Fact-checked by Grok 2 weeks ago

Old quantum theory

Old quantum theory refers to the collection of theoretical developments in physics from roughly 1900 to 1925 that sought to resolve inconsistencies between and emerging experimental evidence for atomic-scale phenomena, by introducing ad hoc quantization rules for and other quantities while retaining much of . This transitional framework, often termed the "old" quantum theory to distinguish it from , originated with Max Planck's 1900 hypothesis that is emitted or absorbed in discrete packets () to explain , a later extended by in 1905 to itself as consisting of (photons) in his explanation of the . A pivotal advancement came in 1913 with Niels Bohr's model of the , which posited that electrons orbit the in stable, quantized "stationary states" with discrete energy levels, transitioning between them by absorbing or emitting specific quanta of energy, thereby accounting for the discrete lines observed in atomic emission spectra. extended this in 1916 by incorporating elliptical orbits and relativistic corrections, introducing action-angle variables and quantization conditions that successfully predicted the of lines, such as the splitting due to relativistic effects quantified by the \alpha \approx 1/137. These models relied on the correspondence principle, proposed by Bohr, which required quantum predictions to approach classical results in the limit of large quantum numbers, providing a bridge between the old and classical theories. Despite these successes—particularly in modeling the and explaining phenomena like the (electric field splitting of spectral lines) and (magnetic field splitting)—old quantum theory struggled with multi-electron atoms, molecular spectra, and the continuous energy distributions in scattering processes, revealing its inconsistencies and incompleteness by the early 1920s. Efforts to systematize it, such as through semiclassical approximations like the WKB method (named after Wentzel, Kramers, and Brillouin in 1926), highlighted the need for a more rigorous foundation. Ultimately, the framework paved the way for , with Werner Heisenberg's 1925 and Erwin Schrödinger's 1926 wave mechanics resolving its limitations by fully abandoning classical trajectories in favor of probabilistic wave functions and operators.

Historical Development

Origins in Classical Problems

In the late , the classical of encountered a profound crisis known as the . Experimental spectra of from a blackbody showed a peak intensity at a finite wavelength that shifted with , followed by a rapid decline at shorter wavelengths. , relying on the wave nature of light and the , led to the Rayleigh-Jeans law, which derives the spectral u(\nu, T) = \frac{8\pi \nu^2 k T}{c^3}, where \nu is the , T the , k Boltzmann's , and c the . This expression arises from treating the electromagnetic modes in the as classical oscillators, each contributing an kT independent of , resulting in the number of modes per per interval being \frac{8\pi \nu^2}{c^3}. At high frequencies, however, the law predicts a divergent approaching infinity in the ultraviolet region, grossly overestimating the observed radiation and implying an absurd infinite total for the . Another classical puzzle emerged in the measurement of specific heat capacities for solid elements. The Dulong-Petit law, established in 1819 through calorimetric experiments on various metals, asserted that the molar heat capacity at constant volume C_v is approximately $3R (where R is the gas constant, about 25 J/mol·K), reflecting the classical expectation that each atom vibrates as a three-dimensional harmonic oscillator with six quadratic degrees of freedom (three kinetic, three potential), each contributing \frac{1}{2}kT per atom via equipartition. This held well at room temperature for many solids. Yet, precise measurements at lower temperatures revealed systematic deviations: for instance, the heat capacity of diamond fell far below $3R even at moderate cooling, approaching zero as temperatures neared absolute zero, contradicting the temperature-independent classical prediction and suggesting that atomic vibrations do not behave as expected under classical equipartition. Observations of the further challenged . In 1887, noted that ultraviolet shining on a metal in a apparatus reduced the voltage required for sparking, indicating that the ejected charged particles from the surface, though he interpreted it as a effect. Subsequent experiments by in 1902 demonstrated that the kinetic energy of these photoelectrons was proportional to the of the incident above a material-specific threshold, while varying the light's intensity only affected the number of electrons emitted, not their individual energies. This frequency dependence defied classical expectations, where as a should impart energy proportional to intensity regardless of , akin to heating a surface. Ernest Rutherford's 1911 model of the atom, based on alpha-particle scattering experiments, depicted a tiny, dense positive at the center orbited by electrons to maintain charge neutrality, much like planets . Classical electrodynamics, however, rendered this configuration unstable: electrons in circular orbits experience centripetal acceleration and thus should radiate electromagnetic waves continuously, per Larmor's formula for accelerating charges, leading to rapid energy loss. The orbital radius would decrease, causing the electrons to spiral inward and collapse into the in roughly $10^{-8} seconds, incompatible with the observed of atoms and spectra.

