Fact-checked by Grok 2 weeks ago

Photon

A photon is the fundamental quantum of , representing the smallest indivisible unit of and all other forms of the , from radio waves to gamma rays. It is a massless, chargeless that always travels at the in vacuum, approximately 3 × 10^8 meters per second. Photons exhibit wave-particle duality, behaving both as particles and waves, with their energy given by E = hν, where h is Planck's constant and ν is the of the associated electromagnetic wave. Key properties of the photon include its , calculated as p = E/ (where c is the ), and its intrinsic angular , or , of 1 in units of ħ (reduced Planck's ), resulting in helicity states of ±1 due to its massless nature. Photons mediate the electromagnetic force between charged particles through exchange, enabling interactions such as absorption and emission that underpin phenomena like the and atomic spectra. Unlike massive particles, photons cannot be at rest and maintain their individuality even in superposition, contributing to quantum effects like in experiments such as the double-slit setup. The concept of the photon emerged in the early , with introducing quantized energy packets in 1900 to resolve puzzles, and formalizing photons in 1905 to explain the , for which he was awarded the 1921 . This quantum description reconciled with particle-like behavior, laying the foundation for (QED), the theory describing photon interactions. Today, photons are central to technologies including lasers, solar cells, and , where their properties enable via and precise information processing.

Physical Properties

Energy and Momentum

The energy of a photon is given by E = h \nu, where h is Planck's constant and \nu is the of the associated electromagnetic wave. This relation originates from Max Planck's 1900 hypothesis, which resolved the in by assuming that the energy of oscillators in the cavity walls is quantized in discrete units of h \nu, rather than continuously variable. extended this quantization to itself in 1905, proposing that light consists of discrete energy packets, or photons, each carrying energy E = h \nu, to explain the . The \mathbf{p} of a photon is \mathbf{p} = \frac{h}{\lambda} \hat{\mathbf{k}}, where \lambda is the , h is Planck's , and \hat{\mathbf{k}} is vector in the direction of propagation; equivalently, p = \frac{E}{c}, with c the in . Einstein first derived the momentum relation theoretically in 1905 and 1917 by considering the pressure exerted by light on a mirror, but experimental confirmation came from Arthur Compton's 1923 experiments, where X-rays interacting with electrons showed wavelength shifts consistent with photon momentum transfer. For photons, the relation E = p c implies zero rest mass, as the relativistic energy-momentum equation E^2 = (p c)^2 + (m_0 c^2)^2 simplifies to this form when the rest mass m_0 = 0, consistent with massless particles traveling at speed c in special relativity. Photons in the visible spectrum, corresponding to wavelengths of approximately 400–700 nm, have energies ranging from about 1.8 eV (red light) to 3.1 eV (violet light). In contrast, X-ray photons typically carry energies on the keV scale, from roughly 0.5 keV to over 100 keV, enabling their penetration of materials. In vacuum, photons propagate without dispersion, meaning the phase velocity v_p = \frac{\omega}{k} and group velocity v_g = \frac{d \omega}{d k} are both equal to c, where \omega is the angular frequency and k is the wave number; this dispersionless behavior ensures that wave packets maintain their shape during propagation.

Polarization and Spin

Photons possess an intrinsic spin angular momentum of magnitude S = \hbar, where \hbar is the reduced Planck's constant, characteristic of spin-1 bosons in quantum electrodynamics (QED). Due to their massless nature and the transversality requirement of electromagnetic fields in vacuum, photons exhibit only two possible helicity states: +1 (right-handed) and -1 (left-handed), corresponding to the projection of spin along the direction of propagation. The zero-helicity state is forbidden, as it would imply a longitudinal polarization incompatible with Maxwell's equations for free-space propagation. Polarization states of photons can be described in terms of linear or circular basis. Linear polarization corresponds to superpositions of the \pm 1 helicity states with equal amplitudes but phase differences of 0 or \pi, while circular polarization aligns directly with the pure helicity eigenstates. For partially polarized light, the Stokes parameters provide a complete characterization: S_0 for total intensity, S_1 and S_2 for linear polarization components along orthogonal axes, and S_3 for circular polarization. These parameters, originally classical, extend naturally to quantum descriptions of photon ensembles. In interactions such as or , angular momentum is conserved, with the photon's contributing to changes in the system's total . For instance, in atomic transitions, a \sigma^+ (right-circularly polarized) photon carries +\hbar , inducing a \Delta m = +1 change in the electron's , as observed in selection rules for electric dipole transitions. This conservation underlies phenomena like , where chiral media selectively absorb or rotate polarized light based on matching. The transverse nature of photon fields ensures that only these two degrees of freedom are physically realizable in , prohibiting longitudinal components. Experimental verification of photon spin and polarization includes observations in photoionization processes, where the ejected electron's angular distribution reflects the incident photon's helicity, as demonstrated in studies of alkali atoms. Similarly, in atomic transitions, polarization-dependent yields in excitation spectra confirm the \pm 1 helicity transfer, such as in helium metastable states via spin-flip mechanisms. These measurements align with QED predictions for transverse spin states.

Mass Constraints

In (QED), the photon serves as the for the U(1) electromagnetic symmetry group, and local gauge invariance mandates that it possesses zero rest mass to preserve the theory's renormalizability and ward identities. A nonzero photon mass would break this gauge symmetry, introducing a longitudinal mode and violating the in . Experimental constraints on the photon rest mass have progressively tightened over decades, confirming its massless nature to extraordinary precision. Early laboratory tests, such as the Plimpton-Lawton experiment measuring deviations in electrostatic forces, established an upper limit of approximately $10^{-44} g (or \sim 10^{-14} eV/c^2). Subsequent refinements, including torsion balance measurements of violations in the 1970s, improved this to < 10^{-18} eV/c^2, as summarized in comprehensive reviews of electromagnetic tests. Modern particle accelerator experiments, such as those at electron-positron colliders searching for anomalous dispersion in photon-mediated processes, further corroborate these bounds by showing no evidence of massive photon propagation up to energies of several GeV, yielding limits consistent with < 10^{-18} eV/c^2. Astrophysical observations provide even stricter constraints through the absence of frequency-dependent dispersion in electromagnetic waves over cosmic distances. For instance, analysis of fast radio bursts (FRBs) propagating across billions of light-years imposes an upper limit of m_\gamma \leq 3.1 \times 10^{-51} kg (\sim 10^{-54} kg in mass units), far below laboratory sensitivities. These bounds arise because a massive photon would cause lower-frequency signals to lag behind higher-frequency ones, an effect not observed in pulsar timing or galactic magnetic field dynamics. A nonzero photon mass would imply a speed v < c in vacuum, contradicting special relativity's postulate that massless particles travel at the invariant speed c. Precision measurements, including laser ranging and cavity-stabilized oscillators, confirm that photons propagate at c to within $10^{-15} relative deviation, ruling out any mass-induced velocity shift. For a massless photon, the dispersion relation simplifies to \omega = c k where \omega is the angular frequency and k is the wave number, ensuring nondispersive propagation across all frequencies.

History and Nomenclature

Historical Development

In the mid-19th century, James Clerk Maxwell developed the classical theory of electromagnetism, unifying electricity and magnetism into a set of equations that described light as transverse electromagnetic waves propagating through space at a constant speed, without any particulate nature. These waves were continuous and deterministic, arising from oscillating electric and magnetic fields, as outlined in Maxwell's 1865 paper, which predicted the existence of such waves but treated energy as infinitely divisible. The concept of discrete energy packets emerged in 1900 when Max Planck introduced his quantum hypothesis to resolve the ultraviolet catastrophe in blackbody radiation, proposing that energy is emitted or absorbed in quanta with E = h \nu, where h is Planck's constant and \nu is the frequency. Although Planck viewed this as a mathematical artifice for oscillators in the radiating body rather than for the radiation itself, it marked the first break from classical continuity. In 1905, Albert Einstein extended Planck's idea to light itself, postulating that electromagnetic radiation consists of localized particles—or "light quanta"—to explain the photoelectric effect, where electrons are ejected from a metal surface only above a threshold frequency, with kinetic energy proportional to h\nu minus a work function. This particle-like interpretation earned Einstein the 1921 Nobel Prize in Physics. The term "photon" for these quanta was later coined by Gilbert N. Lewis in 1926, shifting from Einstein's "light quantum." Experimental confirmation came in 1923 with Arthur Compton's scattering experiments, where X-rays interacting with electrons produced wavelength shifts matching the predictions of particle collisions, solidifying the wave-particle duality of light. Compton's Nobel Prize in 1927 recognized this evidence for photons as having both momentum p = h/\lambda and energy. In 1927, Paul Dirac formulated a relativistic quantum theory of the electron that incorporated photons as quantized excitations of the electromagnetic field, treating radiation absorption and emission probabilistically through interactions between electrons and the field. This laid groundwork for a consistent quantum description of light-matter interactions. Following World War II, quantum electrodynamics (QED) was rigorously developed in the late 1940s by Sin-Itiro Tomonaga, Julian Schwinger, and , who resolved infinities in earlier calculations using renormalization techniques, providing a covariant framework where photons mediate electromagnetic forces between charged particles. Their independent but equivalent formulations, unified by , earned Tomonaga, Schwinger, and Feynman the 1965 Nobel Prize, establishing QED as the most precise theory in physics.

