Fact-checked by Grok 2 weeks ago

Atomic form factor

The atomic form factor, also known as the atomic factor, is a fundamental quantity in physics that quantifies the of X-rays (or other electromagnetic waves) by the of an isolated atom, typically modeled as the of the atom's spherically symmetric distribution. It depends on the q = \frac{4\pi \sin(\theta/2)}{\lambda}, where \theta is the and \lambda is the , with the f(0) equaling the Z (the total number of electrons) at zero and approaching zero for large angles due to destructive interference among scattered waves from distributed electrons. In X-ray crystallography and diffraction analysis, the atomic form factor plays a central role in calculating the structure factor F(\mathbf{G}), which determines the intensity of diffracted beams from a crystal lattice by summing contributions from all atoms in the unit cell: F(\mathbf{G}) = \sum_j f_j(\mathbf{G}) e^{i \mathbf{G} \cdot \mathbf{r}_j}, where j indexes the atoms, f_j is the form factor for the j-th atom, and \mathbf{r}_j is its position. This enables the determination of atomic positions, crystal structures, and material properties, with intensities proportional to |F(\mathbf{G})|^2. The form factor is often approximated numerically using sums of Gaussian functions, f(G) = \sum_i a_i e^{-b_i G^2} + c, with coefficients tabulated for elements up to high atomic numbers. Near absorption edges, the atomic form factor includes anomalous dispersion corrections, comprising real (f') and imaginary (f'') components that account for energy-dependent and , enhancing phase contrast in techniques like multiple-wavelength anomalous (MAD) for solving. Comprehensive tables of form factors, coefficients, and cross-sections are available for elements from (Z=1) to (Z=92) across energies from 1 eV to 433 keV, supporting applications in , , and . Analogous form factors exist for and , though the atomic form factor primarily refers to the case due to its electron-density sensitivity.

General Principles

Definition and Significance

The atomic form factor, also known as the atomic scattering factor, quantifies the amplitude of a scattered wave produced by an isolated atom when interacting with incident radiation such as X-rays, electrons, or neutrons. It serves as a measure of the atom's power and depends on the scattering angle as well as the type of probe used, reflecting the of scattering centers within the atom. This plays a crucial role in experiments for determining atomic and molecular structures in , where it modulates the intensity of diffracted beams by accounting for the collective from all electrons (or other scatterers) in the atom. For example, as the scattering angle increases, the form factor diminishes due to destructive among the waves scattered by electrons at different positions within the atom, thereby influencing the observed diffraction pattern's and . The concept emerged in the early 20th century amid the foundational work on by and William Lawrence Bragg in the 1910s, who demonstrated how atomic arrangements produce diffraction patterns, laying the groundwork for understanding atomic-level . For X-rays, the atomic form factor is typically dimensionless and given in units of electrons, with its value at zero scattering angle, denoted f(0), equal to the Z, representing the total number of electrons available for .

Mathematical Formulation

The atomic form factor originates from the classical Thomson scattering for a free electron, which has a scattering amplitude of −r_e (where r_e is the classical electron radius), independent of scattering angle for low energies. For bound electrons in an atom, the form factor f(\mathbf{q}) accounts for phase differences due to their spatial distribution, given by the coherent sum f(\mathbf{q}) = \sum_{j=1}^Z \exp(i \mathbf{q} \cdot \mathbf{r}_j), such that the total scattering amplitude is −r_e f(\mathbf{q}). Here, \mathbf{q} is the momentum transfer vector with magnitude q = (4\pi / \lambda) \sin \theta, \lambda is the wavelength of the incident radiation, and \theta is half the scattering angle; in the continuum limit, this becomes the Fourier transform of the electron density distribution \rho(\mathbf{r}): f(\mathbf{q}) = \int \rho(\mathbf{r}) \exp(i \mathbf{q} \cdot \mathbf{r}) \, d^3\mathbf{r}, where \rho(\mathbf{r}) represents the charge density for X-ray or electron scattering (or nuclear/magnetic density for neutrons). This formulation assumes an isolated atom, neglecting interatomic interference effects that are accounted for separately by the structure factor in crystalline materials, and treats the atom as a collection of independent scatterers under the first Born approximation. Spherical symmetry of the atomic density is often invoked, justified by the approximate isotropy of isolated atoms, allowing \rho(\mathbf{r}) = \rho(r). Under this assumption, the angular integral over the exponential phase factor simplifies, yielding: f(q) = 4\pi \int_0^\infty \rho(r) r^2 \frac{\sin(qr)}{qr} \, dr. The density \rho(r) is normalized such that \int \rho(\mathbf{r}) , d^3\mathbf{r} = Z for electron scattering, ensuring |f(0)| = Z, the atomic number, which corresponds to forward scattering where all electrons contribute in phase. These expressions hold under the kinematical approximation, valid for small scattering angles where q is modest and multiple scattering is negligible; at high energies or large angles, the form factor breaks down due to relativistic effects or the need for dynamical diffraction theory. In applications to crystalline scattering, the atomic form factor contributes to the overall intensity via I \propto |F|^2, where F is the structure factor summing form factors over lattice sites.