Planck's Quantum Hypothesis

In 1900, Max Planck developed his quantum hypothesis as a means to reconcile theoretical predictions with experimental observations of spectra. Drawing on the concept of linear oscillators modeling the emitting and absorbing elements within the blackbody cavity, Planck posited that the associated with each oscillator of frequency \nu could only take on discrete values E = n h \nu, where n is a non-negative and h is a fundamental constant of nature, later termed Planck's constant, with a value of approximately $6.626 \times 10^{-34} joule seconds. This assumption marked a departure from , where was treated as continuous, and introduced the idea of energy quanta, or indivisible packets of proportional to frequency. To derive the distribution of energy among these oscillators in thermal equilibrium, Planck invoked Ludwig Boltzmann's 1877 statistical interpretation of , which relates S to the probability W of a system's state via S = k \ln W, where k is Boltzmann's . Assuming the oscillators exchange in discrete quanta \epsilon = h \nu, he calculated the most probable distribution by maximizing the , treating the number of ways to distribute p indistinguishable elements among N distinguishable oscillators as a combinatorial problem: W = \frac{(p + N - 1)!}{p! (N - 1)!}. This led to the average per oscillator \langle E \rangle = \frac{h \nu}{e^{h \nu / kT} - 1}, where T is the temperature, resolving the classical by ensuring that high-frequency modes receive vanishingly small at finite temperatures. Building on this average energy, Planck formulated the spectral energy density u(\nu, T) of the field. Considering the density of modes in the cavity, derived from electromagnetic wave theory as \frac{8 \pi \nu^2}{c^3} d\nu per unit volume for frequencies between \nu and \nu + d\nu (with c the ), he obtained the full expression: u(\nu, T) = \frac{8 \pi h \nu^3}{c^3} \frac{1}{e^{h \nu / kT} - 1}. This accurately matched experimental data across all frequencies, unlike the classical Rayleigh-Jeans approximation, which it recovers in the limit of low frequencies or high temperatures (h \nu \ll kT). Planck initially viewed his quantization as a mathematical expedient rather than a fundamental physical reality, expressing reluctance to abandon classical continuity and later describing it as "an act of despair" in a purely formal sense. He presented the hypothesis on December 14, 1900, to the and published a detailed account the following year, establishing the cornerstone of old quantum theory despite his own reservations about its implications.

Bohr's Model of the Atom

In 1913, developed a semi-classical model for the , building on Rutherford's 1911 discovery of the , which revealed that atoms consist of a dense, positively charged core surrounded by electrons. This nuclear model faced a classical instability problem, as orbiting electrons would continuously radiate electromagnetic energy and spiral into the nucleus, contradicting the observed stability of atoms. Bohr's approach addressed this by incorporating discrete quantum conditions, motivated in part by the empirical formula for hydrogen spectral lines discovered in 1885, which suggested a regular pattern in emission wavelengths. He presented his ideas in a trilogy of papers published in the that year, starting with "On the Constitution of Atoms and Molecules" in July. Bohr introduced two fundamental postulates to resolve the classical radiation dilemma. The first postulate defined stationary states: electrons occupy discrete, non-radiating orbits around the , each with a fixed , during which no is emitted despite the involved in . The second postulate concerned transitions between these states: an jumping from a higher- state to a lower one emits a with \nu such that the difference \Delta E satisfies \Delta E = h\nu, where h is Planck's ; absorption occurs analogously for upward transitions. These rules selectively quantized exchanges with radiation while allowing within orbits. To determine the allowed orbits, Bohr imposed a quantization condition on the angular momentum L of the electron-nucleus system: L = n \hbar, where n is a positive integer (the principal ) and \hbar = h / 2\pi. For the , modeled as a single of mass m_e and charge -e orbiting a fixed proton of charge +e, the centripetal force is provided by the Coulomb attraction: \frac{m_e v^2}{r} = \frac{e^2}{4\pi\epsilon_0 r^2}, where v is the orbital speed and r the radius. The angular momentum is L = m_e v r, so substituting the quantization gives m_e v r = n \hbar. Solving these equations yields the quantized radii r_n = n^2 a_0, where the Bohr radius a_0 = \frac{4\pi\epsilon_0 \hbar^2}{m_e e^2} \approx 0.529 \times 10^{-10} m defines the ground-state size (n=1). The total energy E in a combines kinetic and potential terms: E = \frac{1}{2} m_e v^2 - \frac{e^2}{4\pi\epsilon_0 r}. Substituting the expressions for v and r from the force balance and quantization leads to the discrete energy levels E_n = -\frac{13.6 \text{ [eV](/page/EV)}}{n^2} for n = 1, 2, \dots. The negative sign indicates bound states, with E_1 = -13.6 as the ; higher n approach zero from below. Transition frequencies then follow from \nu_{n'm} = \frac{|E_{n'} - E_m|}{h}, reproducing the (visible lines for transitions to n=2) and predicting unobserved (Lyman) and series, later confirmed experimentally. This model marked a pivotal departure from , successfully explaining hydrogen's while highlighting the need for further quantum refinements.

