Einstein coefficients
The Einstein coefficients are a set of three fundamental parameters in quantum mechanics and spectroscopy that describe the probabilities per unit time for the absorption, stimulated emission, and spontaneous emission of photons during transitions between two energy levels in atoms or molecules.[1] The coefficient B12 governs the rate of absorption from the lower energy state (level 1) to the upper state (level 2) under the influence of incident radiation, B21 describes the rate of stimulated emission from level 2 back to level 1 induced by the same radiation field, and A21 quantifies the rate of spontaneous emission from level 2 to level 1 in the absence of external stimulation.[2] These coefficients were introduced by Albert Einstein in his 1917 paper "Zur Quantentheorie der Strahlung" (On the Quantum Theory of Radiation), where he applied thermodynamic arguments to the equilibrium between matter and radiation to reconcile classical electromagnetic theory with quantum principles, particularly in explaining the Planck blackbody spectrum.[3] Einstein postulated that the stimulated emission process, previously unrecognized, must occur alongside absorption to maintain detailed balance in thermal equilibrium, leading to the prediction of what would later enable the development of lasers and masers.[1] The coefficients are interrelated through fundamental physical relations derived from statistical mechanics and quantum electrodynamics: specifically, B21 = (g1/g2) B12, where g1 and g2 are the statistical degeneracies of the respective energy levels, ensuring symmetry in the absorption and stimulated emission processes adjusted for level populations.[4] Additionally, the spontaneous emission coefficient is linked to the stimulated one by A21 = (8π h ν³ / c³) B21, where ν is the transition frequency, h is Planck's constant, and c is the speed of light; this relation connects atomic transition rates directly to the spectral energy density of blackbody radiation.[5] These relations have been experimentally verified and form the basis for calculating radiative lifetimes, oscillator strengths, and line intensities in spectroscopy across fields like astrophysics, plasma physics, and quantum optics.[2]Background on Atomic Transitions
Spectral Lines
Spectral lines are discrete wavelengths of light emitted or absorbed by atoms when electrons transition between quantized energy levels. These lines arise from the specific energy differences between atomic orbitals, producing sharp features in otherwise continuous spectra.[6] In the early 19th century, Joseph von Fraunhofer observed hundreds of dark absorption lines in the spectrum of sunlight, mapping their positions with high precision using improved prisms. These features, now known as Fraunhofer lines, were initially enigmatic but marked the beginning of systematic spectral analysis. Later, in 1859, Gustav Kirchhoff provided the key insight by demonstrating that these dark lines correspond to absorption by chemical elements in the Sun's cooler outer atmosphere, matching the bright emission lines produced when the same elements are heated in laboratory flames. This work established the atomic origin of spectral lines and laid the foundation for spectroscopy as a tool for elemental identification.[7][8] The underlying quantum model explains spectral lines through the relation between energy levels and photon frequencies. The energy difference \Delta E between two atomic levels determines the frequency \nu of the emitted or absorbed radiation via \Delta E = h \nu, where h is Planck's constant. This equation, rooted in the quantization of energy, predicts the precise wavelengths of lines for each transition in an atom.[9] Spectral lines manifest in two primary forms depending on the physical conditions. Emission lines appear as bright features against a dark background in the spectra of hot, low-density gases where excited atoms radiate photons directly. In contrast, absorption lines show as dark gaps in a continuous spectrum when light from a hot source passes through cooler, intervening gas that selectively absorbs photons at matching wavelengths. These distinctions follow from Kirchhoff's laws of spectroscopy, which describe the conditions for continuous, emission, and absorption spectra.[8] In spectroscopy, spectral lines act as unique signatures of atomic structure, enabling the identification of elements in distant stars, nebulae, and laboratory samples without direct contact. The wavelengths and relative strengths of lines reveal details about electron configurations and energy level spacings, providing insights into atomic composition and environmental conditions. The intensities of these lines are governed by the rates of atomic transitions quantified by Einstein coefficients.[10][11]Emission and Absorption Processes
