Fact-checked by Grok 2 weeks ago

First quantization

First quantization is the original formulation of that quantizes the dynamics of a fixed number of particles by replacing classical position and momentum variables with non-commuting operators in a , leading to wave functions that evolve according to the . This approach, developed between 1925 and 1926, marked the transition from classical to quantum descriptions of matter and radiation for non-relativistic systems with a predetermined particle count. The historical origins trace to Werner Heisenberg's 1925 introduction of , which used arrays to represent observables and enforced commutation relations like [q, p] = i\hbar to capture quantum uncertainty. Shortly after, in 1926, formulated wave mechanics, deriving the time-dependent i\hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi by analogy to classical wave equations and applying boundary conditions to quantize energy levels, such as in the . Paul contributed pivotal insights in 1925–1926 by formalizing quantization as a systematic mapping from classical brackets to quantum commutators, bridging matrix and wave approaches in his transformation theory. These efforts culminated in the probabilistic interpretation by , where |\psi|^2 represents probability density, underpinning measurable outcomes like position and momentum eigenvalues from Hermitian operators. In practice, first quantization applies to single- or few-particle systems, solving the many-body i\hbar \frac{\partial \Psi}{\partial t} = \hat{H} \Psi(\mathbf{r}_1, \dots, \mathbf{r}_N, t) with Hamiltonians incorporating , potentials, and interactions, while enforcing symmetrization (for bosons) or antisymmetrization (for fermions) to account for identical particles. This framework excels in non-relativistic contexts, such as spectra, molecular vibrations, and quantum tunneling in , but becomes cumbersome for variable particle numbers or relativistic effects, where it requires explicit handling of statistics and excludes vacuum fluctuations. Distinguished from , which reformulates in using to describe arbitrary particle occupations and field theories, first quantization operates in a fixed-dimensional and treats fields classically, limiting its scope to systems without particle creation or annihilation. Despite these limitations, first quantization remains foundational for understanding core quantum principles like superposition, entanglement, and the Heisenberg uncertainty relation \sigma_x \sigma_p \geq \hbar/2, derived from operator non-commutativity.

Overview and Basics

Definition of First Quantization

First quantization refers to the foundational approach in that transforms the classical description of point particles, characterized by definite trajectories in , into a probabilistic framework using wave functions. These wave functions represent the of the system and evolve in time according to the , providing probabilities for measurement outcomes rather than precise paths. This method applies primarily to non-relativistic systems with a fixed number of particles, treating them as excitations of a wave rather than classical points. The term "first quantization" originated retrospectively to distinguish this particle-based quantization from the later development of for quantum fields, highlighting its role as the initial step in quantizing mechanical systems. It was introduced in contrast to the field quantization procedures pioneered by Dirac and in the late , which extend the formalism to relativistic contexts and variable particle numbers. At its core, first quantization promotes classical variables to operators acting on a , where the is encoded in wave functions. The expectation values of these operators yield measurable observables, which recover classical values in the correspondence limit of large quantum numbers or \hbar \to 0, ensuring consistency with for macroscopic scales. As an introductory example, consider position and momentum: in , these are real numbers specifying a particle's state, but in first quantization, they become operators \hat{q} and \hat{p} whose action on wave functions encodes the and inherent to quantum behavior.

