Fact-checked by Grok 2 weeks ago

Scattering

Scattering is the physical process by which waves or particles deviate from their original upon interacting with an , potential, or medium, resulting in a redistribution of their , , or both. This phenomenon is fundamental across classical and quantum physics, encompassing interactions such as particles colliding with targets or electromagnetic waves encountering . In classical scattering, typically involving point particles or rigid bodies, the interaction is described using central potentials, where a approaches a stationary with an impact b—the between their asymptotic paths—and emerges at a scattering angle θ. Key quantities include the differential cross section dσ/dΩ, which measures the probability of scattering into a , and the total cross section σ, representing the effective area intercepted by the . conserves , as seen in hard-sphere collisions or by Coulomb potentials, which famously revealed the nuclear structure of atoms in 1911 experiments. Quantum scattering theory extends this framework to wave-like particles, solving the time-independent for scattering states with asymptotic forms featuring incoming plane waves and outgoing spherical waves. The f(θ, φ) determines the differential cross section via |f|², while the optical theorem relates the total cross section to the forward , ensuring unitarity. Elastic processes maintain particle identity and internal states, whereas involves energy transfer, such as or . Scattering manifests in diverse contexts, including electromagnetic wave deflection by particles—altering direction and without —as in atmospheric scattering that produces skies via processes for wavelengths much larger than scatterers. In high-energy physics, experiments analyze scattering cross sections to probe fundamental forces and discover particles, while in , techniques like scattering reveal atomic structures. Other types include for comparable particle sizes to wavelengths and for inelastic vibrational energy shifts.

Basic Concepts

Definition and Overview

Scattering is a fundamental physical process in which propagating or particles are deflected from their original trajectories upon interacting with irregularities in a medium, other particles, or potential fields, leading to a redistribution of their or . This phenomenon occurs across classical and quantum regimes, where in the classical interpretation, it arises from geometric or electromagnetic interactions, while quantum scattering involves probabilistic perturbations. represents one common type, where the incident particle or wave retains its post-interaction, though inelastic processes can also occur with energy transfer. A pivotal advancement came in 1911 with Ernest Rutherford's gold foil experiment, where alpha particles were observed scattering at large angles from thin gold foil, providing direct evidence for the and establishing scattering as a key probe for subatomic structure. This development shifted scattering from qualitative to quantitative , influencing modern experimental methodologies. Illustrative examples abound in natural and experimental settings; for instance, scatters off atmospheric molecules, preferentially redirecting shorter wavelengths to produce the observed during daylight. In high-energy physics, particle collisions in accelerators like those at exploit scattering to investigate fundamental interactions, where beams of protons or electrons collide, and the resulting deflections reveal properties of quarks and other subatomic entities. Such processes underscore scattering's role in probing scales from to . A prerequisite for comprehending scattering, particularly in quantum contexts, is wave-particle duality, the principle that entities like electrons and photons exhibit both wave-like and particle-like localized behaviors depending on the observation. This duality underpins the probabilistic nature of quantum scattering amplitudes, enabling unified descriptions across wave optics and particle collisions, though detailed mathematical frameworks quantify these effects.

Single versus Multiple Scattering

In scattering processes, single scattering refers to an event where an incident particle or wave interacts with only one scattering center, resulting in a single deflection or redirection. This regime predominates in dilute media, where the of scatterers is sufficiently low that subsequent interactions are negligible, such as in collisions with isolated atoms or passage through thin samples. For instance, a beam interacting with a microscopic particle exemplifies single scattering, allowing direct of the scattering properties of that individual center. In contrast, multiple scattering arises from the cumulative effects of numerous interactions with scattering centers, leading to a of the particle's or wave's and often resulting in diffusive transport. This occurs in denser or thicker media, such as light propagation through thick or atmospheres, where the incident entity undergoes repeated deflections, producing a more uniform, hazy distribution. The transition between these regimes is governed by the comparison of the —the average distance between successive scattering events—to the overall system size. The τ, defined as the ratio of the system thickness to the , quantifies this: single scattering dominates when τ ≪ 1, as the probability of multiple interactions is low, and the Beer-Lambert law provides a good for the . Conversely, when τ ≫ 1, multiple scattering prevails, requiring more complex models to account for the enhanced path lengthening and . Practically, single scattering enables precise measurements in techniques like single-scattering , where the direct signal from individual interactions reveals material properties without . Multiple scattering, however, is essential for modeling in applications such as radiation shielding, where repeated interactions in dense materials significantly contribute to overall and energy deposition.