X-ray Form Factors

Non-resonant Scattering

In non-resonant X-ray scattering, the atomic form factor f_X(\theta) describes the scattering amplitude from the electron cloud of an atom, approximated as the sum over its Z electrons: f_X(\theta) \approx \sum_{j=1}^Z \exp(i \mathbf{q} \cdot \mathbf{r}_j), where \mathbf{q} is the momentum transfer vector with magnitude q = 4\pi \sin\theta / \lambda, \theta is the scattering angle, and \lambda is the X-ray wavelength. This expression is equivalent to the Fourier transform of the spherically symmetric atomic electron density \rho(r): f_X(q) = \int \rho(\mathbf{r}) \exp(i \mathbf{q} \cdot \mathbf{r}) \, d\mathbf{r}. At zero momentum transfer (q = 0), f_X(0) = Z, reflecting the total number of electrons, while it decreases at higher angles due to destructive interference from the distributed electron density. Tabulated values of f_X as a function of \sin\theta / \lambda are available for elements across the periodic table, enabling direct use in . For instance, the International Tables for , Volume C, provide mean atomic scattering factors in electrons for free atoms, where for carbon (Z = 6), f_X(0) = 6 and the value falls to approximately 1 at high scattering angles corresponding to \sin\theta / \lambda \approx 4 Å^{-1}. These tables, derived from numerical relativistic Dirac-Fock computations, account for core and contributions and are widely used for evaluations in refinement. Detailed tabulations, such as those by Chantler, extend this to precise form factors up to high energies, resolving discrepancies in earlier datasets. The independent atom model (IAM) approximates the electron density in the crystal as the superposition of individual atomic densities without bonding effects, leading to the structure factor F(\mathbf{G}) = \sum_j f_j(\mathbf{G}) e^{i \mathbf{G} \cdot \mathbf{r}_j}, where each f_j is the form factor for the j-th atom, often arising from core or valence electrons treated separately. Computational models for f_j often employ Gaussian expansions or Hartree-Fock wave functions to represent \rho(r); for example, Cromer and Mann's numerical Hartree-Fock calculations provide accurate f_X values by integrating radial electron densities from self-consistent field solutions. Gaussian fits, typically as a sum of 4–10 terms, facilitate rapid evaluation in refinement algorithms while maintaining fidelity to quantum mechanical densities. Thermal motion broadens the effective , introducing a Debye-Waller that modulates the : f_X(q, T) = f_X(q) \exp\left( -\frac{B \sin^2\theta}{\lambda^2} \right), where B (in Ų) is the atomic , typically 1–5 for room-temperature , quantifying mean-square atomic displacements. This isotropic approximation assumes harmonic vibrations and is essential for correcting observed intensities in patterns. In practice, f_X values are computed using crystallographic software for refinement, such as the CCP4 suite's SFALL program, which generates factors from atomic coordinates and tabulated . Python libraries, including those in the Gemmi package, offer modular access to atomic computations via interpolated tables or direct density integrations, supporting workflows in high-throughput .

