Fact-checked by Grok 2 weeks ago

Elastic scattering

Elastic scattering is a process in physics where an incident particle or wave is deflected by a without any alteration to the states of either the or the , resulting in the of total in the center-of-mass frame. This contrasts with , where energy is transferred to excite , such as vibrational or rotational modes in molecules or nuclear excitations. In , elastic scattering is characterized by the of both energy and momentum, often analyzed through trajectories determined by the impact parameter and scattering angle, as seen in the Coulomb potential where the differential cross-section follows the Rutherford formula. Quantum mechanically, it involves identical initial and final states for the particles, described by the f_k(\theta, \phi) and differential cross-section \frac{d\sigma}{d\Omega} = |f_k(\theta, \phi)|^2, with applications extending to phase-shift analysis for central potentials. Historically, elastic scattering gained prominence through Ernest Rutherford's 1911 experiments, where alpha particles scattered by gold foils revealed the dense , supporting the planetary model of the atom via the inverse-square interaction. In , it serves as the simplest collision process between nuclei, particularly at low energies near the , where it probes surface interactions and matter distributions without compound nucleus formation. For neutrons, elastic scattering with light nuclei like efficiently moderates their energy in nuclear reactors by transferring through head-on collisions. Beyond particles, elastic scattering applies to electromagnetic waves, such as of light by atmospheric molecules, which explains the blue color of the sky due to wavelength-dependent deflection without photon absorption. In condensed matter, elastic scattering off cores influences and is central to techniques like , where patterns reveal crystalline structures. These processes underpin scattering theory, a cornerstone of , enabling the extraction of interaction potentials from experimental cross-sections via methods like the or coupled-channel calculations.

General Principles

Definition and Characteristics

Elastic scattering is a collision process between two particles or entities in which the total in the center-of-mass frame is conserved, with the resulting only in a change of direction rather than any loss or gain of energy./03%3A_A_Few_Simple_Problems/3.05%3A_Elastic_Scattering) This conservation implies that no internal , such as vibrational or rotational modes in molecules, are excited during the . Key characteristics of elastic scattering include the strict adherence to the laws of conservation of both and linear , ensuring that the magnitudes of the particles' velocities in the center-of-mass frame remain unchanged post-collision, while their directions are altered by the interaction potential. Unlike inelastic processes, elastic scattering involves no creation or absorption of particles and no conversion of kinetic into other forms, such as or . This phenomenon applies not only to point-like particles, such as electrons or nuclei, but also to wave phenomena, including the of by small particles where the remains unaltered. The concept of elastic scattering was first recognized in the 19th-century , where collisions between molecules were modeled as to explain pressure and diffusion without energy dissipation. It received formalization in through Ernest Rutherford's 1911 analysis of alpha-particle by gold foil, which demonstrated large-angle deflections consistent with elastic interactions from a concentrated positive charge. In and molecular contexts, elastic scattering occurs in low-energy collisions where no takes place, such as ground-state atom-atom interactions that merely redirect momenta without promoting electrons to higher orbitals. In contrast, in similar systems involves energy transfer leading to or , altering the internal states of the particles involved./05%3A_Collisions/5.01%3A_Introduction)