General Theory

Mathematical Framework

The scattering cross-section serves as a fundamental measure of the probability that an incident particle interacts with a target, effectively representing the effective area presented by the target for scattering events. In both classical and quantum frameworks, the total cross-section σ quantifies the overall rate and is obtained by integrating the cross-section over all scattering angles: \sigma = \int \frac{d\sigma}{d\Omega} \, d\Omega, where d\sigma / d\Omega describes the angular distribution of scattered particles. This formulation arises naturally from the of particle and is applicable across various scattering regimes, providing a unified for comparing interaction strengths. In quantum mechanics, the differential cross-section is directly linked to the scattering amplitude f(\theta, \phi), with \frac{d\sigma}{d\Omega} = |f(\theta, \phi)|^2, where the amplitude encodes the quantum interference effects governing the scattering process. The scattering amplitude emerges from the asymptotic behavior of the total wave function far from the scattering center: \psi(\mathbf{r}) \sim e^{i \mathbf{k} \cdot \mathbf{z}} + \frac{f(\theta) e^{i k r}}{r} \quad (r \to \infty), with the first term representing the incident plane wave and the second the outgoing spherical wave, assuming elastic scattering where the magnitudes of the initial and final wave vectors are equal, |\mathbf{k}| = |\mathbf{k}'|. This form captures the transition from free propagation to scattered outgoing waves and forms the basis for exact methods like partial wave analysis. The offers a perturbative approach to compute the for weak scattering potentials, approximating f(\theta) as the first-order term in a : f(\theta) = -\frac{\mu}{2\pi \hbar^2} \int V(\mathbf{r}) e^{i \mathbf{q} \cdot \mathbf{r}} \, d^3\mathbf{r}, where \mu is the , V(\mathbf{r}) is the interaction potential, and \mathbf{q} = \mathbf{k} - \mathbf{k}' is the momentum transfer vector with |\mathbf{q}| = 2k \sin(\theta/2). This of the potential provides a simple, analytically tractable estimate valid when higher-order multiple scatterings are negligible, as originally derived in the foundational development of in . In contrast, classical scattering theory yields explicit formulas for specific potentials, such as the interaction between charged particles. For repulsive scattering of a particle with charge Z_1 e incident on a fixed center with charge Z_2 e and E, the Rutherford formula gives the differential cross-section as \frac{d\sigma}{d\Omega} = \left( \frac{Z_1 Z_2 e^2}{4 E} \right)^2 \frac{1}{\sin^4(\theta/2)}, derived from the trajectories under the inverse-square . This result, which diverges at small angles due to long-range effects, remarkably coincides with the quantum mechanical prediction in the high-energy limit and the first for potentials.

Elastic versus Inelastic Scattering

In scattering processes, elastic scattering occurs when the incident particle interacts with a target without any net transfer of kinetic energy to the target's internal , resulting solely in transfer and of the total in the center-of-mass frame. This restricts the possible outcomes, limiting the scattering to redirection of the particle while preserving its energy, which is particularly useful for probing structural properties without altering the target's state. A representative example is , where coherent from atomic planes in a reveals periodic structures through patterns. In contrast, inelastic scattering involves an exchange of between the incident particle and the target, exciting internal states such as vibrational, rotational, or electronic levels, or even leading to . Here, the scattered particle emerges with reduced (or occasionally increased) , violating strict kinetic conservation due to the deposited in the target. Key examples include , an inelastic process where photons interact with molecular vibrations, shifting the scattered light's to provide information on vibrational spectra, and , where photons lose to free electrons, highlighting electronic interactions. Kinematically, elastic scattering confines the to outcomes where and balance without internal , often yielding discrete angular distributions, whereas inelastic processes open a broader , allowing variable losses and a continuum of scattering angles dependent on the . Experimentally, the distinction between elastic and inelastic scattering is achieved using energy-resolved detectors, which capture the sharp, monochromatic peak corresponding to the unchanged of elastically scattered particles, in contrast to the broadened or shifted spectra from inelastic events where is redistributed. This separation is crucial for applications like , as elastic signals dominate while inelastic components reveal dynamic or excited states. Both types contribute to overall in scattering media, with elastic processes primarily redirecting beams and inelastic ones absorbing .

Attenuation Due to Scattering

Attenuation due to scattering refers to the reduction in the intensity of a propagating caused by particles redirecting photons or away from the original direction, without necessarily absorbing . This contributes to the overall of the beam alongside , and is quantified through the . The for scattering, denoted as \mu_s, is defined as the product of the n of scatterers and the scattering cross-section \sigma_s, such that \mu_s = n \sigma_s. The primary law governing this attenuation is the Beer-Lambert-Bouguer law, which describes the exponential decay of beam intensity I over a path length x as I = I_0 e^{-\mu x}, where I_0 is the initial intensity and \mu is the total attenuation coefficient. This law derives from the probability of single scattering events: consider a thin slab of thickness dx containing n \, dx scatterers per unit area; the fractional loss in intensity due to scattering in this slab is dI / I = -\mu_s \, dx = -n \sigma_s \, dx, assuming the probability of interaction is proportional to the number of scatterers and their effective cross-sectional area. Integrating this differential equation yields the exponential form, valid under the single-scattering approximation where the medium is optically thin (\mu x \ll 1), ensuring negligible probability of multiple interactions per photon. In media with both scattering and absorption, the total attenuation coefficient is the sum \mu = \mu_s + \mu_a, where \mu_a = n \sigma_a is the coefficient analogous to \mu_s. This additive form holds because both processes independently remove energy from the forward , though forward scattering approximations are often applied: small-angle forward-scattered may remain within the beam's acceptance angle and thus not contribute fully to measured , requiring corrections to \mu_s based on the scattering . For instance, in narrow-beam geometries, the effective \mu_s excludes forward-peaked contributions to avoid underestimating . The Beer-Lambert-Bouguer law's exponential decay assumes independent single events and breaks down in optically thick media where multiple scattering dominates, leading to deviations such as enhanced forward diffusion rather than simple attenuation. In such regimes (\mu x \gg 1), diffusion theory provides a more accurate model by approximating the radiance as isotropic and solving the diffusion equation \nabla \cdot (D \nabla I) - \mu_a I + S = 0, where D = 1/(3 \mu_s) is the diffusion coefficient (assuming isotropic scattering) and S represents sources; this captures the random-walk propagation of light through repeated scatters, resulting in slower effective attenuation compared to the exponential law.

