Fact-checked by Grok 2 weeks ago

Rutherford_scattering

Rutherford scattering is the elastic scattering of charged particles, such as alpha particles, by the Coulomb field surrounding atomic nuclei, quantitatively described by a formula derived from classical electrodynamics. This phenomenon was experimentally demonstrated in a series of landmark investigations conducted between 1908 and 1913 by Hans Geiger and Ernest Marsden, under the supervision of Ernest Rutherford at the University of Manchester, using beams of alpha particles incident on thin metallic foils, primarily gold. The observations—that the vast majority of particles traversed the foil undeflected, while a small fraction underwent large-angle deflections, including backscattering—provided definitive evidence against J. J. Thomson's plum pudding model of the atom and supported Rutherford's 1911 proposal of a dense, positively charged nucleus at the atom's center, surrounded by orbiting electrons. In the initial 1909 experiment, alpha particles from a radium-based source were directed perpendicularly onto foils of various metals, with scattered particles detected via scintillations on a zinc sulfide screen observed through a low-power microscope. The fraction of particles scattered at angles greater than 90° was approximately 1 in 8,000 for a platinum foil equivalent to 0.00007 cm of air, with the scattering intensity varying with the foil's atomic weight and thickness but saturating beyond a certain depth, indicating interactions within a superficial layer of atoms. Subsequent refinements in 1910 and 1913 quantified the angular distribution, confirming that large deflections resulted from single encounters rather than cumulative small-angle scatterings, and established the inverse fourth-power dependence on the scattering angle. These results aligned closely with Rutherford's theoretical predictions for scattering by a point-like central charge. Rutherford's formula for the differential cross-section per unit solid angle is \frac{d\sigma}{d\Omega} = \left( \frac{Z_1 Z_2 e^2}{4\pi \epsilon_0 \cdot 4 K} \right)^2 \frac{1}{\sin^4(\theta/2)}, where Z_1 and Z_2 are the atomic numbers of the incident particle and target nucleus, e is the elementary charge, \epsilon_0 is the vacuum permittivity, K is the kinetic energy of the incident particle, and \theta is the scattering angle; this expression, derived assuming hyperbolic trajectories under the inverse-square Coulomb force, accurately predicted the observed scattering rates for gold (Z_2 \approx 100) when alpha particles (Z_1 = 2) had energies around 5–7 MeV. The experiments not only validated the nuclear atomic model but also laid the groundwork for later developments in nuclear physics, including ion-beam analysis techniques like Rutherford backscattering spectrometry used today for material characterization.

History

Early observations of scattering

In early 1906, while determining the mass and velocity of alpha particles emitted from radium and actinium at McGill University, Ernest Rutherford observed the scattering of these particles by air. He directed a narrow beam of alpha particles through varying thicknesses of air onto a photographic plate and noted that the beam's image exhibited diffuse edges, indicating a spreading due to multiple small deflections by air molecules. This broadening increased with air density, suggesting that each atomic encounter caused a slight deviation in the particle's path, consistent with the particles possessing a charge and interacting electrostatically with matter. Later in 1906, Rutherford extended these investigations to solid matter by passing alpha particles through thin mica sheets in a vacuum. The beam's image on the photographic plate showed diffusion primarily on the side facing the mica, confirming that scattering occurred upon interaction with the atoms of the solid, rather than solely due to air. These observations, up to about 2 degrees in deflection, challenged simplistic views of atomic transparency and hinted at the need for a model accounting for intra-atomic electric fields, though Rutherford initially interpreted them as cumulative small-angle scatters. By 1908, after moving to the University of Manchester, Rutherford collaborated with Hans Geiger to quantify alpha particle scattering more precisely using metal foils like gold and platinum. Geiger's preliminary measurements revealed deflections larger than those expected from multiple small encounters alone, with some particles deviating by several degrees after passing through thin foils. These findings, obtained via scintillation screens and photographic detection, built on Rutherford's earlier work and motivated systematic angular distribution studies, laying the groundwork for probing atomic structure.