Distinction from Second Quantization

Second quantization refers to the procedure of quantizing classical fields, such as electromagnetic or matter fields, by promoting them to operators that act on a , thereby incorporating to allow for variable particle numbers. In contrast, first quantization treats particles as fundamental entities with a fixed number, describing their states via wave functions in a configuration space . The primary differences lie in their treatment of particle statistics and interactions: first quantization assumes a predetermined number of particles and requires explicit symmetrization or antisymmetrization of wave functions to handle , whereas naturally accommodates indistinguishability through occupation number representations in and simplifies the description of interactions via field operators. This makes particularly suited for systems where particle number is not conserved, such as in processes involving or . First quantization offers simplicity and computational efficiency for systems with a few non-interacting particles, as it avoids the need for operator algebras and directly leverages familiar wave mechanics tools. However, it becomes inefficient and cumbersome for many-body systems due to the exponential growth in the dimensionality of the and difficulties in treating strong interactions or relativistic effects, where particle number fluctuations are significant. A illustrative example is the : in first quantization, the system is described by a single-particle satisfying the , yielding discrete energy levels without inherent structure; in , the same oscillator is represented using on a , naturally incorporating vacuum fluctuations and multi-mode extensions for identical particles. First quantization proves inadequate for quantum field theories, where relativistic invariance and locality demand a framework that unifies particles and fields, leading to the adoption of —as exemplified in , where and fields are quantized to describe scattering and processes.

Mathematical Preliminaries

Canonical Quantization Procedure

The procedure provides a systematic method to transition from a classical formulation to its quantum counterpart by promoting classical variables to operators on a , guided by algebraic rules that preserve the structure of classical brackets. Developed primarily by , this approach emphasizes the correspondence between classical dynamics and quantum evolution through operator commutators. Central to Dirac's prescription is the replacement of classical Poisson brackets with quantum commutators, scaled by the reduced Planck's constant. Specifically, for any two classical functions A(q, p) and B(q, p) of q and p, the \{A, B\} is mapped to the commutator of the corresponding operators as [\hat{A}, \hat{B}] / i\hbar, where \hbar = h / 2\pi and h is Planck's constant. This rule ensures that the quantum , derived from the , mirror the classical Hamilton's equations in form. For instance, the of an operator \hat{A} follows i\hbar d\hat{A}/dt = [\hat{A}, \hat{H}], analogous to the classical dA/dt = \{A, H\}. The basic dynamical variables are quantized via specific operator representations: the position q becomes the multiplication operator \hat{q} \psi(q) = q \psi(q), while the momentum p is represented as the differential operator \hat{p} = -i\hbar d/dq in the position basis. These lead to the fundamental canonical commutation relations, which underpin the algebra of quantum observables. In one dimension, [\hat{q}, \hat{p}] = i\hbar; in three dimensions, this generalizes to [\hat{x}_i, \hat{p}_j] = i\hbar \delta_{ij} for i, j = 1, 2, 3, with all other commutators vanishing. These relations are postulated as the quantum analog of the classical \{q, p\} = 1. When quantizing more complex classical s involving products of non-commuting operators, ambiguities arise in the operator ordering. To resolve these and ensure the resulting is Hermitian (), proposed a symmetric ordering scheme, known as Weyl ordering, where products like q p are symmetrized as (\hat{q} \hat{p} + \hat{p} \hat{q})/2. This method, derived from group-theoretic considerations, provides a unique quantization map for polynomials in variables and maintains under transformations. A illustrative example is the quantization of the classical H = p^2 / 2m + (1/2) m \omega^2 q^2, where m is and \omega is . Applying the procedure yields the quantum \hat{H} = \hat{p}^2 / 2m + (1/2) m \omega^2 \hat{q}^2, with the ordering ambiguity resolved via Weyl symmetrization if needed, though the simple form suffices here due to the quadratic nature. This operator acts on wave functions in the , leading to energy eigenvalues (n + 1/2) \hbar \omega for integer n \geq 0. The procedure adheres to the correspondence principle, ensuring consistency with classical mechanics in the semiclassical limit. As \hbar \to 0, the expectation values of quantum operators \langle \hat{A} \rangle approach the classical values A(q, p), and quantum dynamics reduce to classical trajectories, validating the quantization rules for macroscopic scales.