Kinematics of Elastic Collisions

In elastic collisions, the kinematics are governed by the conservation of both linear momentum and kinetic energy, assuming non-relativistic particles and no internal excitation. For two particles with initial momenta \vec{p}_1 and \vec{p}_2, and final momenta \vec{p}_1' and \vec{p}_2', momentum conservation requires \vec{p}_1 + \vec{p}_2 = \vec{p}_1' + \vec{p}_2'. Kinetic energy conservation, expressed in terms of magnitudes for particles of masses m_1 and m_2, yields \frac{p_1^2}{2m_1} + \frac{p_2^2}{2m_2} = \frac{p_1'^2}{2m_1} + \frac{p_2'^2}{2m_2}. These laws constrain the possible post-collision trajectories, ensuring that the relative velocity along the line of impact reverses while the component perpendicular to it remains unchanged. To simplify the analysis, particularly for scattering problems where one particle (say, the target) is initially at rest, the center-of-mass (CM) frame is often employed. In this frame, the total momentum is zero, achieved by transforming to velocities \vec{v}_1' = \vec{v}_1 - \vec{v}_{CM} and \vec{v}_2' = \vec{v}_2 - \vec{v}_{CM}, where \vec{v}_{CM} = \frac{m_1 \vec{v}_1 + m_2 \vec{v}_2}{m_1 + m_2}. Here, the particles approach each other with equal and opposite momenta, and after an elastic collision, they recede with the same magnitudes but deflected by a scattering angle \theta_{CM} relative to their initial direction in the CM frame. This frame highlights the symmetry of the interaction, as the magnitudes of the momenta remain unchanged post-collision due to . The observed scattering angles in the laboratory (lab) frame, where the target is at rest, differ from those in the CM frame, necessitating a transformation. For the incident particle of mass m_1 scattering off a stationary target of mass m_2, the relation is \tan \theta_{lab} = \frac{\sin \theta_{CM}}{\cos \theta_{CM} + m_1 / m_2}, derived from vector addition of the CM velocities to the lab frame velocity of the CM. This kinematic mapping is crucial for interpreting experimental data, as detectors typically measure lab-frame angles. When m_1 > m_2, the incident particle cannot be scattered by more than a maximum \theta_{max} = \sin^{-1}(m_2 / m_1) in the lab frame, arising from the of . To see this, consider the post-collision momenta: the target's momentum \vec{p}_2' must lie within a bounded by the incident direction, and the incident particle's deflection is limited by the condition that \vec{p}_1' cannot exceed the where \sin \theta_{lab} = m_2 / m_1, obtained by maximizing \theta_{lab} subject to the equations. For equal masses (m_1 = m_2), the maximum lab-frame scattering for the incident particle is $90^\circ, and in a , the entire is transferred to the , with the incident particle coming to .

Classical Description

Scattering in Central Potentials

In classical mechanics, elastic scattering occurs when a particle interacts with a central potential V(r) that depends only on the radial distance r from the scattering center. The force derived from this potential is \vec{F} = -\nabla V(r), which is conservative and spherically symmetric, ensuring that both energy E and angular momentum \vec{L} are conserved during the interaction. Due to the central nature of , the particle's lies in a plane, and for repulsive potentials where E > V(r) for all r, the unbound motion results in paths that approach from , deflect by a \theta, and recede to . Angular momentum conservation implies L = \mu v_\infty b, where \mu is the , v_\infty is the initial speed at , and b is the impact parameter—the perpendicular distance between the initial velocity vector and the line to the scattering center. Equivalently, b = L / p, with p = \mu v_\infty the initial linear . The relationship between the impact parameter b and the scattering angle \theta is given by the classical scattering function: \theta(b) = \pi - 2b \int_{r_{\min}}^\infty \frac{dr}{r^2 \sqrt{1 - \frac{V(r)}{E} - \frac{b^2}{r^2}}}, where r_{\min} is the , determined as the largest root of $1 - V(r_{\min})/E - b^2/r_{\min}^2 = 0. This arises from integrating the orbital equation in polar coordinates, using the conserved quantities to express the deflection in terms of the . For specific potentials, the scattering function simplifies. In hard-sphere scattering, where the potential is infinite for r < a (with a the sphere radius) and zero otherwise, the trajectory reflects specularly at the surface, yielding \theta = \pi - 2 \sin^{-1}(b/a) for b \leq a, and no scattering (\theta = 0) for b > a. For the Coulomb potential, which is repulsive and proportional to $1/r, the trajectories are conic sections (hyperbolas), providing a preview of the exact hyperbolic paths analyzed in detail under ; the scattering angle increases with potential strength and decreases with energy or impact parameter. In attractive potentials, the scattering function \theta(b) can exhibit non-monotonic behavior, leading to rainbow scattering where d\theta/db = 0 at certain b, corresponding to a caustic in the trajectory density and an enhanced intensity at the rainbow angle.