Scattering in Quantum Mechanics

Born Approximation

The provides a perturbative approach to calculating the in for particles interacting through a weak potential, building on the general mathematical framework of the . Introduced by in 1926 in his foundational work on quantum collision processes, it simplifies the solution of the time-independent by treating the potential as a small to the free-particle . This method is particularly useful for high-energy scattering where multiple partial waves contribute, contrasting with exact non-perturbative techniques like that are better suited for low energies. The derivation starts from the time-independent for the total \psi(\mathbf{r}): \nabla^2 \psi + k^2 \psi = \frac{2m}{\hbar^2} V(\mathbf{r}) \psi, where k^2 = 2mE / \hbar^2, E is the incident , m is the , and V(\mathbf{r}) is the scattering potential assumed to be central and short-ranged. The unperturbed solution is the incident \psi_0(\mathbf{r}) = e^{i \mathbf{k}_i \cdot \mathbf{r}}, with |\mathbf{k}_i| = k. In the first-order , \psi on the right-hand side is replaced by \psi_0, and the equation is solved using the outgoing G(\mathbf{r}, \mathbf{r}') = - \frac{1}{4\pi} \frac{e^{ik|\mathbf{r} - \mathbf{r}'|}}{|\mathbf{r} - \mathbf{r}'|}. The asymptotic form of the scattered wave in the far field is \psi(\mathbf{r}) \sim e^{i k z} + f(\theta) \frac{e^{ikr}}{r}, where the is given by f(\theta) = -\frac{m}{2\pi \hbar^2} \int V(\mathbf{r}') e^{i \mathbf{q} \cdot \mathbf{r}'} \, d^3\mathbf{r}', with the momentum transfer \mathbf{q} = \mathbf{k}_i - \mathbf{k}_f and |\mathbf{k}_f| = k for elastic scattering. This expression represents the Fourier transform of the potential V(\mathbf{r}') evaluated at spatial frequency \mathbf{q}/(2\pi), providing an intuitive interpretation: the scattering amplitude measures how much the potential's Fourier components contribute to momentum transfer \hbar \mathbf{q}. The approximation is valid when the potential is weak compared to the kinetic energy, specifically when |V(\mathbf{r})| \ll \frac{\hbar^2 k^2}{m} throughout the interaction region, ensuring the perturbation does not significantly distort the incident wave. It also holds well at high incident energies, where the de Broglie wavelength is short relative to the potential's range, minimizing multiple scattering effects; a quantitative criterion for a potential well of depth V_0 and range a is \frac{2m V_0 a^2}{\hbar^2} \ll 1. For stronger potentials or low energies, the approximation breaks down, as higher-order terms become comparable or the series diverges. In applications to potential scattering, the Born approximation has been widely used to model nuclear interactions via the Yukawa potential V(r) = -\frac{\beta e^{-\mu r}}{r}, which approximates the exchange of mesons between nucleons. The resulting scattering amplitude is f(\theta) = \frac{2m \beta}{\hbar^2 (q^2 + \mu^2)}, leading to a differential cross section \frac{d\sigma}{d\Omega} = |f(\theta)|^2 that peaks at forward angles and decreases with scattering angle \theta, consistent with experimental nucleon-nucleon scattering data at intermediate energies. This form reduces to the Rutherford formula in the limit \mu \to 0, validating its use for screened Coulomb-like forces in nuclear physics. Higher-order terms in the Born series, obtained by iteratively substituting the wave function back into the , provide corrections for stronger potentials; the second-order amplitude, for instance, involves a over the potential and can improve accuracy for the Yukawa case but often suffers from issues due to oscillatory integrals or effects in singular potentials. Despite these limitations, the remains a for analytic insights into scattering processes.

Partial Wave Analysis

Partial wave analysis provides an exact framework for solving quantum scattering problems with central potentials by decomposing the incident into spherical waves characterized by quantum number l. This method leverages the of the potential to separate the into radial and angular parts, allowing the scattering to be expressed as a sum over partial waves. Each partial wave contributes independently to the total , with the phase shift \delta_l encoding the effect of the potential on that angular momentum channel. The total wave function \psi(\mathbf{r}) for an incident plane wave along the z-axis can be expanded in terms of Legendre polynomials as \psi(\mathbf{r}) = \sum_{l=0}^{\infty} (2l+1) i^l P_l(\cos \theta) \frac{u_l(r)}{r}, where P_l(\cos \theta) are the Legendre polynomials, and u_l(r) is the radial wave function for angular momentum l. Far from the scattering center, where the potential is negligible, the asymptotic form of the radial function is u_l(r) \sim \sin\left(kr - \frac{l\pi}{2} + \delta_l\right), with k the wave number and \delta_l the phase shift induced by the potential. This phase shift arises from matching the interior solution (inside the potential) to the free spherical Bessel functions outside, ensuring continuity and differentiability at the boundary. The differential scattering cross section is determined by the scattering amplitude f(\theta), given by the partial wave sum f(\theta) = \frac{1}{2ik} \sum_{l=0}^{\infty} (2l+1) \left( e^{2i \delta_l} - 1 \right) P_l(\cos \theta). For elastic scattering without absorption, the phase shifts are real (\operatorname{Im} \delta_l = 0), and the total cross section follows from integrating |f(\theta)|^2 over angles, yielding \sigma = \frac{4\pi}{k^2} \sum_{l=0}^{\infty} (2l+1) \sin^2 \delta_l. In practice, phase shifts can be approximated using methods like the for weak potentials. At low energies, where k \to 0, higher partial waves (l \geq 1) contribute negligibly due to the centrifugal barrier, and s-wave (l=0) scattering dominates. The s-wave phase shift admits the effective range expansion k \cot \delta_0 = -\frac{1}{a} + \frac{1}{2} r_0 k^2 + O(k^4), where a is the scattering length and r_0 the , parameters that characterize the low-energy interaction strength and range, respectively. This , derived from the analytic properties of the scattering amplitude near threshold, is particularly useful for extracting potential parameters from experimental data. Resonances occur when a partial wave phase shift \delta_l passes rapidly through \pi/2 as increases, signaling a quasi-bound state where the particle is temporarily trapped by the potential before escaping. This behavior manifests as a sharp peak in the cross section for that partial wave, \sigma_l \propto 4\pi (2l+1)/k^2 \sin^2 \delta_l, and is a hallmark of near-threshold bound states or short-lived intermediates in scattering processes.