Anomalous Dispersion

The anomalous form factor modifies the atomic scattering factor for X-rays near absorption edges, expressed as f = f_0 + f' + i f'', where f_0 is the non-resonant () atomic scattering factor, f' is the real dispersive correction that is typically negative and reduces the effective , and f'' is the positive imaginary absorptive component related to photoelectric . This modification arises physically when the photon energy approaches atomic absorption edges, such as K- or L-edges, where it matches electronic transitions from inner shells, leading to resonant phase shifts in the scattered wave and increased absorption. The real and imaginary parts f' and f'' are interconnected through Kramers-Kronig relations, which derive them from the energy-dependent absorption cross-section \mu(E), ensuring causality in the atomic response. Tabulated values of f' and f'' are available in databases computed using relativistic methods, such as the Cromer-Liberman approach or the NIST Form Factor Database, providing energy-specific corrections for elements across the periodic table. For example, at the K-edge (E \approx 8.98 keV), typical values are f' \approx -5 and f'' \approx 20, illustrating the significant dispersive reduction and absorptive enhancement near . In applications, anomalous dispersion enables solving the phase problem in through techniques like multiple-wavelength anomalous (MAD), where differences in scattering factors at distinct energies provide phase information via \Delta f = f'(E_2) - f'(E_1). This method has become a standard for de novo structure determination of macromolecules, leveraging sources for tunable wavelengths near edges. These corrections are valid primarily near absorption edges, within about \Delta E / E < 10\%, where resonant effects dominate; far from edges, especially for hard X-rays, the anomalous terms become negligible compared to f_0.

Electron Form Factors

Elastic Scattering Amplitude

The elastic scattering amplitude for electrons interacting with atoms is described within the first Born approximation as the Fourier transform of the atomic electrostatic potential V(\mathbf{r}), given by f_e(\mathbf{q}) = -\frac{m_e}{2\pi \hbar^2} \int V(\mathbf{r}) e^{i \mathbf{q} \cdot \mathbf{r}} \, d^3\mathbf{r}, where m_e is the electron mass, \hbar is the reduced Planck's constant, and \mathbf{q} is the momentum transfer vector with magnitude q = \frac{4\pi}{\lambda} \sin(\theta/2), \lambda being the de Broglie wavelength and \theta the scattering angle. For neutral atoms, the potential arises from the nuclear charge screened by the electron cloud, leading to an equivalent form f_e(q) = \frac{2 m_e e^2}{\hbar^2 q^2} [Z - f_X(q)], where Z is the atomic number, e is the elementary charge, and f_X(q) is the X-ray atomic form factor representing the Fourier transform of the electron density. This expression, known as the non-relativistic Mott-Bethe formula, highlights how electrons probe the atomic potential directly, in contrast to X-rays which scatter from the charge density. In the Thomas-Fermi approximation, which models the electron density statistically for high-Z atoms, the screened Coulomb potential yields an analytic form for the scattering amplitude approximately as f_e(q) \approx \frac{Z}{(1 + (q a_0 / 2)^2)^2}, where a_0 is the Bohr radius adjusted by a screening parameter dependent on Z. This approximation captures the exponential screening of the nuclear potential at distances beyond the Thomas-Fermi screening length, roughly $0.885 a_0 Z^{-1/3}, leading to a modified Rutherford scattering where the amplitude decreases more gradually with q compared to point-charge scattering. The differential elastic cross-section is then \frac{d\sigma}{d\Omega} = |f_e(q)|^2, which, due to the longer de Broglie wavelength of electrons relative to X-rays at typical energies, allows probing of atomic structure at larger scattering angles before the amplitude diminishes significantly. Relativistic corrections to the non-relativistic form become important at higher energies, introducing the formula to account for spin-orbit interactions and Dirac effects: f_\text{Mott}(q) = -\frac{Z \alpha}{2 \sin^2(\theta/2)} \left[ 1 + \beta^2 \sin^2(\theta/2) + \text{higher-order terms} \right], where \alpha = e^2 / (\hbar c) \approx 1/137 is the and \beta = v/c is the electron velocity in units of the ; detailed expansions address and screening. Unlike scattering, which is insensitive to light elements due to weak contrast in electron density, electron elastic scattering via the excels in imaging light atoms in (TEM), enabling atomic-resolution structural determination in materials like carbon-based nanostructures. Tabulated values of f_e(q) for practical computations are available in standard references, such as the International Tables for Crystallography, where for (Z = 79) at 100 keV incident energy, the remains significant (above 10% of the forward value) up to scattering angles of approximately 20–30 mrad, supporting high-angle in scanning TEM.