Differential and Total Cross-Sections

In classical elastic scattering, the differential cross-section \frac{d\sigma}{d\Omega} quantifies the probability density for particles to be scattered into a specific solid angle d\Omega, derived from the classical trajectories governed by central potentials. Particles with impact parameters between b and b + db are deflected through angles between \theta and \theta + d\theta, where \theta(b) is determined from the orbital equation of motion. The incident flux intercepted by the annular area $2\pi b \, db yields the scattered flux into d\Omega = 2\pi \sin\theta \, d\theta, resulting in the formula \frac{d\sigma}{d\Omega} = b \left| \frac{db}{d\theta} \right| \frac{1}{\sin \theta}. This expression links the geometric impact parameter distribution to the angular scattering probability. The total cross-section \sigma, representing the overall effective scattering area, is obtained by integrating over all directions: \sigma = \int \frac{d\sigma}{d\Omega} \, d\Omega = 2\pi \int_0^\pi b \left| \frac{db}{d\theta} \right| \sin \theta \, d\theta. Both quantities have units of area (e.g., barns in contexts) and characterize the interaction strength independently of the beam intensity. For a hard-sphere potential with scatterer a, the relation \theta = \pi - 2 \arcsin(b/a) gives an isotropic \frac{d\sigma}{d\Omega} = a^2/4, and thus \sigma = \pi a^2, matching the geometric projection and remaining energy-independent as long as b < a leads to contact. For an inverse-square repulsive potential V(r) = k/r with k > 0, such as the classical Coulomb repulsion between charged particles, the impact parameter is b = (k/(2E)) \cot(\theta/2), where E is the initial kinetic energy in the center-of-mass frame. Substituting yields the differential cross-section \frac{d\sigma}{d\Omega} = \left( \frac{k}{4E} \right)^2 \frac{1}{\sin^4 (\theta/2)}, which diverges at small \theta due to the long-range nature of the force, causing the total cross-section \sigma to diverge logarithmically from contributions at large b. This small-angle dominance previews the Rutherford scattering formula in nuclear physics. In applications to , such as the of dilute classical gases, the standard total cross-section is modified to a transport cross-section that weights by momentum transfer: \sigma_{tr} = \int (1 - \cos \theta) \frac{d\sigma}{d\Omega} \, d\Omega. This integral emphasizes backscattering (\theta \approx \pi) over grazing collisions and enters the Chapman-Enskog expansion for the shear \eta \propto \sqrt{m T}/\sigma_{tr}, where m is the and T the .

