Fact-checked by Grok 2 weeks ago

Structure factor

The structure factor \mathbf{F}_{hkl} is a mathematical function that describes the amplitude and phase of a wave diffracted from a crystalline material by lattice planes with Miller indices h, k, and l. It encapsulates the collective scattering contributions from all atoms in the unit cell, depending on their positions and scattering factors, and serves as the central quantity in diffraction-based structure analysis across X-ray, neutron, and electron methods. Mathematically, the structure factor is expressed as \mathbf{F}_{hkl} = \sum_j f_j \exp\left[2\pi i (h x_j + k y_j + l z_j)\right], where the sum is over all atoms j in the unit cell, f_j is the atomic scattering factor (which approximates the number of electrons for X-rays or related to nuclear properties for neutrons), and (x_j, y_j, z_j) are the fractional coordinates of atom j. This complex-valued quantity can be decomposed into real and imaginary parts: F_{hkl} = A_{hkl} + i B_{hkl}, with A_{hkl} = \sum_j f_j \cos[2\pi (h x_j + k y_j + l z_j)] and B_{hkl} = \sum_j f_j \sin[2\pi (h x_j + k y_j + l z_j)]. The diffracted intensity for each reflection is proportional to |F_{hkl}|^2, but direct measurement yields only the amplitude |F_{hkl}|, while phases must be inferred, posing the well-known phase problem in crystallography. In practice, the structure factor determines the presence and intensity of diffraction peaks, revealing symmetries and systematic absences that aid in identifying crystal systems—for instance, in face-centered cubic lattices, reflections are allowed only when h, k, l are all even or all odd, yielding F_{hkl} = 4f for permitted peaks and zero otherwise. Beyond atomic resolution, it enables Fourier synthesis to reconstruct electron or nuclear density maps, facilitating the determination of molecular structures, bond lengths, and material properties in fields from materials science to biology. Variations like partial structure factors extend its use to disordered or amorphous systems, where they describe average scattering from subsets of atoms or components.

Fundamentals

Definition and Physical Significance

The structure factor, denoted as S(\mathbf{q}), represents the Fourier transform of the pair correlation function describing the distribution of atomic or electron densities within a material, thereby quantifying the collective amplitude and phase of waves scattered coherently from these density distributions. This function captures the interference effects arising from the spatial arrangement of scatterers, distinguishing it from single-particle scattering contributions. In elastic scattering processes, where incident waves such as X-rays, neutrons, or electrons interact with matter without energy loss, the structure factor determines the modulation of scattered intensity based on the scattering vector \mathbf{q}, providing a direct probe of microscopic structural features. Physically, the structure factor encodes essential information about material organization, including lattice periodicity in crystals, density fluctuations in liquids, and short-range atomic ordering in amorphous solids, enabling the inference of interatomic distances, coordination numbers, and overall structural motifs from diffraction patterns. For periodic structures, sharp peaks in S(\mathbf{q}) at reciprocal lattice vectors reveal long-range order, while broadening or diffuse scattering in disordered systems highlights local variations and correlations. This makes it indispensable for characterizing phase transitions, defects, and nanoscale heterogeneities across diverse condensed matter systems. The origins of the structure factor concept trace back to early 20th-century X-ray crystallography, initiated by Max von Laue's 1912 demonstration that crystals diffract X-rays, confirming their wave nature and periodic atomic lattice. William Henry Bragg and William Lawrence Bragg further advanced this in 1913 by formulating Bragg's law, which linked diffraction angles to interplanar spacings and laid the groundwork for interpreting scattering intensities through atomic arrangements. Post-1940s developments extended its application to neutron scattering, with pioneering experiments by Enrico Fermi and collaborators in 1944 at the CP-3 reactor enabling studies of light elements and magnetic structures inaccessible to X-rays. Similarly, electron scattering techniques evolved in the mid-20th century to probe surface and thin-film structures, broadening the utility of structure factor analysis.

Basic Mathematical Formulation

The scattering vector \mathbf{q} is defined as \mathbf{q} = \mathbf{k}_f - \mathbf{k}_i, where \mathbf{k}_i and \mathbf{k}_f are the wavevectors of the incident and scattered radiation, respectively, with |\mathbf{k}_i| = |\mathbf{k}_f| for elastic scattering processes. This vector \mathbf{q} determines the momentum transfer to the sample and encodes the spatial scale probed by the scattering experiment, with |\mathbf{q}| = \frac{4\pi}{\lambda} \sin(\theta) in terms of wavelength \lambda and scattering angle $2\theta. The basic mathematical formulation of the static structure factor S(\mathbf{q}) describes the coherent elastic scattering from a collection of N scattering centers located at positions \mathbf{r}_j. It is expressed as S(\mathbf{q}) = \frac{1}{N} \sum_{j=1}^N \sum_{k=1}^N \exp\left[i \mathbf{q} \cdot (\mathbf{r}_j - \mathbf{r}_k)\right], which is mathematically equivalent to the squared modulus form S(\mathbf{q}) = \frac{1}{N} \left| \sum_{j=1}^N \exp\left(i \mathbf{q} \cdot \mathbf{r}_j\right) \right|^2. This expression captures the interference effects arising from the relative phases of waves scattered by different centers, normalized by the number of scatterers to yield a dimensionless quantity that approaches 1 at high \mathbf{q} (uncorrelated scattering). To account for the intrinsic scattering properties of individual atoms or nuclei, the formulation is extended by incorporating scattering amplitudes specific to the probe. For X-ray scattering, each term in the sum is weighted by the atomic form factor f_j(\mathbf{q}), which represents the scattering from the electron cloud of atom j and depends on the scattering angle: S(\mathbf{q}) = \frac{1}{N} \left| \sum_{j=1}^N f_j(\mathbf{q}) \exp\left(i \mathbf{q} \cdot \mathbf{r}_j\right) \right|^2, where f_j(0) = Z_j (the atomic number) at zero angle, decreasing with |\mathbf{q}| due to the finite size of the electron distribution. In neutron scattering, constant scattering lengths b_j (isotope- and spin-dependent) replace f_j, simplifying the expression since b_j is independent of \mathbf{q}. The structure factor S(\mathbf{q}) is typically computed as an ensemble average \langle S(\mathbf{q}) \rangle over configurations or time to incorporate statistical fluctuations and separate coherent (position-correlated) from incoherent (self-scattering) contributions, with the incoherent part adding a flat background of unity. This averaging ensures S(\mathbf{q}) reflects equilibrium structural correlations in the sample, such as pair distribution functions via Fourier transform.