Relativistic and Inelastic Effects

At high energies, relativistic effects must be incorporated into the electron atomic form factor to account for the Dirac of the , particularly in scattering from the nuclear potential. The relativistic amplitude provides the point-nucleus limit, given by f = \frac{Z e^{2}}{16 \pi \epsilon_{0} E} \left[ \frac{1}{\sin^{2}(\theta/2)} - i \beta \frac{\cos(\theta/2)}{\sin^{2}(\theta/2)} + \text{spin terms} \right], where \beta = v/c is the electron velocity in units of the speed of light, E is the incident kinetic electron energy, and the imaginary term arises from spin-orbit coupling. This formulation, derived from the Dirac equation, enhances the scattering cross-section compared to non-relativistic Rutherford scattering, with the increase more pronounced for heavy elements due to stronger Coulomb fields. Inelastic effects introduce energy loss \omega during , modifying the to include excitations such as plasmons and inner-shell transitions. The imaginary part of the inelastic , \operatorname{Im}[f_{\text{inel}}(q, \omega)], is connected to the dynamic S(q, \omega) through the , which relates it to the imaginary part of the response and ensures between absorption and emission processes. This allows quantification of energy dissipation, with S(q, \omega) capturing the of fluctuations in the atomic electron cloud. Corrections to the elastic f_e(q) arise from and , incorporated via the Hartree-Fock-Slater (HFS) model, which approximates the many-electron potential with a statistical term. For greater precision, especially at high Z and relativistic speeds, the Dirac-Hartree-Fock (DHF) method solves the self-consistently, yielding improved f_e(q) by including spin-orbit interactions and orbital contraction. These approaches refine the beyond the independent approximation, reducing errors in momentum-space densities. In high-voltage (TEM) operating at 300 kV, relativistic effects significantly influence , boosting f_e(q) by a factor of approximately 1.2 for light atoms at typical angles due to increased effective and wave function contraction. Experimental and tabulated data, such as those from the EEDL97 library, support these corrections for validating simulations in materials analysis. However, the first underlying many calculations fails at low energies below 10 keV, where exact partial-wave methods are required, and multiple in condensed matter deviates from the isolated atom model, necessitating dynamical theories.

Neutron Form Factors

Nuclear Coherent Scattering

The neutron atomic form factor arising from nuclear coherent scattering is approximated by the bound coherent scattering length b_\text{coh}, which provides an isotropic and angle-independent description of the scattering amplitude f_n \approx b_\text{coh} to first order for low momentum transfers q. This parameter quantifies the coherent elastic interaction between thermal neutrons and the nucleus, enabling interference effects that reveal atomic structure in scattering experiments. The interaction stems from the short-range strong force, with the scattering length b_\text{coh} determined experimentally from low-energy scattering measurements and tabulated for practical use per . Theoretical contributions from resonances can be approximated using relations derived from the optical , but experimental values are preferred due to complex structure effects. Accounting for the finite nuclear size, the momentum-dependent form factor is given by f_n(\mathbf{q}) = b_\text{coh} \int \rho_\text{nuc}(\mathbf{r}) e^{i \mathbf{q} \cdot \mathbf{r}} \, d^3\mathbf{r}, where \rho_\text{nuc}(\mathbf{r}) is the density distribution. For typical q values and nuclear radii R_\text{nuc} \sim 1{-}5 , this approximates to f_n(q) \approx b_\text{coh} \exp\left( -q^2 R_\text{nuc}^2 / 6 \right), yielding a nearly constant since q R_\text{nuc} \ll 1. Values of b_\text{coh} vary irregularly with isotope due to differences in nuclear structure, enabling contrast variation techniques; for instance, protium (^1H) has b_\text{coh} = -3.74 , deuterium (^2H) has +6.67 , and heavier elements like natural have ~+10.3 , while many heavy nuclei range from ~+4 to +12 . Coherent scattering is separated from incoherent (spin-dependent) contributions by measuring total cross sections and spin statistics. Tabulated data, compiled in resources like the NIST Neutron Data compilation (based on ) and Koester tables, are expressed in femtometers (; 1 fm = 10^{-15} m), with typical magnitudes |b| ~ 10^{-12} cm. These properties make nuclear coherent scattering ideal for neutron diffraction studies of materials with light elements, such as locating positions in biological structures where the large negative b_\text{coh} of protium provides strong contrast against heavier s.