Quantum Mechanical Framework

Scattering Amplitude and Wave Functions

In quantum mechanical treatments of elastic scattering, the system is described by the time-independent Schrödinger equation for a particle incident on a scattering potential V(\vec{r}): -\frac{\hbar^2}{2m} \nabla^2 \psi(\vec{r}) + V(\vec{r}) \psi(\vec{r}) = E \psi(\vec{r}), where E = \frac{\hbar^2 k^2}{2m} is the energy of the incident particle with wave number k. The unperturbed solution, representing an incident plane wave propagating along the z-direction, is \psi_i(\vec{r}) = e^{i k z}. The total wave function is then expressed as \psi(\vec{r}) = \psi_i(\vec{r}) + \psi_s(\vec{r}), where \psi_s(\vec{r}) is the scattered wave. Far from the scattering region (as r \to \infty), the scattered wave takes the asymptotic form \psi_s(\vec{r}) \sim f(\theta) \frac{e^{i k r}}{r}, with f(\theta) denoting the scattering amplitude, which depends on the scattering angle \theta between the incident and outgoing directions. This form captures the outgoing spherical wave nature of the scattered particle, modulated by the amplitude f(\theta). The scattering amplitude f(\theta) encapsulates the quantum mechanical probability for scattering into direction \hat{r}. The differential cross-section, which gives the probability per unit solid angle for scattering into \theta, is directly related to the modulus squared of the amplitude: \frac{d\sigma}{d\Omega} = |f(\theta)|^2. This relation generalizes the classical differential cross-section to the quantum regime, where interference effects are inherently included through the wave function. Integrating over all angles yields the total cross-section \sigma = \int |f(\theta)|^2 d\Omega. The asymptotic behavior ensures that the flux of scattered particles matches experimental observables in far-field detectors. To solve for \psi(\vec{r}) and thus f(\theta), the Lippmann-Schwinger equation provides an integral formulation equivalent to the Schrödinger equation. It expresses the total wave function in terms of the incident wave and the potential: \psi(\vec{r}) = \phi(\vec{r}) + \frac{2m}{\hbar^2} \frac{1}{4\pi} \int G(\vec{r}, \vec{r}') V(\vec{r}') \psi(\vec{r}') d^3 r', where \phi(\vec{r}) = e^{i k z} is the incident plane wave, and G(\vec{r}, \vec{r}') = -\frac{e^{i k |\vec{r} - \vec{r}'|}}{|\vec{r} - \vec{r}'|} is the outgoing Green's function for the Helmholtz equation (\nabla^2 + k^2) G = \delta(\vec{r} - \vec{r}'). This equation is particularly useful for perturbative solutions, as it resums multiple scattering events through the integral term. The scattering amplitude can be extracted from the asymptotic limit of this equation. The formalism was developed by Lippmann and Schwinger in 1950 as a variational approach to dynamic problems in quantum mechanics. A key perturbative method to approximate f(\theta) is the first-order , obtained by replacing \psi(\vec{r}') in the Lippmann-Schwinger equation with the incident wave \phi(\vec{r}'). This yields f(\theta) \approx -\frac{m}{2\pi \hbar^2} \int e^{-i \vec{q} \cdot \vec{r}} V(\vec{r}) \, d^3 r, where \vec{q} = \vec{k} - \vec{k}' is the momentum transfer vector with |\vec{k}| = |\vec{k}'| = k and \vec{k}' = k \hat{r}, so q = 2 k \sin(\theta/2). The integral represents the of the potential, linking the scattering amplitude directly to its spatial structure. This approximation was introduced by in 1926 as part of early quantum . The Born approximation holds for weak potentials where |V(\vec{r})| \ll E, ensuring higher-order terms in the Born series are negligible, and for slowly varying potentials where the potential changes little over the de Broglie wavelength \lambda = 2\pi / k, corresponding to q times the potential range being small. These conditions are typically met in high-energy scattering or dilute systems, but fail for strong or rapidly oscillating potentials, such as in low-energy atomic collisions.

Partial Wave Analysis

Partial wave analysis expands the scattering amplitude in terms of eigenfunctions of angular momentum, offering an exact solution for elastic scattering from short-range central potentials. This technique decomposes the incident plane wave into spherical waves, each with definite orbital angular momentum quantum number l, allowing the scattering problem to be reduced to solving independent radial equations. The resulting phase shifts encode the effect of the potential on each partial wave, enabling computation of the differential and total cross-sections without approximations for finite-range interactions. Building on the general form of the scattering amplitude f(\theta) from the quantum mechanical framework, the partial wave expansion is given by f(\theta) = \frac{1}{2ik} \sum_{l=0}^\infty (2l+1) (e^{2i\delta_l} - 1) P_l(\cos \theta), where P_l(\cos \theta) are the Legendre polynomials and \delta_l are the real phase shifts for each partial wave. The phase shifts \delta_l arise from matching the asymptotic form of the radial wave function u_l(r) \sim \sin(kr - l\pi/2 + \delta_l) to the solution of the radial Schrödinger equation, -\frac{\hbar^2}{2\mu} \frac{d^2 u_l}{dr^2} + \left[ \frac{\hbar^2 l(l+1)}{2\mu r^2} + V(r) \right] u_l = \frac{\hbar^2 k^2}{2\mu} u_l, with V(r) the short-range potential vanishing beyond some finite distance. For large r, the phase shift quantifies the deviation from free-particle behavior due to the potential, and only a finite number of terms contribute significantly when ka \ll 1, where a is the range of the potential. The total elastic cross-section follows directly from integrating the differential cross-section |f(\theta)|^2, \sigma = \frac{4\pi}{k^2} \sum_{l=0}^\infty (2l+1) \sin^2 \delta_l. This sum converges rapidly for low energies, as higher-l phase shifts vanish. The optical theorem provides a complementary relation, linking the total cross-section to the forward scattering amplitude: \sigma = \frac{4\pi}{k} \Im f(0). In the partial wave expansion, substituting \theta = 0 yields f(0) = \frac{1}{2ik} \sum_{l=0}^\infty (2l+1) (e^{2i\delta_l} - 1), confirming the theorem's consistency with unitarity. This result, first derived for quantum scattering in by Peierls and Placzek in the context of reactions, underscores the conservation of probability flux in wave . At low energies, where k \to 0, centrifugal barriers suppress contributions from l \geq 1, leaving s-wave (l=0) dominant; the cross-section then approaches $4\pi a^2, with a the defined by \delta_0 \approx -ka. More precisely, the s-wave phase shift admits the expansion k \cot \delta_0 = -\frac{1}{a} + \frac{1}{2} r_0 k^2 + O(k^4), where r_0 is the , characterizing the potential's shape beyond the . This , introduced by Bethe in 1949 to analyze neutron-proton data, allows extraction of low-energy parameters from experimental phase shifts without assuming a specific potential form. The foundations of were laid in the 1920s by Faxén and Holtsmark, who applied it to in atomic potentials, with subsequent refinements by others in the early quantum era.