Terminology and Notation

The term "photon" was coined by American chemist in a letter published in Nature on December 18, 1926, where he proposed it to describe a hypothetical new atom involved in the transmission of radiant energy. derived the name from the Greek word "phōs" (φῶς), meaning light, combined with the suffix "-on," used for elementary particles such as the and , to emphasize its role as a discrete unit of light. This nomenclature marked a shift from earlier terminology, as and had referred to such entities as "light quanta" (Lichtquanten) in their foundational work on and the , focusing on quantized energy packets rather than named particles. In modern physics, the photon is conventionally denoted by the Greek letter γ (gamma), a symbol likely originating from its association with gamma rays, the high-energy electromagnetic radiation discovered in the late 19th century. Related constants include h, which relates photon energy to frequency via E = hν, and the reduced Planck's constant ħ = h/2π, often used in quantum mechanical expressions involving angular momentum. These notations are standardized across and texts. Photons are distinguished as "real" or "virtual" based on whether they satisfy the on-shell condition, where their four-momentum p obeys p² = m²c² (with m = 0 for photons, implying E² = p²c²). Real photons are observable, propagating freely and conserving energy and momentum at detection, whereas virtual photons are off-shell intermediaries in interactions, such as in processes, where they do not obey this relation and cannot be directly observed. As an electrically neutral boson in the Standard Model, the photon is its own antiparticle, meaning no distinct antiphoton exists; its quantum numbers (zero charge, zero lepton number) remain unchanged under particle-antiparticle conjugation. Photon properties are often quantified in practical units suited to atomic and optical scales: energy in electronvolts (eV), where visible light photons range from about 1.77 eV (red) to 3.1 eV (violet), and wavelength in nanometers (nm), spanning 400–700 nm for the visible spectrum.

Wave-Particle Duality

Duality Evidence

The photoelectric effect provides compelling evidence for the particle nature of light. When light strikes a metal surface, electrons are ejected only if the light's frequency exceeds a material-specific threshold, regardless of intensity; below this frequency, no emission occurs, even with high-intensity light. This threshold corresponds to the energy required to overcome the metal's work function, with each photon's energy given by E = h\nu, where h is and \nu is the frequency. The instantaneous nature of electron emission further refutes classical wave theory, which predicts energy spreading over the surface and a delay proportional to intensity; instead, emission begins immediately upon illumination, consistent with discrete photon absorption. The Compton effect demonstrates the photon as a particle with momentum through X-ray scattering off electrons. In 1923, Arthur Compton observed that scattered X-rays exhibit a wavelength shift depending on scattering angle, described by \Delta \lambda = \frac{h}{m_e c} (1 - \cos \theta), where m_e is the electron mass, c is the speed of light, and \theta is the scattering angle. This shift arises from conservation of energy and momentum in a collision between the incident photon and a free electron at rest, treating the photon as a particle with momentum p = h / \lambda. The angular dependence and magnitude match quantum predictions, not classical Thomson scattering, which assumes no wavelength change. Electron recoil spectra confirm the particle-like transfer of momentum. Experiments with single photons in a double-slit setup reveal the wave nature through interference patterns that build probabilistically. When photons are attenuated to emit one at a time and detected individually, each registers as a discrete particle at a specific screen position, yet over many trials, the accumulation forms an interference fringe pattern characteristic of wave superposition. This pattern emerges only if both slits are open, indicating the photon's probability amplitude passes through both paths simultaneously before collapsing to a particle-like detection. Early low-intensity versions by G. I. Taylor in 1909 hinted at this, but modern setups using attenuated lasers confirm no multi-photon coincidences interfere. The Davisson-Germer experiment on electron diffraction established wave properties for matter, providing an analogy for photons; conversely, photon diffraction experiments, such as those using single photons in grating or slit arrays, exhibit wave-like de Broglie interference despite corpuscular detection. For instance, single-photon interference in a shows path indistinguishability leading to wave superposition, mirroring electron results but affirming light's duality. Atom interferometry extends this duality to massive particles, while for photons, analogous demonstrations occur in multi-path optical interferometers or integrated photonic circuits, where single photons exhibit interference patterns upon detection. Modern single-photon detectors, such as avalanche photodiodes or superconducting nanowire devices, enable precise corpuscular detection while preserving wave interference evidence. In delayed-choice quantum eraser experiments, single photons are detected one-by-one, confirming particle localization, yet erasing which-path information restores the full interference pattern, quantifying the duality trade-off. These detectors achieve near-unity quantum efficiency, ruling out classical explanations and verifying the probabilistic wave function governs detection statistics. High-fidelity sources, like heralded photons from spontaneous parametric down-conversion, ensure true single-photon states, with interference visibility exceeding 99% in balanced setups.

Uncertainty Relations

The Heisenberg uncertainty relations for photons stem from the fundamental commutation relations in the quantized electromagnetic field, reflecting the wave-particle duality inherent to light quanta. These relations impose limits on the simultaneous knowledge of conjugate variables, such as position and momentum or energy and time, adapted to the relativistic and massless nature of photons. Unlike massive particles, photons lack a straightforward position operator, so uncertainties are often defined using the center-of-energy or wavepacket spreads, ensuring rigorous applicability in . The position-momentum uncertainty principle for photons takes the form \Delta x \Delta p \geq \frac{\hbar}{2} in one dimension, where \Delta x is the standard deviation of the photon's transverse position in a wavepacket and \Delta p is the corresponding momentum spread. In the infinite-momentum frame or paraxial approximation relevant to optical beams, this bound aligns with nonrelativistic quantum mechanics, though full three-dimensional treatments yield a modified lower bound of approximately \frac{3}{2} \hbar due to spin contributions and transversality constraints. This relation has profound implications for photon localization: attempting to confine a photon to a spatial region much smaller than its wavelength \lambda (where \Delta x \sim \lambda / 4\pi) requires a momentum uncertainty \Delta p \gtrsim \hbar / \Delta x, leading to significant diffraction and delocalization effects that prevent perfect particle-like detection without wave-like spreading. The energy-time uncertainty principle, \Delta E \Delta t \geq \frac{\hbar}{2}, applies to the photon's energy spread \Delta E = \hbar \Delta \omega and the time over which the photon is observed or its wavepacket duration \Delta t. This manifests in spectral broadening, where the frequency uncertainty \Delta \omega determines the linewidth of the photon's spectrum; for instance, a shorter coherence time \Delta t (e.g., from finite excited-state lifetimes in emission processes) broadens the linewidth \Delta \omega \geq 1/(2 \Delta t), limiting the spectral purity of single photons. In emission contexts, this relation governs the natural linewidth, providing a fundamental lower limit on how sharply a photon's frequency can be defined relative to its temporal extent. For the angular form related to polarization, the photon's intrinsic spin-1 nature leads to uncertainties among the components of its angular momentum operator. Specifically, the transverse spin components satisfy \Delta S_x \Delta S_y \geq \frac{1}{2} |\langle S_z \rangle| \hbar, arising from the commutation relations [S_x, S_y] = i \hbar S_z. Here, S_z = \pm \hbar corresponds to circular polarization helicities, but measuring linear polarization (along x or y) introduces complementary uncertainty, preventing simultaneous precise knowledge of orthogonal polarization bases. This spin-induced uncertainty contributes to the noncommutativity of position operators in three dimensions, elevating the overall position-momentum bound beyond the scalar case. These relations are grounded in the Fourier transform duality between conjugate variables: the spatial wavepacket \psi(x) and its momentum-space counterpart \tilde{\psi}(p) obey \Delta x \Delta p \geq \frac{\hbar}{2} via the Fourier pair, with equality for Gaussian profiles. Similarly, the temporal envelope and frequency spectrum satisfy \Delta t \Delta \omega \geq \frac{1}{2}, linking wavepacket duration to frequency chirp or broadening. In applications to photon wavepackets, minimal-uncertainty states such as achieve the equality \Delta x \Delta p = \frac{\hbar}{2} in the transverse plane, forming fundamental modes in lasers and fibers where the beam waist w_0 and far-field divergence \theta satisfy w_0 \theta \approx \lambda / (2\pi). These states optimize information-carrying capacity while respecting uncertainty limits, essential for quantum communication and imaging.