Magnetic Scattering

Magnetic scattering of s by atoms occurs through the dipole interaction between the 's intrinsic , \mu_n = -1.91 \mu_N (with \mu_N the ), and the generated by the s and orbital currents of atomic electrons. This contrasts with nuclear scattering by providing sensitivity to electronic magnetism, enabling probes of atomic-scale magnetic distributions. The interaction is proportional to \vec{\sigma}_n \cdot \vec{B}, where \vec{\sigma}_n is the Pauli and \vec{B} is the local from the . The differential cross-section for unpolarized neutrons in magnetic scattering is given by \frac{d\sigma}{d\Omega} \propto (\gamma_n r_0)^2 |f_\text{mag}(\mathbf{q})|^2 \sin^2 \alpha, where \gamma_n is the neutron gyromagnetic ratio, r_0 the classical electron radius, \mathbf{q} the momentum transfer, f_\text{mag}(\mathbf{q}) the magnetic form factor, and \alpha the angle between \mathbf{q} and the local magnetization direction. The \sin^2 \alpha dependence arises from the transverse nature of the magnetic interaction operator, which projects perpendicular to \mathbf{q}. The magnetic form factor itself is expressed as f_\text{mag}(\mathbf{q}) = g \mu_B \sum_j \langle \mathbf{S}_j^\perp \rangle F_\text{mag}(q), where g is the Landé g-factor, \mu_B is the Bohr magneton, \langle \mathbf{S}_j^\perp \rangle is the component of the spin expectation value perpendicular to \mathbf{q} for the j-th electron, and F_\text{mag}(q) is the Fourier transform of the atomic magnetization density, normalized such that F_\text{mag}(0) = 1. This formulation captures both the atomic structure factor from electron positions and the q-dependent smearing due to the spatial extent of the magnetic electrons. For paramagnetic or ferromagnetic atoms, an atomic approximation simplifies F_\text{mag}(q) under the dipole limit as F_\text{mag}(q) \approx \frac{\int j_1(qr) M(r) r \, dr}{M(0)}, where j_1(x) = \frac{\sin x}{x^2} - \frac{\cos x}{x} is the spherical Bessel function of the first order, and M(r) is the radial magnetization density. This approximation holds when q is small compared to the inverse atomic size, emphasizing the dipolar character of the neutron-electron coupling; F_\text{mag}(q) decays rapidly with increasing q (typically as $1/q^2 at large q) because the electron cloud extends over ~1 Å, blurring the phase factors in the Fourier transform. More precise calculations incorporate higher multipoles via radial integrals of electron wavefunctions, but the dipole term dominates for most transition metal ions. Prominent examples occur in transition metals with unpaired d-electrons, such as iron (), where the atomic is approximately 2 \mu_B (Bohr magnetons). For , f_\text{mag}(q) remains substantial up to q \approx 0.5 Å^{-1}, allowing resolution of magnetic correlations before significant falloff; beyond this, it drops to ~10% of its q=0 value due to the ~0.5–1 Å radial extent of 3d orbitals. Such form factors are routinely tabulated from polarized diffraction data on paramagnetic salts or ferromagnetic compounds, providing benchmarks for theoretical models of electron correlations. In distinction to nuclear coherent scattering (which provides a q-independent baseline from the nuclear potential), magnetic scattering is inherently vectorial and orientation-dependent, vanishing entirely for non-magnetic isotopes or atoms lacking net spin polarization. This selectivity has made it essential for elucidating complex magnetism, such as antiferromagnetic ordering in transition metal oxides, where interference between nuclear and magnetic amplitudes reveals spin alignments.