Extensions by Sommerfeld and Collaborators

In 1916, extended Niels Bohr's atomic model by incorporating elliptical electron orbits, motivated by discrepancies in that the circular orbit assumption could not explain. Using action-angle variables from , Sommerfeld quantized the radial and azimuthal motions separately, introducing two quantum numbers: the principal quantum number n for the total action and the k (now denoted as \ell) for the angular motion. This allowed electrons to occupy elliptical paths with varying eccentricity, where the energy depended on both n and k, reducing to Bohr's circular orbits in the limit k = n. The foundation of this extension was the Wilson-Sommerfeld quantization rule, proposed independently by William Wilson in 1915 and refined by Sommerfeld, which required the action integral for each degree of freedom to satisfy \oint p \, dq = n h, where h is Planck's constant and the integral is taken over a full cycle. For a two-dimensional orbit, this implied that the area enclosed in phase space was n h, or equivalently $2\pi n \hbar with \hbar = h / 2\pi. Sommerfeld's collaboration with Wilson emphasized applying this rule to atomic spectra, building on Wilson's earlier work on heat radiation and quantization during the mid-1910s. This approach provided a more general framework for multi-dimensional systems beyond Bohr's one-dimensional radial quantization. To account for relativistic effects, Sommerfeld incorporated special relativity into the electron's dynamics, deriving a fine structure correction to the energy levels. The resulting energy expression is approximately E_{n,k} \approx E_n \left[1 + \frac{\alpha^2}{n^2} \left(\frac{n}{k} - \frac{3}{4}\right)\right], where E_n is the non-relativistic Bohr energy and \alpha is the fine-structure constant, introducing the dimensionless parameter \alpha = e^2 / (4\pi \epsilon_0 \hbar c) \approx 1/137. This formula predicted the splitting of spectral lines into closely spaced components, explaining the observed fine structure in hydrogen spectra. Sommerfeld's work from 1916 to the 1920s, detailed in his influential book Atombau und Spektrallinien (1919), partially succeeded in ordering elements in the periodic table by assigning quantum numbers to electron shells, though it struggled with multi-electron atoms. Sommerfeld's model successfully interpreted the doublet structure in alkali metal spectra, attributing the splitting to transitions between states differing in the azimuthal quantum number k, such as k = n-1 and k = n-2, modulated by relativistic corrections. For X-ray spectra, the theory provided a theoretical underpinning for Henry Moseley's empirical law (1913–1914), where the frequency \nu of K-series lines scales as \sqrt{\nu} \propto (Z - b)^2, with Z the atomic number and b \approx 1 a screening constant; Sommerfeld derived this from quantized inner-shell transitions in multi-electron atoms. These extensions marked significant progress in the old quantum theory during the 1910s, bridging classical mechanics and emerging quantum ideas, though limitations emerged with complex spectra.

Fundamental Principles

Energy Quantization Rules

In old quantum theory, the foundational concept of energy quantization posited that physical systems possess discrete energy levels, known as stationary states, rather than a continuum of possible energies as in . This idea originated with Max Planck's 1900 hypothesis for , where he proposed that the energy of a could only take on discrete values E = n h \nu, with n an integer, h Planck's constant, and \nu the oscillator's frequency. Unlike , where energy could vary continuously, this quantization ensured stability and resolved the by limiting high-frequency energy contributions. Early formulations, including Planck's, did not incorporate , allowing the ground state to have zero energy in some models. Niels Bohr extended this principle to atomic systems in his 1913 model of the , introducing the postulate of stationary states where electrons occupy orbits with quantized energies E_n, and transitions between these states involve the emission or absorption of with \nu satisfying \Delta E = h \nu. Here, \Delta E represents the energy difference between two stationary states, and the \nu is directly tied to this discrete change, marking a departure from classical electrodynamics where accelerating charges would radiate continuously. This rule applied broadly in old quantum theory to bound systems, emphasizing that only specific energy values were permissible, leading to line spectra rather than continuous emission. For periodic motions in old quantum theory, an additional condition linked the quantized energies to the system's dynamics: the of the emitted or absorbed corresponds to the classical orbital in the of large quantum numbers, ensuring consistency with observed spectral lines. This "old quantum condition" was ad hoc, applied heuristically to various systems; for harmonic oscillators, energies were multiples of h \nu, while for in orbital motion, it took values of h / 2\pi. These rules distinguished old quantum theory from classical predictions by imposing discreteness without a unified framework, relying instead on system-specific postulates to match experimental data like spectra.