Photon Statistics and Thermal Behavior

Bose-Einstein Distribution

Photons, as quanta of the electromagnetic field, possess an intrinsic spin of 1, classifying them as bosons with integer spin angular momentum. This integer spin implies that photons are indistinguishable particles that obey , allowing multiple photons to occupy the same quantum state without restriction, in contrast to fermions which adhere to the Pauli exclusion principle. In quantum statistical mechanics, the behavior of non-interacting bosons like photons is described within the grand canonical ensemble, where the system exchanges both energy and particles with a reservoir. The average occupation number \langle n \rangle for a single mode of energy \varepsilon = h\nu is derived as \langle n \rangle = \frac{1}{e^{\beta (\varepsilon - \mu)} - 1}, where \beta = 1/kT, k is Boltzmann's constant, T is temperature, and \mu is the chemical potential. For photons, the chemical potential \mu = 0 because their number is not conserved; photons can be freely created or annihilated in thermal processes due to the massless nature of the particles, leading to a variable particle number in equilibrium. Thus, the occupation number simplifies to \langle n \rangle = \frac{1}{e^{h\nu / kT} - 1}. This distribution contrasts with the classical Maxwell-Boltzmann statistics, which assumes distinguishable particles and yields \langle n_s \rangle = e^{-\beta (\varepsilon_s - \mu)} in the low-occupancy limit where \langle n \rangle \ll 1, applicable at high temperatures or low densities but failing for quantum effects in photon gases. A direct consequence of this Bose-Einstein distribution is Planck's law for blackbody radiation, where the spectral energy density u(\nu, T) in frequency space is given by u(\nu, T) = \frac{8\pi h \nu^3}{c^3} \langle n \rangle = \frac{8\pi h \nu^3 / c^3}{e^{h\nu / kT} - 1}, accounting for the two polarization states and the density of modes in a cavity. This formula emerges from integrating the occupation number over the photon density of states, originally derived by treating light quanta as indistinguishable entities. The zero rest mass of photons ensures that equilibrium is maintained without a fixed particle count, enabling the thermal radiation spectrum observed in nature.

Photon Gas in Equilibrium

A photon gas in thermal equilibrium, such as that realized in blackbody radiation within a cavity, exhibits thermodynamic properties derived from the of massless bosons. The total energy density u of this gas is given by u = a T^4, where a = \frac{4\sigma}{c} is the radiation constant and \sigma = \frac{\pi^2 k_B^4}{60 \hbar^3 c^2} is the , with k_B the , \hbar the , and c the . This relation quantifies the energy per unit volume as proportional to the fourth power of the temperature T, reflecting the relativistic nature of photons and their quadratic dispersion relation. Due to the isotropic distribution and zero rest mass of photons, the pressure P exerted by the gas satisfies P = \frac{u}{3}, analogous to that of any gas of ultrarelativistic massless particles. The entropy S of the photon gas is S = \frac{4}{3} \frac{U}{T}, where U = u V is the total internal energy and V the volume, leading to S \propto V T^3. In cosmological contexts, such as the expansion of the universe, this implies that the photon gas undergoes adiabatic evolution where entropy is conserved, resulting in the temperature scaling as T \propto 1/a with scale factor a; this governs the cooling of the cosmic microwave background (CMB) radiation. The number density n of photons in equilibrium scales as n \propto T^3, specifically n = \frac{2 \zeta(3)}{\pi^2} \left( \frac{k_B T}{\hbar c} \right)^3 with Riemann zeta function \zeta(3) \approx 1.202. Consequently, the average energy per photon is approximately $2.7 k_B T, obtained as the ratio of energy density to number density. Wien's displacement law further characterizes the spectrum, stating that the frequency \nu_{\max} at which the spectral energy density peaks is proportional to T, with \nu_{\max} \approx 5.879 \times 10^{10} T Hz for T in kelvin.

Emission Processes

Spontaneous Emission

Spontaneous emission refers to the probabilistic process in which an atom, molecule, or quantum system in an excited state decays to a lower energy state by emitting a single , occurring randomly in time, direction, and phase without any external radiation field to stimulate it. This fundamental quantum optical phenomenon arises from the interaction between the system's electric dipole moment and the vacuum fluctuations of the electromagnetic field, leading to an irreversible decay characterized by an exponential probability distribution. The process is essential for understanding incoherent light generation in atomic and molecular systems. The rate of spontaneous emission is quantified by the Einstein A coefficient, introduced by to describe the transition probability per unit time from an upper state (denoted 2) to a lower state (1). For electric dipole-allowed transitions in free space under the dipole approximation, the spontaneous emission rate Γ is given by \Gamma = A_{21} = \frac{\omega^3 \mu^2}{3 \pi \epsilon_0 \hbar c^3}, where ω is the angular frequency of the transition, μ = |⟨1| \mathbf{d} |2⟩| is the magnitude of the transition dipole moment (with \mathbf{d} = -e \mathbf{r} the electric dipole operator), ε₀ is the vacuum permittivity, ħ is the reduced Planck constant, and c is the speed of light. This expression highlights the dependence on the cube of the frequency, emphasizing that higher-energy transitions decay more rapidly. The coefficient A_{21} connects phenomenological rate equations to microscopic quantum interactions, enabling the prediction of emission probabilities in dilute gases and isolated systems. The microscopic origin of this rate emerges from time-dependent perturbation theory applied to the atom-field Hamiltonian, where the emission couples the discrete excited state to the continuum of photon modes. Using , the transition rate is calculated as the squared matrix element of the interaction Hamiltonian times the density of final states, summed over all possible photon wavevectors and polarizations. This yields the as the total decay rate into the vacuum, treating the electromagnetic field quantum mechanically while assuming weak coupling to avoid strong-field effects. The derivation assumes the (valid when the wavelength greatly exceeds atomic size) and neglects magnetic dipole or higher-order electric quadrupole contributions, which are typically much weaker. The inverse of the spontaneous emission rate defines the natural radiative lifetime of the excited state, τ = 1/Γ, representing the average time before decay occurs. This finite lifetime introduces an inherent uncertainty in the energy of the transition, manifesting as the natural linewidth of the emitted photon's frequency spectrum, Δω = 1/τ (full width at half maximum in angular frequency). The linewidth arises from the Fourier transform of the exponentially decaying wavefunction amplitude, broadening the otherwise sharp atomic resonance and setting a fundamental limit on spectral resolution in atomic spectroscopy. For typical optical transitions, lifetimes range from nanoseconds to microseconds, corresponding to linewidths of megahertz to gigahertz. In free space, the total emission is isotropic when averaging over randomly oriented dipoles, as in a thermal vapor, due to the uniform availability of all photon modes. However, for a coherently oriented dipole (e.g., in an aligned ensemble or single molecule), the angular intensity distribution follows the classical dipole radiation pattern, I(θ) ∝ sin²θ, where θ is the angle from the dipole axis; emission is zero along the axis and maximum in the equatorial plane. This distribution reflects the transverse nature of electromagnetic waves and influences applications like directional light sources. Quantum mechanical selection rules dictate which transitions can occur via electric dipole spontaneous emission, prohibiting or allowing decays based on conservation of angular momentum and parity. For atomic systems, the rules require a change in orbital angular momentum quantum number Δl = ±1, total angular momentum ΔJ = 0, ±1 (but not 0 to 0), and magnetic quantum number Δm_J = 0, ±1, with parity change required (odd to even or vice versa). These rules, derived from the vanishing of the dipole matrix element for forbidden transitions, explain why certain excited states (e.g., 2s to 1s in hydrogen) decay via slower magnetic dipole or two-photon processes rather than rapid electric dipole emission. Spontaneous emission plays a central role in luminescence processes: fluorescence involves prompt radiative decay from the lowest excited singlet state to the ground singlet state, typically on nanosecond timescales, while phosphorescence arises from slower triplet-to-singlet intersystem crossing followed by forbidden radiative decay from the triplet state, lasting milliseconds to seconds due to weaker spin-orbit coupling.