References

  1. [1]
    Atomic form factors
    The atomic form factor is the Fourier transform of the electron density of an atom. It is assumed that the electron density is spherically symmetric.
  2. [2]
    NIST Atomic Form Factors: Form factors and standard definitions
    The (x-ray) atomic form factor f is the resonant scattering amplitude of x-rays by charge (primarily electron) density.
  3. [3]
    Atomic Form Factor and Crystal Structure Factor
    The amplitude of x-ray scattering by an atom is called atomic form factor. Consider x-ray scattering by an atom. Each electron of the atom contributes to the x ...
  4. [4]
    X-Ray Form Factor, Attenuation, and Scattering Tables | NIST
    Detailed tabulation of atomic form factors, photoelectric absorption and scattering cross section, and mass attenuation coefficients for Z = 1-92 from E = 1-10 ...
  5. [5]
    Atomic scattering factor - Online Dictionary of Crystallography
    Nov 8, 2017 · A measure of the scattering power of an isolated atom. Also known as the atomic form factor. The scattering factor depends on the scattering amplitude of an ...
  6. [6]
    X-ray Crystallography: One Century of Nobel Prizes
    Oct 30, 2014 · In one century, at least 25 Nobel Prizes have been awarded for work directly or indirectly involving crystallography.The Beginnings · The First Crystal Structures of... · Data Treatment and...
  7. [7]
    Calculation of atomic scattering factors - RuppWeb
    A plot of scattering factor f in units of electrons vs. sin(theta)/lambda shows this behavior. Note that for zero scattering angle the value of f equals the ...
  8. [8]
    [PDF] Review: Thomson Scattering
    Aug 26, 2010 · for an entire atom, integrate to get the atomic form factor f o(Q):. −rof o(Q) = −ro. Z ρ(r)eiQ·rd3r the total atomic scattering factor is.
  9. [9]
    [PDF] Chapter 3 Scattering from an Atom
    Oct 29, 2013 · The atomic scattering factor represents how an atom scatters X-ray. It is dominated by the structure factor of an atom, which is sometimes ...
  10. [10]
    SFALL (CCP4: Supported Program)
    This program calculates structure factors and X-ray gradients for refinement using inverse and forward fast Fourier techniques.
  11. [11]
    Scattering - Gemmi 0.7.4-dev documentation
    Gemmi has functions to calculate structure factor by direct summation of structure factors from individual atoms. This route is not commonly used in ...Missing: MOTIF software
  12. [12]
    Relativistic Calculation of Anomalous Scattering Factors for X Rays
    Sep 1, 1970 · Anomalous scattering factors Δf′ and Δf′′ have been calculated relativistically for Cr, Fe, Cu, Mo, and Ag α Kα radiations for the atoms Li through Cf.
  13. [13]
    Anomalous X-Ray Scattering - an overview | ScienceDirect Topics
    At wavelengths where specific atoms strongly absorb i.e. K, L or M absorption edges, the X-ray scattering form factor changes due to anomalous dispersion. Both ...
  14. [14]
    Comparison between Ti anomalous x-ray scattering factors obtained ...
    One of these techniques uses the Kramers–Kronig (K-K) dispersion relationship to integrate absorption-coefficient data in order to obtain the real anomalous ...
  15. [15]
    Phase Determination by Multiple-Wavelength X-Ray Diffraction
    A novel x-ray diffraction technique, multiple-wavelength anomalous dispersion (MAD) phasing, has been applied to the de novo determination of an unknown ...
  16. [16]
    Review Advances in multiple wavelength anomalous diffraction ...
    Multiple wavelength anomalous diffraction (MAD) phasing has advanced from an esoteric technique used in only a few favorable cases to the method of choice for ...
  17. [17]
  18. [18]
    The Coulomb Scattering of Relativistic Electrons by Nuclei | Phys. Rev.
    The cross section for the Coulomb scattering of relativistic electrons by atomic nuclei has been evaluated. The exact results obtained by Mott have been ...