Geiger–Marsden experiment

The Geiger–Marsden experiment, conducted between 1909 and 1913 at the University of Manchester under the supervision of Ernest Rutherford, involved bombarding thin metal foils with alpha particles to investigate their scattering behavior. In their 1909 paper, Hans Geiger and Ernest Marsden initially observed that a small fraction of alpha particles incident on a metal plate were deflected such that they emerged from the side of incidence, indicating diffuse reflection. This work built on earlier qualitative observations of alpha particle deflections through thin sheets like mica, aiming to quantify the scattering process quantitatively. The experimental apparatus consisted of a vacuum chamber to minimize air scattering, a radioactive source of alpha particles (typically radium emanation or polonium), and a thin scattering foil, often gold leaf approximately 0.4 micrometers thick. A narrow beam of alpha particles was directed at the foil, and scattered particles were detected using a zinc sulfide scintillation screen viewed through a microscope; scintillations were counted manually over extended periods, sometimes exceeding 100,000 events per measurement. Angles of deflection were varied from 5° to 150° by adjusting the position of the detector relative to the foil and source, with the setup enclosed in a light-tight cylindrical metal box to enhance observation accuracy. Experiments were performed with foils of different metals (e.g., aluminum, silver, gold) and thicknesses, as well as varying alpha particle velocities by selecting sources with different energies. Key observations revealed that alpha particles could be scattered through large angles, including nearly 180° backscattering, which contradicted expectations from the prevailing Thomson plum pudding atomic model that predicted only small deflections. The number of particles scattered at a given angle θ followed an approximately 1/sin⁴(θ/2) dependence, tested across angles from 5° to 150°. Scattering intensity was proportional to foil thickness for thin foils and to the square of the atomic weight of the scatterer, while decreasing as the inverse fourth power of the alpha particle velocity. For instance, the fraction of alpha particles scattered at 45° by a gold foil equivalent to 1 mm of air was measured as 3.7 × 10⁻⁷, implying a central atomic charge roughly half the atomic weight in electron charge units. These results provided empirical evidence for Rutherford's hypothesis of a concentrated positive charge within the atom, as the large-angle scattering could only be explained by close encounters with a massive, nuclear core. The experiments confirmed the inverse-square law of electrostatic repulsion for alpha particles interacting with atomic nuclei and allowed estimation of nuclear charge (e.g., approximately 77 electron charges for gold) and size, indicating the nucleus is about 10,000 times smaller than the atom. Subsequent refinements in 1910 and 1913 directly tested predictions from Rutherford's 1911 nuclear model, establishing the scattering laws as foundational to atomic structure theory.