Derivation

Derivation of the Structure Factor S(q)

The derivation of the structure factor S(\mathbf{q}) begins with the scattering amplitude in the first Born approximation, which is applicable to elastic scattering processes where the incident wave interacts weakly with the sample, neglecting higher-order multiple scattering effects. For X-ray or neutron scattering, the scattering amplitude A(\mathbf{q}) is proportional to the Fourier transform of the scattering length density \rho(\mathbf{r}), given by A(\mathbf{q}) \propto \int \rho(\mathbf{r}) \exp(i \mathbf{q} \cdot \mathbf{r}) \, d\mathbf{r}, where \mathbf{q} = \mathbf{k}_i - \mathbf{k}_f is the scattering vector with |\mathbf{q}| = (4\pi/\lambda) \sin(\theta/2), \lambda is the wavelength, and \theta is the scattering angle. This form arises from the kinematic approximation, assuming plane-wave incident and scattered waves, and is valid for systems where the potential is weak compared to the incident energy. For a system of N discrete atoms or scattering centers with positions \mathbf{r}_j (assuming identical atomic form factors for simplicity, or incorporating them separately), the total scattering amplitude becomes the sum over individual contributions: A(\mathbf{q}) = \sum_{j=1}^N \exp(i \mathbf{q} \cdot \mathbf{r}_j). The measured intensity I(\mathbf{q}) is proportional to the squared modulus |A(\mathbf{q})|^2, averaged over thermal ensembles if necessary. Normalizing by the number of scatterers yields the structure factor: S(\mathbf{q}) = \frac{1}{N} \left| \sum_{j=1}^N \exp(i \mathbf{q} \cdot \mathbf{r}_j) \right|^2 = \frac{1}{N} \sum_{j=1}^N \sum_{k=1}^N \exp[i \mathbf{q} \cdot (\mathbf{r}_j - \mathbf{r}_k)]. This double sum separates into a coherent part, \left| \frac{1}{N} \sum_j \exp(i \mathbf{q} \cdot \mathbf{r}_j) \right|^2, capturing interference between different atoms, and an incoherent (self) part, \frac{1}{N} \sum_j 1 = 1, representing single-atom scattering without positional correlations. The derivation assumes elastic scattering (static positions or time-averaged) and isotropic averaging over orientations for powders or liquids. In the continuum limit for dense systems like liquids or amorphous materials, the positions \mathbf{r}_j are treated statistically, with \rho(\mathbf{r}) = \sum_j \delta(\mathbf{r} - \mathbf{r}_j). The structure factor then relates to the density-density correlation function, specifically the pair distribution function g(\mathbf{r}), which describes the probability of finding a particle at \mathbf{r} relative to another at the origin. The Fourier transform yields S(\mathbf{q}) = 1 + \rho \int [g(\mathbf{r}) - 1] \exp(i \mathbf{q} \cdot \mathbf{r}) \, d\mathbf{r}, where \rho = N/V is the average number density, the "1" accounts for self-correlations, and the integral captures distinct pair correlations. This form assumes a homogeneous, isotropic fluid under the single-scattering (Born) approximation, neglecting anharmonic or many-body effects beyond pairwise.

Connection to Density and Correlation Functions

The structure factor S(\mathbf{q}) provides a direct measure of density fluctuations in a system, linking scattering experiments to statistical mechanics through the fluctuation-dissipation theorem. In this context, it is expressed as S(\mathbf{q}) = \frac{1}{N} \left< |\delta \rho(\mathbf{q})|^2 \right>, where N is the number of particles, \delta \rho(\mathbf{q}) is the Fourier component of the local density deviation \delta \rho(\mathbf{r}) = \rho(\mathbf{r}) - \left< \rho \right>, and the angular brackets indicate an ensemble average over thermal fluctuations (for \mathbf{q} \neq 0, \delta \rho(\mathbf{q}) = \rho(\mathbf{q}) since \left< \rho(\mathbf{q}) \right> = 0). This relation highlights how S(\mathbf{q}) captures the amplitude of collective density modes at wavevector \mathbf{q}, with the forward scattering limit S(\mathbf{q} \to 0) corresponding to the normalized variance of particle number fluctuations in a subvolume. Note that N = \left< \rho \right> V, where V is the system volume. A more detailed statistical interpretation emerges from expanding S(\mathbf{q}) in terms of the pair correlation function g(\mathbf{r}), which describes the probability of finding two particles separated by \mathbf{r} relative to a random distribution. The structure factor is given by S(\mathbf{q}) = 1 + \rho \int \left[ g(\mathbf{r}) - 1 \right] \exp(i \mathbf{q} \cdot \mathbf{r}) \, d\mathbf{r}, where \rho is the average density. This Fourier transform representation connects S(\mathbf{q}) to real-space structural correlations in liquids and amorphous systems, with the term g(\mathbf{r}) - 1 quantifying deviations from ideal gas behavior due to interparticle interactions. In liquid theory, g(\mathbf{r}) is obtained from molecular simulations or integral equation approximations, allowing S(\mathbf{q}) to be computed as a diagnostic of short- and long-range order. At long wavelengths (\mathbf{q} \to 0), the structure factor relates thermodynamic properties via the compressibility equation, S(0) = \rho k_B T \kappa_T, where k_B is Boltzmann's constant, T is the temperature, and \kappa_T is the isothermal compressibility. This equation, derived from the grand canonical ensemble, equates the zero-wavevector limit of density correlations to the system's susceptibility to volume changes under pressure, providing a bridge between microscopic structure and macroscopic thermodynamics such as equation-of-state data. For ideal gases, \kappa_T = 1/(\rho k_B T) yields S(0) = 1, while interactions in dense liquids typically suppress S(0) < 1. Further insight into these correlations is provided by the Ornstein-Zernike equation, which decomposes the total pair correlation h(\mathbf{r}) = g(\mathbf{r}) - 1 into direct contributions c(\mathbf{r}) and indirect chains mediated by the medium: h(\mathbf{r}) = c(\mathbf{r}) + \rho \int c(\mathbf{r}') h(|\mathbf{r} - \mathbf{r}'|) \, d\mathbf{r}'. In Fourier space, this yields S(\mathbf{q}) = [1 - \rho \tilde{c}(\mathbf{q})]^{-1}, where \tilde{c}(\mathbf{q}) is the Fourier transform of the direct correlation function. The function c(\mathbf{r}), which decays more rapidly than h(\mathbf{r}), encodes irreducible two-body interactions and serves as input for closure approximations in liquid theory, enabling predictions of S(\mathbf{q}) from potential models.

Perfect Crystals

Units and Notation

In crystallography and scattering theory, the scattering vector \mathbf{q} is defined as the difference between the wavevectors of the scattered and incident beams, \mathbf{q} = \mathbf{k}_f - \mathbf{k}_i, with magnitude |\mathbf{q}| = 4\pi \sin\theta / \lambda, where \theta is half the scattering angle and \lambda is the radiation wavelength. This vector is commonly expressed in units of inverse angstroms (Å⁻¹) for practical measurements or in reciprocal lattice units (rlu) for indexing relative to the crystal lattice. For the static structure factor S(\mathbf{q}), which quantifies scattering intensity normalized by the number of scatterers, the value is dimensionless, reflecting the average correlation of atomic positions. For perfect crystals, the structure factor F_{hkl} corresponds to diffraction peaks at reciprocal lattice vectors \mathbf{G}_{hkl}, where \mathbf{q} = \mathbf{G}_{hkl} satisfies the Laue condition. The Miller indices (hkl) are integers denoting the family of lattice planes, with the reciprocal lattice vector given by \mathbf{G}_{hkl} = 2\pi (h \mathbf{a}^* + k \mathbf{b}^* + l \mathbf{c}^*), where \mathbf{a}, \mathbf{b}, \mathbf{c} are the direct lattice basis vectors and the reciprocal basis vectors are \mathbf{a}^* = (\mathbf{b} \times \mathbf{c}) / V, \mathbf{b}^* = (\mathbf{c} \times \mathbf{a}) / V, \mathbf{c}^* = (\mathbf{a} \times \mathbf{b}) / V, with V the unit cell volume (using the physics convention incorporating $2\pi for Fourier consistency). In X-ray diffraction, F_{hkl} has units of electrons, as it sums contributions from atomic form factors f_j \approx Z_j (atomic number) in the forward limit. For neutron diffraction, F_{hkl} is in units of femtometers (fm), summing nuclear scattering lengths b_j (typically 2–15 fm), such that |F_{hkl}|^2 yields coherent scattering cross-sections in barns (1 barn = 10⁻²⁴ cm²). The structure factor F_{hkl} is generally a complex quantity, F_{hkl} = |F_{hkl}| \exp(i \phi_{hkl}), where the magnitude |F_{hkl}| determines diffraction intensity via I_{hkl} \propto |F_{hkl}|^2, and the phase \phi_{hkl} encodes positional information essential for structure reconstruction. Phase factors arise from the exponential term \exp(i \mathbf{G}_{hkl} \cdot \mathbf{r}_j) in the summation over atomic positions \mathbf{r}_j. Crystal symmetry, described by space groups, imposes constraints: equivalent reflections have identical |F_{hkl}| but related phases, while systematic absences occur for specific (hkl) (e.g., h + k + l odd in body-centered lattices), rendering F_{hkl} = 0. Thermal motion attenuates scattering amplitudes through the Debye-Waller factor, which multiplies each atomic form factor f_j by \exp(-B \sin^2 \theta / \lambda^2), where B is the isotropic temperature factor (typically 0.2–0.8 Ų, increasing with temperature) and accounts for mean-square atomic displacements \langle u^2 \rangle via B = 8\pi^2 \langle u^2 \rangle. This factor is real and less than unity, broadening and reducing peak intensities without altering peak positions. For anisotropic cases, a tensor form replaces B, but the isotropic approximation suffices for introductory notation.