Hilbert Space and Wave Functions

In first quantization, the quantum state of a single non-relativistic particle is described by a normalized vector in the separable Hilbert space \mathcal{H} = L^2(\mathbb{R}^3), consisting of complex-valued square-integrable functions on three-dimensional Euclidean space. This space is equipped with the inner product \langle \psi | \phi \rangle = \int_{\mathbb{R}^3} \psi^*(\mathbf{r}) \phi(\mathbf{r}) \, d^3\mathbf{r}, which induces a norm \|\psi\| = \sqrt{\langle \psi | \psi \rangle} measuring the "size" of states and ensuring completeness for infinite-dimensional systems. The Hilbert space structure allows for the representation of quantum superpositions as linear combinations of basis states, embodying the principle that any state can be expressed as \psi = \sum c_n |n\rangle where \sum |c_n|^2 = 1. The wave function \psi(\mathbf{r}, t) represents the state in the position basis, where \psi(\mathbf{r}, t) = \langle \mathbf{r} | \psi(t) \rangle gives the amplitude for finding the particle at position \mathbf{r} at time t. According to the , the probability density of measuring the particle at \mathbf{r} is |\psi(\mathbf{r}, t)|^2, with the normalization condition \int_{\mathbb{R}^3} |\psi(\mathbf{r}, t)|^2 \, d^3\mathbf{r} = 1 ensuring total probability unity. In the momentum basis, the wave function is obtained via the \tilde{\psi}(\mathbf{p}, t) = \frac{1}{(2\pi\hbar)^{3/2}} \int_{\mathbb{R}^3} \psi(\mathbf{r}, t) e^{-i\mathbf{p}\cdot\mathbf{r}/\hbar} \, d^3\mathbf{r}, allowing equivalent descriptions in either representation. Time evolution of the state is governed by unitary operators on the , specifically |\psi(t)\rangle = U(t) |\psi(0)\rangle where U(t) = e^{-i\hat{H}t/\hbar} and \hat{H} is the operator, preserving the norm \|\psi(t)\| = \|\psi(0)\| and thus probability conservation. The extends to general measurements: for an with eigenbasis |\phi_n\rangle, the probability of outcome n is |\langle \phi_n | \psi \rangle|^2, highlighting the probabilistic nature of quantum predictions. A representative example is the , where an initial \psi(\mathbf{r}, 0) = \left(\frac{2\alpha}{\pi}\right)^{3/4} e^{-\alpha r^2} e^{i\mathbf{k}_0 \cdot \mathbf{r}} (with \alpha > 0) evolves such that its spatial width \sigma(t) \approx \sqrt{\sigma_0^2 + (\hbar t / 2m\sigma_0)^2} increases quadratically with time, illustrating due to the superposition of plane waves with different momenta. This spreading underscores the irreversible loss of initial localization in first quantization, contrasting classical point-particle trajectories.

Historical Context

Precursors to Quantum Mechanics

The foundations of , which preceded the development of , were laid in the through the formulation. In this framework, the dynamics of a are described by Hamilton's equations, which govern the evolution of q and momenta p via the function H(q, p, t): \frac{dq}{dt} = \frac{\partial H}{\partial p}, \quad \frac{dp}{dt} = -\frac{\partial H}{\partial q}. These equations define trajectories in , a multidimensional where each point represents a unique state of the system, allowing for a complete deterministic description of motion under conservative forces. This classical picture successfully explained macroscopic phenomena but encountered severe limitations when applied to microscopic systems, particularly in and . One prominent failure was the in , where the Rayleigh-Jeans law predicted an infinite at high frequencies, diverging from experimental observations of finite . Similarly, classical electrodynamics applied to Rutherford's 1911 nuclear model of the atom forecasted instability: orbiting electrons would continuously radiate energy, spiraling into the nucleus in fractions of a second, contradicting the observed persistence of atomic structures. These inconsistencies motivated the introduction of quantization. In 1900–1901, resolved the blackbody problem by proposing that energy is emitted and absorbed in discrete quanta, with E = n h f, where n is an integer, h is Planck's constant, and f is the frequency, fitting experimental spectra without infinities. Building on this, in 1905 explained the by treating light as discrete packets (quanta, later called photons) with energy E = h f, accounting for the threshold frequency dependence of electron emission. Niels Bohr's 1913 atomic model incorporated quantization to stabilize Rutherford's structure, postulating stationary electron orbits with quantized L = n \hbar (where \hbar = h / 2\pi), an approximation later refined to \sqrt{l(l+1)} \hbar in full ; transitions between orbits emitted discrete radiation, matching spectral lines. This initiated the of the 1910s–1920s, which applied ad hoc rules to periodic systems, such as the Wilson-Sommerfeld condition \oint p \, dq = n h for action integrals over closed paths, extending Bohr's ideas to elliptical orbits and multi-electron atoms but remaining inconsistent for non-integrable systems.