Theoretical description

Classical derivation

The classical derivation of Rutherford scattering treats the interaction between an incident charged particle and a target nucleus as a two-body problem under a repulsive Coulomb potential, solved using conservation laws in classical mechanics. The nucleus is modeled as a point charge Ze fixed at the origin, where Z is the atomic number and e is the elementary charge, while the incident particle carries charge ze (with z = 2 for an alpha particle) and has initial kinetic energy T = \frac{1}{2} m v_\infty^2, where m is its mass and v_\infty its speed at infinity. The force is given by Coulomb's law: F(r) = \frac{z Z e^2}{4\pi \epsilon_0 r^2}, directed radially outward. The trajectory of the incident particle is a hyperbola, as the effective potential V_\mathrm{eff}(r) = \frac{z Z e^2}{4\pi \epsilon_0 r} + \frac{l^2}{2 m r^2} (with angular momentum l = m v_\infty b, where b is the impact parameter) yields unbound motion for E > 0. The scattering angle \theta relates to b via conservation of energy and angular momentum. At the distance of closest approach r_\mathrm{min}, the radial velocity is zero, leading to r_\mathrm{min} = \frac{D}{2} \left( 1 + \sqrt{1 + \left( \frac{2 b}{D} \right)^2 } \right), where D = \frac{z Z e^2}{4\pi \epsilon_0 T} is twice the distance of closest approach for a head-on collision (b = 0, \theta = \pi). The angle is then \theta = \pi - 2 \psi, where \sin \psi = \frac{1}{\sqrt{1 + (2 b / D)^2}}, yielding b = \frac{D}{2} \cot(\theta / 2). To obtain the scattering probability, consider the differential cross section \frac{d\sigma}{d\Omega}, which measures the effective area per unit solid angle for scattering into angle \theta. The number of particles scattered into d\Omega is proportional to the annular area $2\pi b \, db incident on the target, related to the solid angle by d\Omega = 2\pi \sin \theta \, d\theta. Differentiating b(\theta) gives \frac{db}{d\theta} = -\frac{D}{4 \sin^2 (\theta / 2)}, so \frac{d\sigma}{d\Omega} = b \left| \frac{db}{d\theta} \right| / \sin \theta = \left( \frac{D}{4} \right)^2 \frac{1}{\sin^4 (\theta / 2)}. This is the Rutherford formula, predicting a $1/\sin^4 (\theta / 2) angular dependence and $1/T^2 energy scaling. This derivation assumes single scattering events dominate, valid for thin targets where multiple interactions are negligible, and neglects nuclear recoil and relativistic effects. It successfully explained Geiger and Marsden's observations of large-angle deflections, supporting a concentrated nuclear charge.

Rutherford formula

The Rutherford formula provides the differential cross-section for the elastic scattering of charged particles by the Coulomb field of a point-like nucleus, derived classically assuming a hyperbolic trajectory under inverse-square repulsion. It quantifies the probability of scattering into a given solid angle and was pivotal in interpreting the Geiger–Marsden experiments. In the classical treatment, an incident particle of charge Ze, mass m, and initial kinetic energy E scatters off a stationary target nucleus of charge Ze'. The trajectory is a hyperbola with the nucleus at the external focus, and the scattering angle \theta relates to the impact parameter b (the perpendicular distance from the initial trajectory to the nucleus) via b = \frac{Z Z' e^2}{8\pi \epsilon_0 E} \cot\left(\frac{\theta}{2}\right), where \epsilon_0 is the vacuum permittivity. This relation arises from conservation of energy and angular momentum, yielding the asymptotes of the hyperbola that determine the deflection. The differential cross-section \frac{d\sigma}{d\Omega}, which gives the effective area per unit solid angle for scattering into d\Omega, follows by considering the annular area $2\pi b \, db corresponding to scattering angles between \theta and \theta + d\theta: \frac{d\sigma}{d\Omega} = \left( \frac{Z Z' e^2}{8\pi \epsilon_0 E} \right)^2 \frac{1}{\sin^4(\theta/2)}. Here, d\Omega = 2\pi \sin\theta \, d\theta, and the transformation from db to d\theta produces the characteristic \sin^{-4}(\theta/2) dependence, emphasizing large forward scattering at small angles. This formula assumes non-relativistic speeds, a point-like target much heavier than the projectile, and neglects screening by atomic electrons. Originally formulated in cgs units without \epsilon_0, the expression matched experimental scattering rates from thin foils, where the number of particles scattered per unit time into d\Omega at distance r is I \cdot n t \cdot \frac{d\sigma}{d\Omega} \cdot \frac{d\Omega}{r^2}, with I the incident flux, n the target density, and t the foil thickness. The formula's success confirmed the nuclear charge concentration, as diffuse charge distributions would yield softer angular dependence.