Stimulated Emission

Stimulated emission is a quantum process in which an incoming photon interacts with an excited atom or molecule, triggering the release of an identical photon from the upper energy state to a lower one. This phenomenon, predicted by in 1917 as part of his quantum theory of radiation, contrasts with spontaneous emission by being induced and directional, leading to coherent amplification of light. In Einstein's framework, the rate of stimulated emission transitions from an upper energy level 2 to a lower level 1 is given by W_{21}^{\text{stim}} = B_{21} \rho(\nu), where B_{21} is the Einstein coefficient for stimulated emission and \rho(\nu) is the energy density of the radiation field at frequency \nu. This rate depends on the population of the excited state and the incident radiation intensity, enabling exponential growth in photon number under suitable conditions. Einstein also derived the fundamental relation between the spontaneous emission coefficient A_{21} and the stimulated emission coefficient B_{21}: \frac{A_{21}}{B_{21}} = \frac{8\pi h \nu^3}{c^3}, linking the two processes through blackbody radiation principles and ensuring consistency with thermodynamic equilibrium. For net amplification via stimulated emission to occur, a population inversion must be achieved, where the number of atoms in the upper state exceeds that in the lower state (N_2 > N_1), overcoming the natural tendency toward thermal equilibrium. Without inversion, absorption dominates, but with it, the medium exhibits optical gain. The emitted photon in stimulated emission is indistinguishable from the stimulating one, matching its phase, direction, and polarization, which fosters coherence essential for applications like lasers. In laser operation, sustained requires the round-trip gain in the to exceed losses from absorption, scattering, and output coupling, defining the condition. Below threshold, occurs but without net output; above it, coherent light emerges. Historically, Einstein's prediction remained theoretical until the 1950s, when Charles Townes and colleagues demonstrated the first (microwave amplification by of ) using ammonia in 1954. This paved the way for optical lasers, with achieving the first in 1960, realizing Einstein's vision on a practical scale.

Quantum Field Theory Framework

Electromagnetic Field Quantization

The quantization of the classical represents a foundational step in , transforming the continuous field variables into discrete s that create and annihilate photons as quanta of the field. This procedure, known as , applies rules to the in the Coulomb gauge, expanding the field into a sum over normal modes corresponding to free-space plane waves. Each mode behaves like a , with creation a_k^\dagger adding a photon in state \mathbf{k} and \lambda, and annihilation a_k removing one. The electric and are expressed through this mode expansion. The operator is given by \mathbf{E}(\mathbf{r}, t) = i \sum_{\mathbf{k}, \lambda} \sqrt{\frac{\hbar \omega_k}{2 \epsilon_0 V}} \left[ \hat{\epsilon}_{\mathbf{k}\lambda} a_{\mathbf{k}\lambda} e^{i(\mathbf{k} \cdot \mathbf{r} - \omega_k t)} - \hat{\epsilon}_{\mathbf{k}\lambda}^* a_{\mathbf{k}\lambda}^\dagger e^{-i(\mathbf{k} \cdot \mathbf{r} - \omega_k t)} \right], where V is the quantization volume, \omega_k = c |\mathbf{k}| is the , and \hat{\epsilon}_{\mathbf{k}\lambda} is the orthogonal to \mathbf{k}. The follows similarly from \mathbf{B} = \nabla \times \mathbf{A}, with the \mathbf{A}(\mathbf{r}, t) expanded analogously without the i factor. These expressions ensure the fields satisfy in free space and exhibit transverse for each mode. The operators obey bosonic commutation relations [a_{\mathbf{k}\lambda}, a_{\mathbf{k}'\lambda'}^\dagger] = \delta_{\mathbf{k}\mathbf{k}'} \delta_{\lambda\lambda'}, while [a_{\mathbf{k}\lambda}, a_{\mathbf{k}'\lambda'}] = [a_{\mathbf{k}\lambda}^\dagger, a_{\mathbf{k}'\lambda'}^\dagger] = 0. The number operator for each mode is N_{\mathbf{k}\lambda} = a_{\mathbf{k}\lambda}^\dagger a_{\mathbf{k}\lambda}, with eigenvalues representing the photon occupation number in that mode. These relations stem from imposing equal-time commutation rules on the canonical variables \mathbf{A} and its conjugate \mathbf{\Pi} = -\epsilon_0 \partial_t \mathbf{A}, mirroring the quantization of mechanical oscillators. The resulting for the free quantized is H = \sum_{\mathbf{k}, \lambda} \hbar \omega_k \left( a_{\mathbf{k}\lambda}^\dagger a_{\mathbf{k}\lambda} + \frac{1}{2} \right), where the zero-point term \sum_{\mathbf{k}, \lambda} \frac{1}{2} \hbar \omega_k accounts for the of the , even in the absence of real photons. This infinite arises from the nonzero expectation value of H in the |0\rangle, defined by a_{\mathbf{k}\lambda} |0\rangle = 0 for all modes, and manifests as zero-point fluctuations that produce measurable effects like the Casimir force. The sum diverges due to the continuum of modes, requiring regularization in interacting theories. To manage infinities in expectation values involving the field operators, places all operators to the left of operators, such that \langle 0 | : \mathbf{E}^2 : | 0 \rangle = 0, subtracting the fluctuations explicitly. This technique reveals the physical origin of the as arising from the electron's interaction with these virtual fluctuations, shifting the 2S state of by approximately 1058 MHz relative to the Dirac prediction. While the momentum-basis modes \mathbf{k} describe plane-wave-like photons with definite energy-momentum, practical applications often require localized wave packets, prompting transformations to real-space photon operators. These are obtained by superposing momentum modes via transforms, yielding operators \hat{b}^\dagger(\mathbf{r}) that create photons centered at \mathbf{r} with a spatiotemporal profile, preserving commutation relations but enabling descriptions of propagating photon wave packets in experiments. This basis shift highlights the duality between delocalized field modes and localized particle-like excitations.

Gauge Boson Role

In (QED), the serves as the mediating the electromagnetic interaction, arising from the local U(1) gauge invariance of the theory. The QED , which encodes this symmetry, is given by \mathcal{L} = -\frac{1}{4} F_{\mu\nu} F^{\mu\nu} + \bar{\psi} (i D_\mu \gamma^\mu - m) \psi, where F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu is the strength tensor, A_\mu is the photon field, \psi represents the Dirac field for charged fermions, m is the fermion mass, \gamma^\mu are the Dirac matrices, and the is D_\mu = \partial_\mu - i e A_\mu with e the . This form ensures invariance under local phase transformations \psi \to e^{i \alpha(x)} \psi and A_\mu \to A_\mu + \frac{1}{e} \partial_\mu \alpha(x), where \alpha(x) is an arbitrary spacetime-dependent function. The photon's masslessness is a direct consequence of this U(1) gauge , which prohibits a mass term (1/2) m_\gamma^2 A_\mu A^\mu in the , as it would break the invariance. In the full , the photon remains massless because the U(1)_Y subgroup associated with is unbroken by the , unlike the SU(2)_L × U(1)_Y structure that gives to the W and Z bosons; the photon field is a linear combination orthogonal to the massive states and thus acquires no Higgs coupling. Perturbative calculations in , such as those involving Feynman diagrams, treat the photon as the exchange particle for electromagnetic forces; for instance, the tree-level diagram for electron-electron yields the Coulomb potential V(r) = -\frac{e^2}{4\pi \epsilon_0 r} through the in the non-relativistic limit. QED's predictive power relies on to handle infinities in higher-order diagrams, where the bare coupling e is replaced by the physical \alpha = \frac{e^2}{4\pi \epsilon_0 \hbar c} \approx 1/137.036, a dimensionless parameter quantifying the interaction strength. This preserves the theory's consistency, with gauge invariance ensuring finite results after absorbing divergences into redefined parameters. To maintain unitarity (probability conservation) and (no signaling) in quantum calculations, is essential; the Lorenz gauge \partial^\mu A_\mu = 0 is commonly employed, introducing a that respects these principles while leaving physical amplitudes gauge-invariant.