Key Formulations in the 1920s

The development of first quantization began with Werner Heisenberg's introduction of in 1925, where physical observables such as and were represented by infinite arrays, or matrices, that do not commute under multiplication. This formulation emphasized observable quantities and their transitions, avoiding unobservable classical trajectories, and incorporated the fundamental commutation relation between x and p operators, [x, p] = i\hbar, which implies the limiting simultaneous precision in measuring conjugate variables. Shortly thereafter, and formalized Heisenberg's ideas in their 1925 paper, establishing the commutation relations as a cornerstone of the theory and proving their consistency with the old quantum conditions for periodic systems. They demonstrated that these relations could be derived from a quantum analog of the , providing a systematic framework for calculating transition amplitudes and energy levels in atomic spectra. In 1926, Erwin Schrödinger proposed an alternative wave mechanics approach, representing quantum states by wave functions \psi in configuration space and deriving the time-dependent Schrödinger equation, i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi, where \hat{H} is the Hamiltonian operator. This equation emerged from extending Louis de Broglie's wave-particle duality to a relativistic Hamilton-Jacobi equation, treating particles as wave packets whose interference patterns govern their dynamics. In a subsequent 1926 paper, Schrödinger demonstrated the mathematical equivalence between his wave mechanics and Heisenberg's matrix mechanics, providing an early unification of the two approaches. John von Neumann established a more rigorous mathematical equivalence between matrix and wave mechanics in his 1927 papers, proving an isomorphism between the two formulations through the for operators in . This unification showed that both approaches describe the same physical predictions, with matrix elements corresponding to integrals over wave functions, solidifying first quantization as a coherent framework. Paul Dirac contributed to bridging these formulations with his 1927 transformation theory, which generalized quantum states as vectors in an abstract space and allowed representations in either matrix or wave forms via unitary transformations. In the same work, Dirac further refined the canonical quantization procedure by systematically replacing classical Poisson brackets \{q, p\} with commutators [q, p]/i\hbar, providing a correspondence principle that extended first quantization to more complex systems. As an extension of first quantization, and introduced early ideas of in 1929, treating field amplitudes as operators to handle radiation and particle creation, though this marked a departure toward . These formulations resolved key debates on the completeness of , as later highlighted in the 1935 Einstein-Podolsky-Rosen paradox, which questioned whether first quantization fully described physical reality but affirmed its foundational role post-1928.