Quantum mechanical treatment

Born approximation

The Born approximation provides a perturbative approach to quantum scattering problems, particularly useful when the scattering potential is relatively weak compared to the incident particle's kinetic energy. Developed by Max Born in 1926, it approximates the scattering amplitude by replacing the full wave function in the Lippmann-Schwinger equation with the incident plane wave, yielding a first-order expression in the potential. In scattering theory, the first Born approximation for the scattering amplitude f(\mathbf{k}', \mathbf{k}) from initial wave vector \mathbf{k} to final \mathbf{k}' (with |\mathbf{k}| = |\mathbf{k}'| = k) is f(\mathbf{k}', \mathbf{k}) = -\frac{m}{2\pi \hbar^2} \int d^3\mathbf{r}\, e^{-i \mathbf{k}' \cdot \mathbf{r}} V(\mathbf{r}) e^{i \mathbf{k} \cdot \mathbf{r}} = -\frac{m}{2\pi \hbar^2} \tilde{V}(\mathbf{q}), where m is the reduced mass, \mathbf{q} = \mathbf{k} - \mathbf{k}' is the momentum transfer (q = 2k \sin(\theta/2), with \theta the scattering angle), and \tilde{V}(\mathbf{q}) is the Fourier transform of the potential V(\mathbf{r}). The differential cross section is then d\sigma / d\Omega = |f|^2. This approximation holds when the potential does not significantly distort the incident wave in the scattering region, typically requiring |V| \ll \hbar^2 k^2 / m. For Rutherford scattering, the potential is the repulsive Coulomb interaction between an incident charged particle (charge Z_1 e) and a fixed nucleus (charge Z_2 e): V(r) = \frac{Z_1 Z_2 e^2}{4\pi \epsilon_0 r}. The Fourier transform of this potential is \tilde{V}(\mathbf{q}) = \frac{Z_1 Z_2 e^2}{\epsilon_0 q^2}, leading to the scattering amplitude f(\theta) = -\frac{m Z_1 Z_2 e^2}{2\pi \epsilon_0 \hbar^2 q^2} = -\frac{m Z_1 Z_2 e^2}{8\pi \epsilon_0 \hbar^2 k^2 \sin^2(\theta/2)}. The resulting differential cross section is \frac{d\sigma}{d\Omega} = |f(\theta)|^2 = \left( \frac{Z_1 Z_2 e^2}{16\pi \epsilon_0 E} \right)^2 \frac{1}{\sin^4(\theta/2)}, where E = \hbar^2 k^2 / 2m is the incident kinetic energy. This expression exactly reproduces the classical Rutherford scattering formula derived from hyperbolic trajectories in the Coulomb field. Despite this agreement, the Born approximation is not strictly valid for the pure Coulomb potential due to its long-range nature, which causes significant distortion of the wave function far from the scatterer and leads to infrared divergences in higher-order terms. The total cross section diverges logarithmically from small-angle scattering, a feature shared with the classical case but requiring careful regularization (e.g., via a screened Yukawa potential with screening length taken to infinity). The exact quantum solution for Coulomb scattering, obtained using parabolic coordinates and confluent hypergeometric functions, confirms the same differential cross section but reveals additional effects like phase shifts and potential orbiting in attractive cases. For practical applications in nuclear physics, the Born result serves as a reliable benchmark at high energies where quantum corrections are small.