Advanced Interactions

In (QED), virtual photons play a crucial role in mediating electromagnetic interactions as off-shell particles, meaning their does not satisfy the on-shell condition p^2 = 0 for real photons. These virtual photons typically carry space-like four-momenta (q^2 < 0), enabling the instantaneous transfer of force between charged particles without violating conservation laws over short distances, as governed by the Heisenberg uncertainty principle. This off-shell nature allows virtual photons to propagate between interacting particles, such as in electron-electron scattering, where they facilitate the exchange without being directly observable. In contrast, time-like virtual photons (q^2 > 0) appear in processes like , where a virtual photon with sufficient can decay into a particle-antiparticle pair, such as an electron-positron pair above the of $2m_e c^2 \approx 1.022 MeV. These time-like photons arise in perturbative diagrams, contributing to phenomena like the Breit-Wheeler process in high-energy collisions, and their properties are probed in experiments such as virtual . Hadronic light-by-light represents a higher-order process modified by (QCD), where two photons fuse via virtual loops within hadronic intermediates, rather than purely leptonic loops. This effect, dominant in low-energy regimes, involves the photons coupling to quark-antiquark pairs that form or mesons, contributing significantly to observables like the muon's anomalous a_\mu. calculations have quantified these amplitudes, revealing non-perturbative dynamics that have been crucial in achieving agreement between and experiment in measurements, as shown by the 2025 updates to the prediction. Recent computations, incorporated in the 2025 Theory Initiative White Paper, have refined these contributions, leading to consistency between the experimental value and the prediction within uncertainties. The electromagnetic self-energy of hadrons, arising from the surrounding photon cloud in , provides a small but measurable correction to their es. For the proton, this photon cloud—comprising virtual photon fluctuations around its charged constituents—contributes to the electromagnetic component of the mass, estimated at approximately 0.76 MeV in the proton-neutron mass splitting, or about 0.08% of the proton's total of 938 MeV. This , computed via relations and simulations, highlights the interplay between perturbative and non-perturbative QCD binding. Two-photon fusion into hadrons, \gamma \gamma \to hadrons, exemplifies advanced beyond pure , with thresholds dictated by the lightest stable hadronic states. The lowest threshold occurs for \pi^0 \pi^0 production at a center-of-mass energy of $2m_{\pi^0} c^2 \approx 270 MeV, while higher thresholds, such as for \eta \eta or multi-pion states, probe vector meson dominance and resonances up to several GeV. Experimental studies at colliders like LEP have mapped these cross sections, revealing QCD scaling behaviors and resonances near 1-2 GeV. Delbrück scattering involves the elastic interaction of a real photon with a through electron-positron pairs polarized by the nuclear Coulomb field. Predicted in the 1930s and verified in the 1950s, this process occurs via box diagrams in , where the photon scatters off the virtual pair without real pair creation below the 1.022 MeV threshold, leading to amplitude contributions that interfere with nuclear . High-precision measurements at energies up to tens of MeV have confirmed its role in testing in strong fields, with applications to astrophysical photon propagation.

Behavior in Matter

Propagation and Absorption

Photons propagate through at the c \approx 3 \times 10^8 m/s without energy loss due to , though their stretches due to the of , resulting in cosmological quantified by z = \Delta \lambda / \lambda. In media, photon is characterized by the n, defined as the ratio of the speed of light in to the v in the medium, so n = c / v and v = c / n. This reduction in phase velocity arises from the interaction of the with the medium's electrons, altering the wave's without altering the photon's intrinsic , which remains E = h \nu dependent on \nu. Absorption occurs when photons transfer energy to matter, reducing beam intensity exponentially according to the absorption coefficient \alpha, as described by the Beer-Lambert law: I = I_0 e^{-\alpha x}, where I is the intensity after distance x and I_0 is the initial intensity. The coefficient \alpha quantifies the medium's opacity at a given frequency, with higher values indicating stronger absorption; for example, in optical materials like glass, \alpha is low in the visible range (\sim 10^{-3} cm^{-1}), allowing transmission over centimeters. Resonant absorption is particularly efficient near atomic transition frequencies, where the photon's energy matches the energy difference between atomic levels, leading to excitation; the absorption cross-section \sigma scales with the square of the transition dipole moment |\mu|^2, as \sigma \propto |\mu|^2, enhancing absorption by orders of magnitude at line centers. In metals, photon absorption is pronounced due to free electron interactions, characterized by the skin depth \delta = 2 / \alpha (where \alpha is the intensity absorption coefficient), the distance over which the electromagnetic field amplitude penetrates before attenuating to $1/e of its surface value (corresponding to intensity attenuating to $1/e^2). For instance, at optical frequencies, \delta in is on the order of nanometers, confining absorption to the surface and enabling applications like reflective coatings. occurs at interfaces between media when photons incident from a higher-index medium (n_1 > n_2) exceed the \theta_c = \sin^{-1}(n_2 / n_1), resulting in complete reflection without transmission loss, as the evanescent wave decays exponentially in the lower-index medium. This phenomenon, distinct from which deflects photons without net energy loss, underpins optic waveguides by trapping light through repeated reflections.

Scattering Phenomena

Scattering phenomena involving photons occur when these interact with , resulting in deflection or redirection of the photon's path without complete , often accompanied by exchange in inelastic cases. These processes are fundamental to understanding in media ranging from gases to solids and plasmas, influencing phenomena like and high-energy particle interactions. Elastic scattering preserves the photon's , while transfers a portion to the target particle or system, altering the photon's . Rayleigh scattering describes the elastic interaction of photons with particles much smaller than the , such as atmospheric molecules, where the scattering cross-section σ scales inversely with the fourth power of the , σ ∝ 1/λ⁴. This dependence arises from the induced oscillation in the scatterer, leading to re-radiation of the photon in a new direction with the same frequency. In Earth's atmosphere, by and oxygen molecules preferentially scatters shorter wavelengths more efficiently than longer ones, resulting in the observed color of the daytime as is diffused across the viewing angles. Compton scattering is an inelastic process where a photon collides with a free or loosely bound , transferring and to the while the photon scatters at an angle θ with a wavelength shift given by Δλ = ( / m_e c) (1 - cos θ), where is Planck's constant, m_e is the , and c is the . This shift, known as the , is independent of the incident and highlights the particle-like nature of photons, with the maximum shift occurring at θ = 180° for backscattering. The process is prominent for and gamma-ray photons in materials with low , where the can be treated as quasi-free. In the low-energy limit of , where photon energies are much less than the rest mass energy (hν ≪ m_e c²), the interaction reduces to , the classical elastic of photons by free s or positrons. The Thomson cross-section is σ_T = (8π/3) (α² ħ² / m_e² c²), where α is the and ħ is the reduced Planck's constant, yielding a value of approximately 6.65 × 10^{-25} cm². In the classical treatment for unpolarized incident light, the differential cross-section varies with the scattering angle θ, and the scattered intensity is proportional to (1 + cos² θ)/2 relative to the incident direction. For polarized incident light, the angular distribution depends on the polarization state. dominates in astrophysical plasmas, such as the interactions with free s. Raman scattering involves inelastic interactions of photons with molecular vibrations or phonons in solids, where the scattered photon experiences a shift corresponding to the of the vibrational . In Stokes Raman scattering, the photon loses to excite a vibrational state, shifting to a lower (longer ), while anti-Stokes scattering involves energy gain from an already excited , increasing the frequency. The shift Δν is typically on the order of 10²–10⁴ cm⁻¹, directly probing molecular structure without requiring absorption, and the process is weak, with cross-sections about 10^{-3} to 10^{-6} times that of elastic .

