Fact-checked by Grok 2 weeks ago

Einstein coefficients

The Einstein coefficients are a set of three fundamental parameters in and that describe the probabilities per unit time for the , , and of photons during transitions between two levels in atoms or molecules. The coefficient B12 governs the rate of from the lower (level 1) to the upper (level 2) under the influence of incident radiation, B21 describes the rate of from level 2 back to level 1 induced by the same radiation field, and A21 quantifies the rate of from level 2 to level 1 in the absence of external stimulation. 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. 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. 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. 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. 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.

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. In the early , 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 , were initially enigmatic but marked the beginning of systematic . Later, in 1859, 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 as a tool for elemental identification. 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. 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 , which describe the conditions for continuous, , and spectra. In , spectral lines act as unique signatures of structure, enabling the identification of in distant , nebulae, and samples without direct contact. The wavelengths and relative strengths of lines reveal details about configurations and spacings, providing insights into composition and environmental conditions. The intensities of these lines are governed by the rates of transitions quantified by Einstein coefficients.

Emission and Absorption Processes

In , the interaction between matter and occurs through three primary processes: , , and . 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. Absorption takes place when an atom in a lower state encounters a with h\nu matching the between two levels, leading to to the higher state. This process is probabilistic, with the transition rate increasing with the intensity at the resonant , reflecting the atom's inherent to the . Spontaneous emission involves an excited atom decaying randomly to a lower state, releasing a of h\nu in the process. Unlike , this occurs without external prompting, driven solely by the instability of the ; the emitted photon's direction, phase, and are random, resulting in incoherent . The probability of spontaneous emission is characterized by a fixed transition rate for a given atomic system, independent of surrounding . Stimulated emission arises when an incoming of energy h\nu interacts with an in an , prompting the to drop to the lower while emitting a second identical to the first in , , and propagation direction. This induced process amplifies the incident radiation coherently, as the two photons emerge in lockstep, contrasting with the randomness of . The likelihood of stimulated emission mirrors that of but applies to de-excitation, scaling with the density. Classically, descriptions of radiation-matter interactions, such as those in the Rayleigh-Jeans law for blackbody spectra, treated as continuously distributed and waves as classical fields, leading to the where predicted diverges to infinity at short wavelengths, contradicting observations. This failure highlighted the need for a quantum , where is quantized in units, as pioneered by Planck and extended by Einstein to encompass transitions and -based and .

Formal Definition of Einstein Coefficients

Spontaneous Emission Coefficient

The spontaneous emission coefficient, denoted as A_{21}, was introduced by in 1917 as part of his foundational work on the of , where he analyzed the statistical equilibrium between matter and to derive . In this context, A_{21} characterizes the probability of a or transitioning from an upper energy state 2 to a lower state 1 by emitting a photon without external stimulation from the field. 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. Its units are s^{-1} (inverse seconds), reflecting its role as a rate constant. Physically, A_{21} determines the natural lifetime of the , given by \tau = 1 / A_{21}, which represents the average time an atom remains in the upper state before decaying via , assuming no other decay channels. This lifetime is crucial for understanding the duration of in isolated atoms or dilute gases. In non-equilibrium conditions, such as in astrophysical plasmas or laboratory discharges, A_{21} governs the contribution of to the overall intensity of spectral lines, where the emission rate is proportional to A_{21} times the population of the upper level. This process complements but dominates in low-radiation-density environments.

Stimulated Emission and Absorption Coefficients

The Einstein coefficients B_{21} and B_{12} quantify the rates of and , respectively, for a two-level interacting with a field. Introduced by in his 1917 paper on the quantum theory of , these coefficients represent the transition probability per unit time per unit for an atom in the upper state (level 2) to emit a or in the lower state (level 1) to absorb one, driven by the incident . 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 of atoms in the upper level, and \rho(\nu) is the of the per unit interval at \nu. Analogously, the absorption rate is B_{12} \rho(\nu) N_1, with N_1 the in the lower level. These rates vanish in the absence of , distinguishing stimulated processes from , which occurs independently of the external field. In systems without level degeneracy, B_{12} = B_{21}, reflecting the between and 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. The coefficients are typically defined using frequency \nu, as in Einstein's original , 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 contexts. In some modern treatments, particularly in , 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.