Partial wave analysis

In quantum mechanical scattering theory, partial wave analysis provides a powerful framework for treating central potentials by expanding the incident plane wave into eigenstates of angular momentum, labeled by the quantum number l. The scattered wave is then described by phase shifts \delta_l in each partial wave, which encode the effect of the potential on the radial wave function. The scattering amplitude takes the form f(\theta) = \frac{1}{2ik} \sum_{l=0}^{\infty} (2l + 1) (e^{2i \delta_l} - 1) P_l(\cos \theta), where k is the wave number, \theta is the scattering angle, and P_l are Legendre polynomials. The differential cross section is |f(\theta)|^2. This expansion is particularly useful for short-range potentials, where higher l waves are unaffected, but for the long-range Coulomb potential V(r) = \frac{Z_1 Z_2 e^2}{4\pi \epsilon_0 r} (repulsive for like charges), the analysis requires special care due to the slow decay, leading to non-standard asymptotic behavior and divergent sums over l. For Rutherford scattering, the partial waves are characterized by Coulomb phase shifts \sigma_l = \arg \Gamma(l + 1 + i \eta), where \eta = \frac{Z_1 Z_2 e^2}{4\pi \epsilon_0 \hbar v} is the Sommerfeld parameter, with v the incident velocity. These phase shifts arise from the exact solution of the radial Schrödinger equation for the Coulomb potential, first derived by Gordon in 1928 using separation in parabolic coordinates. The S-matrix elements are S_l = e^{2i \sigma_l}, but the standard partial wave sum for the amplitude diverges because all partial waves contribute significantly, reflecting the infinite range. However, an analytic summation of the series yields the closed-form scattering amplitude f_C(\theta) = -\frac{\eta}{2k} \frac{\Gamma(1 + i \eta)}{\Gamma(1 - i \eta)} \exp\left[-i \eta \ln \sin^2(\theta/2)\right] \csc^2(\theta/2), which, upon taking the modulus squared, reproduces the classical Rutherford differential cross section \frac{d\sigma}{d\Omega} = \left( \frac{Z_1 Z_2 e^2}{16\pi \epsilon_0 E} \right)^2 \frac{1}{\sin^4(\theta/2)}, with E the incident energy. This equivalence holds in the semiclassical limit of high energy or large impact parameter, confirming the validity of the classical picture despite the quantum treatment. The total cross section diverges due to the forward-scattering singularity (\theta \to 0), a feature shared with the classical result, arising from the infinite number of contributing partial waves. In practice, for realistic experiments involving wave packets rather than plane waves, the partial wave sum converges, revealing small quantum deviations from the Rutherford formula, particularly at low energies and small angles, such as a "shadow zone" where scattering probability vanishes. These effects, predicted using wave-packet regularization, become observable for parameters like \eta \approx 800 (e.g., 3.8 keV alpha particles on gold), with angular widths on the order of 0.6°. Nonetheless, for most historical and applied contexts of Rutherford scattering, the partial wave approach reinforces the exact match to the classical formula, underscoring the robustness of the point-charge model.

Experimental methods

Setup and apparatus

The Geiger–Marsden experiments on alpha particle scattering employed a series of apparatuses that evolved from simple configurations for observing small-angle deflections to more precise setups capable of measuring large-angle scattering. In the initial 1909 experiment, the apparatus consisted of a glass tube containing purified radium emanation as the alpha particle source, sealed at one end with a thin mica window equivalent to about 1 cm of air in thickness to allow particle emission. The tube was positioned such that a narrow beam of alpha particles struck a thin metal foil reflector, typically gold or platinum, placed approximately 1 cm away. Scattered particles were detected using a zinc sulfide screen mounted behind a lead plate to block direct beam particles, with scintillations observed and counted through a low-power microscope focused on a defined area of the screen. This setup operated in air at atmospheric pressure, allowing qualitative observation of diffuse reflection at angles up to about 90 degrees, though quantitative precision was limited by the fixed geometry. Subsequent refinements, detailed in Geiger's 1908 apparatus description and extended in later work, introduced collimation to produce a narrower pencil of alpha particles and improved shielding to reduce background scintillations. The source remained radium-based, but diaphragms were added to define the beam, and the entire system was enclosed to minimize scattering from air molecules. By 1910, experiments incorporated variable foil thicknesses and distances, with the scintillation screen movable to assess scattering distribution. The most advanced setup, used in the 1913 experiments to quantify large-angle deflections, featured a robust cylindrical metal box serving as a vacuum chamber to eliminate air scattering. Inside, a fixed alpha source— a thin-walled glass tube (1 mm diameter) filled with low-pressure radium emanation—was mounted alongside the scattering foil, typically gold leaf about 0.4 μm thick, separated by 2.5 cm. A diaphragm collimated the alpha beam onto the foil, and a rotatable microscope with an attached zinc sulfide screen (radius ~1.6 cm) allowed observation of scintillations at angles from 5° to 150° by rotating the assembly around the source-foil axis. The chamber, exhausted via a tube and sealed with a ground-glass plate, was affixed to a graduated circular platform for precise angular control, with the box lined in paper and equipped with aluminum screens to suppress stray particles from walls. Counts of over 100,000 scintillations per run were recorded manually, correcting for source decay and geometry. This configuration enabled systematic variation of foil material, thickness, and distance, confirming the inverse square law dependence of scattering on foil thickness and the angular distribution.