Applications and Modern Uses

Technological Implementations

Photons play a central role in photovoltaic devices, where cells convert incident photon energy into electrical power through the in p-n s. In these structures, photons with energy exceeding the material's bandgap are absorbed, exciting electrons from the valence band to the conduction band and generating electron-hole pairs. The built-in at the p-n separates these charge carriers, with electrons drifting to the n-type region and holes to the p-type region, thereby producing a that can be extracted as usable . Silicon-based cells, the most common type, operate via this mechanism, with theoretical efficiency limited by fundamental thermodynamic and recombination processes. The Shockley-Queisser limit establishes the maximum efficiency for single-junction solar cells under standard solar illumination, reaching approximately 33% for 's 1.12 eV bandgap due to losses from sub-bandgap transmission, thermalization of excess , and radiative recombination. This limit, derived from principles considering and carrier statistics, highlights the trade-off between absorption and voltage output, guiding material selection and device optimization in . Practical efficiencies approach 25-27% in high-performance cells, underscoring the impact of non-radiative recombination and material quality on real-world performance. Light-emitting diodes (LEDs) and lasers harness of photons to produce light from diodes, enabling efficient, coherent sources for illumination and signaling. In LEDs, electrons injected across a p-n junction recombine with holes, releasing photons through spontaneous and , with wavelengths determined by the bandgap (e.g., gallium arsenide phosphide for red light). The first visible-spectrum LED, demonstrated in 1962 using gallium arsenide phosphide, marked the beginning of practical light sources, now ubiquitous in displays and lighting with external quantum efficiencies exceeding 50%. Lasers extend this principle by achieving population inversion in the active region of a semiconductor diode, amplifying stimulated emission to generate coherent, monochromatic light with low divergence. Semiconductor lasers, first realized in 1962 using gallium arsenide junctions, rely on optical feedback from cleaved facets or gratings to sustain lasing, producing outputs from infrared to visible wavelengths. These devices power applications like optical data storage and telecommunications, with threshold currents reduced to milliamperes through heterostructure designs that confine carriers and photons effectively. Fiber optic systems utilize photons for high-speed data transmission by guiding light through silica cores via total internal reflection at the core-cladding interface, where the refractive index difference confines the beam with minimal loss. In communication networks, photons modulated by lasers traverse waveguides, enabling bandwidths up to terabits per second over distances exceeding hundreds of kilometers. Modern single-mode fibers achieve attenuation below 0.2 dB/km at 1550 nm, primarily due to optimized glass purity that minimizes Rayleigh scattering and infrared absorption by OH groups. This low-loss propagation, combined with wavelength-division multiplexing, underpins global internet infrastructure and supports dense, long-haul photon-based signaling. Charge-coupled device (CCD) sensors detect individual photons in imaging applications by converting their energy into electron charge packets within a pixel array of photodiodes. Each photon absorption generates photoelectrons proportional to the incident flux, which are then shifted and read out to form digital images, enabling high-fidelity capture in astronomy and microscopy. The spatial resolution of such photon-counting systems is fundamentally constrained by the diffraction limit, approximately λ/2 for the Airy disk radius in incoherent illumination, where λ is the photon wavelength. This limit arises from wave optics, dictating that features smaller than half the wavelength cannot be resolved without advanced techniques like super-resolution. CCDs, with quantum efficiencies up to 90%, thus balance photon sensitivity with this optical bound for applications from consumer cameras to scientific instruments. Photolithography employs ultraviolet (UV) and extreme ultraviolet (EUV) photons to pattern nanoscale features on semiconductor wafers, forming the intricate circuits of integrated chips. In this process, deep UV (DUV) light from excimer lasers or mercury lamps exposes a photoresist layer through a mask for nodes above ~7 nm, while EUV light at 13.5 nm wavelength, generated by laser-produced tin plasma sources, is used for advanced nodes below 10 nm, inducing chemical changes that define conductive paths and transistors with critical dimensions below 10 nm. The photon's energy breaks molecular bonds in the resist, enabling selective etching or deposition to transfer the pattern onto the substrate, with resolution scaling inversely with wavelength per the Rayleigh criterion. This technique, evolved from contact printing to projection systems, has driven Moore's Law by achieving sub-5 nm nodes through EUV single patterning and High-NA EUV systems introduced in high-volume manufacturing in 2025, enhancing photon utilization for finer features.

Quantum Optics and Information

In quantum optics, photons serve as ideal carriers for quantum information due to their weak interactions and ability to maintain over long distances. Quantum information processing leverages the photon's , path, or temporal modes to encode qubits, enabling protocols that exploit superposition and entanglement. Single-photon sources are essential for such applications, as they provide the non-classical light required for reliable quantum operations. Two prominent methods for generating single photons include (SPDC) in nonlinear crystals, where a pump photon splits into a correlated pair, with one photon heralding the presence of the other, and semiconductor s, which emit single photons upon resonant . Quantum dot sources have achieved brightness exceeding 65% and indistinguishability over 99%, making them suitable for scalable quantum networks. SPDC-based heralded sources, while probabilistic, offer high purity and are widely used in proof-of-principle experiments. Photon entanglement, a cornerstone of quantum information, is routinely generated via SPDC in type-II crystals, producing polarization-entangled Bell states such as |\Psi^-\rangle = \frac{1}{\sqrt{2}} (|HV\rangle - |VH\rangle), where H and V denote horizontal and vertical polarizations. These states violate Bell's inequalities, demonstrating non-locality; for instance, experiments with SPDC photons have achieved violations exceeding 2.7, confirming quantum predictions over local realism. Such entangled photons enable secure communication through (QKD). The protocol, proposed by Bennett and , encodes bits in photon polarization states—using rectilinear (H/V) or diagonal bases—sent over optical fibers or free space. Alice randomly prepares single photons in one of four states, and Bob measures in a randomly chosen basis; matching bases yield the raw key, with eavesdropping detected via error rates exceeding the quantum bit error rate threshold of 11%. Polarization encoding ensures security via the , as any interception disturbs the quantum states. Linear optical quantum computing (LOQC) utilizes photons for universal quantum gates, as outlined in the Knill-Laflamme-Milburn (KLM) scheme, which employs beam splitters, phase shifters, single-photon sources, and detectors to implement nonlinear interactions probabilistically. The teleports conditional sign shifts using ancillary photons and post-selection on detection outcomes, achieving fault-tolerant computation with success probabilities approaching 1% per gate, scalable with improved resources. This approach overcomes the limitations of linear optics, where two-photon interference alone cannot produce entanglement deterministically. The profoundly impacts photonic systems, prohibiting perfect replication of arbitrary unknown photon states, such as superpositions of . Formally, for non-orthogonal states |\psi\rangle and |\phi\rangle, no unitary operation can produce clones | \psi \psi \rangle and | \phi \phi \rangle without distinguishing them first. This underpins QKD security and necessitates measurement-based schemes; upon detection, the photon's wavefunction collapses, projecting it into a definite state (e.g., H or V), destroying superposition but enabling probabilistic readout in LOQC. As of 2025, photonic approaches have advanced to practical systems, such as Quandela's , the most powerful photonic quantum computer deployed in , equipped with over 100 squeezed-light qubits for hybrid applications. Initiatives like DARPA's Quantum Benchmarking program have propelled companies including Photonic Inc. toward industrially useful photonic quantum devices, demonstrating progress in scalable error-corrected computation using photons.