Scattering in Electrodynamics

Rayleigh Scattering

Rayleigh scattering describes the of electromagnetic waves by particles much smaller than the of the incident , specifically in the regime where the scatterer size parameter a \ll \lambda / (2\pi), with a being the particle radius and \lambda the . This approximation treats the scatterer as an induced oscillating electric , neglecting higher-order multipoles. The process is dominant for molecular-scale scatterers, such as air molecules, under visible illumination. In this dipole approximation, the scatterer's response to the incident \mathbf{E}_\mathrm{inc} is characterized by its \alpha, which quantifies the induced \mathbf{p} = \alpha \mathbf{E}_\mathrm{inc}. For a small , the is given by \alpha = 4\pi \epsilon_0 a^3 \frac{\epsilon_r - 1}{\epsilon_r + 2}, where \epsilon_0 is the and \epsilon_r is the of the . This expression, derived from the Clausius-Mossotti for dilute systems, assumes the particle is homogeneous and non-absorbing. The scattered field arises from the radiation of this induced oscillating dipole. For a time-harmonic incident field with frequency \omega = 2\pi c / \lambda, the far-field of the dipole radiation in the radiation zone (r \gg \lambda) is \mathbf{E}_\mathrm{sc} = \frac{k^2}{4\pi \epsilon_0} \frac{[\mathbf{n} \times \mathbf{p}] \times \mathbf{n}}{r} e^{i(kr - \omega t)}, where \mathbf{p} is the , \mathbf{n} is the unit in the of , k = 2\pi / \lambda = \omega / c, c is the , and r is the distance from the scatterer. The \sin\theta angular dependence emerges from the cross-product terms, with \theta the angle between the incident field and the scattering , peaking at 90° to the incident . Since \mathbf{p} \propto \alpha \mathbf{E}_\mathrm{inc} and \mathbf{E}_\mathrm{inc} \propto E_0 e^{-i\omega t}, the scattered I_\mathrm{sc} \propto |\mathbf{E}_\mathrm{sc}|^2 \propto |\alpha|^2 \omega^4 / r^2. The differential scattering cross-section, which measures the scattered power per unit solid angle normalized by the incident , follows as \frac{d\sigma}{d\Omega} = \frac{k^4}{16\pi^2 \epsilon_0^2} |\alpha|^2 \sin^2\theta \propto \frac{|\alpha|^2}{\lambda^4} \sin^2\theta. For non-absorbing dielectrics, |\alpha|^2 \propto a^6, yielding d\sigma / d\Omega \propto a^6 / \lambda^4. This strong inverse fourth-power dependence on wavelength explains why shorter () wavelengths scatter more efficiently than longer () ones by a factor of approximately (700/400)^4 \approx 9.4 (roughly 10) for violet versus light in the . In Earth's atmosphere, by and oxygen molecules preferentially scatters sunlight, rendering the daytime sky while allowing direct redder sunlight to reach observers. Applications of Rayleigh scattering span , where it underpins the explanation of sky color and patterns observed since the . In , Rayleigh lidar systems exploit molecular scattering to profile atmospheric density and temperature in clean air, serving as a baseline for air quality monitoring by distinguishing molecular signals from aerosol-induced in polluted conditions. Rayleigh contributions also drive wavelength-dependent in optical propagation, with \mu \propto 1/\lambda^4, influencing clear-sky models.