Data collection and analysis

In the Geiger-Marsden experiments, data collection relied on manual observation of scintillations produced by scattered alpha particles impacting a zinc sulfide screen. A narrow beam of alpha particles from a radium emanation source was directed onto a thin foil target, such as gold, within a vacuum chamber to minimize air scattering. The screen, mounted on a rotatable arm, was positioned at various angles relative to the incident beam, typically from 5° to 150°. Observers used a low-power microscope to count individual flashes of light (scintillations) on the screen, with each flash corresponding to the arrival of a single alpha particle; exposure times varied from minutes to hours to accumulate sufficient counts, often exceeding 100,000 scintillations per dataset to ensure statistical reliability. Corrections were applied during collection to account for systematic errors, including the exponential decay of the radioactive source, background radiation from beta and gamma rays (which produced weaker, non-resolvable scintillations), and spurious particles from chamber walls or secondary scattering. For instance, in measurements with silver foil, counts at 30° yielded 2500 scintillations after subtracting an estimated 10-20% background. The total number of incident particles was determined separately by direct scintillation counting without the foil. Analysis involved tabulating the number of scattered particles N(\theta) as a function of scattering angle \theta, normalized by the incident flux and solid angle subtended by the screen. Plots of N(\theta) versus \theta revealed a sharp decrease with increasing angle, with the data fitting the form N(\theta) \propto \frac{1}{\sin^4(\theta/2)}, confirming Rutherford's theoretical prediction for single scattering events by a point-like Coulomb center. Further verification showed proportionality to foil thickness (for thin foils where multiple scattering was negligible) and to the square of the atomic number Z of the target, as tested across elements like aluminum (Z=13), silver (Z=47), and gold (Z=79). Velocity dependence was assessed using sources of varying alpha energies, yielding N \propto v^{-4}, where v is the particle speed. These relations were quantified through least-squares fitting to logarithmic plots, establishing the experiment's precision within 5-10% error margins. Modern recreations of the experiment employ digital detectors, such as silicon surface barrier diodes or cloud chambers, to automate counting and improve resolution, but the foundational manual scintillation method remains a benchmark for understanding large-angle scattering statistics. Analysis in these updates often uses Monte Carlo simulations to model multiple scattering contributions, ensuring the single-scattering regime dominates for foils thinner than 1 μm.

Significance and applications

Impact on atomic model

The Geiger-Marsden experiments of 1909, conducted under Ernest Rutherford's supervision at the University of Manchester, involved bombarding thin gold foil with alpha particles from a radioactive source and observing their scattering patterns using a zinc sulfide screen. These experiments revealed that while most alpha particles passed through the foil undeflected, approximately 1 in 20,000 particles were deflected by angles greater than 90 degrees, with some backscattering directly toward the source. This observation starkly contradicted J.J. Thomson's plum pudding model of the atom, which posited a uniform distribution of positive charge throughout a diffuse sphere with embedded electrons, predicting only small-angle deflections from multiple cumulative encounters rather than rare large-angle scatters. In his seminal 1911 paper published in the Philosophical Magazine, Rutherford analyzed these results and proposed a revolutionary atomic model to account for the large deflections. He concluded that the atom must contain a tiny, dense, positively charged core—later termed the nucleus—concentrated in a volume less than 10^{-13} meters in diameter, comprising nearly all the atom's mass and positive charge, with electrons orbiting at a distance. Rutherford estimated the nuclear charge for gold as approximately 100 times the elementary charge e, roughly proportional to the square root of the atomic weight. This nuclear model explained the scattering as resulting from close encounters between alpha particles and the nucleus via Coulomb repulsion, where particles follow hyperbolic trajectories, rather than diffuse interactions. The nuclear model fundamentally transformed atomic theory, shifting from Thomson's diffuse structure to a compact central nucleus surrounded by orbiting electrons, akin to a planetary system. This framework resolved the scattering anomalies but introduced challenges, such as electron stability against electromagnetic radiation, which later prompted Niels Bohr's 1913 quantum refinements. By establishing the nucleus as the atom's core, Rutherford's interpretation laid the groundwork for modern nuclear physics and chemistry, enabling subsequent discoveries like isotopic variations and nuclear reactions.