References

  1. [1]
    DOE Explains...Photons - Department of Energy
    Photons are the smallest possible particles of electromagnetic energy and therefore also the smallest possible particles of light. Photons can travel at the ...
  2. [2]
    Photons - The Quantum Atlas
    One photon is the smallest flicker of light allowed, meaning that a single photon can't be divided any further. Photons also have some characteristic properties ...
  3. [3]
    29.4 Photon Momentum – College Physics - UCF Pressbooks
    Photons have momentum, given by p = h λ , where λ is the photon wavelength. · Photon energy and momentum are related by p = E c , where E = h f = h c / λ for a ...
  4. [4]
    Spin in Quantum Mechanics, with Relativity, and the Dirac Equation
    Feb 10, 2013 · Photons, which are massless spin 1 particles, can only have their spin along or opposite their direction of motion, and are then referred to ...
  5. [5]
    Magnetic Field is Made of Photons | Physics Van | Illinois
    The electromagnetic interaction is mediated by the constant exchange of photons from one charged object to another.
  6. [6]
    [PDF] Planck, the Quantum, and the Historians - csbsju
    In late 1900, the German theoretical physicist Max Planck derived an expression for the spectrum of black-body radiation. That derivation was the first step ...
  7. [7]
    [PDF] Einstein's Proposal of the Photon Concept-a Translation
    Of the trio of famous papers that Albert Einstein sent to the Annalen der Physik in 1905 only the paper proposing the photon concept has been unavailable in ...
  8. [8]
    [PDF] A Quantum Theory of the Scattering of X-Rays by Light Elements
    The experimental support of the theory indicates very convinc- ingly that a radiation quantum carries with it directed momentum as well as energy. Emphasis has ...Missing: photon | Show results with:photon
  9. [9]
    [PDF] Chapter 2: Relativistic Mechanics Introduction
    That would mean. E = pc and hence, that m = 0, i.e., it would be a massless particle. So massless particles travel at the speed of light, and particles that ...<|separator|>
  10. [10]
    Visible Light - UCAR Center for Science Education
    Red photons of light carry about 1.8 electron volts (eV) of energy, while each blue photon transmits about 3.1 eV. Visible light's neighbors on the EM ...
  11. [11]
    For Educators - HEASARC - NASA
    Aug 24, 2022 · X-rays usually range in energy from around 0.5 keV up to around 1000 keV. Like line emission, continuum X-ray emission involves charged ...
  12. [12]
    Group Velocity - RP Photonics
    The group velocity of light in a medium is defined as the inverse of the derivative of the wavenumber with respect to angular optical frequency.
  13. [13]
    [PDF] Lecture 17 - Purdue Physics department
    call these helicity states of a spin 1 particle “transverse” polarization. A massive spin 1 particle can also have helicity 0. We call this “longitudinal” ...Missing: transversality | Show results with:transversality
  14. [14]
    [PDF] Electroweak Unification and the W and Z Bosons - Particle Physics
    A real (i.e. not virtual) massless spin-1 boson can exist in two transverse polarization states, a massive spin-1 boson also can be longitudinally polarized.
  15. [15]
    Symmetries and conservation of spin angular momentum, helicity ...
    Spin angular momentum exhibits purely spatial symmetry, rendering it a conserved quantity even in the presence of time modulation.Abstract · Article Text · INTRODUCTION · POLARIZATION-DEPENDENT...
  16. [16]
    Quantum Entanglement and the Two-Photon Stokes Parameters
    Oct 30, 2001 · Abstract: A formalism for two-photon Stokes parameters is introduced to describe the polarization entanglement of photon pairs.
  17. [17]
    The Feynman Lectures on Physics Vol. III Ch. 18: Angular Momentum
    The conservation of angular momentum (among other things) will determine the polarization and angular distribution of the emitted photons.
  18. [18]
    Angular momentum transfer from photon polarization to an electron ...
    Jul 16, 2019 · The angular momentum is conserved in the optical transition such that a photon with angular momentum −1 (+1) for σ− (σ+) polarization, creates ...
  19. [19]
    Experimental confirmation of photon-induced spin-flip transitions in ...
    The results indicate that a substantial contribution to the observed MY in the zero field is due to the SO coupling of SESs involved in the radiative cascade ...
  20. [20]
    Observation of photon-nucleus angular-momentum transfer in the ...
    Jun 11, 2019 · Here, we report the theoretical and experimental observation of the transfer of the spin angular momentum of the photon into the nuclear orbital ...
  21. [21]
    [PDF] 6. Quantum Electrodynamics - DAMTP
    The massless vector field Aµ has 4 components, which would naively seem to tell us that the gauge field has 4 degrees of freedom. Yet we know that the photon ...
  22. [22]
    The QED LaGrangian and Gauge Invariance
    The LaGrangian is no longer gauge invariant. Gauge invariance implies zero mass photons and even maintains the massless photon after radiative corrections.
  23. [23]
    Introduction
    The Plimpton-Lawton experiment of 1936, reinterpreted according to these principles, yields a limit on the photon rest mass of 10-44 g. They performed an ...
  24. [24]
    Photon and graviton mass limits | Rev. Mod. Phys.
    Mar 23, 2010 · The lower limit on the photon Compton wavelength has improved by four orders of magnitude, to about one astronomical unit.
  25. [25]
    [PDF] γ (photon) - Particle Data Group
    May 31, 2024 · 2WANG 23B use fast radio burst photon mass dependent dispersion relation to determine an upper limit of the photon mass. 3BONETTI 17 uses ...
  26. [26]
    Bounding the photon mass with cosmological propagation of ... - arXiv
    Mar 29, 2021 · The bound on the photon mass will be tighter in the near future with increment in the number of FRBs, more accurate measurement of the redshift ...
  27. [27]
    Experimental Basis of Special Relativity
    In these experiments you will find tests using electromagnetic and nuclear measurements (including both strong and weak interactions). Gravitational tests are ...
  28. [28]
    VIII. A dynamical theory of the electromagnetic field - Journals
    Leung P (2014) On Maxwell's discovery of electromagnetic waves and the gauge condition, European Journal of Physics, 10.1088/0143-0807/36/2/025002, 36:2 ...Missing: 1860s | Show results with:1860s
  29. [29]
    [PDF] The Thermal Radiation Formula of Planck (1900) - arXiv
    Feb 12, 2004 · The fundamental lesson is that the quantum postulate introduces dis- creteness and therefore justifies atomicity. Namely the old hypothesis of.
  30. [30]
    A Quantum Theory of the Scattering of X-rays by Light Elements
    Jan 27, 2025 · Quantum Milestones, 1923: Photons Are Real. Published 27 January, 2025. Arthur Compton's results convinced most skeptics that in some ...
  31. [31]
    The quantum theory of the emission and absorption of radiation
    The quantum theory of dispersion. P. A. M. Dirac, Proceedings A, 1927. On the theory of quantum mechanics. Paul Adrien Maurice Dirac, Proceedings A, 1926. The ...
  32. [32]
    [PDF] Photon: history, mass, charge - arXiv
    Feb 13, 2006 · “History” briefly describes the emer- gence and evolution of the concept of photon during the first two decades of the 20th century. “Mass” ...
  33. [33]
    This Month in Physics History | American Physical Society
    He was also the first to coin the word “photon” to describe the unit of light in a December 18, 1926 letter to Nature. ... Lewis has a street in his hometown of ...
  34. [34]
    [PDF] Photon: New light on an old name - arXiv
    Abstract: After G. N. Lewis (1875-1946) proposed the term “photon” in 1926, many physicists adopted it as a more apt name for Einstein's light quantum.
  35. [35]
    The quantum leap | Nature Materials
    May 1, 2010 · Only in 1926 was the word 'photon' introduced, by the chemist Gilbert Lewis. His theory of a "hypothetical new atom that is not light but plays ...
  36. [36]
    What Is a Photon? Definition and Facts - Science Notes
    Dec 14, 2022 · The symbol hν refers to photon energy, where h is Planck's constant and the Greek letter nu (ν) is the photon frequency. Another symbol is hf, ...
  37. [37]
    [PDF] Feynman Diagrams. - UCLA Physics & Astronomy
    Particles that are not on-shell are said to be off-shell or virtual. 2 Virtual particles are not real: they can never be measured directly.
  38. [38]
    [PDF] Forces via exchange of particles - UF Physics
    The virtual particles are said to be “off their mass shells”. The real particle's E and p must be within the 3- dimentional surface (shell) in the four ...
  39. [39]
    Antiphotons? - Physics Van - University of Illinois Urbana-Champaign
    Some particles are their own antiparticles, notably the force carriers like photons, the Z boson, and gluons, which mediate the electromagnetic force, the weak ...
  40. [40]
    Photon Energies and the Electromagnetic Spectrum | Physics
    Since many wavelengths are stated in nanometers (nm), it is also useful to know that hc = 1240 eV · nm. These will make many calculations a little easier. All ...
  41. [41]
  42. [42]
    [PDF] Lecture 10: Quantum Statistical Mechanics
    Bose-Einstein statistics applies to particles of integer spin, like photons, phonons, vibrational modes, 4He atoms (=2 neutrons, 2 protons and 2 electron) ...
  43. [43]
    [PDF] LECTURE 13 Maxwell–Boltzmann, Fermi, and Bose Statistics
    Now we can calculate the mean occupation number ns = −. 1 β. ∂ lnZ. ∂εs ... As in the Bose–Einstein case, µ is called the chemical potential. This is ...Missing: zero | Show results with:zero
  44. [44]
    None
    ### Summary of Bose-Einstein Distribution for Photons
  45. [45]
    [PDF] Derivation of Planck's Law of Radiation by Satyendranath Bose*
    The idea, introduced by Bose in 1924 for photons, was general- ized to atoms in 1924–25 by Einstein. These ideas later gave rise to Bose-Einstein statistics (BE ...
  46. [46]
    [PDF] ON THE QUANTUM THEORY OF RADIATION