Systems in First Quantization

One-Particle Systems

In first quantization, the behavior of a single particle subject to a potential V(\mathbf{r}) is described by solving the time-independent for stationary states: \hat{H} \psi(\mathbf{r}) = E \psi(\mathbf{r}), where the operator is \hat{H} = -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}), \psi(\mathbf{r}) is the wave function, E is the eigenvalue, m is the particle mass, and \hbar is the reduced Planck's constant. This eigenvalue problem yields quantized levels and corresponding spatial probability distributions |\psi(\mathbf{r})|^2, providing physical insights into bound states and confinement effects. For a free particle where V(\mathbf{r}) = 0, the solutions are plane waves of the form \psi(\mathbf{r}, t) = A \exp\left[i (\mathbf{k} \cdot \mathbf{r} - \omega t)\right], with the dispersion relation E = \frac{\hbar^2 k^2}{2m} and \omega = E / \hbar, reflecting the particle's momentum \mathbf{p} = \hbar \mathbf{k}. These delocalized waves illustrate the wavelike nature of matter, as proposed by de Broglie, and highlight the absence of discrete energies in unbounded systems. A foundational example of confinement is the infinite square well potential, defined as V(x) = 0 for $0 < x < a and V(x) = \infty otherwise in one dimension. The normalized wave functions are \psi_n(x) = \sqrt{\frac{2}{a}} \sin\left(\frac{n \pi x}{a}\right), \quad n = 1, 2, 3, \dots, with discrete energy levels E_n = \frac{n^2 \pi^2 \hbar^2}{2 m a^2}. These standing waves vanish at the boundaries, demonstrating quantization due to the finite spatial domain and enabling calculations of transition probabilities between levels. The hydrogen atom provides a three-dimensional illustration, where the potential is V(r) = -\frac{e^2}{4\pi \epsilon_0 r} (in atomic units, simplified to -1/r). Separation of variables in spherical coordinates leads to the radial Schrödinger equation, solved using associated Laguerre polynomials, yielding quantum numbers n (principal), l (orbital angular momentum, $0 \leq l < n), and m (magnetic, -l \leq m \leq l). The bound-state energies depend only on n: E_n = -\frac{13.6 \, \mathrm{eV}}{n^2}. The angular momentum operators are defined as \hat{\mathbf{L}} = \hat{\mathbf{r}} \times \hat{\mathbf{p}}, with eigenvalues \sqrt{l(l+1)} \hbar for \hat{L}^2 and m \hbar for \hat{L}_z, explaining the degeneracy and spectroscopic structure observed in atomic spectra. Quantum tunneling exemplifies non-classical penetration through barriers, as in Gamow's 1928 model of alpha decay. For an alpha particle confined in a nuclear potential well (r < R, where V(r) < E) that encounters a Coulomb barrier (V(r) > E for r > R), the transmission probability P \propto \exp\left(-2 \int \sqrt{2m (V(r) - E)} / \hbar \, dr \right) allows escape, predicting decay rates that match Geiger-Nuttall law observations for heavy nuclei. This application underscores first quantization's success in nuclear processes.