Phase Space Quantization

In old quantum theory, phase space quantization provided a systematic prescription for discretizing the states of classical mechanical systems by imposing conditions on integrals over trajectories. The core idea, known as the Wilson-Sommerfeld rule, required that the action variables—defined for periodic motions—take on discrete values multiples of the reduced Planck's constant. This approach extended Niels Bohr's earlier quantization for circular orbits to more general, separable systems with multiple . The action variable J for a one-dimensional periodic system is given by J = \frac{1}{2\pi} \oint p \, dq, where the integral is taken over one complete cycle of the coordinate q and its conjugate momentum p. The quantization condition then states that J = n \hbar, with n a positive integer and \hbar = h / 2\pi the reduced Planck's constant, or equivalently, \oint p \, dq = n h. For systems separable into multiple coordinates q_i, the rule generalizes to independent quantization of each action variable: J_i = n_i \hbar (or \oint p_i \, dq_i = n_i h) for integers n_i. This formulation ensured that allowed states corresponded to closed orbits in phase space enclosing areas that were integer multiples of h. Historically, the rule emerged independently in 1915 from William Wilson, who applied it to derive selection rules for spectral lines in his analysis of radiation and atomic spectra, and from Jun Ishiwara, who in his paper "Universelle Bedeutung des Wirkungsquantums" proposed a similar phase integral condition as a universal principle for quantizing periodic motions, including applications to the . provided a more rigorous foundation in , deriving the condition from the classical concept of adiabatic invariants—quantities like J that remain constant under slow, parameter-varying perturbations to the . 's work, building on earlier suggestions by regarding the invariance of action integrals, allowed the rule to be applied to elliptical orbits and relativistic corrections in atomic models. Ishiwara's contributions, though less emphasized in Western literature, paralleled these developments and highlighted the rule's generality for non-circular paths. Applications of the rule to simple systems illustrate its implications. For the classical , the trajectory is an with area $2\pi E / \omega, yielding J = E / \omega = n \hbar, or E = n h \nu where \nu = \omega / 2\pi is the ; this recovers Max Planck's quantization for oscillators. For a rigid rotator, the azimuthal is J_\phi = \frac{1}{2\pi} \oint p_\phi \, d\phi = L, the , quantized as L = n \hbar. In bound systems, these conditions naturally imply discrete levels as a consequence.

Correspondence Principle

The correspondence principle, introduced by in 1923, asserts that for sufficiently large quantum numbers, the frequencies and intensities of quantum transitions coincide with the components of the classical radiation emitted by the corresponding periodic motion. This asymptotic agreement ensures that recovers classical electrodynamics in regimes where ħ becomes negligible relative to the action scales involved. In the context of old quantum theory, the principle provided a guiding to extend initial postulates, such as those in Bohr's atomic model, by linking discrete quantum jumps to continuous classical oscillations. Bohr formulated the correspondence principle to address inconsistencies and ambiguities in applying quantization rules to complex systems, where multiple possible transitions could occur without clear predictions for their likelihood or spectral positions. By requiring quantum results to match classical limits for high quantum numbers n ≫ 1, it offered a for selecting valid theoretical extensions, such as in the Bohr-Sommerfeld quantization of non-separable Hamiltonians. For large n, the quantized areas encompass extensive classical orbits, enabling direct comparison of quantum stationary states with classical periodic trajectories. A primary application of the principle concerns selection rules for allowed transitions. In the , the electric from an orbiting charge dominates via its , which corresponds to quantum transitions satisfying Δn = ±1 between adjacent states with large n. This arises because higher multipole or contributions become negligible, restricting quantum jumps to those mimicking the primary classical . The principle also guided the prediction of transition intensities through a semiclassical approach. Quantum probabilities for allowed transitions were taken proportional to the squares of the classical Fourier coefficients associated with the dipole moment's time variation in the corresponding classical orbit. This method allowed quantitative estimates of relative line strengths in spectra, bridging the gap between discrete quantum emissions and the continuous classical power spectrum without invoking full wave mechanics.

Applications to Simple Systems

Harmonic Oscillator

In old quantum theory, the was treated by quantizing its energy levels as discrete multiples of the fundamental quantum E_n = n h \nu, where n = 0, 1, 2, \dots, h is Planck's constant, and \nu is the classical frequency of oscillation. This quantization rule followed directly from the adiabatic invariance of \oint p \, dq = n h, which for the yields evenly spaced levels without a . Einstein applied this model to the vibrational modes of atoms in solids, viewing each as an independent . To compute thermal properties, the average energy \langle E \rangle was obtained using Boltzmann statistics over the discrete levels: \langle E \rangle = \frac{\sum_{n=0}^{\infty} E_n e^{-E_n / kT}}{\sum_{n=0}^{\infty} e^{-E_n / kT}} = \frac{h \nu}{e^{h \nu / kT} - 1}, where k is Boltzmann's constant and T is temperature. In the high-temperature limit where h \nu \ll kT, this approximates to \langle E \rangle \approx kT, recovering the classical result from equipartition, which assigns kT/2 to each quadratic term in the energy (kinetic and potential, totaling kT per one-dimensional oscillator). This quantization resolved discrepancies in the specific heat of solids. Classically, the at constant volume C_V for a solid with N atoms is $3Nk (Dulong-Petit law), independent of , from $3N oscillators each contributing k. However, experiments showed C_V decreasing at low temperatures. Einstein's 1907 model assumed all $3N oscillators share a single \nu, yielding C_V = 3Nk \left( \frac{h \nu}{kT} \right)^2 \frac{e^{h \nu / kT}}{\left( e^{h \nu / kT} - 1 \right)^2}. This formula matches the classical $3Nk at high T but approaches zero as T \to 0, explaining the observed low-temperature behavior.

