J^2 = j(j+1) and J_z = -j, \dots, j, illustrating how geometric quantization classifies unitary representations via orbit methods.

References

  1. [1]
    On the theory of quantum mechanics - Journals
    Dirac Paul Adrien Maurice. 1926On the theory of quantum mechanicsProc. R. Soc. Lond. A112661–677http://doi.org/10.1098/rspa.1926.0133. Section. Abstract ...
  2. [2]
    PAM Dirac and the discovery of quantum mechanics - AIP Publishing
    Mar 1, 2011 · In short, Born and Jordan and Dirac independently discovered canonical quantization, and thereby transformed Heisenberg's scheme into a complete ...
  3. [3]
    [PDF] 4 Canonical Quantization
    Canonical quantization associates classical canonical pairs (q, p) with operators q and p in quantum mechanics, obeying commutation relations [q, p] = i/h.
  4. [4]
    [PDF] Canonical Quantization C6, HT 2016
    Canonical quantization promotes classical coordinates and momenta to operators, using commutation relations to move from classical to quantum theory.
  5. [5]
    [PDF] Canonical Quantization and Application to the Quantum Mechanics ...
    Classical mechanics is internally consistent. No amount of examination of. Newton's Laws as an abstract system will lead you to quantum mechanics.
  6. [6]
    [PDF] PRINCIPLES QUANTUM MECHANICS
    An important concept in general dynamical theory is the Poisson. Bracket ... quantum mechanics which are describable in terms of canonical coordinates ...
  7. [7]
    Zur Quantenmechanik einfacher Bewegungstypen
    Download PDF · Zeitschrift für Physik. Zur Quantenmechanik einfacher ... Cite this article. Kennard, E.H. Zur Quantenmechanik einfacher Bewegungstypen.
  8. [8]
    [PDF] 1913 On the Constitution of Atoms and Molecules
    This paper is an attempt to show that the application of the above ideas to Rutherford's atom-model affords a basis for a theory of the constitution of atoms.
  9. [9]
    [PDF] Quantization Conditions, 1900–1927 - PhilSci-Archive
    Mar 9, 2020 · Using this procedure, Sommerfeld was able to generalize the circular orbits of Niels Bohr's hydrogen atom (selected through quantization of the ...
  10. [10]
    [PDF] On quantum-theoretical reinterpretation of kinematic and ...
    In this paper, we will attempt to achieve a quantum-theoretical mechanics that is based exclusively upon relationships between quantities that are observable, ...Missing: matrix | Show results with:matrix
  11. [11]
    [PDF] The 1925 Born and Jordan paper “On quantum mechanics” - ISY
    Heisenberg made the breakthrough in his historic 1925 paper, “Quantum-theoretical reinterpretation of kinematic and mechanical relations.”2 Heisenberg's bold ...
  12. [12]
    [PDF] 1926-Schrodinger.pdf
    The paper gives an account of the author's work on a new form of quantum theory. §1. The Hamiltonian analogy between mechanics and optics. §2. The analogy is to ...
  13. [13]
    [PDF] Relation between the Quantum Mechanics of Heisenberg, Born ...
    MATRIX MECHANICS AND WAVE MECHANICS. 47 may interest us from the surmise that ... ever, the following supplement to the proof of equivalence given above ...
  14. [14]
    [PDF] Quantization: History and Problems - arXiv
    Feb 16, 2022 · Abstract. In this work, I explore the concept of quantization as a mapping from classical phase space functions to quantum operators.
  15. [15]
    The fundamental equations of quantum mechanics - Journals
    Dirac Paul Adrien Maurice. 1925The fundamental equations of quantum mechanicsProc. R. Soc. Lond. A109642–653http://doi.org/10.1098/rspa.1925.0150. Section.
  16. [16]
    [PDF] The Fundamental Equations of Quantum Mechanics
    The Fundamental Equations of Quantum Mechanics. Author(s): P. A. M. Dirac. Source: Proceedings of the Royal Society of London. Series A, Containing Papers of ...
  17. [17]
    [PDF] The Principles of Quantum Mechanics
    The growth of the use of transformation theory, as applied first to relativity and later to the quantum theory, is the essence of the new method in theoretical.
  18. [18]
    [PDF] THE LAGRANGIAN IN QUANTUM MECHANICS. By PAM Dirac.
    P. A. M. Dirac. The Lagrangian in Quantum Mechanics. 65 to take over the classical Lagrangian equations in any very. ~iirect way. These equations involve ...
  19. [19]
    [PDF] Lecture 3 Canonical Quantization
    This issue is tackled by Dirac, which derives a relativistic equation for the wave-function that is first order in both time and space derivatives. But this ...
  20. [20]
    [PDF] ANGULAR MOMENTUM - MIT OpenCourseWare
    Dec 16, 2013 · A key property of central potential problems is that the angular momentum operators commute with the Hamiltonian. Central potential ...
  21. [21]
    [PDF] Quantum mechanics in more than one-dimension - TCM
    4.2 Angular momentum. 4.2.1 Commutation relations. Following the usual canonical quantization procedure, the angular momentum operator is defined by L = r×p ...
  22. [22]
    [PDF] Quantum Physics I, Lecture Note 22 - MIT OpenCourseWare
    May 4, 2016 · We get two Schrödinger equations. The first equation tells us that the center of mass moves as a free particle of mass M. Thus, the CM energy is ...<|separator|>
  23. [23]
    None
    ### Summary of Hydrogen Atom Content from Lecture 7
  24. [24]
    [PDF] 1 The postulates of quantum mechanics - WPI
    The symmetrization of the classical physical quantity is necessary to en- sure that the resulting operator A is Hermitian. Assume, for example, that A = xp. In ...