Optical Scattering Phenomena

Rayleigh Scattering

refers to the elastic scattering of electromagnetic waves by particles whose dimensions are much smaller than the of the incident radiation, typically in the regime where the particle radius a satisfies a \ll \lambda, with \lambda being the . This process is dominated by the approximation, wherein the induced in the particle oscillates at the frequency of the incident field, reradiating energy without change in . The theory was first developed by Lord Rayleigh in 1871 to explain the and color of skylight, attributing the phenomenon to scattering by atmospheric molecules. The total scattering cross-section in this regime is given by \sigma = \frac{8\pi}{3} \left( \frac{2\pi}{\lambda} \right)^4 \frac{\alpha^2}{(4\pi \epsilon_0)^2}, where \alpha is the polarizability of the particle. For a small dielectric sphere, the polarizability takes the form \alpha = 4\pi \epsilon_0 a^3 \frac{\epsilon - 1}{\epsilon + 2}, with \epsilon the relative permittivity of the sphere. This \lambda^{-4} dependence implies that shorter wavelengths (e.g., blue light) are scattered more efficiently than longer ones (e.g., red light), leading to the blue coloration of the daytime sky as viewed from Earth. The theory also accounts for the reddish hues of sunsets, where longer path lengths through the atmosphere preferentially scatter out shorter wavelengths, leaving red light to reach the observer. Additionally, Rayleigh scattering underlies the Tyndall effect observed in colloidal suspensions, where visible light is scattered by small particles in a medium. The differential scattering cross-section exhibits angular dependence proportional to $1 + \cos^2 \theta for unpolarized incident light, where \theta is the relative to the incident ; this results in stronger forward and backward scattering compared to the sides. From a quantum mechanical perspective, describes the elastic scattering of photons by atoms or molecules without or , treatable via the first for weak scattering potentials. This approximation, relying on the derived from the incident , aligns with the classical dipole radiation in the low-energy limit and connects to the broader quantum framework of elastic processes.

Mie Scattering

Mie theory provides an exact solution to for the elastic scattering of electromagnetic waves by a homogeneous, isotropic whose size is comparable to the of the incident . Developed by Gustav Mie in , the theory expands the incident, internal, and scattered fields in terms of and applies boundary conditions at the sphere's surface to determine the scattering coefficients a_n and b_n, which represent the electric and magnetic multipole contributions of order n, respectively. The angular distribution of the scattered intensity is described by the amplitude functions S_1(\theta) and S_2(\theta), which account for the components perpendicular and parallel to the plane: S_1(\theta) = \sum_{n=1}^\infty \frac{2n+1}{n(n+1)} \left( a_n \pi_n(\cos\theta) + b_n \tau_n(\cos\theta) \right) S_2(\theta) = \sum_{n=1}^\infty \frac{2n+1}{n(n+1)} \left( a_n \tau_n(\cos\theta) + b_n \pi_n(\cos\theta) \right) where \pi_n(\cos\theta) and \tau_n(\cos\theta) are angular functions derived from . These functions capture the differences in for the two states, leading to complex effects that deviate from the simple approximation seen in smaller-particle regimes. The total scattering cross-section is then given by \sigma_s = \frac{2}{k^2} \sum_{n=1}^\infty (2n+1) \Re(a_n + b_n), where k is the wave number. A key parameter in Mie theory is the size parameter x = ka = 2\pi a / \lambda, where a is the sphere's radius and \lambda is the ; this governs the relative importance of , , and . For large x, the cross-section approaches approximately twice the geometric cross-section \pi a^2, a phenomenon known as the extinction paradox, arising from the constructive of diffracted waves in the forward direction. In applications such as aerosol optics, Mie theory explains atmospheric phenomena like forward patterns—bright backscattering due to surface waves—and structures from internal reflections and refractions in spherical droplets.