Free Particle in One Dimension

In old quantum theory, the free particle in one dimension represents a straightforward case of bounded motion, where a particle of mass m is confined between impenetrable walls at x = 0 and x = L, with zero potential energy U = 0 inside the interval. Classically, the particle travels back and forth with constant momentum magnitude p = \sqrt{2mE}, reversing direction upon hitting the walls. The Bohr-Sommerfeld quantization rule, which generalizes Bohr's original condition to arbitrary periodic orbits, requires that the action integral over one complete cycle equals an integer multiple of Planck's constant: \oint p \, dx = n h, where n = 1, 2, 3, \dots and h is Planck's constant. For this system, the integral evaluates to $2 p L = n h, since the momentum is constant in magnitude over the round trip distance $2L. Solving for the momentum yields p = \frac{n h}{2 L}, and substituting into the classical energy-momentum relation gives the discrete energy levels E_n = \frac{n^2 h^2}{8 m L^2}. This result arises directly from the area enclosed by the classical trajectory being n times the fundamental area h, consistent with the semiclassical approach of old quantum theory. Notably, the energy E_1 = \frac{h^2}{8 m L^2} > 0, implying a that forbids the classical minimum of zero, where the particle would be at rest. This feature highlights a key departure from , as the quantization enforces standing-wave-like conditions on the particle's motion despite the absence of an explicit wave picture in the theory at this stage. The model found application in early attempts to understand electron behavior in solids, such as treating conduction electrons as free particles confined within the dimensions of a metal, providing a quantized basis for specific heat and electrical conductivity estimates. In one dimension, it illustrates boundary-induced quantization for confined systems more broadly, contrasting with unbounded free motion where no discrete levels exist.

Linear Potential

In the old quantum theory, the linear potential U(x) = F x (with F > 0) models the behavior of a particle under a constant force, such as in a uniform gravitational or electric field. Classically, for a particle of mass m and total energy E, the motion is confined between turning points x_1 = 0 (assuming a hard reflecting wall for bounded periodic motion) and x_2 = E/F, where the kinetic energy vanishes at the upper turning point. The momentum is p(x) = \sqrt{2m (E - F x)}, and the trajectory is parabolic in time due to the constant acceleration F/m, with the particle accelerating toward x = 0 and reflecting instantaneously at the wall. The quantization of energy levels in this potential followed the Bohr-Sommerfeld rule, which posits that the action integral over the full classical orbit equals an integer multiple of Planck's constant h. In the original formulation of the old quantum theory, this condition is $2 \int_{x_1}^{x_2} \sqrt{2m (E - F x)} \, dx = n h, where n = 1, 2, 3, \dots (later approximations in semiclassical methods adjust this to (n + 1/2) h). Evaluating the integral gives \frac{4}{3} \frac{\sqrt{2m} \, E^{3/2}}{F} = n h, leading to the energy levels E_n \approx \left[ \frac{(n h F)^2}{2m} \right]^{1/3} (up to numerical prefactors of order unity). This semiclassical spectrum captures the essential n^{2/3} scaling and approximates the exact quantum mechanical solutions, where the wave functions are Airy functions and the energies correspond to the zeros of the Airy function \mathrm{Ai}(z) = 0, with z \propto - (2m F / \hbar^2)^{1/3} E. A key application of this quantization arose in early treatments of the , where a is subjected to an external , introducing a linear -e F z to the Coulomb potential. In 1916, and Paul Epstein separately applied the Bohr-Sommerfeld conditions in , separating the into two independent equations with effective linear potentials in the variables \xi and \eta. The resulting quantized action integrals \int p_\xi \, d\xi = n_\xi h and \int p_\eta \, d\eta = n_\eta h (along with the angular part n_\phi h) yielded energy shifts linear in the field strength F, \Delta E \propto F n (n_\xi - n_\eta), successfully explaining the observed splitting of Balmer spectral lines. This approach provided seminal sketches for the Stark effect in Rydberg atoms, highlighting how high-principal-quantum-number states exhibit Airy-like level spacings under strong fields, though it struggled with selection rules and intensities.

Rigid Rotator

In the rigid rotator model, the rotational dynamics of a are approximated by treating the nuclei as point masses separated by a fixed , yielding a I. Classically, the is given by \mathbf{L} = I \boldsymbol{\omega}, where \boldsymbol{\omega} is the vector, and the associated rotational is E = \frac{L^2}{2I}. Within old quantum theory, quantization of the rigid rotator was achieved by applying the Bohr-Sommerfeld phase integral condition to the angular coordinates, quantizing the azimuthal component of as m \hbar where m = 0, \pm 1, \pm 2, \dots, leading to discrete levels E_m = \frac{m^2 \hbar^2}{2I} (degenerate for \pm m \neq 0). For large |m|, this approximates the quantum form \frac{l(l+1) \hbar^2}{2I} with l \approx |m|. Transitions between these levels obey the selection rule \Delta m = \pm 1, derived from the conservation of during the interaction with and aligned with Bohr's . At sufficiently high temperatures, where kT \gg \frac{\hbar^2}{2I}, the quantized rigid rotator model predicts a rotational contribution to the specific heat of diatomic gases that approaches the classical equipartition value of k per (or R per for two rotational ), successfully accounting for experimental observations of molecular .