In atomic physics, the interaction between matter and electromagnetic radiation occurs through three primary processes: absorption, spontaneous emission, and stimulated emission. These mechanisms describe how atoms transition between discrete energy levels, exchanging energy with photons of frequency \nu, and form the basis for understanding radiative transitions that produce spectral lines.[12] Absorption takes place when an atom in a lower energy state encounters a photon with energy h\nu matching the difference between two energy levels, leading to excitation to the higher state. This process is probabilistic, with the transition rate increasing with the radiation intensity at the resonant frequency, reflecting the atom's inherent susceptibility to the electromagnetic field.[13] Spontaneous emission involves an excited atom decaying randomly to a lower energy state, releasing a photon of energy h\nu in the process. Unlike absorption, this decay occurs without external prompting, driven solely by the instability of the excited state; the emitted photon's direction, phase, and polarization are random, resulting in incoherent light. The probability of spontaneous emission is characterized by a fixed transition rate for a given atomic system, independent of surrounding radiation.[12][14] Stimulated emission arises when an incoming photon of energy h\nu interacts with an atom in an excited state, prompting the atom to drop to the lower energy level while emitting a second photon identical to the first in frequency, phase, and propagation direction. This induced process amplifies the incident radiation coherently, as the two photons emerge in lockstep, contrasting with the randomness of spontaneous emission. The likelihood of stimulated emission mirrors that of absorption but applies to de-excitation, scaling with the photon density.[12][13] Classically, descriptions of radiation-matter interactions, such as those in the Rayleigh-Jeans law for blackbody spectra, treated energy as continuously distributed and waves as classical fields, leading to the ultraviolet catastrophe where predicted energy diverges to infinity at short wavelengths, contradicting observations. This failure highlighted the need for a quantum framework, where energy is quantized in discrete photon units, as pioneered by Planck and extended by Einstein to encompass discrete atomic transitions and photon-based absorption and emission.[15][14]Formal Definition of Einstein Coefficients
Spontaneous Emission Coefficient
The spontaneous emission coefficient, denoted as A_{21}, was introduced by Albert Einstein in 1917 as part of his foundational work on the quantum theory of radiation, where he analyzed the statistical equilibrium between matter and radiation to derive Planck's law.[16] In this context, A_{21} characterizes the probability of a molecule or atom transitioning from an upper energy state 2 to a lower state 1 by emitting a photon without external stimulation from the radiation field.[17] Formally, A_{21} is defined as the transition probability per unit time for spontaneous emission from the upper level 2 to the lower level 1.[18] Its units are s^{-1} (inverse seconds), reflecting its role as a rate constant.[17] Physically, A_{21} determines the natural lifetime of the excited state, given by \tau = 1 / A_{21}, which represents the average time an atom remains in the upper state before decaying via spontaneous emission, assuming no other decay channels.[18] This lifetime is crucial for understanding the duration of excited states in isolated atoms or dilute gases. In non-equilibrium conditions, such as in astrophysical plasmas or laboratory discharges, A_{21} governs the contribution of spontaneous emission to the overall intensity of spectral lines, where the emission rate is proportional to A_{21} times the population of the upper level.[17] This process complements stimulated emission but dominates in low-radiation-density environments.Stimulated Emission and Absorption Coefficients
The Einstein coefficients B_{21} and B_{12} quantify the rates of stimulated emission and absorption, respectively, for a two-level atomic system interacting with a radiation field. Introduced by Albert Einstein in his 1917 paper on the quantum theory of radiation, these coefficients represent the transition probability per unit time per unit spectral energy density for an atom in the upper state (level 2) to emit a photon or in the lower state (level 1) to absorb one, driven by the incident radiation.[16] The stimulated emission rate, expressed as the number of transitions per unit time per unit volume, is B_{21} \rho(\nu) N_2, where N_2 is the population density of atoms in the upper level, and \rho(\nu) is the energy density of the radiation per unit frequency interval at frequency \nu. Analogously, the absorption rate is B_{12} \rho(\nu) N_1, with N_1 the population density in the lower level. These rates vanish in the absence of radiation, distinguishing stimulated processes from spontaneous emission, which occurs independently of the external field.[16] In systems without level degeneracy, B_{12} = B_{21}, reflecting the symmetry between absorption and stimulated emission processes. For degenerate levels, the relation generalizes to g_1 B_{12} = g_2 B_{21}, where g_1 and g_2 are the degeneracies of the lower and upper levels.[16] The coefficients are typically defined using frequency \nu, as in Einstein's original formulation, with SI units of \mathrm{m}^3 \mathrm{J}^{-1} \mathrm{s}^{-2} when \rho(\nu) has units of \mathrm{J} \mathrm{m}^{-3} \mathrm{Hz}^{-1}; in cgs units, they are \mathrm{cm}^3 \mathrm{erg}^{-1} \mathrm{s}^{-2}. Equivalent cgs formulations exist for historical atomic physics contexts. In some modern treatments, particularly in quantum optics, angular frequency \omega = 2\pi \nu is employed, requiring B(\omega) = 2\pi B(\nu) to preserve the invariance of the product B \rho under the transformation \rho(\omega) = \rho(\nu) / 2\pi.[19][20]Thermodynamic Relations