Structure Factor F_hkl for Infinite Crystals

For infinite perfect crystals, the structure factor F_{hkl} quantifies the amplitude and phase of the scattered wave from the unit cell for a specific reflection indexed by Miller indices h, k, and l, assuming the Laue condition is met whereby the scattering vector equals a reciprocal lattice vector \mathbf{G}_{hkl} = 2\pi (h \mathbf{a}^* + k \mathbf{b}^* + l \mathbf{c}^*). This condition arises from the summation over all lattice points in an infinite crystal, which yields delta functions \delta_{\mathbf{q}, \mathbf{G}} at reciprocal lattice points, confining diffraction to those discrete positions. The mathematical definition of F_{hkl} is given by the sum over all atoms n in the unit cell: F_{hkl} = \sum_n f_n \exp \left[ 2\pi i (h x_n + k y_n + l z_n) \right] where f_n is the atomic scattering factor for atom n (dependent on the scattering angle and atom type), and (x_n, y_n, z_n) are the fractional coordinates of atom n within the unit cell. This expression represents the coherent interference of waves scattered from each atom, with the exponential term accounting for phase shifts due to atomic positions relative to the origin. In a Bravais lattice crystal, the total scattering separates into the lattice contribution (enforcing the Laue condition via S_{hkl} = \delta_{\mathbf{q}, \mathbf{G}_{hkl}}, which is effectively infinite at allowed points for an ideal infinite crystal) and the motif or basis contribution, simplifying F_{hkl} to S_{hkl} \times \sum_j f_j \exp(i \mathbf{G}_{hkl} \cdot \mathbf{r}_j), where the sum is over atoms j in the basis at positions \mathbf{r}_j. The intensity of the (hkl) reflection is then proportional to the squared modulus of this structure factor, I_{hkl} \propto |F_{hkl}|^2, modulated by geometric factors such as multiplicity (number of equivalent reflections) and the Lorentz-polarization correction in experimental setups. Symmetry elements in the space group can lead to systematic absences, where F_{hkl} = 0 for certain indices, resulting in missing reflections. For instance, a twofold screw axis along \mathbf{c} (e.g., $2_1) causes F_{hkl} = 0 unless l is even, due to the translational component introducing destructive interference; similarly, a c-glide plane perpendicular to \mathbf{a} yields absences when h + l is odd. These absences stem directly from the phase factors in the structure factor sum becoming zero under the symmetry operations, aiding in space group determination without full structure solution.

Examples in Three Dimensions

In three-dimensional crystals, the structure factor F_{hkl} for infinite perfect lattices reveals systematic selection rules determined by the atomic basis within the unit cell, leading to allowed and forbidden reflections that reflect the symmetry of the structure. These examples illustrate how the phase differences from atom positions result in constructive or destructive interference for specific Miller indices (hkl). For the body-centered cubic (BCC) structure, which consists of atoms at the corners and one at the body center (positions: (0,0,0) and (1/2,1/2,1/2)), the structure factor is given by
F_{hkl} = f \left[ 1 + e^{i \pi (h + k + l)} \right],
where f is the atomic scattering factor. This simplifies to F_{hkl} = 2f when h + k + l is even and F_{hkl} = 0 when h + k + l is odd, due to destructive interference in the latter case. For instance, reflections like (110) and (200) are allowed, while (100) and (111) are forbidden.
The face-centered cubic (FCC) structure features atoms at the corners and face centers (positions: (0,0,0), (1/2,1/2,0), (1/2,0,1/2), (0,1/2,1/2)), yielding
F_{hkl} = 4f
if h, k, l are all even or all odd (unmixed indices), and F_{hkl} = 0 otherwise, as mixed indices produce phase cancellation. Allowed reflections include (111) and (200), whereas (100) and (211) are absent, highlighting the symmetry-imposed extinctions.
In the diamond cubic structure, common to elements like silicon and germanium, the lattice is FCC with a two-atom basis at (0,0,0) and (1/4,1/4,1/4). The structure factor is
F_{hkl} = 8f
when h + k + l = 4n (where n is an integer), F_{hkl} = 0 for h + k + l = 4n \pm 2, and for the cases where h + k + l is odd ($4n \pm 1), it takes the form F_{hkl} = 4f (1 \pm i), resulting in |F_{hkl}|^2 = 32 f^2. These rules arise from the combined FCC selection (all even or all odd indices) and the basis phase shift, forbidding reflections like (200) while allowing (111) and (220).
The zincblende structure, adopted by compounds like GaAs and ZnS, is analogous to diamond but with distinct atom types on the basis (e.g., cation at (0,0,0), anion at (1/4,1/4,1/4)). The structure factor becomes
F_{hkl} = 4 (f_\text{cation} + f_\text{anion} e^{i \pi (h + k + l)/2})
for unmixed indices (h, k, l all even or all odd), yielding |F_{hkl}|^2 = 16 (f_\text{cation} + f_\text{anion})^2 when h + k + l ≡ 0 \pmod{4}, $16 (f_\text{cation} - f_\text{anion})^2 when ≡ 2 \pmod{4}, and $16 (f_\text{cation}^2 + f_\text{anion}^2) when h + k + l is odd, with F_{hkl} = 0 for mixed indices. This leads to intensity variations dependent on the scattering factor difference, with forbidden reflections mirroring FCC.
For the cesium chloride (CsCl) structure, a simple cubic lattice with atoms at (0,0,0) and (1/2,1/2,1/2) of different types, the structure factor is
F_{hkl} = f_\text{Cs} + f_\text{Cl} e^{i \pi (h + k + l)},
resulting in F_{hkl} = f_\text{Cs} + f_\text{Cl} for h + k + l even and f_\text{Cs} - f_\text{Cl} for odd, with no inherent zeros but reduced intensity when the difference is small. Reflections like (100) and (110) are thus observable, unlike in BCC.
Hexagonal close-packed (HCP) structures, such as those in magnesium and zinc, use four-index Miller-Bravais notation (hkil) with i = -(h + k) to account for the three-fold basal symmetry. The unit cell has two identical atoms at (0,0,0) and (2/3, 1/3, 1/2), giving
F_{hkil} = f \left[ 1 + e^{2\pi i (2h/3 + k/3 + l/2)} \right].
Allowed reflections satisfy specific conditions: |F_{hkil}|^2 = 4f^2 for l even and $3n = h - k (e.g., (0002), (11\bar{2}0)); |F_{hkil}|^2 = f^2 for l even and $3n \pm 1 = h - k (e.g., (10\bar{1}0)); |F_{hkil}|^2 = 0 for l odd and $3n = h - k (e.g., (0001)); and |F_{hkil}|^2 = 3f^2 for l odd and $3n \pm 1 = h - k (e.g., (10\bar{1}1)). These rules produce characteristic diffraction patterns distinguishing HCP from cubic phases.