Nuclear and Particle Physics Applications

Rutherford Scattering

Rutherford scattering describes the elastic scattering of charged particles, such as alpha particles, by the potential of atomic nuclei, assuming point-like charges and non-relativistic energies. This process arises from the repulsive electrostatic interaction between the incident particle with charge Z_1 e and the target nucleus with charge Z_2 e, governed by the potential V(r) = \frac{Z_1 Z_2 e^2}{4\pi \epsilon_0 r}. The scattering trajectory is a in , reflecting the and in this central force field. The seminal experiments conducted by and in 1911, under Ernest Rutherford's supervision, involved bombarding thin gold foils with alpha particles from a radioactive source and observing their deflection patterns via scintillation screens. These observations revealed that while most particles passed through undeflected, a small fraction scattered at large angles, up to nearly 180 degrees, which contradicted the prevailing of the and supported the existence of a compact, positively charged . Rutherford's analysis of these results in 1911 led to the formulation of the nuclear model of the , where the occupies a tiny volume at the center, surrounded by mostly empty space. In the gold foil experiments, the scattering deviated from pure Rutherford predictions at small angles due to screening effects from the atomic electrons, which modify the effective field at larger distances. The differential cross-section for , derived classically from the hyperbolic orbit, 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 E is the of the incident particle and \theta is the scattering angle in the center-of-mass frame. This formula emerges from solving the for the $1/r potential, where the scattering angle \theta relates to the impact parameter b via \theta = \pi - 2\psi, with \psi being the asymptotic angle of the determined by . Integrating over the impact parameter yields the cross-section, which peaks strongly at small angles due to the long-range nature of the force. Remarkably, the same expression is obtained quantum mechanically using the first for the same potential. The impact parameter b and are connected by b = \frac{Z_1 Z_2 e^2}{8\pi \epsilon_0 E} \cot(\theta/2), illustrating that corresponds to large impact parameters, while s (\theta = 180^\circ, b=0) result in backscattering. For a , the d is d = \frac{Z_1 Z_2 e^2}{4\pi \epsilon_0 E}, marking the turning point where the particle's is fully converted to . In the 1911 alpha-gold experiments, typical energies of several MeV yielded d on the order of tens of femtometers, comparable to dimensions, confirming the nucleus's compactness. The total cross-section \sigma = \int \frac{d\sigma}{d\Omega} d\Omega diverges to infinity because the integrand behaves as $1/\theta^4 for small \theta, reflecting the infinite range of the unscreened interaction. In reality, the electron cloud screens the charge, rendering the potential effectively short-ranged and yielding a finite cross-section, with screening cutting off contributions at impact parameters around ~100 fm. This screening was evident in the Geiger-Marsden data as reduced yields at very small angles compared to pure Rutherford predictions.