Atomic and Molecular Models

Bohr-Sommerfeld Hydrogen Atom

The Bohr-Sommerfeld model extended the original Bohr atomic model by incorporating elliptical orbits for the in a , applying the general quantization rules of old quantum theory to both radial and angular motions. In this framework, the is quantized as L = k \hbar, where k is the (an integer) and \hbar = h / 2\pi with h being Planck's constant. The radial motion is quantized via the action integral \oint p_r \, dr = n_r h, where n_r is the radial (a non-negative integer) and p_r is the radial momentum. The total quantum number is then n = n_r + k, allowing for a range of elliptical orbits within each labeled by n. In the non-relativistic limit, the energy levels derived from these quantization conditions for the hydrogen atom are identical to those in Bohr's original circular orbit model, given by E_n = -\frac{m e^4}{2 \hbar^2 n^2}, where m is the electron mass and e is the elementary charge; this independence from k implies degeneracy among the elliptical orbits within a given n. Sommerfeld's analysis solved the classical equations of motion for the Coulomb potential under these constraints, yielding the same binding energies as Bohr's 1913 derivation but accommodating more general trajectories. Bohr's circular orbits correspond to the special case where k = n and n_r = 0. To account for observed spectral fine structure, Sommerfeld introduced relativistic corrections to the electron's motion, modifying the quantization conditions and lifting the degeneracy by introducing a dependence on k. The corrected energy levels become E_{n,k} = E_n \left[ 1 + \frac{\alpha^2}{n^2} \left( \frac{n}{k} - \frac{3}{4} \right) \right], where \alpha = e^2 / (4\pi \epsilon_0 \hbar c) \approx 1/137 is the (with \epsilon_0 the and c the ); this splitting explains the multiplet structure in hydrogen spectral lines. The model successfully predicts the frequencies of , such as the Balmer (visible) and Lyman () series, through transitions between levels: \nu = R \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right), where R = \frac{m e^4}{8 \epsilon_0^2 h^3 c} is the (approximately $1.097 \times 10^7 m^{-1}) and n_1 < n_2 are principal quantum numbers. This formula matches experimental observations quantitatively, with the fine-structure corrections accounting for small deviations in line positions.

Multi-Electron Atoms

In the old quantum theory, the treatment of multi-electron atoms extended the by approximating the potential experienced by valence electrons as a screened , where inner electrons partially shield the nuclear charge, leading to an effective potential of the form V(r) = -\frac{Z_{\text{eff}} e^2}{r}, with Z_{\text{eff}} = Z - \sigma and \sigma as an empirical screening constant determined from spectral data. This approach, introduced by , treated inner electrons as a core that reduced the effective nuclear charge seen by outer electrons, allowing approximate quantization rules to be applied similar to the , though exact solutions were intractable due to electron-electron repulsions. Electrons in multi-electron atoms were assigned quantum numbers following the old quantum framework, with the principal quantum number n = 1, 2, 3, \dots defining energy shells and the azimuthal quantum number k = 1, 2, \dots, n (corresponding to angular momentum quantum number l = k - 1, up to n-1) specifying subshells within each shell. Bohr proposed that shells fill according to stability criteria, with a maximum of $2n^2 electrons per shell, but adjusted empirically to match observed capacities like 2 for n=1, 8 for n=2, and 18 for n=3, reflecting elliptical orbits and penetration effects where outer electrons could interact more closely with the nucleus. This shell-filling scheme provided a partial rationale for the periodic table, attributing chemical periodicity to the completion of stable closed shells that minimized energy and dictated valence through the number of electrons in the outermost subshell, though the rules remained ad hoc, relying on empirical correlations with atomic spectra and chemical properties rather than rigorous derivation. For instance, alkali metals were modeled with a single valence electron in a higher n shell screened to behave like hydrogen, explaining their similar spectra. In the presence of a weak external magnetic field, the Zeeman effect in multi-electron atoms was described by the interaction of the orbital magnetic moment with the field, yielding an energy shift \Delta E = \mu_B B m_l, where \mu_B = \frac{e \hbar}{2m_e} is the , B is the field strength, and m_l is the magnetic quantum number (-l, \dots, l), assuming a Landé g-factor of 1 for pure orbital motion without spin considerations. This normal Zeeman splitting applied primarily to the valence electron's orbit, consistent with in the old quantum picture, though it failed to account for anomalous splittings observed in complex atoms.