Mie Scattering

Mie scattering refers to the analytical solution of for the electromagnetic scattering of a by a homogeneous, isotropic whose size is comparable to the of the incident . This theory provides an exact description for spherical particles, extending beyond the approximations valid for much smaller or larger particles. It is particularly relevant for understanding interactions with atmospheric aerosols, droplets, and colloidal suspensions, where particle diameters range from submicron to tens of micrometers. The solution begins with the expansion of the electromagnetic fields in spherical coordinates using , which separate the fields into transverse electric () and transverse magnetic (TM) modes. For an incident , the scattered fields are expressed as infinite series of these harmonics, with coefficients determined by matching boundary conditions at the sphere's surface— of the tangential electric and magnetic fields. The scattering coefficients a_n (for TM modes) and b_n (for TE modes) are given by: a_n = \frac{m \psi_n(x) \psi_n'(mx) - \psi_n(mx) \psi_n'(x)}{m \psi_n(x) \xi_n'(mx) - \xi_n(mx) \psi_n'(x)}, \quad b_n = \frac{\psi_n(x) \psi_n'(mx) - m \psi_n(mx) \psi_n'(x)}{\psi_n(x) \xi_n'(mx) - m \xi_n(mx) \psi_n'(x)} where \psi_n and \xi_n are Riccati-Bessel functions, x = 2\pi a / \lambda is the size parameter (with a the sphere radius and \lambda the wavelength), and m is the complex refractive index of the sphere relative to the surrounding medium. These coefficients fully characterize the scattered field amplitudes for each multipole order n. Key observables in Mie theory include the extinction efficiency Q_\mathrm{ext}, which quantifies the total cross-section for scattering plus normalized by the geometric area \pi a^2: Q_\mathrm{ext} = \frac{2}{x^2} \sum_{n=1}^\infty (2n+1) \mathrm{Re}(a_n + b_n). The scattering efficiency Q_\mathrm{sca} follows analogously from the imaginary parts, while the asymmetry parameter g describes forward-backward scattering balance. For non-absorbing particles (m real), Q_\mathrm{ext} = Q_\mathrm{sca}, and the theory predicts oscillations in efficiency curves due to between diffracted and surface-reflected waves. In the Rayleigh limit (x \ll 1), Mie theory reduces to the dipole approximation of , where higher-order terms vanish and a_1 dominates, yielding Q_\mathrm{sca} \propto x^4. Conversely, for large particles (x \gg 1), the solution approaches geometric , with Q_\mathrm{ext} \approx 2 due to around the sphere contributing equally to shadow scattering; and patterns emerge from ray-tracing interpretations of the series terms. These asymptotic behaviors bridge small-particle and large-particle . Computationally, evaluating Mie coefficients requires summing the series until convergence, which is rapid for x < 100 (typically n_\mathrm{max} \approx x + 4x^{1/3} + 2) but demands careful handling of spherical Bessel functions to avoid numerical instability, often using logarithmic derivatives or upward recurrence relations. Software implementations, such as those in Python's PyMieScatt or MATLAB toolboxes, facilitate rapid calculation of size distributions and polarization effects. Applications abound in atmospheric science, where Mie theory models the angular distribution of scattered sunlight in clouds to retrieve droplet sizes, and in aerosol optics for remote sensing via lidar, as validated by comparisons with laboratory measurements of polystyrene spheres. Convergence issues arise for very large x > 10^4, prompting hybrid methods combining series with asymptotic expansions.

Other Applications

Acoustic Scattering

Acoustic scattering refers to the redirection of sound waves upon encountering obstacles in a fluid medium, such as air or water, governed by the principles of wave propagation in acoustics. The fundamental equation describing acoustic wave propagation is the linear wave equation for pressure p: \nabla^2 p - \frac{1}{c^2} \frac{\partial^2 p}{\partial t^2} = 0, where c is the speed of sound in the medium. For time-harmonic waves of the form p(\mathbf{r}, t) = \Re \{ \psi(\mathbf{r}) e^{-i \omega t} \}, with angular frequency \omega, this reduces to the Helmholtz equation: \nabla^2 \psi + k^2 \psi = 0, where k = \omega / c is the wavenumber. The total field is expressed as the sum of the incident plane wave and the scattered field, which satisfies the Sommerfeld radiation condition at infinity to ensure outgoing waves. This framework adapts classical wave scattering theory to scalar acoustic fields in fluids, analogous to the scattering amplitude in quantum mechanics but without particle-specific quantum effects. A key approximation for computing acoustic scattering from rigid obstacles is the Kirchhoff approximation, which assumes high-frequency incidence where the wavelength is much smaller than the obstacle size. For rigid bodies, the boundary conditions are either the pressure-release condition p = 0 (soft scatterer) or the normal velocity condition \partial p / \partial n = 0 (hard scatterer) on the surface. The scattered field is then approximated by integrating over the illuminated surface, yielding the far-field amplitude as: \psi(\mathbf{r}) \approx -\frac{e^{i k r}}{r} \int_S \left( \psi_i(\mathbf{x}) \frac{\partial e^{-i k \hat{r} \cdot \mathbf{x}}}{\partial n} - e^{-i k \hat{r} \cdot \mathbf{x}} \frac{\partial \psi_i(\mathbf{x})}{\partial n} \right) dS, where \psi_i is the incident field, S is the scatterer surface, and \hat{r} is the observation direction. This method provides efficient estimates for scattering cross sections from complex geometries like rough surfaces or vehicles. For spherical scatterers, partial wave analysis offers an exact series expansion in spherical harmonics, similar to quantum scattering but applied to scalar potentials. At low frequencies, where k a \ll 1 with a the characteristic scatterer size, the scattering is dominated by and contributions from volume and surface effects, respectively. The scattering length a_s, defined as the effective low-frequency of the phase shift in the s-wave, characterizes the strength of isotropic scattering, leading to the total scattering cross section \sigma = 4 \pi a_s^2. For a rigid , a_s equals the , yielding \sigma = 4 \pi a^2, four times the geometric cross section due to destructive interference in the forward direction. Acoustic scattering principles underpin diverse applications, including for sonar systems, where scattering from submarines or seafloor features informs target detection and imaging algorithms. In , scattering models optimize noise barriers along highways, reducing propagated sound by altering diffraction patterns at barrier edges. imaging relies on scattering from tissue inhomogeneities to generate contrast in echograms, enabling non-invasive diagnostics.