Elastic Scattering in Hadronic Interactions

Elastic scattering in hadronic interactions refers to processes where two hadrons, such as protons or pions, collide and emerge intact with their internal quantum numbers conserved, exchanging primarily through force without producing new particles. This phenomenon is a key probe of the non-perturbative regime of (QCD), particularly at low momentum transfers (|t| < 1 GeV²), where perturbative methods fail. At high energies (√s ≳ 10 GeV), elastic scattering exhibits diffractive patterns characterized by a forward peak in the differential cross section dσ/dt, reflecting the spatial extent of the hadronic interaction. Theoretically, high-energy elastic hadron scattering is described within , where the is dominated by Pomeron exchange, a Regge trajectory representing the exchange of multiple gluons. The Pomeron intercept α_P(0) ≈ 1.08 leads to a slow rise in the total cross section σ_tot ∝ s^{α_P(0)-1}, consistent with experimental observations of . The elastic amplitude can be parameterized as T_el(s,t) ∝ i s^{α_P(t)} exp(B t / 2), with the slope parameter B increasing as B ≈ B_0 + α' ln(s/s_0), capturing the energy dependence of the interaction . For unitarization at higher energies, eikonal models are employed, expressing the amplitude in impact parameter space as T(s,b) = i [1 - exp(-Ω(s,b)/2)], where the opacity Ω(s,b) accounts for multiple effects and leads to a "black disk" limit at asymptotic energies, with the disk growing as R ∝ ln s. Experimentally, elastic scattering has been studied from the CERN Intersecting Storage Rings (ISR) at √s ≈ 50 GeV to the Large Hadron Collider (LHC), where the TOTEM collaboration measures proton-proton elastic events using Roman Pots to detect intact protons at small angles. At √s = 7 TeV, the total cross section is σ_tot = 98.6 ± 2.2 mb, with the elastic cross section σ_el ≈ 25.3 mb. At higher energies, such as √s = 13 TeV, TOTEM measured σ_tot = 110.6 ± 3.4 mb and σ_el = 31.0 ± 1.7 mb. There is a characteristic dip in dσ/dt at |t| ≈ 0.5 GeV² shifting to smaller |t| with increasing energy due to interference effects. The ratio ρ = Re T / Im T ≈ 0.10-0.14 at t=0 indicates a predominantly imaginary amplitude, supporting Pomeron dominance. These measurements validate Regge-eikonal models and constrain QCD-inspired descriptions of the hadron's transverse profile. Seminal contributions include the Donnachie-Landshoff parameterization, which fits data across decades of energy.