Thermodynamic Relations

Principle of Detailed Balancing

The principle of detailed balancing states that, in , the net rate of transitions between any two energy levels of an atomic system is zero, implying that the rate of upward transitions () precisely equals the rate of downward transitions (total ). This condition ensures no net change in the population of the levels over time, reflecting the of the processes under . For a two-level atomic system, with level 1 as the lower energy state and level 2 as the upper, 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 of level 1, B_{12} is the coefficient, and \rho(\nu) is the spectral energy density of the radiation field at \nu. The total emission rate from level 2 to 1 comprises N_2 A_{21} and N_2 B_{21} \rho(\nu), where N_2 and A_{21} are the and 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]. In thermal equilibrium, the populations N_1 and N_2 follow the , 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), where h is Planck's constant, k is the , and T is the . Albert 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 by linking atomic transitions to a discrete energy exchange of h\nu. 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.

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. Consider two levels: the lower level 1 with E_1 and degeneracy g_1, and the upper level 2 with 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. Under isotropic radiation, the equilibrium condition balances and emission rates: N_1 B_{12} \rho(\nu) = N_2 (B_{21} \rho(\nu) + A_{21}), where \rho(\nu) is the spectral density at \nu. In the limit of high 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}. 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 , rearranging the balance equation produces \rho(\nu) = \frac{A_{21} / B_{21}}{\exp(h\nu / kT) - 1}. 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 ; this relation is derived by imposing consistency with the known spectral energy density of thermal radiation. 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. 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 regime.

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. The is directly related to the Einstein coefficient for stimulated 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 , e the , \epsilon_0 the , h Planck's constant, \nu the frequency, and g_1, g_2 the statistical weights (degeneracies) of the lower and upper levels, respectively; this relation holds in 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 rate to match the classical cross-section for a single . 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 ), the sum of oscillator strengths over all possible final states j equals the number of non-relativistic electrons Z available for : \sum_j f_{ij} = Z. This rule, derived from the commutator of the with the , ensures conservation of the total transition strength and serves as a benchmark for the accuracy of computed or measured values across an atom's ; for example, in (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 lines in or spectra under controlled conditions, such as in optically thin or beams, where line strength is proportional to gf (with g the degeneracy factor). Alternatively, they can be inferred from radiative \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 for interpreting stellar spectra and abundance determinations, as well as in plasma diagnostics to infer densities and temperatures from observed line ratios. For instance, precise f-values for iron lines enable modeling of photospheric conditions.

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 for atoms compared to visible light wavelengths of 400–700 . In this regime, the interaction between the and the system is dominated by the electric dipole (E1) term in the of the , as higher-order terms like or electric contributions are suppressed by factors of (atomic size / ). 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 of the . 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. 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). 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.
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. 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}. The dipole matrix element also connects to the classical oscillator strength f \propto |\mu_{21}|^2, bridging quantum and semiclassical descriptions.

Applications in Radiation Theory

Derivation of Planck's Law

The derivation of from the Einstein coefficients relies on the conditions established by the principle of detailed balancing, where the rates of and stimulated emission balance the in a field. Consider two energy levels with degeneracies g_1 and g_2, and populations N_1 and N_2 (with N_1 > N_2 in equilibrium for the lower and upper levels, respectively), related by the Boltzmann factor N_2 / N_1 = (g_2 / g_1) \exp(-h\nu / kT). In , the transition upward due to equals the total downward : N_1 B_{12} \rho(\nu) = N_2 A_{21} + N_2 B_{21} \rho(\nu), with the relation g_1 B_{12} = g_2 B_{21}. Rearranging yields \rho(\nu) = \frac{N_2 A_{21}}{B_{21} (N_1 \frac{g_2}{g_1} - N_2)} (using B_{12} = \frac{g_2}{g_1} B_{21}). Substituting the population ratio gives \frac{N_2}{N_1 \frac{g_2}{g_1} - N_2} = \frac{1}{\exp(h\nu / kT) - 1}, so \rho(\nu) = \frac{A_{21}}{B_{21}} \cdot \frac{1}{\exp(h\nu / kT) - 1}. The degeneracies cancel, yielding an expression independent of g_i. The relation between coefficients, A_{21} / B_{21} = 8\pi h \nu^3 / c^3, follows from thermodynamic arguments equating the high-temperature (low-frequency) limit of the quantum \rho(\nu) to the classical Rayleigh-Jeans result \rho(\nu) = (8\pi \nu^2 / c^3) kT, where the mode density of the radiation field times the average thermal energy per mode kT matches the form after substituting the occupation number. Substituting this yields the full Planck form: \rho(\nu, T) = \frac{8\pi h \nu^3}{c^3} \cdot \frac{1}{\exp(h\nu / kT) - 1}. This expression describes the energy density per unit frequency interval in blackbody radiation. Einstein's derivation in 1917 provided a heuristic foundation for quantizing the radiation field, interpreting the result as arising from discrete photons occupying cavity modes, with the average occupation number per mode given by $1 / (\exp(h\nu / kT) - 1). This work built on his 1905 light-quantum hypothesis and resolved the ultraviolet catastrophe—the classical prediction of infinite energy density at high frequencies—by confirming Planck's 1900 spectral law through atomic transition probabilities.