Nuclear and Particle Scattering

Nuclear and particle scattering encompasses processes where subatomic particles interact with atomic nuclei or other particles, revealing fundamental structures and forces at microscopic scales. A seminal example is , observed in experiments conducted by and under Ernest Rutherford's direction between 1909 and 1913, where alpha particles from radioactive sources were directed at thin gold foil, and their deflection patterns indicated a concentrated positive charge within the atom. This scattering confirmed the existence of a dense , overturning the of the atom by demonstrating that most alpha particles passed undeflected while a small fraction scattered at large angles, consistent with Coulomb repulsion from a point-like positive charge. The classical differential cross-section for this process, derived from Rutherford's analysis, is given by \frac{d\sigma}{d\Omega} = \left( \frac{Z_1 Z_2 e^2}{8\pi \epsilon_0 E} \right)^2 \frac{1}{\sin^4(\theta/2)}, where Z_1 and Z_2 are the atomic numbers of the projectile and target, e is the elementary charge, \epsilon_0 is the vacuum permittivity, E is the kinetic energy of the incident particle, and \theta is the scattering angle; this formula accurately predicted the observed angular distribution. In the realm of quantum electrodynamics, Compton scattering represents a key inelastic process involving the interaction of photons with electrons, first discovered by Arthur Holly Compton in 1923 through experiments scattering X-rays off graphite and measuring the wavelength shift in the scattered radiation. This effect provided crucial evidence for the particle nature of light, as the scattered photon's wavelength increases according to the formula \Delta \lambda = \frac{h}{m_e c} (1 - \cos \theta), where h is Planck's constant, m_e is the electron mass, c is the speed of light, and \theta is the scattering angle, reflecting the conservation of energy and momentum in a photon-electron collision treated as particle-particle scattering. Compton's observations, which deviated from classical Thomson scattering predictions, earned him the 1927 Nobel Prize in Physics and underscored the quantum mechanical description of light-matter interactions. Deep inelastic scattering (DIS) experiments at the Stanford Linear Accelerator Center (SLAC) in the late 1960s and 1970s, led by researchers including Jerome Friedman, Henry Kendall, and Richard Taylor, probed the internal structure of protons by accelerating electrons to high energies and colliding them with hydrogen or deuterium targets. These experiments revealed that protons are composed of point-like constituents, interpreted through Richard Feynman's parton model, where the scaling behavior of structure functions indicated quasi-free scattering off fractionally charged partons (later identified as quarks). The results, which showed deviations from elastic scattering and supported the idea of dynamically confined quarks, laid the groundwork for quantum chromodynamics (QCD), the theory of the strong interaction, and earned the SLAC team the 1990 Nobel Prize in Physics. In contemporary high-energy physics, proton-proton scattering at the (LHC) has enabled the discovery of the in 2012 by the ATLAS and collaborations, where the particle was produced via gluon fusion or vector boson fusion processes within the collisions and decayed into observable final states like diphotons or four leptons. This breakthrough confirmed the mechanism for electroweak in the , with the Higgs mass measured at approximately 125 GeV. Additionally, scattering processes in experiments, such as those at T2K and , detect neutrino interactions in near and far detectors to measure oscillation parameters, revealing mixing angles and mass differences that indicate neutrinos have non-zero mass and challenge aspects of the . These modern applications continue to test QCD predictions and search for through precise scattering cross-section measurements.