References

  1. [1]
    Elastic Scattering - an overview | ScienceDirect Topics
    Elastic scattering is a common mechanism by which fast neutrons lose their energy when they interact with atomic nuclei of low atomic number, such as hydrogen ( ...
  2. [2]
  3. [3]
    [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 ...
  4. [4]
    Elastic and inelastic scattering for the system: Target and projectile ...
    Nov 20, 2015 · Elastic scattering is the simplest process which can occur in the collision of two nuclei. At low energies, it is a surface process and thus ...
  5. [5]
    Elastic and Inelastic Scattering - Richard Fitzpatrick
    The low-energy elastic scattering cross-section is again independent of the incident particle velocity (which is proportional to $ k$ ), whereas the inelastic ...
  6. [6]
    8.3 Elastic and Inelastic Collisions - Physics | OpenStax
    Mar 26, 2020 · The law of conservation of momentum is very useful here, and it can be used whenever the net external force on a system is zero. Figure 8.6 ...
  7. [7]
    [PDF] Scattering - UCSB Physics
    In the case that a collision between two particles conserves total energy, we typically refer to the scattering as being elastic. Before we begin drawing ...
  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 ...Missing: elastic | Show results with:elastic
  9. [9]
    [PDF] collide - JILA
    Elastic collisions are those in which the atom or molecule remains in the same atomic eigenstate (although its momentum is changed), while inelastic collisions ...
  10. [10]
    [PDF] 8.01SC S22 Chapter 15: Collision Theory - MIT OpenCourseWare
    Jun 15, 2022 · The angle Θ cm between the incoming and outgoing velocities is called the center-of-mass scattering angle. 15-24. Page 26. 15.7.2 Scattering in ...
  11. [11]
    Scattering in the Laboratory Frame - Richard Fitzpatrick
    Equations (378)-(385) enable us to relate the particle energies and scattering angles in the laboratory frame to those in the center of mass frame. In ...
  12. [12]
    [PDF] Maximum Angle of Deflection - Feynman Lectures
    When cos θ is extremal θ is extremal, thus to find the maximum angle of deflection, differentiate the right side of (6) with respect to v1 , and solve for zero.
  13. [13]
    None
    **Summary of Classical Scattering in Central Potentials (Section 1.2)**
  14. [14]
    [PDF] Chapter 3 Two Body Central Forces - Rutgers Physics
    the maximum angle of scattering. θ θ. Another unusual effect occurs when θ(b) becomes 0 or π for some nonzero value of b, with db/dθ finite. Then dσ/dΩ blows ...
  15. [15]
    [PDF] Scattering
    Figure 11.2 - Classical elastic hard-sphere scattering. b = R cos(θ/2), yielding the scattering angle as a function of impact parameter for what is known as ' ...
  16. [16]
  17. [17]
  18. [18]
    [PDF] 2. Kinetic Theory - DAMTP
    The purpose of this section is to lay down the foundations of kinetic theory, starting from the Hamiltonian description of 1023 particles, and ending with ...
  19. [19]
    [PDF] arXiv:1404.5476v2 [physics.chem-ph] 8 Oct 2014
    Oct 8, 2014 · Physically motivated expressions for the transport cross sections describing classical scattering in the Lennard-Jones potential are proposed. ...
  20. [20]
    [PDF] Born approximation validity conditions - UMD Physics
    This is equivalent to the condition / V dt h we derived when treating the scattering by first order time-dependent perturbation theory.
  21. [21]
    [PDF] Quantum theory of Rayleigh scattering - arXiv
    Abstract: We develop a quantum theory of atomic Rayleigh scattering. Scattering is considered as a relaxation of incident photons from a selected mode of ...
  22. [22]
    [PDF] SAND78-6018-Mie-1908-translation.pdf - ScattPort
    The first group represents the waves due to elec- trical oscillations of the sphere; it is characterized by: E, 0, -0. The second group represents the waves due ...
  23. [23]
    [PDF] MATLAB Functions for Mie Scattering and Absorption - OMLC
    Jun 8, 2002 · The key parameters for Mie calculations are the Mie coefficients an and bn to com- pute the amplitudes of the scattered field, and cn and dn ...
  24. [24]
    Mie Scattering Function — miepython 3.0.2 documentation
    Mie scattering describes the special case of the interaction of light passing through a non-absorbing medium with a single embedded spherical object.
  25. [25]
    [PDF] Lorenz-Mie Scattering
    Oct 15, 2020 · Mie scattering is a solution method for light scattering from spheres, applicable for any size parameter, but “Mie regime” ∼ 0.1 <x< 100.Missing: a_n b_n
  26. [26]
    Mie Theory Examples - Ocean Optics Web Book
    Jan 6, 2021 · The standard explanation for the extinction paradox is that the “extra” scattering is due to diffraction by the particle. The electric field of ...
  27. [27]
    [PDF] Mie Theory 1908-2008 - ScattPort
    In 1908 Gustav Mie published his paper describing a theoretical approach to calculate the light scattering by gold colloids. Today, a 100 years later, the ...
  28. [28]
    [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 ...
  29. [29]
    [PDF] 1 Rutherford Formula in Classical Mechanics - Dr. Ian Balitsky
    Remarkable fact about the Rutherford formula is that for Coulomb scattering Classical. Mechanics, Quantum Mechanics, and Quantum Field Theory all lead to the ...
  30. [30]
    [PDF] The Laws of Deflexion of α Particles Through Large Angles
    Geiger and E. Marsden, PM, 25, 604. 1913. The Laws of Deflexion of α Particles Through Large. Angles. H. Geiger and E. Marsden. (Received 1913). In a former ...
  31. [31]
    Scattering from a Screened Coulomb Potential
    Without screening, the total Coulomb scattering cross section is infinite because the range of the force in infinite. The potential energy then is. \begin ...
  32. [32]
    [PDF] 10. Scattering Theory - DAMTP
    There's a surprise when we try to calculate the total cross-section σT . We find that it's infinite! This is because the Coulomb force is long range. The ...Missing: screening | Show results with:screening
  33. [33]
    [PDF] 20. High Energy Soft QCD and Diffraction - Particle Data Group
    May 31, 2024 · The classic example is the elastic scattering of hadrons on nuclei (see e.g. [16]), which manifests an angular distribution with a series of ...
  34. [34]
  35. [35]
  36. [36]