Implications for Blackbody Radiation

The introduction of Einstein coefficients provided a quantum mechanical framework that resolved key paradoxes in classical radiation theory, particularly the predicted by the Rayleigh-Jeans law, where classical equipartition of energy implied infinite energy at high frequencies. By balancing (governed by the B coefficient) with (governed by the A coefficient), the model ensures that the of follows Planck's , which converges at short wavelengths and matches experimental observations without . This balance demonstrates how quantum transitions prevent the unphysical accumulation of high-frequency modes, marking a foundational shift from classical to quantum descriptions of . A direct consequence of the stimulated emission process encoded in the B coefficient is the principle underlying laser operation, where population inversion in an atomic medium—achieved by pumping more atoms to an than the —allows to dominate over and . This amplification of coherent light arises because incoming photons trigger identical photons from excited atoms, leading to in intensity within a resonant . Einstein's formulation laid the theoretical groundwork for this phenomenon, enabling the development of devices that produce monochromatic, directional beams essential for applications in , , and . In , the Einstein coefficients underpin the interpretation of the () as relic from the early universe, with its precisely matching at a of approximately 2.725 . The balance of absorption, , and in ensures the CMB's near-perfect blackbody form, providing evidence for the hot model and allowing constraints on cosmological parameters like the universe's expansion history. Observations confirm deviations from a classical are negligible, validating the quantum statistical nature of the photon gas. Experimental validations of using Einstein coefficients have relied on cavity measurements, where the predicted spectral from the coefficients aligns with observed intensities across wavelengths. Early tests by Lummer and Pringsheim, using electrically heated cavities up to 1600°C, confirmed Planck's law's between Wien's short-wavelength and Rayleigh-Jeans long-wavelength limits, while Rubens and Kurlbaum's measurements further verified the law's accuracy. Modern experiments, such as COBE/FIRAS, have measured the CMB spectrum to a of 0.005%, directly supporting the coefficient-derived predictions without deviations exceeding instrumental limits. These tests underscore the universality of the model for enclosed systems. The Einstein coefficient framework also reveals modern extensions in quantum statistics, where the equilibrium distribution derived for photons follows Bose-Einstein statistics, treating them as indistinguishable bosons with zero . This emerges naturally from the between transition rates, explaining phenomena like bunching and the absence of a for radiation. Such insights have influenced treatments of bosonic fields in and condensed matter systems, highlighting the coefficients' role in unifying thermal and quantum descriptions.