Examples in One and Two Dimensions

In one dimension, consider a perfect monatomic chain of atoms spaced by lattice constant a. The structure factor S(\mathbf{q}) is given by the sum over all lattice sites, S(q) = \sum_{n=-\infty}^{\infty} \exp(i q n a), which, for an infinite chain, yields a series of delta functions at the reciprocal lattice points q = 2\pi m / a, where m is an integer. The intensity at these points is proportional to the square of the atomic scattering factor f, reflecting constructive interference from identical atoms. When the one-dimensional chain includes a basis of multiple atoms per unit cell, the structure factor incorporates phase differences from their relative positions. For a basis with atoms at fractional coordinates x_j along the chain direction, it becomes F_h = \sum_j f_j \exp(2\pi i h x_j), where h is the Miller index corresponding to the reciprocal lattice vector $2\pi h / a. This can lead to destructive interference and systematic absences for specific h values if the phases cancel, such as F_h = 0 for odd h in a two-atom basis separated by a/2. Extending the notation from three dimensions by setting irrelevant indices to zero, the two-dimensional square lattice with lattice constant a has structure factor F_{hk} = \sum_j f_j \exp[2\pi i (h x_j + k y_j)], where x_j and y_j are the fractional coordinates of atoms in the unit cell. For a primitive square lattice with a single atom at the origin, F_{hk} = f for all integers h, k, allowing reflections at all reciprocal lattice points without systematic absences. In cases with additional symmetry, such as a base-centered square lattice, absences occur for reflections where h + k is odd due to the structure factor vanishing from destructive interference. For the two-dimensional primitive hexagonal lattice, formed by basis vectors of equal length at 120° to each other, the structure factor takes the form F_{hk} = \sum_j f_j \exp[2\pi i (h x_j + k y_j)], adapted to the hexagonal coordinate system using Miller-Bravais indices where the third index i = -(h + k). All integer h, k are permitted in the primitive case with no systematic absences for a single-atom basis. Compared to three dimensions, one- and two-dimensional perfect crystals exhibit theoretically infinite long-range positional order, but their diffraction patterns show inherently weaker enforcement of correlations across the structure, leading to less sharp reciprocal lattice spots in practice due to the reduced geometric constraints and increased susceptibility to fluctuations.

Imperfect Crystals

Finite-Size Effects

In perfect infinite crystals, the structure factor manifests as delta functions at reciprocal lattice vectors, producing sharp diffraction peaks. However, real crystals are finite in size, leading to a truncation of the lattice sum in the structure factor expression. For a finite crystal, the total scattering amplitude near a reciprocal lattice point \mathbf{G}_{hkl} is given by A(\mathbf{q}) = F_{hkl} \sum_{m} \exp[i 2\pi (\mathbf{q} - \mathbf{G}_{hkl}) \cdot \mathbf{R}_m], where the sum is over the N unit cell positions \mathbf{R}_m within the crystal volume, and F_{hkl} is the unit-cell structure factor. This finite sum replaces the infinite crystal's delta function with a broadened distribution, typically of sinc-squared form for simple geometries, resulting in peak widths \Delta q \approx 2\pi / L along each dimension, where L is the crystal dimension in that direction. The broadening arises from the shape transform of the crystal, which is the Fourier transform of the crystal's geometric envelope. For a cubic crystal of side length L, the intensity near \mathbf{G}_{hkl} approximates I(\mathbf{q}) \propto N^2 \left[ \frac{\sin(\pi (\mathbf{q} - \mathbf{G}_{hkl}) \cdot \mathbf{L}/2)}{\pi (\mathbf{q} - \mathbf{G}_{hkl}) \cdot \mathbf{L}/2} \right]^2, where \mathbf{L} is the vector of dimensions, yielding a full width at half maximum \Delta q \sim 2\pi / L and peak intensity scaling as N^2 for fully coherent scattering, with N = L^3 / v and v the unit cell volume. This effect is most pronounced in nanocrystals, where L < 100 nm leads to detectable broadening, as quantified by the Scherrer equation for the angular width \beta \approx K \lambda / (L \cos \theta), with shape factor K \approx 0.9 and \lambda the X-ray wavelength. Surface-to-volume ratio effects further reduce overall intensity relative to the infinite case, as only interior planes contribute coherently. In the Laue construction, finite size extends reciprocal lattice points into rods (rel rods) along directions perpendicular to the crystal faces, with length inversely proportional to the corresponding dimension L. The Ewald sphere then intersects these rods over a finite range, producing diffuse scattering streaks or broadened spots rather than points, particularly evident in small or thin crystals. For mosaic crystals, comprising coherent domains of size L with an orientation spread \eta (typically 0.1–1°), additional angular broadening \Delta \theta \approx \eta convolves with the shape-induced width, as modeled by Warren's mosaic block theory, where domain misorientations mimic a polycrystalline aggregate. This combined broadening is separable via profile analysis, with size effects dominating for L < 100 nm and mosaic effects for larger but imperfect crystals.

Disorder of the First Kind

Disorder of the first kind, originally termed replacement disorder by André Guinier, describes static random substitutions of atoms in a crystal lattice that maintain the overall long-range translational order while introducing compositional variations at individual sites. This type of disorder is characteristic of substitutional alloys, such as binary systems A_{1-x}B_x, where atoms A and B occupy equivalent lattice positions randomly, with occupancy probabilities (1-x) and x, respectively. In structural refinement, such random occupations are modeled using average atomic scattering factors, f = (1-x)f_A + x f_B, which yield the structure factor F_{hkl} for the average lattice. The effect on the structure factor S(q) manifests as an averaging of the Bragg reflections, where the intensity at reciprocal lattice vectors G = (hkl) is proportional to |<F_{hkl}>|^2, with <F_{hkl}> denoting the ensemble average over configurations. However, local deviations from the average structure due to differences in atomic scattering lengths produce diffuse scattering distributed throughout reciprocal space. In cases of pure compositional disorder without size mismatch, this results in Laue monotonic scattering, a uniform background intensity proportional to x(1-x)|f_A - f_B|^2. When substituting atoms have significantly different atomic radii, local strain fields arise around each substitutional defect, leading to long-range distortions that couple to the lattice vibrations. These strains produce asymmetric diffuse scattering centered near the Bragg peaks, known as Huang scattering, which extends over a region in reciprocal space scaled by the inverse of the defect-induced displacement field. The Huang scattering intensity I_H(q) near a reciprocal lattice point G can be approximated as I_H(q) \propto |F_G|^2 \left( \frac{\Delta V}{V} \right)^2 \frac{1}{(q \cdot e)^2}, where \Delta V/V is the relative volume change per defect, e is the polarization vector, and q is the deviation from G; this form highlights the 1/|q|^2 decay characteristic of dipole-like strain fields. A classic example occurs in binary alloys like Cu-Zn (brass), where random substitution leads to reduced intensities of fundamental reflections based on the average structure, accompanied by Huang diffuse scattering lobes elongated along directions sensitive to the size mismatch between Cu and Zn atoms. In the rocksalt superstructure, such as disordered NaCl-like alloys with 50% occupancy of two atom types on each sublattice, the superlattice reflections at half-integer indices are extinguished in the average structure, with their intensity redistributed as diffuse scattering at those positions due to random occupations.