Molecular Rotations and Vibrations

In old quantum theory, the vibrations of diatomic molecules were treated as motion in an anharmonic potential well, approximated by forms such as the Kratzer potential or later the Morse potential, leading to quantized energy levels roughly given by E_v = (v + \frac{1}{2}) h \nu, where v is the vibrational quantum number and \nu is the fundamental frequency; however, early applications often ignored the zero-point energy term \frac{1}{2} h \nu. The anharmonicity introduced deviations from equal spacing in vibrational levels, enabling the observation of overtones (\Delta v > 1) and providing a limit to the number of bound states near the dissociation energy. Rotational motion was quantized separately using the Bohr-Sommerfeld condition for angular momentum, yielding energy levels E_J = \frac{J^2 h^2}{8 \pi^2 I}, where J is an integer rotational quantum number and I is the moment of inertia about the center of mass. In the coupled rotation-vibration model, the vibrational motion modulates the internuclear distance, making the effective moment of inertia amplitude-dependent and thus introducing a coupling term; this results in a vibrational-state-dependent rotational constant B_v = B_e - \alpha (v + \frac{1}{2}), where B_e = \frac{h}{8 \pi^2 c I_e} is the equilibrium rotational constant, c is the speed of light, and \alpha reflects the centrifugal stretching and vibrational averaging effects. The total energy for a rotating is then approximately E_{v,J} = E_v + B_v J(J+1), though old theory used the J^2 form for ; for low vibrational , this approaches the rigid rotator levels in the v = 0 limit. Transitions between these levels, governed by electric dipole selection rules derived via the correspondence principle, produce structured band spectra: the P branch (\Delta J = -1) appears at lower frequencies, the R branch (\Delta J = +1) at higher frequencies, and the Q branch (\Delta J = 0) in the band center for certain cases like parallel transitions. These models successfully predicted the fine rotational structure within vibrational bands observed in infrared spectra of molecules like HCl and CO, as well as ultraviolet electronic-vibrational bands, offering early quantitative interpretations of molecular spectra and validating the quantization of non-electronic degrees of freedom.

Bridge to Wave Mechanics

De Broglie's Hypothesis

In 1923, Louis de Broglie first proposed his hypothesis of wave-particle duality for matter in three short articles published in the Comptes Rendus de l'Académie des Sciences, with the full development presented in his 1924 doctoral thesis Recherches sur la théorie des quanta that all particles of matter exhibit wave-particle duality, mirroring the dual nature already attributed to light through Einstein's concept of light quanta. Drawing inspiration from special relativity and his brother Maurice de Broglie's experimental work on X-ray interference, de Broglie argued that a moving particle is inherently linked to a periodic phase wave propagating through space-time. This hypothesis aimed to unify the corpuscular and undulatory aspects of physical entities, providing a foundational step beyond the ad hoc quantization rules of old quantum theory. The central element of de Broglie's hypothesis is the relation between the p of a particle and the \lambda of its associated , given by \lambda = \frac{h}{p}, where h is Planck's ; equivalently, the can be expressed as p = h / \lambda. De Broglie derived this by considering the invariance of in relativistic transformations, positing that the particle's corresponds to the wave's . For massive particles like electrons, this becomes significant at scales, where p is small enough for \lambda to be comparable to interatomic distances. Applying the hypothesis to the Bohr atomic model, de Broglie interpreted the quantized orbits as stationary states where the electron forms a around the . The condition for stability requires the orbit's circumference to equal an integer multiple of the de Broglie wavelength: $2\pi r_n = n \lambda, with n as the principal quantum number and r_n the orbital radius. Substituting \lambda = h / (m v), where m is the and v its speed, yields the Bohr quantization m v r_n = n \hbar, with \hbar = h / 2\pi, thus deriving the old theory's quantization rule from rather than postulate. This picture explained the discrete stability of orbits without invoking arbitrary assumptions. De Broglie's matter waves also accounted for scattering phenomena, extending the wave-particle duality to explain effects like , where the shift in wavelength during photon-electron collisions implies a corpuscular interaction consistent with associated waves for both entities. The hypothesis predicted diffraction of electron beams, dramatically verified by the 1927 Davisson-Germer experiment, in which low-energy scattered from a produced patterns matching de Broglie wavelengths, confirming wave-like for particles. These insights bridged old quantum theory toward a fully wave-based by suggesting underlying wave dynamics for all quantized systems.

Kramers' Transition Matrix

In 1924, Hendrik Kramers, working as an assistant at Niels Bohr's Institute for Theoretical Physics in Copenhagen, developed a method to compute transition probabilities within the framework of old quantum theory, building on discussions at the institute regarding the correspondence principle. This approach represented a significant step toward formalizing quantum transitions without full wave mechanics, predating Heisenberg's matrix mechanics by introducing matrix-like elements for probabilities. Kramers' work originated from his earlier studies in Leiden under Paul Ehrenfest but matured through collaboration with Bohr, culminating in key publications that year. Kramers proposed that the semiclassical transition amplitude between quantum states could be obtained from the Fourier components of the corresponding classical motion, aligning with Bohr's to bridge classical and quantum descriptions. In this method, the matrix elements representing transitions, such as those for the , were evaluated as the Fourier coefficients of the classical coordinate's periodic orbit, ensuring that high-quantum-number behaviors matched classical patterns. This decomposition allowed for a systematic of probabilities for virtual and real transitions, treating the atom as a collection of oscillators tied to allowed energy levels. A central result was Kramers' quantum dispersion formula, which expressed the atomic polarizability \alpha(\omega) in terms of transition matrix elements: \alpha(\omega) = \sum_j \frac{|\langle i | \mathbf{d} | j \rangle|^2}{E_j - E_i - \hbar \omega} + \sum_j \frac{|\langle i | \mathbf{d} | j \rangle|^2}{E_j - E_i + \hbar \omega}, where \langle i | \mathbf{d} | j \rangle denotes the dipole matrix element between states i and j, E_j - E_i is the energy difference, and the sums run over virtual transitions. This formula generalized the classical Lorentz dispersion relation to quantum theory, incorporating absorption lines at Bohr frequencies while using correspondence-derived matrix elements for off-resonant terms. Kramers applied this matrix method to compute scattering and absorption intensities, such as in and line strengths, by evaluating elements via the classical prescription for systems like the . For , the transition probability was proportional to the squared element times a density-of-states factor, enabling predictions of spectral intensities that matched experimental dispersion data. These calculations, later refined in the Bohr-Kramers-Slater theory, highlighted the role of probabilistic s in processes without invoking for virtual photons.