Principle of Detailed Balancing
The principle of detailed balancing states that, in thermal equilibrium, the net rate of transitions between any two energy levels of an atomic system is zero, implying that the rate of upward transitions (absorption) precisely equals the rate of downward transitions (total emission). This condition ensures no net change in the population of the levels over time, reflecting the microscopic reversibility of the processes under thermodynamic equilibrium.[16] For a two-level atomic system, with level 1 as the lower energy state and level 2 as the upper, the principle manifests in the equality of transition rates involving the Einstein coefficients. The absorption rate from level 1 to 2 is given by N_1 B_{12} \rho(\nu), where N_1 is the population of level 1, B_{12} is the absorption coefficient, and \rho(\nu) is the spectral energy density of the radiation field at frequency \nu. The total emission rate from level 2 to 1 comprises spontaneous emission N_2 A_{21} and stimulated emission N_2 B_{21} \rho(\nu), where N_2 and A_{21} are the population and spontaneous emission coefficient for level 2, respectively. Thus, detailed balancing requires: N_1 B_{12} \rho(\nu) = N_2 \left[ A_{21} + B_{21} \rho(\nu) \right]. [21][16] In thermal equilibrium, the populations N_1 and N_2 follow the Boltzmann distribution, accounting for the degeneracies g_1 and g_2 of the levels: \frac{N_2}{N_1} = \frac{g_2}{g_1} \exp\left( -\frac{h\nu}{kT} \right), [16][21] where h is Planck's constant, k is the Boltzmann constant, and T is the temperature. Albert Einstein introduced this principle in his 1917 analysis to reconcile classical radiation theory with quantum hypotheses, demonstrating that equilibrium demands both spontaneous and stimulated emission processes; this argument provided key evidence for the quantization of electromagnetic radiation by linking atomic transitions to a discrete energy exchange of h\nu.[16] While the principle of detailed balancing generalizes to any reversible microscopic process in equilibrium—such as chemical reactions or particle collisions—its application here is confined to radiative transitions in atomic systems, where it constrains the interplay between matter and radiation. This condition establishes foundational relations among the Einstein coefficients without presupposing the form of the equilibrium radiation spectrum.[21][16]Equilibrium Conditions
In thermal equilibrium, the principle of detailed balancing ensures that the rates of upward and downward transitions between two energy levels are equal, leading to specific relations among the Einstein coefficients.[16] Consider two levels: the lower level 1 with energy E_1 and degeneracy g_1, and the upper level 2 with energy E_2 = E_1 + h\nu and degeneracy g_2. The population of the levels follows Maxwell-Boltzmann statistics, so N_2 / N_1 = (g_2 / g_1) \exp(-h\nu / kT), where k is Boltzmann's constant and T is the temperature.[16] Under isotropic radiation, the equilibrium condition balances absorption and emission rates: N_1 B_{12} \rho(\nu) = N_2 (B_{21} \rho(\nu) + A_{21}), where \rho(\nu) is the spectral energy density at frequency \nu.[16] In the limit of high temperature or large \rho(\nu), spontaneous emission becomes negligible compared to stimulated processes, yielding N_1 B_{12} = N_2 B_{21}. Substituting the Boltzmann population ratio gives g_1 B_{12} = g_2 B_{21}, or equivalently, B_{12} = (g_2 / g_1) B_{21}.[16] This adjustment accounts for degeneracy effects, ensuring the coefficients reflect the statistical weights of the levels when g_1 \neq g_2. For the full equilibrium including spontaneous emission, rearranging the balance equation produces \rho(\nu) = \frac{A_{21} / B_{21}}{\exp(h\nu / kT) - 1}.[16] To connect this to blackbody radiation, the form of \rho(\nu) matches Planck's law only if A_{21} / B_{21} = 8\pi h \nu^3 / c^3, where c is the speed of light; this relation is derived by imposing consistency with the known spectral energy density of thermal radiation.[16] The assumptions of isotropic radiation and Maxwell-Boltzmann population statistics underpin these derivations, treating the radiation field as uniform and the atomic ensembles as classical in their thermal distribution.[16] This framework verifies consistency with classical limits at low frequencies, where h\nu \ll [kT](/page/KT). The exponential approximates to $1 + h\nu / [kT](/page/KT), so \rho(\nu) \approx (A_{21} / B_{21}) ([kT](/page/KT) / h\nu). Substituting the A_{21}/B_{21} relation yields \rho(\nu) \approx (8\pi \nu^2 [kT](/page/KT) / c^3), recovering the Rayleigh-Jeans law for the ultraviolet catastrophe regime.[16]Related Physical Quantities
Oscillator Strengths