Disorder of the Second Kind

Disorder of the second kind encompasses thermal and dynamic effects in crystals that primarily dampen the intensities of Bragg peaks in the structure factor without introducing peak splitting, arising from atomic vibrations and related motions. These effects are distinct from static positional irregularities and manifest as a reduction in coherent scattering amplitude due to the time-averaged positions of atoms deviating from their ideal lattice sites. The primary manifestation is through thermal diffuse scattering (TDS), where inelastic scattering from phonons contributes to diffuse intensity around reciprocal lattice points, while the elastic Bragg scattering is attenuated. The attenuation of Bragg peak intensities is quantified by the Debye-Waller factor, which multiplies the ideal structure factor squared as |F|^2 \exp(-2W), where W = \langle u^2 \rangle q^2 / 3 for isotropic vibrations in three dimensions, with \langle u^2 \rangle the mean-square atomic displacement and q the scattering vector magnitude. This factor originates from the Fourier transform of the thermally averaged atomic positions, effectively smearing the electron density and reducing phase coherence. Originally derived by Debye for the interference effects of X-rays with heat motion, the factor was refined by Waller to account for quantum harmonic vibrations, showing that \langle u^2 \rangle depends on temperature via phonon occupation. In practice, TDS arises as the complementary intensity, representing one-phonon processes that transfer energy and momentum without contributing to sharp Bragg reflections. Phonon contributions to the scattering are captured by the dynamic structure factor S(\mathbf{q}, \omega), which extends the static structure factor to include energy transfer \omega. For one-phonon processes in harmonic crystals, the inelastic scattering term near a reciprocal lattice vector \mathbf{G} is proportional to |\mathbf{G} \cdot \mathbf{e}|^2 / \omega \, [n(\omega) + 1/2 \pm 1/2], summed over phonon branches and wavevectors, where \mathbf{e} is the phonon polarization vector, \omega the phonon frequency, and n(\omega) the Bose-Einstein occupation factor; the +1/2 term corresponds to Stokes (energy gain) and anti-Stokes (energy loss) processes. This formulation, central to neutron and X-ray inelastic scattering, reveals how lattice vibrations modulate the structure factor, with the elastic part damped by the Debye-Waller factor and inelastic parts forming the TDS wings. The full S(\mathbf{q}, \omega) integrates over multiphonon contributions at higher orders, but the one-phonon approximation dominates near Bragg peaks. In cases involving defects or diffusive motions within the crystal lattice, dynamic disorder introduces quasi-elastic scattering, characterized by a Lorentzian broadening of the elastic peak in S(\mathbf{q}, \omega). This broadening, with half-width proportional to the diffusion coefficient D as \Gamma = \hbar D q^2, reflects over-damped or relaxational dynamics without discrete energy transfers, often observed in ionic conductors or defect-laden crystals via neutron spectroscopy. Such effects combine with thermal vibrations, enhancing the overall damping while preserving the average lattice periodicity. At high temperatures, the behavior approaches the classical limit described by the Einstein model, where each atom vibrates independently in a harmonic potential with frequency \omega_E. Here, \langle u^2 \rangle = 3 k_B T / m \omega_E^2, making the Debye-Waller factor linearly dependent on temperature, W \approx (k_B T q^2) / (m \omega_E^2), and the TDS intensity scales as T times the static structure factor. This classical approximation holds when k_B T \gg \hbar \omega_E, simplifying calculations for elevated-temperature diffraction and aligning with the equipartition theorem for vibrational energy.

Disordered Systems

Structure Factor in Liquids

In liquids, the structure factor S(\mathbf{q}) characterizes the spatial correlations between particles in a dense fluid, serving as the Fourier transform of the pair correlation function g(r). Unlike in crystals, where sharp Bragg peaks dominate, the structure factor in liquids exhibits a broad, oscillatory form reflecting short-range order and the absence of long-range periodicity. This function is central to interpreting scattering experiments, such as neutron or X-ray diffraction, and encapsulates thermodynamic properties through its limiting behaviors. For an ideal gas of non-interacting particles, the structure factor simplifies to S(q) = 1 across all wavevectors q, indicating no correlations beyond random thermal motion. In dense liquids, interactions introduce deviations, with S(q) modulating based on particle packing and potential energies. At high q (short distances), S(q) approaches 1, reflecting the dominant self-scattering term, but small oscillations persist due to short-range order from excluded volume effects and weak interatomic forces. These oscillations decay with increasing q, diminishing the influence of correlations. In the low-q limit (long wavelengths), S(q \to 0) = \rho k_B T \kappa_T, where \rho is the number density, k_B is Boltzmann's constant, T is temperature, and \kappa_T is the isothermal compressibility; this relation arises from density fluctuations and the fluctuation-dissipation theorem. A prototypical model for simple liquids is the hard-sphere fluid, where particles interact via repulsive cores without attraction. The Percus-Yevick approximation provides an analytical solution for the direct correlation function, yielding S(q) with pronounced oscillations that mimic packing effects, such as a principal peak corresponding to the average interparticle distance and subsequent minima and secondary peaks from nearest-neighbor shells. This approximation accurately captures the structure for moderate densities but underestimates compressibility at high packing fractions. For real liquids, S(q) shows similar features tailored to molecular interactions. In liquid water at ambient conditions, neutron scattering reveals a broad first peak at q \approx 2.0 Å^{-1} , attributed to the tetrahedral coordination in the first hydration shell, with weaker oscillations at higher q from hydrogen bonding. In liquid metals like sodium or aluminum, the structure factor displays a sharp principal peak at q \approx 2.5–$3.0 Å^{-1} , signaling dense packing of the first coordination shell, alongside a low-q rise consistent with metallic compressibility. These peaks provide direct insight into local atomic arrangements, bridging microscopic correlations to macroscopic properties.

Structure Factor in Polymers

In polymeric materials, the structure factor describes the spatial arrangement of monomer units along flexible chains, influencing scattering patterns in techniques such as small-angle X-ray scattering (SAXS) and small-angle neutron scattering (SANS). For a single isolated polymer chain modeled as a Gaussian chain, the structure factor, often referred to as the form factor P(q), is given by the Debye function: P(q) = \frac{2}{x^2} \left( x - 1 + e^{-x} \right), where x = q^2 R_g^2 and R_g is the radius of gyration. At intermediate scattering vectors q (where $1/R_g \ll q \ll 1/l and l is the monomer length), this approximates to P(q) \approx 2/(q^2 R_g^2), reflecting the random walk statistics of the chain and a power-law decay in scattering intensity. This regime highlights the coil-like conformation, with deviations occurring for real chains due to excluded volume effects or stiffness. In polymer solutions or melts, the total structure factor S(q) combines the single-chain form factor P(q) with an interchain structure factor S_{\text{inter}}(q), accounting for correlations between different chains: S(q) \approx P(q) \cdot S_{\text{inter}}(q). In dilute solutions, S_{\text{inter}}(q) \approx 1, reducing to the isolated chain case, while in semidilute solutions or melts, interchain interactions suppress large-scale fluctuations, leading to a plateau at low q and enhanced scattering at higher q due to screening. The random phase approximation (RPA) provides a framework for S_{\text{inter}}(q), particularly in blends or multicomponent systems, where it captures composition fluctuations. This decomposition allows separation of intra- and interchain contributions, revealing how concentration and chain entanglement affect overall morphology. Crystalline polymers exhibit ordered structures, such as lamellar stacks formed by chain folding, which produce distinct features in the structure factor. In SAXS patterns, meridional reflections along the chain direction correspond to the lamellar thickness (typically 10-20 nm), while equatorial reflections arise from lateral packing of lamellae or chain stems. These peaks indicate long-range periodicity, with the structure factor modulated by the electron density contrast between crystalline and amorphous regions. For example, in polyethylene, SAXS reveals a long period of about 20 nm from lamellar stacking, alongside wide-angle reflections from the orthorhombic unit cell. Amorphous polymers, lacking long-range order, display a broad scattering halo in the structure factor due to short-range correlations from local chain packing and van der Waals interactions, typically centered at q \approx 1.5 \, \AA^{-1} with no sharp Bragg peaks. This halo reflects pairwise monomer distances of around 0.4-0.5 nm, arising from conformational preferences rather than crystalline registry. In polystyrene, SANS profiles show this amorphous halo dominating the mid-q range, with low-q upturn from chain coils, illustrating how local order persists without global periodicity.