  19. [19]
    Inelastic scattering from core electrons: A multiple scattering approach
    Jul 21, 2005 · The real-space multiple-scattering approach is applied to model nonresonant inelastic scattering from deep core electron levels over a broad energy spectrum.
  20. [20]
    Behaviour of hartree-fock and hartree-fock-slater (Xα) atomic form ...
    Mar 14, 2005 · The atomic form factor and the radial electron density in momentum space, from the solutions of the α-parameterized Hartree-Fock-Slater (Xα) ...
  21. [21]
    [PDF] Relativistic effects in elastic scattering of electrons in TEM
    Modern transmission electron microscopes (TEM) typically operate at acceleration voltages between 80 and 300 kV leading to projectile electron velocities ...
  22. [22]
    epdl97 - IAEA Nuclear Data Services
    Nov 19, 2012 · With databases: EPDL Photon Interaction Data. EADL Atomic Relaxation Data. EEDL Electron Interaction Data. EXDL Excitation Data.Missing: form | Show results with:form
  23. [23]
    Neutron scattering lengths and cross sections
    Oct 13, 2021 · Select the element, and you will get a list of scattering lengths and cross sections. All of this data was taken from the Special Feature section of neutron ...Neutron Scattering Lengths... · Isotope · Li.html · Fe.html
  24. [24]
    [PDF] Special Feature Neutron scattering lengths and cross sectioirn
    The latter are, by convention, tabulated for k = 3.494 A--', which corresponds to a wavelength h = 1.798 A, an energy E = 25.30 meV, or a velocity v = 2200 mls.
  25. [25]
    [PDF] Summary of coherent neutron scattering length - INIS-IAEA
    The scattering length of the neutron-nucleus system is the basic quantity which describes the strength and the character of the interaction of low energy ...
  26. [26]
    The nuclear radius - Book chapter - IOPscience
    For light nuclei such as 6Li the charge density distribution is Gaussian, and the corresponding form factor is also Gaussian: ... The measurement of the neutron ...
  27. [27]
    [PDF] An Introduction to Neutron Scattering
    A uniform scattering length density only gives forward scattering (Q=0), thus SANS measures deviations from average scattering length density. • If ρ is the SLD ...
  28. [28]
    Neutron scattering lengths and cross sections
    A knowledge of the scattering lengths and the corresponding scattering and absorption cross sections of the elements.
  29. [29]
    Where is the Hydrogen? | Elements - GeoScienceWorld
    Jun 1, 2021 · Neutron diffraction can locate hydrogen sites in mineral structures, reveal any static or dynamic hydrogen disorder, help define the libration ...Comparison Of X-Ray And... · Bragg Scattering · Hydrogen In Mineral...
  30. [30]
    On the Magnetic Scattering of Neutrons | Phys. Rev.
    This paper elaborates on magnetic scattering of neutrons, discussing interaction functions, coherent/incoherent scattering, and the cross section due to  ...
  31. [31]
    [PDF] Chapter 1 MAGNETIC NEUTRON SCATTERING.
    Much of our understanding of the atomic-scale magnetic structure and the dynamical properties of solids and liquids was gained from neutron- scattering studies.<|control11|><|separator|>
  32. [32]
    Form factors in magnetic scattering of thermal neutrons
    This lecture addresses the concept of form factor in magnetic scattering of thermal neutrons, analyzing its meaning, discussing its measurement by polarized ...
  33. [33]
    Magnetic form factor of metallic iron and nickel as seen by inelastic ...
    Oct 1, 1981 · The magnetic form factors of iron and nickel have been studied by means of coherent inelastic vibrational scattering of polarized neutrons ...
  34. [34]
    Magnetic Form Factors - Institut Laue-Langevin
    Magnetic Form Factors · Definition of form factors · Analytical approximation · Tables of Form Factors · <j0> Form factors for 3d transition elements and their ...