References

  1. [1]
    Einstein A and B Coefficients
    Einstein postulated on thermodynamic grounds that the probability for spontaneous emission, A, was related to the probability of stimulated emission, B.
  2. [2]
    Einstein Coefficients - AstroBaki - CASPER
    Dec 15, 2018 · Einstein coefficients describe the absorption and emission of photons via electronic transitions in atoms.Einstein Coefficients · 3 Stimulated Emission, B 21... · 1 Einstein Relations among...
  3. [3]
    [PDF] On the Quantum Theory of Radiation
    Einstein, Beiträge zur Quantentheorie. Verh. der D. Physikal. Ges. 16. (1914) 820. 1d A. Einstein, Strahlungs-Emission und Absorption nach der Quantentheorie.Missing: Strahlung | Show results with:Strahlung
  4. [4]
    Einstein Coefficients - Astronomy 505
    There are three Einstein coefficients, B 12, B 21, and A 21, which are used to calculate the probability per unit time for emission or absorption of a photon.
  5. [5]
    [PDF] Lecture 5: Quantitative Emission/Absorption
    Einstein theory of radiation. ▫. Einstein coefficients of radiation. State 2. State 1. Transition probability/s of process per atom in state 1 or 2. Spontaneous.
  6. [6]
    Spectral Lines - RP Photonics
    Spectral lines are emission or absorption lines specific to substances, used for identification and concentration measurement.
  7. [7]
    Joseph von Fraunhofer
    In 1814 Joseph von Fraunhofer noticed that the light from the sun does not give a continuous spectrum. By using an unusually good prism, Fraunhofer observed a ...
  8. [8]
    Kirchhoff's Laws and Spectral Analysis
    1. A hot dense gas or solid object produces a continuous spectrum with no dark spectral lines. · 2. A hot diffuse gas produces bright emission lines. · 3. A cool ...
  9. [9]
    Hydrogen energies and spectrum - HyperPhysics
    The Bohr model for an electron transition in hydrogen between quantized energy levels with different quantum numbers n yields a photon by emission with quantum ...Missing: source | Show results with:source
  10. [10]
    Spectra and What They Can Tell Us
    From spectral lines astronomers can determine not only the element, but the temperature and density of that element in the star. The spectral line also can tell ...
  11. [11]
    Atomic Spectroscopy - Spectral Lines | NIST
    Oct 3, 2016 · Spectral lines in atomic spectroscopy are related to selection rules, intensities, transition probabilities, and line strengths, including ...
  12. [12]
    Einstein predicts stimulated emission - American Physical Society
    First, Einstein proposed that an excited atom in isolation can return to a lower energy state by emitting photons, a process he dubbed spontaneous emission.
  13. [13]
    Einstein's Model Of Light And Changing The Physics Of Empty Space
    Feb 28, 2019 · Einstein argued, it must be possible for light to remove energy from atoms by triggering the emission of additional light.
  14. [14]
    [PDF] Einstein's contributions to atomic physics
    Apr 29, 2009 · Einstein's 1917 paper on 'The quantum theory of radiation' [8] was the first formulation of a number of different physical processes. It was the ...Missing: summary | Show results with:summary
  15. [15]
  16. [16]
    [PDF] ON THE QUANTUM THEORY OF RADIATION
    By postulating some hypotheses on the emission and absorption of radiation by molecules, which suggested themselves from quantum theory, I was able to show that.Missing: coefficient A21
  17. [17]
    [PDF] Einstein coefficients, cross sections, f values, dipole moments ... - arXiv
    In 1917 Einstein introduced A and B coefficients to describe spontaneous emission and induced absorption and emission. The Einstein A coefficient is defined in.
  18. [18]
    [PDF] E + hυ
    Define the spontaneous emission coefficient, A. 21. , as probability per unit time that the system in level 2 drops to level 1 with emission of a photon. Note ...
  19. [19]
    Einstein coefficients, cross sections, f values, dipole moments, and ...
    EINSTEIN COEFFICIENTS. In 1917 Einstein introduced A and B coefficients to de- scribe spontaneous emission and induced absorption and. emission. The Einstein A ...
  20. [20]
    A semiclassical optics derivation of Einstein's rate equations
    Oct 1, 2012 · The results show that the coefficient for stimulated processes (absorption and emission) matches Einstein's B coefficient ... angular frequency ...
  21. [21]
    [PDF] Einstein coefficients
    Definitions Einstein coefficients. We consider here processes of emission and absorption of line radiation in atoms. In a formal sense, these can be ...
  22. [22]
    [PDF] 17. Precision Oscillator Strength and Lifetime Measurements
    71). Precision measurements of oscillator strengths and lifetimes also provide stringent tests of atomic structure calculations. These quantities are very ...
  23. [23]
    [PDF] Lecture 17 Electric dipole approximation; Einstein A and B ...
    Quantum Mechanics. Lecture 17. Electric dipole approximation;. Einstein A and B coefficients;. Spontaneous Emission;. Selection Rules. Page 2. A quick recap.
  24. [24]
    [PDF] Einstein
    In this way he eliminates the probability coefficients from his equilibrium equation and obtains the Planck formula for the energy density of black-body ...
  25. [25]
    [PDF] Einstein and the Quantum - arXiv
    Oct 24, 2005 · Since this radiation law does not agree with experiments, and theoretically suffers from “ultraviolet catastrophe” (i.e. infinite total energy), ...
  26. [26]
    [PDF] On the Law of Distribution of Energy in the Normal Spectrum
    The recent spectral measurements made by O. Lummer and E. Pringsheim1, and even more notable those by H. Rubens and F. Kurlbaum2, which together confirmed ...<|control11|><|separator|>
  27. [27]
    [PDF] The Cosmic Microwave Background Spectrum from the Full COBE1 ...
    The FIRAS spectrum of the cosmic microwave background radiation agrees with a blackbody spectrum to high accuracy. The CMBR monopole and dipole spectra are the.Missing: nature. | Show results with:nature.