Applications in nuclear physics

Rutherford scattering plays a pivotal role in nuclear physics by enabling the determination of nuclear sizes and charge distributions through elastic scattering of charged projectiles, such as alpha particles, from atomic nuclei. In these experiments, the measured differential cross section is compared to the point-like Rutherford formula; deviations arise when the projectile's trajectory is influenced by the finite nuclear size and the short-range strong interaction, allowing extraction of the nuclear radius. This radius is typically parameterized as R = r_0 A^{1/3}, with r_0 \approx 1.2 fm and A the mass number, providing a benchmark for nuclear density models. For instance, scattering of high-energy alpha particles from heavy targets like ^{208}Pb at fixed angles reveals deviations at impact parameters around 7.4 fm and closest approaches of about 12.9 fm, yielding a nuclear radius of approximately 7.1 fm, which aligns with empirical expectations for the charge distribution in heavy nuclei. These early systematic studies of alpha scattering cross sections as a function of energy and angle established the groundwork for quantifying nuclear dimensions and validated the nuclear model against diffuse charge distributions. At low momentum transfers, the elastic scattering cross section can be analyzed using \left( \frac{d\sigma}{d\Omega} \right) = \left( \frac{d\sigma}{d\Omega} \right)_R |F(q)|^2, where \left( \frac{d\sigma}{d\Omega} \right)_R is the Rutherford cross section and F(q) is the nuclear form factor, representing the Fourier transform of the charge density \rho(r). Fitting experimental data with quantum mechanical treatments, such as the first Born approximation, extracts \rho(r), often modeled as a Fermi distribution \rho(r) = \frac{\rho_0}{1 + \exp\left( (r - R)/a \right)} with skin thickness a \approx 0.5 fm. This approach has been applied to study charge radii across isotopes, revealing variations due to nuclear shell effects and neutron-proton asymmetries.