References

  1. [1]
    [PDF] Scattering
    Scattering is the redirection of radiation out of its original direction, usually due to interactions with molecules and particles.
  2. [2]
    [PDF] Introduction to Scattering Theory
    Scattering theory is time-independent perturbation theory applied to a continuous spectrum, aiming to solve the full energy-eigenstate problem.
  3. [3]
    [PDF] Scattering - UCSB Physics
    Scattering is when particles, due to a repulsive potential, move away from each other after approaching, and never meet again.
  4. [4]
    [PDF] Quantum Physics III Chapter 7: Scattering - MIT OpenCourseWare
    Feb 7, 2019 · We will focus on elastic scattering of particles without spin in the nonrelativistic approximation. We will also assume that the interaction ...
  5. [5]
    [PDF] Scattering - NASA Technical Reports Server (NTRS)
    Scattering is the modification of electromagnetic energy's direction and polarization when a particle is introduced into a beam of light.
  6. [6]
    [PDF] Introduction to Scattering Theory and Scattering from Central Force ...
    Scattering is a fundamental process, often elastic, where particles gain/lose kinetic energy by interacting with a potential, and is important in particle  ...
  7. [7]
    The Science of Color - Smithsonian Libraries
    In the 1660s, English physicist and mathematician Isaac Newton began a series of experiments with sunlight and prisms. He demonstrated that clear white ...
  8. [8]
    [PDF] LXXIX. The scattering of α and β particles by matter and the structure ...
    To cite this Article Rutherford, E.(1911) 'LXXIX. The scattering of α and β particles by matter and the structure of the atom', Philosophical Magazine ...
  9. [9]
    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 ...
  10. [10]
    Wave-Particle Duality – University Physics Volume 3
    Wave-particle duality means particles and radiation can act as both particles and waves depending on experimental conditions.
  11. [11]
    Scattering: An Overview - AZoOptics
    Aug 9, 2024 · Types of Scattering · Rayleigh Scattering · Mie Scattering · Tyndall Scattering · Raman Scattering.<|control11|><|separator|>
  12. [12]
    Multiple scattering limit in optical microscopy
    ### Summary of Multiple Scattering Limit, Mean Free Path, and Transition from Single to Multiple Scattering
  13. [13]
    [PDF] Multiple scattering of light and some of its observable consequences
    If the medium is uniform, the optical depth, which is dimensionless, is z(x+B), and may be interpret- ed as the depth in units of total mean free path 1/(x+B).
  14. [14]
  15. [15]
    Single-scattering spectroscopy for the endoscopic analysis of ...
    We report on the development of an optical-fiber-based diagnostic tool that is sensitive to single-scattering events close to the fiber-optic probe tip.
  16. [16]
    Revealing enhanced X-Ray radiation shielding of 2D layered ...
    Oct 13, 2023 · The X-ray shielding by these 2D structure is explained to be a result of the multiple scattering and interactions (reflection) of the photons ...
  17. [17]
    [PDF] On the quantum mechanics of collision processes.
    “Zur Quantenmechanik der Stoßvorgänge,” Zeit. Phys. 37 (1926), 863-867. On the quantum mechanics of collision processes. [Preliminary announcement (. 1. )] By ...
  18. [18]
    Elastic and Inelastic Scattering - Richard Fitzpatrick
    Elastic scattering conserves particle number, while inelastic scattering does not, and may involve absorption or emergence of particles.
  19. [19]
    Elastic and Inelastic Scattering | nuclear-power.com
    In an elastic scattering reaction, there is no energy transferred into nuclear excitation. Besides, in an inelastic scattering reaction, nuclear excitation ...
  20. [20]
  21. [21]
    [PDF] arXiv:1509.08695v1 [physics.atom-ph] 29 Sep 2015
    Sep 29, 2015 · We stress that for elastic scattering the initial and final atomic states are the same, while for inelastic scattering the final atomic state ...
  22. [22]
    [PDF] Compton Scattering from Quantum Electrodynamics
    Jan 17, 2014 · Compton scattering is the inelastic scattering of a photon with an electically charged particle, first discovered in 1923 by Arthur Compton [1].
  23. [23]
    [PDF] Aspects of electron scattering, the elastic, and the inelastic. - arXiv
    Sep 19, 2023 · We talked about inelastic scattering electrons. But electrons can elastically scatter from nuclie, change direction, maintaining their kinetic ...
  24. [24]
    Effects of multiple elastic and inelastic scattering on energy-resolved ...
    For geometries where the incoming beam is glancing, the contrast first gradually increases with energy loss and decreases slowly for losses larger than 100 eV.
  25. [25]
    Separation of the inelastic and elastic scattering in time-of-flight ...
    The pulsed beam and the time-of-flight data acquisition enable the separation of elastic and inelastic scattering from hydrogenous samples.
  26. [26]
    2.3: Absorption, Scattering and Attenuation Coefficients
    Jul 14, 2021 · The atomic (or molecular) absorption, scattering and extinction coefficients are respectively α/N, σ/N and ε/N, where N is the number density.
  27. [27]
    [PDF] Atmospheric Transmission Beer's Law
    Mar 31, 2010 · • This is due to the extreme asymmetry in forward-backward scattering by cloud droplets, and the dominance of single (not multiple) scattering.
  28. [28]
    The Single-Scattering Approximation - Ocean Optics Web Book
    May 3, 2021 · Solution (4a) is simply the Lambert-Beer law: the initial unscattered radiance decays exponentially with optical depth. Using (BC1) to ...
  29. [29]
    [PDF] Multiple scattering as a diffusion process - Dynamic Graphics Project
    Multiple scattering in participating media is generally a complex phenomenon. In the limit of an optically thick medium, i.e., when the mean free path of ...
  30. [30]
    Zur Quantenmechanik der Stoßvorgänge | Zeitschrift für Physik A ...
    Durch eine Untersuchung der Stoßvorgänge wird die Auffassung entwickelt, daß die Quantenmechanik in der Schrödingerschen Form nicht nur.
  31. [31]