Structure Factor in Glasses and Amorphous Materials

In glasses and amorphous materials, the static structure factor S(q) captures frozen structural correlations akin to those in the supercooled liquid state, where atomic dynamics are arrested upon vitrification, preserving short- and medium-range order without long-range periodicity. This results in a diffuse scattering pattern with characteristic peaks reflecting pairwise atomic distributions, similar to liquids at short distances but lacking diffusive motion that would otherwise broaden features over time. The pair correlation function g(r), derived via Fourier transform from S(q), shows pronounced oscillations at short r (e.g., nearest-neighbor distances) that decay more slowly than in simple liquids due to the rigidity of the network. A hallmark of the structure factor in these materials is the first sharp diffraction peak (FSDP) appearing at intermediate q values (typically 0.5–2 Å⁻¹), indicative of medium-range order spanning 5–10 Å, such as quasi-periodic arrangements of coordination polyhedra or voids in the atomic network. The FSDP arises from chemical and spatial correlations, often modeled as pre-peaks in partial structure factors, and its position and intensity shift with composition or density, providing insights into topological constraints. For instance, in silica glass (v-SiO₂), the total S(q) exhibits an FSDP at approximately 1.5 Å⁻¹ linked to tetrahedral SiO₄ units and Si-Si correlations over medium range, alongside a principal peak at ~2.4 Å⁻¹ corresponding to nearest-neighbor O-O distances around 2.6 Å. These features distinguish amorphous silica from crystalline polymorphs like quartz, where long-range order sharpens peaks. The boson peak manifests as an excess in low-q scattering intensity in the structure factor, associated with terahertz vibrational modes that deviate from Debye-like behavior, contributing to anomalous low-temperature specific heat. This excess, observed around q \approx 1 Å⁻¹ or lower, stems from quasi-localized modes hybridized with plane waves, reflecting structural heterogeneity and disorder in the amorphous matrix. In network glasses like silica, the boson peak correlates with the FSDP, as both probe similar length scales of disorder, with simulations showing that network connectivity modulates the peak's prominence. Paracrystalline models describe the structure factor of glasses as arising from finite-sized, distorted lattices with weak positional correlations, leading to Lorentzian-broadened peaks instead of delta-function Bragg reflections. In silica glass, this is conceptualized as small microparacrystals (e.g., 3 netplane layers) with a high distortion parameter g \approx 12\%, where lattice fluctuations smear out long-range order, reproducing the observed diffuse halo and FSDP as remnants of underlying tetrahedral packing. Unlike liquids, where thermal motion continuously disrupts correlations, the paracrystalline framework in glasses emphasizes static topological disorder, with g(r) showing damped oscillations that align with experimental neutron and X-ray data.

Applications and Extensions

Experimental Measurement Techniques

The structure factor S(\mathbf{q}) is experimentally determined through scattering experiments using X-rays, neutrons, or electrons, where the scattered intensity is proportional to the square of the structure factor modulated by atomic form factors and other experimental factors. These techniques probe the Fourier transform of the pair correlation function, providing insights into atomic arrangements across various length scales. X-ray diffraction is the most common method due to its accessibility, while neutron and electron scattering complement it for specific material properties. In X-ray diffraction, laboratory sources such as rotating anode generators or sealed tubes produce characteristic radiation (e.g., Cu Kα at 1.54 Å wavelength) suitable for routine measurements on single crystals or powders, but they offer limited flux and resolution compared to synchrotron sources. Synchrotron radiation provides tunable, high-brilliance X-rays with fluxes orders of magnitude higher, enabling studies of weak scattering signals, small samples, or time-resolved dynamics; it is particularly advantageous for powder diffraction where broad q-coverage is needed. Single-crystal X-ray diffraction measures discrete Bragg peaks to extract structure factors via integrated intensities, while powder diffraction captures Debye-Scherrer rings from polycrystalline samples, averaging over orientations. Area detectors like CCD or pixel arrays record 2D patterns, which are processed to yield I(q) \propto |F(\mathbf{q})|^2, from which S(\mathbf{q}) is derived after corrections for polarization, Lorentz factors, and multiplicity. Neutron scattering employs thermal or cold neutrons from reactor sources, which provide steady-state beams for high-resolution measurements, or spallation sources, which generate pulsed neutrons via proton bombardment for time-of-flight (TOF) spectrometers that access wide energy and q-ranges efficiently. Reactors excel in continuous flux for precise powder or single-crystal studies, whereas spallation sources support broader dynamic range for diffuse scattering. A key advantage is isotopic contrast variation, where substituting isotopes (e.g., H for D) alters coherent scattering lengths, allowing isolation of partial structure factors S_{ab}(q) for multicomponent systems like liquids or alloys through combinations of measurements on isotopically labeled samples. Detectors such as 3He tubes or scintillator arrays capture the scattering cross-section, and S(\mathbf{q}) is obtained from the differential cross-section after normalization to the incident flux and subtraction of incoherent background. Electron diffraction, often performed in transmission electron microscopy (TEM), is ideal for nanoscale samples (e.g., nanocrystals or thin films) where X-rays or neutrons require larger volumes. Selected-area electron diffraction (SAED) or convergent-beam electron diffraction (CBED) probes local structure factors, with diffuse scattering analysis revealing disorder contributions beyond Bragg peaks. High-voltage TEM (100-300 kV) provides short de Broglie wavelengths (~0.02-0.037 Å), enabling high-q resolution, though multiple scattering effects necessitate kinematic approximations or dynamical corrections for accurate S(\mathbf{q}) extraction from intensity patterns. Data processing for all techniques involves azimuthal integration or radial averaging of 2D detector images to obtain the 1D scattering profile I(q), essential for isotropic samples like powders where Debye-Scherrer rings form concentric patterns. Software tools apply geometric corrections, mask beam stops, and normalize to absolute scale using standards (e.g., vanadium for neutrons). The structure factor is then computed as S(q) = I(q) / \langle f(q)^2 \rangle + 1, accounting for self-scattering and form factors, with forward scattering normalization ensuring S(0) relates to compressibility. Resolution limits vary by probe: X-ray techniques typically access q from ~0.01 Å⁻¹ (small-angle, synchrotron SAXS) to 10 Å⁻¹ (wide-angle lab XRD), limited by source brilliance and detector size. Neutron scattering covers similar ranges, with reactors favoring low-q resolution (~0.01-5 Å⁻¹) and spallation extending to higher q via TOF (~0.1-20 Å⁻¹), constrained by neutron flux and absorption. Electron diffraction achieves the highest q (~1-50 Å⁻¹) due to short wavelengths but is limited at low q by sample thickness and spherical aberration, typically probing ~0.5-20 Å⁻¹ in practice.