    Einstein, Strahlungs-Emission und Absorption nach der Quantentheorie. Verh. der D. Physikal. Ges. 18 (1916) 318.
  47. [47]
    Population Inversion – gain, upper laser level - RP Photonics
    Population inversion is a state of a medium where a higher-lying electronic level has a higher population than a lower-lying level.
  48. [48]
    Stimulated Emission – gain, amplification, amplifier, laser
    Stimulated emission is a quantum effect, where photon emission is triggered by other photons. It is essential for the operation of lasers.
  49. [49]
    Laser Threshold - RP Photonics
    The threshold of a laser is the state where the small-signal gain just equals the resonator losses, so that laser emission can just begin.
  50. [50]
    Invention of the Maser and Laser - Physics Magazine
    Jan 27, 2005 · Charles Townes and his colleagues were the first to build a “maser,” which operated in the microwave frequency range. It was the precursor of the laser.
  51. [51]
    Physics History | American Physical Society
    It all started with one lone physicist, Theodore Maiman, who defied the doubts of skeptical colleagues to build the first working laser in 1960.
  52. [52]
    Space-Time Approach to Quantum Electrodynamics | Phys. Rev.
    References (28). R. P. Feynman, Phys. Rev. 76, 749 (1949); R. P. Feynman, Phys. Rev. 74, 939 (1948); R. P. Feynman, Phys. Rev. 74, 1430 (1948); J. Schwinger ...
  53. [53]
    Current advances: The fine-structure constant
    The fine-structure constant α is of dimension 1 (i.e., it is simply a number) and very nearly equal to 1/137. It is the "coupling constant" or measure of the ...Missing: citation | Show results with:citation
  54. [54]
    [PDF] arXiv:1810.04341v3 [quant-ph] 29 Mar 2020
    Mar 29, 2020 · For example Penrose[27] defines a virtual particle as follows: A virtual particle is “off-shell” and “can occur only in the interior of a ...
  55. [55]
    Timelike virtual compton scattering from electron-positron radiative ...
    Feb 10, 2010 · It is the production of hadron pairs from photon pairs where one photon is virtual ... Timelike DVCS in QED: radiative muon pair production. The ...
  56. [56]
    Hadronic light-by-light scattering amplitudes from lattice QCD versus ...
    Oct 2, 2018 · The hadronic contribution to the eight forward amplitudes of light-by-light scattering ( 𝛾 * ⁢ 𝛾 * → 𝛾 * ⁢ 𝛾 * ) is computed in lattice QCD.
  57. [57]
    [PDF] arXiv:1408.6628v1 [nucl-th] 28 Aug 2014
    Aug 28, 2014 · the electromagnetic contribution to the proton-neutron mass splitting was δM γ p−n = 0.76 ± 0.30MeV [13, 14]. The recent work of WCM has ...
  58. [58]
    [PDF] Hadron production in two-photon collisions at LEP-L3 - arXiv
    The γγ → hadrons cross sections unfolded with Phojet at √s= 183 GeV and √s= 189 GeV are compared in Figure 2. The data are compatible within systematic errors.
  59. [59]
    [PDF] Large contribution of virtual Delbrück scattering to the emission of ...
    Oct 9, 2007 · Delbrück scattering is an elastic scattering of a photon in the Coulomb field of a nucleus via a virtual electron loop. The contribution of this ...
  60. [60]
    Cosmological Redshift Topics - UNLV Physics
    Cosmological Redshift and Energy Conservation:​​ Recall photon energy E = hf = hc/λ = 1.2398419739(75) eV-μ/λ_μ. This implies that photons propagating through ...
  61. [61]
    32 Refractive Index of Dense Materials - Feynman Lectures - Caltech
    so, Eq. (32.21) represents a wave with the phase velocity vph=ω/k. The index of refraction n is defined (see Chapter 31, Vol. I) by letting vph=cn. Thus Eq.
  62. [62]
    Refractive Index - StatPearls - NCBI Bookshelf
    May 20, 2023 · The index of refraction (n) is an essential parameter in optics that determines the speed by which light travels through a medium other than a vacuum.
  63. [63]
    The Beer-Lambert Law - Chemistry LibreTexts
    Jan 29, 2023 · The Beer-Lambert law relates the attenuation of light to the properties of the material through which the light is traveling.Example 1 · The Beer-Lambert Law · Example 3 · The Importance of Concentration
  64. [64]
    2.3 The Lambert-Beer law and the absorption coefficient α
    The Lambert-Beer law describes how light intensity decreases exponentially with distance (x) through an absorbing medium, where α is the absorption coefficient.Missing: photon | Show results with:photon
  65. [65]
    Einstein coefficients, cross sections, f values, dipole moments, and ...
    The relationships among various parameters describing the strength of optical transitions in atoms and molecules are reviewed. The application of these ...
  66. [66]
    Electromagnetic Signal Attenuation (Skin Depth) | US EPA
    Jan 24, 2025 · Skin depth is the depth at which the EM signal attenuates to 1/e or approximately a third. Skin depth is typically measured in meters.Missing: absorption | Show results with:absorption
  67. [67]
    Skin Effect | SpringerLink
    Aug 8, 2022 · The skin depth δ [m] is the distance an electromagnetic wave has to travel through that absorbing material until its field strength is reduced to 37% of E 0 or ...<|separator|>
  68. [68]
    1.5: Total Internal Reflection - Physics LibreTexts
    Mar 16, 2025 · Total internal reflection occurs for any incident angle greater than the critical angle , and it can only occur when the second medium has an ...Missing: photons | Show results with:photons<|control11|><|separator|>
  69. [69]
    201. 25.4 Total Internal Reflection - University of Iowa Pressbooks
    Total internal reflection occurs when the incident angle is greater than the critical angle. Snell's law states the relationship between angles and indices of ...
  70. [70]
    Blue Sky and Rayleigh Scattering - HyperPhysics
    The blue color of the sky is caused by the scattering of sunlight off the molecules of the atmosphere. This scattering, called Rayleigh scattering, is more ...
  71. [71]
    [PDF] CHAPTER 23 The Interaction of Light with Matter: I - Scattering
    14). The cross-section for Thomson scattering is the Thomson cross section: σs = σT = 8π. 3 r2 e = 8πe4. 3m2 ec4 ≃ 6.65 × 10−25cm2. Note that this cross ...Missing: formula | Show results with:formula
  72. [72]
    Raman Scattering - HyperPhysics
    Raman scattering produces scattered photons which differ in frequency from the radiation source which causes it, and the difference is related to vibrational ...
  73. [73]
    Raman Techniques: Fundamentals and Frontiers - PMC
    The Raman effect originates from the inelastic scattering of light, and it can directly probe vibration/rotational-vibration states in molecules and materials.
  74. [74]
    [PDF] CERN-69-22.pdf
    Møller scattering is the scattering of two electrons on each other. For historical and practical reasons (storage rings), the process is treated, in this and ...
  75. [75]
    Solar Photovoltaic Cell Basics | Department of Energy
    When the semiconductor is exposed to light, it absorbs the light's energy and transfers it to negatively charged particles in the material called electrons.
  76. [76]
  77. [77]
    Detailed Balance Limit of Efficiency of p‐n Junction Solar Cells
    The maximum efficiency is found to be 30% for an energy gap of 1.1 ev and fc = 1. Actual junctions do not obey the predicted current‐voltage relationship, and ...
  78. [78]
    Milestone-Proposal talk:Invention of the Semiconductor Laser, 1962
    The semiconductor laser first illuminated the world in fall 1962, developed independently at four institutions: GE Research Niskayuna, IBM Research, GE ...
  79. [79]
    Low-Loss Optical Fiber - an overview | ScienceDirect Topics
    Within three years, the fiber attenuation reached 0.2 dB/km at 1550 nm [5], which is the typical attenuation value of today's standard single-mode fiber ...
  80. [80]
    CCD Resolution for Optical Microscopy | Nikon's MicroscopyU
    In diffraction limited optical instruments, such as the microscope, the Abbe limit of optical resolution at an average wavelength of 550 nanometers is 0.22 ...Missing: photon | Show results with:photon
  81. [81]
    Diffraction limit - Scientific Volume Imaging
    The numerical aperture (NA) and the resolution limit is schematically illustrated in figure 1. The limit is basically a result of diffraction processes and the ...Missing: CCD sensors
  82. [82]
    The crucial role of lithography in IC fabrication | imec
    Continuous innovations in photolithography, from EUV to NA-EUV, have enabled the semiconductor industry to produce increasingly powerful chips.
  83. [83]
    High-performance semiconductor quantum-dot single-photon sources
    Nov 7, 2017 · This Review describes progress in the fabrication of semiconductor quantum-dot structures, which are approaching the ideal single-photon ...
  84. [84]
    New High-Intensity Source of Polarization-Entangled Photon Pairs
    Dec 11, 1995 · Abstract. We report on a high-intensity source of polarization-entangled photon pairs with high momentum definition. Type-II noncollinear phase ...Missing: URL | Show results with:URL
  85. [85]
    Experimental Realization of Einstein-Podolsky-Rosen-Bohm ...
    Jul 12, 1982 · Experimental Realization of Einstein-Podolsky-Rosen-Bohm Gedankenexperiment: A New Violation of Bell's Inequalities. Alain Aspect, Philippe ...Missing: original | Show results with:original<|separator|>
  86. [86]
    Quantum cryptography: Public key distribution and coin tossing - arXiv
    Mar 14, 2020 · This is a best-possible quality scan of the original so-called BB84 paper as it appeared in the Proceedings of the International Conference ...Missing: URL | Show results with:URL
  87. [87]
    A scheme for efficient quantum computation with linear optics - Nature
    Jan 4, 2001 · Here we show that efficient quantum computation is possible using only beam splitters, phase shifters, single photon sources and photo-detectors.Missing: URL | Show results with:URL
  88. [88]
    A single quantum cannot be cloned - Nature
    Oct 28, 1982 · We show here that the linearity of quantum mechanics forbids such replication and that this conclusion holds for all quantum systems.