The oscillator strength f_{12}, for a transition from a lower energy level 1 to an upper level 2, is a dimensionless parameter that quantifies the intensity or probability of an atomic or molecular transition in the context of electromagnetic radiation absorption or emission. It originates from the classical model of a damped electron oscillator, where the electron is bound harmonically and driven by an oscillating electric field at the transition frequency \nu. In this analogy, f_{12} represents the effective number of classical electrons contributing to the radiation at that frequency, providing a bridge between classical radiation theory and quantum mechanical transition probabilities. Oscillator strengths typically range from $10^{-3} to 1 for allowed transitions, with values near unity indicating strong lines comparable to classical dipole radiation.[17] The oscillator strength is directly related to the Einstein coefficient for stimulated absorption B_{12} through the equation B_{12} = \frac{e^2}{4 \epsilon_0 m_e h \nu} \frac{g_2}{g_1} f_{12}, where m_e is the electron mass, e the elementary charge, \epsilon_0 the vacuum permittivity, h Planck's constant, \nu the transition frequency, and g_1, g_2 the statistical weights (degeneracies) of the lower and upper levels, respectively; this relation holds in SI units. This connection allows oscillator strengths to be computed from known B_{12} values or vice versa, facilitating the use of empirical data in quantum calculations. The formula underscores how f_{12} scales the quantum transition rate to match the classical absorption cross-section for a single electron.[17] A key theoretical constraint on oscillator strengths is the Thomas-Reiche-Kuhn (TRK) sum rule, which states that for a given initial state i (often the ground state), the sum of absorption oscillator strengths over all possible final states j equals the number of non-relativistic electrons Z available for excitation: \sum_j f_{ij} = Z. This rule, derived from the commutator of the dipole operator with the Hamiltonian, ensures conservation of the total transition strength and serves as a benchmark for the accuracy of computed or measured values across an atom's spectrum; for example, in hydrogen (Z=1), the sum is exactly 1. Violations or modifications occur in relativistic or multi-electron systems but remain approximately valid. Experimentally, oscillator strengths are determined from the relative intensities of spectral lines in absorption or emission spectra under controlled conditions, such as in optically thin plasmas or beams, where line strength is proportional to gf (with g the degeneracy factor). Alternatively, they can be inferred from radiative lifetimes \tau of excited states, since A_{21} = 1/\tau and A_{21} relates to f_{21} via the Einstein relations and the above formula for B. These methods are essential in astrophysics for interpreting stellar spectra and abundance determinations, as well as in plasma diagnostics to infer ion densities and temperatures from observed line ratios. For instance, precise f-values for iron lines enable modeling of solar photospheric conditions.[22]Dipole Approximation
The electric dipole approximation provides a quantum mechanical framework for computing the Einstein coefficients from transition matrix elements, applicable when the wavelength of the emitted or absorbed radiation is much longer than the spatial extent of the atomic or molecular system, typically on the order of 0.1 nm for atoms compared to visible light wavelengths of 400–700 nm.[23] In this regime, the interaction between the electromagnetic field and the system is dominated by the electric dipole (E1) term in the multipole expansion of the Hamiltonian, as higher-order terms like magnetic dipole or electric quadrupole contributions are suppressed by factors of (atomic size / wavelength).[17] This approximation simplifies the perturbation to H' = -\boldsymbol{\mu} \cdot \mathbf{E}, where \boldsymbol{\mu} = -e \sum_i \mathbf{r}_i is the electric dipole operator for electrons and \mathbf{E} is the electric field of the radiation.[23] The coefficients are derived using time-dependent perturbation theory, where the transition probability per unit time follows Fermi's golden rule:R_{i \to f} = \frac{2\pi}{\hbar} \left| \langle f | H' | i \rangle \right|^2 \delta(E_f - E_i - \hbar \omega),
with the density of final states accounted for in the continuum of photon modes.[23] For spontaneous emission, summing over photon polarizations and directions yields the Einstein A coefficient for a transition from upper state |2⟩ to lower state |1⟩ (assuming non-degenerate levels for simplicity):
A_{21} = \frac{\omega^3 |\mu_{21}|^2}{3 \pi \epsilon_0 \hbar c^3},
where \omega = (E_2 - E_1)/\hbar is the transition angular frequency and \mu_{21} = \langle 2 | -e \mathbf{r} | 1 \rangle is the dipole matrix element (in SI units, with |\mu_{21}|^2 often averaged over orientations).[17] For stimulated emission and absorption, the rate depends on the radiation energy density \rho(\omega), leading to the B coefficient:
B_{21} = \frac{\pi |\mu_{21}|^2}{3 \epsilon_0 \hbar^2}.
Degeneracies are incorporated via g_1 B_{12} = g_2 B_{21}, where g_i is the degeneracy of level i, and |\mu_{21}|^2 may include sums over magnetic sublevels.[17] This approach highlights the E1 transitions as the primary mechanism for allowed lines in spectra, with selection rules \Delta l = \pm 1 and \Delta m = 0, \pm 1 arising from the vector nature of the dipole operator.[23] However, the approximation fails for forbidden transitions where \mu_{21} = 0 due to parity or angular momentum conservation, requiring inclusion of weaker M1 or E2 terms that are smaller by factors of \sim 10^{-2} to $10^{-5}.[23] The dipole matrix element also connects to the classical oscillator strength f \propto |\mu_{21}|^2, bridging quantum and semiclassical descriptions.[17]