Relation to Diffraction Patterns

The structure factor plays a central role in interpreting diffraction patterns by determining the positions and relative strengths of observed features, such as spots in single-crystal diffraction or rings in powder patterns. In crystalline materials, diffraction peaks arise at specific scattering vectors \mathbf{q} corresponding to reciprocal lattice points, where the structure factor F_{hkl} is non-zero, reflecting constructive interference from the atomic arrangement. According to Bragg's law, constructive interference occurs when n\lambda = 2d \sin\theta, where n is an integer, \lambda is the wavelength, d is the interplanar spacing, and \theta is the Bragg angle; this condition defines peak positions at q = 4\pi \sin\theta / \lambda, with high-intensity peaks manifesting where the structure factor |F_{hkl}| is large. The law ensures that only waves scattered from parallel lattice planes reinforce at these angles, but the structure factor modulates whether a reflection is allowed or forbidden based on the basis of atoms within the unit cell. The intensity of each diffraction feature is primarily governed by I_{hkl} \propto |F_{hkl}|^2, where F_{hkl} sums the contributions from all atoms in the unit cell, weighted by their atomic scattering factors f_j and positions. Additional factors, including the Lorentz-polarization correction (accounting for detector geometry and beam polarization) and absorption (due to sample thickness and composition), further modulate these intensities, ensuring that observed patterns accurately reflect the underlying atomic structure. In powder diffraction, the random orientations of microcrystallites produce concentric Debye-Scherrer rings, with the azimuthal average over these rings yielding a one-dimensional intensity profile I(q) proportional to the orientationally averaged structure factor S(q). This averaged S(q) relates to the real-space pair correlation function g(r) through a Fourier transform, S(q) = 1 + \rho_0 \int [g(r) - 1] e^{i\mathbf{q}\cdot\mathbf{r}} d\mathbf{r}, where \rho_0 is the average number density, enabling extraction of structural information like interatomic distances from the pattern. Anomalous scattering introduces wavelength-dependent corrections to the atomic scattering factors near X-ray absorption edges, making f_j = f_0 + f' + i f'' complex and site-specific, which alters the structure factor F_{hkl} and produces measurable intensity differences between Bijvoet pairs (hkl and -h-k-l). These near-edge effects, strongest when the X-ray energy matches electronic transitions (e.g., for selenium at \lambda \approx 0.98 Å), allow differentiation of atomic sites and facilitate phase determination in diffraction patterns. Diffraction patterns provide only the magnitudes |F_{hkl}| from measured intensities, as phases \phi_{hkl} are not directly observable, posing the phase problem that prevents straightforward Fourier reconstruction of electron density \rho(\mathbf{r}) = \frac{1}{V} \sum_{hkl} F_{hkl} e^{2\pi i (hx + ky + lz)}. Phases must be recovered using indirect methods, such as direct methods (exploiting probabilistic relations like triplet phase invariants \phi_h + \phi_k - \phi_{h+k} \approx 0) or anomalous dispersion, to fully interpret the structure factor and solve the crystal structure.