References

  1. [1]
    On a diffuse reflection of the α-particles - Journals
    A small fraction of the α-particles falling upon a metal plate have their directions changed to such an extent that they emerge again at the side of incidence.
  2. [2]
    LXXIX. The scattering of α and β particles by matter and the structure ...
    The scattering of α and β particles by matter and the structure of the atom ... https://doi.org/10.1080/14786440508637080 · References · Citations · Metrics ...
  3. [3]
    The scattering of α-particles by matter - Journals
    The present investigation was undertaken with a view to obtain a quantitative measurement of the scattering by determining the most probable angle.
  4. [4]
    On the scattering of the α-particles by matter - Journals
    Professor Rutherford* showed that the image of a narrow slit produced by the α-rays on a photographic plate broadens out when the slit was covered with a thin ...
  5. [5]
    The great scattering experiments | IOPSpark - Institute of Physics
    Geiger and Marsden went on to make a great series of measurements of the deflections of a narrow beam of alpha particles that hits gold leaf in a vacuum.
  6. [6]
    Geiger and Marsden's Alpha deflection article of 1913 - chemteam.info
    The Laws of Deflexion of a Particles through Large Angles1. By Dr. H. GEIGER and E. MARSDEN2. Philosophical Magazine Series 6, Volume 25, Number 148
  7. [7]
    [PDF] Chapter 2 Rutherford Scattering
    This relation is derived using Newton's Second Law of Motion, Coulomb's law for the force between the α-particle and and nucleus, and conservation of angular ...
  8. [8]
    [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 ...
  9. [9]
  10. [10]
    Rutherford Scattering Cross-Section
    The probability per unit time of the particle being scattered into the range of solid angle ${\mit\Omega}$ to ${\mit\Omega}+d{\mit\Omega}$ then the ...
  11. [11]
    Born approximation - Wikipedia
    The Born approximation consists of taking the incident field in place of the total field as the driving field at each point in the scatterer.
  12. [12]
    Born Approximation - Richard Fitzpatrick
    Equation (950) is identical to the classical Rutherford scattering cross-section formula. increases, implying that the Born approximation is more accurate at ...
  13. [13]
    [PDF] Scattering Theory II 1 Born Approximation - 221B Lecture Notes
    Born approximation is the first order approximation in the potential, neglecting O(V^2) corrections, and is good only when scattering is weak.
  14. [14]
    [PDF] 10. Scattering Theory - DAMTP
    It turns out that the. Born approximation is valid for the Yukawa potential in certain regimes, but is never valid for the Coulomb potential! The difficulty ...<|control11|><|separator|>
  15. [15]
    [PDF] Scattering Theory III 1 Partial Wave Analysis - 221B Lecture Notes
    Therefore the scattering occurs only for the l = 0 partial wave, or S-wave. We first analyze the S-wave only, which turns out to be particularly simple.
  16. [16]
    [PDF] arXiv:1704.04305v3 [quant-ph] 3 Oct 2017
    Oct 3, 2017 · Using partial wave analysis, the sum over the angular momentum quantum number, l, does not converge.
  17. [17]
    Alpha Particles and the Atom, Rutherford at Manchester, 1907–1919
    In Rutherford's now-famous paper of May 1911 on the scattering of alpha particles by gold foil, he included this sketch of the hyperbolic path of a particle.
  18. [18]
    [PDF] The Laws of Deflexion of α Particles Through Large Angles
    It will be seen that the laws “single scattering” found in this paper are quite distinct from the laws of “compound scattering” previously deduced by Geiger.
  19. [19]
    LXI. The laws of deflexion of a particles through large angles
    Jun 8, 2010 · The experiments described in the foregoing paper were carried out to test a theory of the atom proposed by Prof. Rutherford.
  20. [20]
    Rutherford Scattering - Galileo
    Rutherford's alpha scattering experiments were the first experiments in which individual particles were systematically scattered and detected.Rutherford as Alpha-Male · Scattering Alphas · Seeing the Nucleus
  21. [21]
    Rutherford's Nucleus Paper of 1911
    The deviation of a particle throughout a considerable angle from an encounter with a single atom will in this paper be called 'single' scattering.Missing: royal | Show results with:royal
  22. [22]
    May, 1911: Rutherford and the Discovery of the Atomic Nucleus
    May 1, 2006 · In the now well-known experiment, alpha particles were observed to scatter backwards from a gold foil. Rutherford's explanation, which he ...
  23. [23]
    The Roles of Thomson and Rutherford in the Birth of Atomic Physics ...
    Jun 28, 2024 · In order to account for this large angle scattering of α particles, Rutherford supposed that the atom consists of a positively charged ...
  24. [24]
    Scattering of Alpha Particles | Rev. Mod. Phys.
    Scattering of Alpha Particles. R. M. Eisberg and C. E. Porter · R. M. Eisberg ... Rev. Mod. Phys. 33, 190 – Published 1 April, 1961. DOI: https://doi.org ...
  25. [25]
    Rutherford Scattering - HyperPhysics
    Scattering of high energy alpha particles at a fixed angle can be used to estimate nuclear radius. Eisberg, R. M. and Porter, C. E., Rev. Mod. Phys. 33, 190 ( ...
  26. [26]
    [PDF] Lecture 2: Nuclear Phenomenology - INPP - Ohio University
    • The radial distribution of charge & matter are probed via scattering. •Electrons are best for the charge radius, because structure-less but charged.