Limitations and Transition

Inconsistencies and Failures

One major shortcoming of the old quantum theory was its inability to quantitatively predict transition probabilities between quantized states, limiting it to qualitative selection rules for lines. For instance, while the theory successfully reproduced the frequencies of lines, it provided no mechanism to calculate the relative intensities or brightness of those lines, relying instead on assumptions. This failure was particularly evident in the spectrum, where Bohr's initial calculation of the ground-state energy was accurate to within 5% of experiment, but subsequent refinements by Sommerfeld and others, including and Pauli's 1922 variational approach, yielded discrepancies of up to 20% or more due to inadequate treatment of interactions and oscillations. The theory's adherence to definite, classical-like orbits for electrons violated the principles of superposition inherent in classical wave descriptions, precluding any between multiple possible paths or orbits in atomic transitions. In classical electrodynamics, radiation from accelerated charges could interfere constructively or destructively depending on phase relations, but the old quantum theory treated orbits as localized and non-overlapping, neglecting the wave nature of particles and thus failing to predict effects in processes like or multi-electron dynamics. This limitation stemmed from the absence of a coherent framework for combining amplitudes from different orbital configurations, leading to inconsistencies when attempting to match classical patterns at high quantum numbers via the correspondence principle. For systems involving identical particles, such as multi-electron atoms, the old quantum theory lacked any principle enforcing exclusion, resulting in erroneous ground-state predictions. In the , early models allowed both s to occupy the lowest 1s orbital simultaneously without regard for indistinguishability, overestimating the and failing to explain the observed and triplet states or the para- and ortho-helium spectra. This oversight arose from neglecting and the need for antisymmetric wave functions, leading to configurations that violated empirical spectral data and chemical properties. Relativistic extensions of the old quantum theory, such as Sommerfeld's incorporation of into the for , encountered deeper inconsistencies beyond hydrogen's spectral splitting. In multi-electron systems and molecular ions like H₂⁺, relativistic corrections predicted unstable ground states—for example, Pauli's calculation yielded -0.52 Rydberg units against the experimental -1.20 Rydberg units—due to unaccounted electron-nucleus collisions and improper handling of quantization (l ≠ 0 exclusion). These issues foreshadowed paradoxes in particle through strong potentials, akin to precursors of the , where states and probability currents defied classical intuitions without a full quantum field description.

Path to Modern Quantum Theory

The pivotal years of 1925 and 1926 marked the culmination of efforts to resolve the quantization rules of old quantum theory, leading to the formulation of through two complementary approaches: wave mechanics and . Building on Louis de Broglie's 1924 hypothesis of matter waves, developed a in late 1925 that described particles as wave functions evolving in space and time. This framework provided an exact solution for the , yielding energy levels and wave functions \psi_{nlm} that matched Bohr's model while introducing quantum numbers l and m. The Compton effect, observed by in 1923, had already bolstered the particle nature of by demonstrating photon-electron collisions with transfer, confirming the quantum of action in experiments and paving the way for unified wave-particle descriptions. Independently, introduced in 1925, representing physical observables as infinite arrays (matrices) whose elements corresponded to transition probabilities between quantum states, inspired by the correspondence principle and dispersion relations from Hendrik Kramers. and formalized this in their 1925 paper, establishing the non-commutativity of position x and momentum p via the relation [x, p] = i \hbar, which directly incorporated Planck's constant and resolved inconsistencies in classical analogies for atomic spectra. This algebraic approach successfully reproduced the hydrogen spectrum without explicit orbits, emphasizing observables over trajectories. The old theory's quantization rules served as a conceptual starting point for both formulations, guiding the search for consistent dynamical laws. Schrödinger's 1926 proof demonstrated the mathematical equivalence of wave and , showing that the eigenvalues of the operator in wave mechanics corresponded to the energy levels from matrix calculations, and that matrix elements could be derived from integrals over wave functions. This unification resolved the apparent rivalry between the intuitive wave picture and the abstract matrix algebra, establishing a single framework for that extended beyond the limitations of semiclassical models. The rapid succession of these developments during 1925–1926, often termed the birth of modern quantum mechanics, transformed physics by providing a predictive theory for atomic phenomena.