References

  1. [1]
    Structure factor - Online Dictionary of Crystallography
    Nov 20, 2017 · The structure factor \mathbf{F}_{hkl} is a mathematical function describing the amplitude and phase of a wave diffracted from crystal lattice planes.
  2. [2]
    1.2. The structure factor - Wiley Online Library
    Introduction. The structure factor is the central concept in structure analysis by diffraction methods. Its modulus is called the structure amplitude.
  3. [3]
    [PDF] Crystal Structure Analysis - Rutgers Physics
    The structure factor gives the amplitude of a scattered wave arising from the atoms with a single primitive cell. STRUCTURE FACTOR. For crystals composed of ...
  4. [4]
    [PDF] Handout 6: Structure Factors
    “Structure Determination by X-ray Crystallography”, Ladd and Palmer, Plenum, 1994. o For any given wave, we can write F = |F|·(cos(αt) + i·sin(αt)) = |F|·eiαt ...Missing: formula | Show results with:formula
  5. [5]
    [PDF] Phase Problem in X-ray Crystallography, and Its Solution
    The mathematical representation of this wave is called a structure factor. Unfortunately, it is only possible to measure the amplitude of the diffraction ...
  6. [6]
    None
    ### Summary of Structure Factor in Diffraction from the Document
  7. [7]
    Structure Factor - an overview | ScienceDirect Topics
    The structure factor is defined as a mathematical expression that represents the sum over positions of particles illuminated by an incident beam, ...
  8. [8]
    Calculating Structure Factors of Protein Solutions by Atomistic ... - NIH
    In essence, the structure factor S ( q ) measured by SAXS and SANS is the Fourier transform of the pair distribution function, which in turn is determined by ...
  9. [9]
    [PDF] Equilibrium correlation functions and Scattering
    In short, the scattered intensity is proportional to the structure factor as shown in Eq. (7.7), and the structure factor is the Fourier transform of the par ...
  10. [10]
    [PDF] PDF
    structure factor S(k) and g(r) are related by Fourier transform, an osciallatory behavior in one function means the other function will also show oscillations.
  11. [11]
    [PDF] CALCULATING PROPERTIES Space correlation functions Radial ...
    The calculation of the structure factor S(k) is achieved via a Fourier transform of the pair distribution function g(r). Equal to the computed structur efactor ...
  12. [12]
    [PDF] Demystifying X-ray Crystallography - stoltz2.caltech.edu
    Jan 5, 2018 · The Development of X-ray Diffraction. 1912. Discovery of X-ray diffraction by crystals. Max von Laue. German physicist won the Nobel. Prize in ...
  13. [13]
    A Brief History of Materials R&D at Argonne National Laboratory ...
    Sep 3, 1996 · Neutron scattering has a long history at Argonne: Enrico Fermi, Bob Sachs and Bill Sturm did pioneering experiments around 1947 at the CP-3 ...
  14. [14]
    [PDF] An Introduction to Neutron Scattering
    Elastic scattering and definition of the structure factor, S(Q). 4. Coherent & incoherent scattering. 5. Inelastic scattering. 6. Magnetic scattering. 7 ...<|control11|><|separator|>
  15. [15]
    [PDF] The influence of accidental deviations of density on the equation of ...
    ORNSTEIN and DJ', F. ZERNIKE. (Communicated by Prof. H. A. LORENTZ). (Communicated in the meetin« or June 24, 1916). In thei!' article in the Encyclopädie ...
  16. [16]
    [PDF] Structure factors II - MIT OpenCourseWare
    With the help of Braggs law and the Ewald construction, we can calculate the place of a reflection on the detector, provided we know the unit cell dimensions.
  17. [17]
    Fifteen years of the Protein Crystallography Station - IUCr Journals
    Oct 17, 2016 · Neutron and X-ray crystallography are complementary methods for structure–function studies of macromolecules, with X-rays useful in the location ...
  18. [18]
    [PDF] Elastic neutron scattering
    Important reduction of refinable parameters! The structure factor is the Fourier transform of the unit cell scattering potential functions. The magnetic form ...
  19. [19]
    Temperature Effects
    fT = f exp(−B sin2θ / λ2). where T is the temperature and B = 8π2<u2> (units ... The B values are often referred to as B factors (Debye-Waller is a ...
  20. [20]
    [PDF] Analytical Techniques
    mathematically infinite, it is realistically a large sum. For an infinite sum, the structure factor will be delta functions at these reciprocal space points.
  21. [21]
    [PDF] MAT SCI 361: Crystallography and Diffraction
    Jan 7, 2025 · This course covers elementary crystallography and basic diffraction theory with a reciprocal space perspective.
  22. [22]
    [PDF] Systematic Absences and Space Group determination
    May 1, 2014 · The structure factor is the amplitude of scattered radiation by a crystal unit cell. Symmetry operations can affect the structure factor and ...
  23. [23]
    [PDF] Crystal Structure Factor Calculation
    The conventional UC has lattice points as the vertices. ❑ There may or may not be atoms located at the lattice points.Missing: Miller | Show results with:Miller
  24. [24]
    Theory of Simple Liquids - ScienceDirect.com
    This chapter focuses on theories for calculating the thermodynamic and structural properties of non-uniform fluids. It shows how useful approximations can be ...
  25. [25]
    Small-angle scattering and the structure of ambient liquid water - NIH
    In particular, we show that Huang et al. have misrepresented standard density fluctuations as “concentration fluctuations” to justify a shape and size of ∼13–14 ...Missing: δρ( | Show results with:δρ(
  26. [26]
    Structure of Liquid Water by Neutron Scattering - Oxford Academic
    A comprehensive analysis of the structure of liquid water was performed by combining neutron and X-ray diffraction data. A formula expressing the structure ...
  27. [27]
    The Structure Factors of Liquid Metals in Low Q Region
    Aug 9, 2025 · The low Q structure factor is a slowly increasing function of Q and its profile is quite similar among the liquid metals presently investigated.
  28. [28]
    Molecular-weight Determination by Light Scattering.
    A gram–Charlier analysis of scattering to describe nonideal polymer conformations. Macromolecules 2024, 57 (20) , 9518-9535.
  29. [29]
    Characterizing polymer structure with small-angle neutron scattering
    May 3, 2021 · For an isolated polymer chain, the angular dependence of the scattered neutrons depends on the form factor P ( Q ) ⁠. The form factor amplitude ...
  30. [30]
    Small-Angle X-Ray and Neutron Scattering by Polymethylene ...
    Mar 1, 1976 · Monte Carlo Prediction of the Structure Factor of Polyethylene in Good and ϑ-Solvents. Macromolecules 1996, 29 (5) , 1721-1727. https://doi ...
  31. [31]
    Thermodynamically consistent random phase approximation for ...
    Jul 16, 2019 · The calculated structure factor of ternary polymer blend is shown to be in good agreement with the small angle neutron scattering experiments.
  32. [32]
    [PDF] Structure Determination of Polymer Nanocomposites by Small Angle ...
    Dec 3, 2010 · volume of one particle, and F2(q) is called the normalized particle form factor (sometimes written P(q)). It is normalized because its low-q ...
  33. [33]
    SAXS Investigation of Hierarchical Structures in Biological Materials
    The sharpness of the meridional reflection relates to the longitudinal coherence length (~1010 nm). The breadth of the reflections in the lateral direction ...
  34. [34]
    Insight into the Structure and Dynamics of Polymers by Neutron ...
    Dec 21, 2020 · Intermolecular correlations are revealed by the 'amorphous halo' in the range of scattering vector Q ≈ 1 Å − 1 (see Figure 1).
  35. [35]
    Computational Reverse-Engineering Analysis of Scattering ...
    Jan 5, 2021 · Computational Reverse-Engineering Analysis for Scattering Experiments for Form Factor and Structure Factor Determination (“P(q) and S(q) ...<|control11|><|separator|>
  36. [36]
    Relationship between the boson peak and first sharp diffraction ...
    Mar 20, 2025 · The FSDP appears as a strong peak at the lowest angle in the structure factor S\left( k \right) of the glass, where k is the wavenumber. The ...
  37. [37]
    [PDF] Relationship between the boson peak and first sharp diffraction ...
    The FSDP appears as a strong peak at the lowest angle in the structure factor S(k) of the glass 30,31, where k is the wavenumber. The reciprocal of FSDP ...
  38. [38]
    Origin of the first sharp diffraction peak in glasses | Phys. Rev. B
    Oct 12, 2015 · ... FSDP in the concentration-concentration partial structure factor of silica glass [6] . Furthermore, it fails to justify the appearance of ...
  39. [39]
    Machine learning molecular dynamics reveals the structural origin of ...
    Nov 16, 2023 · The first sharp diffraction peak (FSDP) in the total structure factor has long been regarded as a characteristic feature of medium-range ...
  40. [40]
    Insights into the origin of the first sharp diffraction peak in ...
    The FSDP and the principal peak of F ( Q ) are located at 1.5 and 2.8 Å − 1 , respectively.
  41. [41]
    Intermediate range order in vitreous silica from a partial structure ...
    Aug 7, 2025 · Intermediate range order in vitreous silica from a partial structure factor analysis. October 2008; Physical Review B 78(14):144204. DOI ...
  42. [42]
    Understanding the emergence of the boson peak in molecular glasses
    Jan 13, 2023 · The boson peak represents a peak in the vibrational density of states corresponding to an excess in the density of states over that expected ...
  43. [43]
    [PDF] Structural Origin of Boson Peak in Glasses - arXiv
    Boson peak, the excess low energy excitations in the terahertz regime, is one of the most unique features of disordered systems and has been linked to many.
  44. [44]
    Paracrystalline Lattice Structure of Silica Glass, α‐ and β‐Crystobalite
    ### Key Points on Paracrystalline Model for Structure Factor in Silica Glass
  45. [45]
    Intermediate range order in vitreous silica from a partial structure ...
    Oct 20, 2008 · In this paper, we report the successful measurement of the partial structure factors of vitreous SiO 2 obtained using the method of isotopic ...
  46. [46]
    Comparative study of conventional and synchrotron X-ray electron ...
    Sep 4, 2023 · Five different electron density datasets obtained from conventional and synchrotron single crystal X-ray diffraction experiments are compared.
  47. [47]
    Recent advances in synchrotron X-ray studies of the atomic ...
    Dec 20, 2024 · Similar to the conventional laboratory X-ray diffraction (XRD), synchrotron XRD works on the same principle of Bragg's law. Fig. 8(a) shows that ...
  48. [48]
    Hydrogen/deuterium isotope effects in water and aqueous solutions ...
    The partial structure factors Sab(Q) are obtained from measurements on samples of different isotopic composition (e.g., with H2O, D2O and mixtures of the ...
  49. [49]
    Analysis of diffuse scattering in electron diffraction data for the ...
    The electron diffraction data contain sharp Bragg reflections and strong diffuse streaks, associated with severe stacking disorder.
  50. [50]
    (PDF) Processing two-dimensional X-ray diffraction and small-angle ...
    A software package for the calibration and processing of powder X-ray diffraction and small-angle X-ray scattering data is presented.
  51. [51]
    [PDF] X-ray and Neutron Scattering - The Schreiber Group
    Feb 16, 2025 · The series of lectures focuses on experimental techniques using X-rays and neutrons. We will discuss differences and similarities between these ...
  52. [52]
    [PDF] Chapter 3 X-ray diffraction • Bragg's law • Laue's condition
    Bragg angle is just the half of the total angle by which the incident beam is deflected. θ θ2. There are different ways of sectioning the crystal into planes, ...<|separator|>
  53. [53]
    None
    ### Summary of Structure Factor and Diffraction Patterns
  54. [54]
    Estimating the structure factors in X-ray diffraction - PMC
    The meaning of the structure factor and how it impacts on the determination of structures are reassessed. A route to obtaining the structure factors is ...
  55. [55]
    [PDF] Anomalous scattering - MIT OpenCourseWare
    Anomalous Scattering. Anomalous scattering causes small but measurable differences in intensity between the reflections hkl and -h-k-l not normally present.
  56. [56]
    None
    ### Summary of the Phase Problem from the Document