Many-Particle Systems

In first quantization, the quantum state of a system consisting of N particles is described by a wave function \psi(\mathbf{r}_1, \mathbf{r}_2, \dots, \mathbf{r}_N, t) defined on a $3N-dimensional configuration space, where each \mathbf{r}_i represents the position of the i-th particle in . This high-dimensional space encodes the joint probabilities for all particle positions, extending the single-particle three-dimensional to capture correlations among particles. The time evolution follows the many-particle i\hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi, where the governs the . The Hamiltonian for a system of N particles takes the form \hat{H} = \sum_{i=1}^N \left( -\frac{\hbar^2}{2m_i} \nabla_i^2 + V_i(\mathbf{r}_i) \right) + \sum_{i < j} V_{ij}(\mathbf{r}_i, \mathbf{r}_j), with kinetic energy terms for each particle of mass m_i, external potentials V_i, and interaction potentials V_{ij} between pairs. For non-interacting particles (V_{ij}=0), the Hamiltonian is separable. When interactions are present, such as Coulomb repulsion in atomic systems, the wave function generally cannot be separated into products of single-particle functions, leading to entangled states that reflect collective behavior. For identical particles, quantum statistics impose symmetry requirements on the wave function under particle exchange. Bosons, with integer spin, have symmetric wave functions satisfying \psi(\dots, \mathbf{r}_j, \dots, \mathbf{r}_k, \dots) = \psi(\dots, \mathbf{r}_k, \dots, \mathbf{r}_j, \dots), allowing arbitrary occupation of quantum states. Fermions, with half-integer spin like electrons, have antisymmetric wave functions \psi(\dots, \mathbf{r}_j, \dots, \mathbf{r}_k, \dots) = -\psi(\dots, \mathbf{r}_k, \dots, \mathbf{r}_j, \dots), enforcing the Pauli exclusion principle that prohibits multiple fermions from occupying the same state. A practical construction for fermionic wave functions is the , formed as the antisymmetrized product of single-particle orbitals \phi_\alpha(\mathbf{r}_i): \psi(\mathbf{r}_1, \dots, \mathbf{r}_N) = \frac{1}{\sqrt{N!}} \det \begin{vmatrix} \phi_1(\mathbf{r}_1) & \phi_1(\mathbf{r}_2) & \cdots & \phi_1(\mathbf{r}_N) \\ \phi_2(\mathbf{r}_1) & \phi_2(\mathbf{r}_2) & \cdots & \phi_2(\mathbf{r}_N) \\ \vdots & \vdots & \ddots & \vdots \\ \phi_N(\mathbf{r}_1) & \phi_N(\mathbf{r}_2) & \cdots & \phi_N(\mathbf{r}_N) \end{vmatrix}. This ensures antisymmetry and approximates the in methods like Hartree-Fock. An illustrative example is the ground state of the helium atom, modeled with two electrons. A simple variational trial wave function \psi(\mathbf{r}_1, \mathbf{r}_2) = \frac{Z'^3}{\pi a_0^3} e^{-Z' (r_1 + r_2)/a_0}, where Z' = 27/16 optimizes the effective nuclear charge, yields an energy of -77.5 eV, close to the exact -79.0 eV and capturing electron correlation approximately. More accurate Hylleraas functions incorporating the interelectron distance r_{12} improve this to within 0.001% of the exact value. Solving the many-particle faces severe computational challenges due to the curse of dimensionality: the configuration space volume grows exponentially as O((L/a)^{3N}) for spacing a and system size L, rendering exact infeasible for large N. Strong interactions exacerbate this, as fails and numerical methods like scale poorly. Relativistic extensions in first quantization, such as multi-particle Klein-Gordon equations for scalar particles, encounter fundamental limitations including densities and unstable vacuum states, necessitating for consistency.