  25. [25]
    [PDF] A completely algebraic solution of the simple harmonic oscillator
    Using this matrix-valued equation of motion and the canonical commutation relation, Heisenberg was able to find the quantized energy levels. The first ...
  26. [26]
    None
    ### Summary of 1D Harmonic Oscillator, Ladder Operators, and Energy Levels
  27. [27]
  28. [28]
    Quantum mechanics as a statistical theory
    Oct 24, 2008 · An attempt is made to interpret quantum mechanics as a statistical theory, or more exactly as a form of non-deterministic statistical dynamics.
  29. [29]
    Quantenmechanik und Gruppentheorie | Zeitschrift für Physik A ...
    § 1. Mathematische Grundbegriffe, die Hermiteschen Formen betreffend. § 2. Der physikalische Begriff des reinen Falles. § 3. Die physikalische Bedeutung der ...
  30. [30]
    On the principles of elementary quantum mechanics - ScienceDirect
    Physica · Volume 12, Issue 7, October 1946, Pages 405-460. Physica. On the principles of elementary quantum mechanics. Author links open overlay panel H.J. ...
  31. [31]
    [PDF] On the Groenewold-Van Hove problem for R - arXiv
    In Groenewold's paper [Gro] a stronger result was claimed: his assertion was that there is no quantization of P, period. This is not what Theorem 2 states. For ...
  32. [32]
    [PDF] a groenewold-van hove theorem for s2
    In a striking paper, Groenewold [Gr] showed that one cannot “consistently” quantize all polynomials in the classical positions qi and momenta pi on R2n. Sub-.Missing: original | Show results with:original
  33. [33]
    On the Groenewold–Van Hove problem for R2n - AIP Publishing
    We discuss the Groenewold–Van Hove problem for R2n, and completely solve it when n=1. We rigorously show that there exists an obstruction to quantizing the ...
  34. [34]
    [PDF] Quantization: History and problems - OSTI.GOV
    Feb 20, 2022 · ... Dirac's initial motivation for introducing the Poisson bracket into ... paper about quantum mechanics, “On the Theory of Quantum Mechanics,” ...
  35. [35]
    Quantization: History and problems - ScienceDirect.com
    In this work, I explore the concept of quantization as a mapping from classical phase space functions to quantum operators.
  36. [36]
    [PDF] Lectures on the Geometry of Quantization - UC Berkeley math
    These notes are based on a course entitled “Symplectic geometry and geometric quantization” taught by Alan Weinstein at the University of California, ...<|control11|><|separator|>
  37. [37]
    [PDF] NP Landsman - Mathematics
    ... N.P. Landsman. Korteweg-de Vries Institute for Mathematics. University of ... (quantization). This monograph draws on two traditions: The algebraic ...
  38. [38]
    [PDF] 8 (1974), No. 5 QUANTIZATION UDC 513-88 F. A. BE
    quantization, under several additional conditions of an algebraic and topological nature. Page 10. 1118. F. A. BEREZIN. §2. Quantization in Kahler manifolds. 1 ...
  39. [39]
    The quantum theory of the emission and absorption of radiation
    The new quantum theory, based on the assumption that the dynamical variables do not obey the commutative law of multiplication, has by now been developed ...
  40. [40]
    Konfigurationsraum und zweite Quantelung | Zeitschrift für Physik A ...
    Fock, V. Konfigurationsraum und zweite Quantelung. Z. Physik 75, 622–647 (1932). https://doi.org/10.1007/BF01344458. Download citation. Received: 10 March 1932.
  41. [41]
    [PDF] On the attraction between two perfectly conducting plates
    -. On the attraction between two perfectly conducting plates. By H. B. G. CASIMIR. (Communicated at the meeting of May 29. 1948.) In a recent paper by POLDER ...
  42. [42]
    [PDF] Quantum dynamics of wave fields I - Neo-classical physics
    “Zur Quantendynamik der Wellenfelder,” Zeit. Phys. 56 (1929), 1-61. On the quantum dynamics of wave fields. By W. Heisenberg in Leipzig and W. Pauli in ...
  43. [43]
    Zur Quantendynamik der Wellenfelder | Zeitschrift für Physik A ...
    Heisenberg, W., Pauli, W. Zur Quantendynamik der Wellenfelder. Z. Physik 56, 1–61 (1929). https://doi.org/10.1007/BF01340129
  44. [44]
    [PDF] Pauli_spin_statistics_1940.pdf - Fisica Fundamental
    In the following paper we conclude for the relativistically invariant wave equation for free particles: From postulate (I), according to which the energy must ...
  45. [45]
    Deformation theory and quantization. I. Deformations of symplectic ...
    March 1978, ... Deformation theory and quantization. I. Deformations of symplectic structures☆. Author links open overlay panel. F Bayen , M Flato, C Fronsdal,
  46. [46]
    [PDF] Deformation Quantization : an introduction
    Abstract. I shall focus, in this presentation of Deformation Quantization, to some math- ematical aspects of the theory of star products.
  47. [47]
    [PDF] States and representations in deformation quantization - arXiv
    Aug 17, 2004 · Based on works of Weyl, Groenewold, Moyal, Berezin and others [12–14,78,108,140] on the physical side and on Gerstenhaber's deformation ...
  48. [48]
    Monodromy and the Bohr-Sommerfeld Geometric Quantization
    We study the linear part of the monodromy of completely integrable Hamiltonian systems via Bohr--Sommerfeld Geometric Quantization.
  49. [49]
    [PDF] Six lectures on Geometric Quantization - arXiv
    May 31, 2023 · 5.3 The 2-sphere, quantization of angular momentum. Finally, as an example of a geometric quantization of a compact manifold, let us con-.