    [PDF] arXiv:2102.13105v1 [quant-ph] 25 Feb 2021
    Feb 25, 2021 · This paper discusses three ways to obtain the Born approximation for Coulomb scattering, including standard, Oppenheimer's, and Landau and ...
  32. [32]
    [PDF] Scattering Theory: Born Series - JuSER
    The focus lies on the introduction of the description of the scattering process in terms of the Hamiltonian of the scattering projectile at a finite range ...
  33. [33]
    [PDF] Copyright cG 2021 by Robert G. Littlejohn
    The Lippmann-Schwinger equation also easily leads to a simple approximation scheme, the Born series, as well as to an intuitive and quantitative criterion for ...
  34. [34]
    On higher Born approximations in potential scattering - Journals
    In this paper, the second Born approximation for scattering of a Dirac electron by a Yukawa potential is calculated by a correct method.
  35. [35]
    [PDF] Born approximation for scattering of wave packets on atoms I ... - arXiv
    Jul 31, 2015 · In the next section we remind the stan- dard Born approximation including well-know formulae for the scattering on the Gaussian potential and on ...
  36. [36]
    [PDF] QUANTUM MECHANICS-Non-Relativistic Theory
    L. D. Landau and E. M. Lifshitz. Institute of Physical Problems USSR. Academy of Sciences. Quantum Mechanics. Pergamon. Pergamon Press. Page 2. QUANTUM ...
  37. [37]
    Theory of the Effective Range in Nuclear Scattering | Phys. Rev.
    The effective range (r0) is a parameter describing neutron scattering by protons, alongside the scattering length at zero energy (a). It is also used in proton ...
  38. [38]
    The Molecular Atmosphere (Chapter 3) - Lidar Engineering
    Feb 16, 2023 · Rayleigh scattering is due to particles much smaller than the wavelength, in the Mie regime the particle size is comparable to the wavelength, ...3.1. 1 Scattering Formalism · 3.3 Molecular Energy Effects · 3.3. 3 Raman Scattering
  39. [39]
    [PDF] The Wave and Helmholtz Equations
    Feb 3, 2006 · It states that the scattered field consists of outgoing waves only. Solutions of the Helmholtz equation which satisfy the radiation condition ...
  40. [40]
    1.3 Acoustics and the Helmholtz Equation - Boundary Element Method
    In this Section it is shown how the time-dependent wave equation governing the acoustic field can be simplified to the Helmholtz equation when harmonic ...
  41. [41]
    [PDF] Scattering of time-harmonic acoustic waves: Helmholtz equation ...
    Feb 28, 2022 · ... sound- soft scattering problems for the homogeneous Helmholtz equation in two dimensions. The Helmholtz equation ∆u + k2u = 0 is relevant ...
  42. [42]
    Acoustic scattering comparison of Kirchhoff approximation to ...
    Sep 11, 2018 · The Helmholtz-Kirchhoff approximation (HKA) is often used for scattering calculations from general sea surfaces.
  43. [43]
    Low‐frequency acoustic scattering by slender bodies of arbitrary ...
    Jul 1, 1992 · The scattering of acoustic waves from soft and hard slender bodies with arbitrary cross sections is considered in the limit ka≪1 and kl∼1, ...
  44. [44]
    [PDF] Acoustic Scattering from a Sphere - BYU Physics and Astronomy
    Nov 24, 2006 · 1. Differential cross section for scattering from an acoustically soft sphere of radius 0.1 wavelengths. 13.
  45. [45]
    An Efficient Acoustic Scattering Model Based on Target Surface ...
    Oct 13, 2020 · 3 Acoustic Scattering Problem. The acoustic scattering problem has special importance in underwater applications including sonar imaging.
  46. [46]
    Acoustic wave scattering by highway noise barriers - AIP Publishing
    Aug 12, 2005 · In this paper we wish to analyze acoustic wave scattering by highway barriers of different shape and height. The sources considered will be ...
  47. [47]
    Ultrasound Physics and Instrumentation - StatPearls - NCBI Bookshelf
    Mar 27, 2023 · A common application to illustrate posterior acoustic enhancement is bladder ultrasound, where time gain compensation must often be ...
  48. [48]
    Rutherford Scattering - Galileo
    Rutherford's alpha scattering experiments were the first experiments in which individual particles were systematically scattered and detected.
  49. [49]
    Rutherford Scattering - HyperPhysics
    rmin= x10^ m = fermi. Rutherford scattering was the first method used to measure the size of nuclei. More precise measurements are made with electron ...Missing: original | Show results with:original
  50. [50]
    [PDF] A Quantum Theory of the Scattering of X-Rays by Light Elements
    Compton, Bull. Nat ... It is well known that for soft X-rays scattered by light elements the total scattering is in accord with Thomson's formula.Missing: discovery | Show results with:discovery
  51. [51]
    Compton Scattering - Nondestructive Evaluation Physics : X-Ray
    Compton effect was first observed by Arthur Compton in 1923 ... The applet below demonstrates Compton scattering as calculated with the Klein-Nishina formula ...
  52. [52]
    [PDF] DEEP INELASTIC SCATTERING - Nobel Prize
    QCD in conjunction with electroweak theory, which describes the interactions of leptons and quarks under the influence of the combined weak and electromagnetic ...<|separator|>
  53. [53]
    [PDF] DEEP INELASTIC SCATTERING
    THE PARTON MODEL. Model by Feynman and Bjorken: Partons ≡ constituents of proton. Today: ☞ charged partons = quarks. ☞ neutral partons = gluons. The model ...
  54. [54]
    The discovery and measurements of a Higgs boson - Journals
    Jan 13, 2015 · In July 2012, the ATLAS and CMS collaborations at CERN's Large Hadron Collider announced the discovery of a Higgs-like boson, a new heavy particle.
  55. [55]
    The Higgs boson: a landmark discovery - ATLAS Experiment
    On 4 July 2012, the ATLAS and CMS experiments at CERN announced that they had independently observed a new particle in the mass region of around 125 GeV: a ...Missing: scattering | Show results with:scattering
  56. [56]
    Joint neutrino oscillation analysis from the T2K and NOvA experiments
    Oct 22, 2025 · Neutrino oscillation remains the most powerful experimental tool for addressing many of these questions, including whether neutrinos violate ...