References

  1. [1]
    A. First and Second Quantization - IuE
    The original form of quantum mechanics is denoted first quantization, while quantum field theory is formulated in the language of second quantization.
  2. [2]
  3. [3]
    First quantization | Nanoelectronics
    ### Summary of First Quantization Definition and Key Aspects
  4. [4]
    [PDF] Quantum Mechanics
    In first-quantization, we ask: Which particle is in which state? In second-quantization, we ask: How many particles are there in every state? The ...
  5. [5]
    Quantization (First, Second) - SpringerLink
    Jul 25, 2009 · H. D. Zeh: There Is no “first” quantization. Physics Letters A 309, 329–334 (2003). Article ADS MathSciNet Google Scholar · Download references ...
  6. [6]
    [PDF] Quantization: History and Problems - arXiv
    Feb 16, 2022 · On November 7 of 1925, however, Dirac's first paper on QM was received, “The Fundamental. Equations of Quantum Mechanics,” in which he proposed ...
  7. [7]
    [PDF] 11 Canonical quantization of classical fields
    In principle, quantization of classical fields is very similar to the quantization of the system of many coupled harmonic oscillators discussed in Section 5.6: ...
  8. [8]
    [PDF] 1 Second quantization
    First quantization is also the most useful formulation in many approaches that treat the full interacting many-body problem. In particular, the quantum Monte.
  9. [9]
    Quantum Field Theory - Stanford Encyclopedia of Philosophy
    Jun 22, 2006 · The term 'second quantization' has often been criticized partly because it blurs the important fact that the single particle wave function \(\ ...
  10. [10]
    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 ...
  11. [11]
    Quantenmechanik und Gruppentheorie | Zeitschrift für Physik A ...
    Download PDF · Zeitschrift für Physik. Quantenmechanik und ... Cite this article. Weyl, H. Quantenmechanik und Gruppentheorie. Z. Physik 46, 1–46 (1927).
  12. [12]
  13. [13]
    An Undulatory Theory of the Mechanics of Atoms and Molecules
    28, 1049 – Published 1 December, 1926. DOI: https://doi.org/10.1103/PhysRev.28.1049. Export Citation. Show metrics. Abstract. The paper gives an account of ...
  14. [14]
  15. [15]
    On the laws of radiation | Proceedings of the Royal ... - Journals
    (2002) Jeans' role in the law of black-body radiation, Physics Education, 10.1088/0031-9120/15/4/318, 15:4, (255-260), Online publication date: 1-Jul-1980.
  16. [16]
    [PDF] LXXIX. The scattering of α and β particles by matter and the structure ...
    It was shown that the number N of the electrons within the atom could be deduced from observations of the scattering of electrified particles. The accuracy of ...Missing: model | Show results with:model
  17. [17]
    [PDF] On the Law of Distribution of Energy in the Normal Spectrum
    where λ is the wavelength, E · dλ represents the volume density of the “black-body” radiation12 within the spectral region λ to λ + dλ, θ represents temperature ...
  18. [18]
    [PDF] Einstein's Proposal of the Photon Concept-a Translation
    Of the trio of famous papers that Albert Einstein sent to the Annalen der Physik in 1905 only the paper proposing the photon concept has been unavailable in ...
  19. [19]
  20. [20]
    [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 ...
  21. [21]
    [PDF] On quantum-theoretical reinterpretation of kinematic and ...
    (1925), 879-893. On the quantum-theoretical reinterpretation of kinematical and mechanical relationships. By W. Heisenberg in Göttingen. (Received on 29 July ...
  22. [22]
    [PDF] On Quantum Mechanics - Neo-classical physics
    34 (1925), 858-888. On Quantum Mechanics. By M. Born and P. Jordan in Göttingen. (Received on 27 September 1925).
  23. [23]
    [PDF] An undulatory theory of the mechanics of atoms and molecules - ISY
    SCHRÖDINGER. ABSTRACT. 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 ...Missing: citation | Show results with:citation
  24. [24]
    [PDF] Von Neumann's 1927 Trilogy on the Foundations of Quantum ... - arXiv
    May 20, 2025 · The object of the present article is to present new, fully annotated translations of these remarkable papers.1 We begin with some brief ...
  25. [25]
    [PDF] The Physical Interpretation of the Quantum Dynamics
    Dirac. in any direction determines the probability of the ,colliding electron (or other body) being scattered in that direetion. Recently Heisenberg ...
  26. [26]
    [PDF] Quantum theory of wave fields II - Neo-classical physics
    On the quantum theory of wave fields, II. By W. Heisenberg in Leipzig and W. Pauli in Zurich. (Received on 7 September 1929). Translated by D. H. ...
  27. [27]
    [PDF] Can Quantum-Mechanical Description of Physical Reality Be
    MAY 15, 1935. PHYSICAL REVIEW. VOLUME 47. Can Quantum-Mechanical Description of Physical Reality Be Considered Complete? A. EINSTEIN, B. PODOLSKY AND N. ROSEN, ...
  28. [28]
    Quantisierung als Eigenwertproblem - Wiley Online Library
    Quantisierung als Eigenwertproblem. E. Schrödinger,. E. Schrödinger. Zürich ... First published: 1926. https://doi.org/10.1002/andp.19263840404. Citations ...
  29. [29]
    [PDF] There is no “first” quantization - arXiv
    According to the present state of the art, the “second” quantization in terms of fields is the first and only one, while particles represent a derived and ...