Fact-checked by Grok 2 weeks ago

Crystallography

Crystallography is the branch of science devoted to the study of molecular and crystalline structures and their properties, focusing on the arrangement of atoms, ions, or molecules in solids that exhibit long-range order. A , in this context, is defined as a where its constituent particles form, on average, a periodic, repeating pattern in , enabling the determination of atomic positions through techniques. The field originated in the late 18th century with René-Just Haüy's observations of crystal geometry and cleavage patterns, which laid the foundations for understanding in natural minerals. Modern crystallography emerged in 1912 when , Walter Friedrich, and Paul Knipping discovered diffraction by crystals, demonstrating that X-rays could reveal atomic-scale structures. This breakthrough was rapidly advanced by and William Lawrence Bragg, who developed methods to interpret diffraction patterns and determine crystal structures, earning them the 1915 . Key techniques in crystallography include X-ray diffraction, which uses X-rays to probe crystal lattices and produce diffraction patterns for structural analysis; neutron diffraction, valuable for locating light atoms like hydrogen; and electron crystallography, applied to thinner samples or non-crystalline materials. Single-crystal methods provide high-resolution atomic models, while analyzes polycrystalline samples for phase identification and refinement. These approaches rely on mathematical principles of , space groups, and transforms to reconstruct three-dimensional structures from two-dimensional data. Crystallography has profound applications across disciplines, underpinning materials science by revealing defect structures and phase transitions in alloys and semiconductors; in chemistry, it confirms molecular geometries and reaction mechanisms; and in biology, it has elucidated protein and nucleic acid structures, such as the double helix structure of DNA, elucidated through key X-ray diffraction studies by Rosalind Franklin and others. In drug discovery, it enables structure-based design by visualizing ligand-protein interactions at atomic resolution, accelerating the development of therapeutics. The field's impact extends to geosciences for mineral identification and to physics for studying quantum materials, with ongoing advances in synchrotron sources and computational modeling, including recent developments as of 2025 such as AI integration for structure prediction and quantum crystallography, enhancing resolution and throughput.

Fundamentals

Crystal Structure Basics

A crystal is defined as a solid material in which the constituent atoms, ions, or molecules are arranged in a highly ordered, repeating three-dimensional , exhibiting long-range positional that extends throughout the entire structure. This periodicity arises from the regular arrangement of particles in a , distinguishing crystals from other solids by their structural coherence over macroscopic distances. In crystals, atomic or molecular packing refers to the spatial arrangement of these particles to minimize empty space, often modeled as in theoretical analyses. The , which indicates the number of nearest neighbors surrounding a given atom, typically reaches 12 in the most efficient close-packed structures, such as hexagonal close-packed (HCP) or cubic close-packed (CCP) arrangements. Packing efficiency, the fraction of the total volume occupied by the particles, achieves a maximum of 74% in these closest-packed configurations for equal-sized spheres, leaving 26% as interstitial voids. Crystals exhibit several key physical properties stemming from their ordered structure, including anisotropy, where properties such as mechanical strength, electrical conductivity, or vary with direction due to the non-uniform atomic bonding. is the tendency to break along specific crystallographic planes of weakness, producing flat, parallel surfaces, as seen in minerals like . describes the external shape or form of the crystal, influenced by growth conditions and the relative development of faces, often appearing as prismatic, tabular, or equant morphologies. In contrast to amorphous solids, which lack long-range order and thus display isotropic properties with irregular surfaces and no distinct cleavage planes, crystals show sharp melting points and well-defined geometric fragments upon breaking. Crystals can be classified into four main types based on the dominant bonding interactions: ionic crystals, such as (NaCl), where cations and anions alternate in a held by electrostatic forces, resulting in high brittleness and melting points; covalent network crystals, like , featuring extensive covalent bonds in a three-dimensional framework, conferring exceptional hardness; metallic crystals, exemplified by , with positive ions in a "sea" of delocalized electrons enabling and ; and molecular crystals, where discrete molecules are linked by weak van der Waals or hydrogen bonds, leading to relatively low melting points. The fundamental building block of a crystal structure is the unit cell, the smallest repeating that, when translated in three dimensions, generates the entire . It is characterized by three edge lengths (a, b, c) and three interaxial (α, β, γ), with the cell V calculated as V = a × b × c × √[1 - cos²α - cos²β - cos²γ + 2 cosα cosβ cosγ]. This asymmetric unit encapsulates the essential and atomic positions, allowing the full crystal to be reconstructed periodically.

Lattice Systems and Bravais Lattices

Crystal lattices provide the foundational geometric framework for understanding the periodic arrangement of atoms in crystalline solids. These lattices are classified into seven crystal systems based on the relationships between the unit cell edge lengths a, b, c and the interaxial angles \alpha (between b and c), \beta (between a and c), and \gamma (between a and b). The triclinic system has no restrictions, with a \neq b \neq c and \alpha \neq \beta \neq \gamma \neq 90^\circ. The monoclinic system features a \neq b \neq c and \alpha = \gamma = 90^\circ, \beta \neq 90^\circ. Orthorhombic imposes a \neq b \neq c and \alpha = \beta = \gamma = 90^\circ. Tetragonal requires a = b \neq c and \alpha = \beta = \gamma = 90^\circ. Trigonal (or rhombohedral) has a = b = c and \alpha = \beta = \gamma \neq 90^\circ. Hexagonal sets a = b \neq c with \alpha = \beta = 90^\circ, \gamma = 120^\circ. Finally, the cubic system mandates a = b = c and \alpha = \beta = \gamma = 90^\circ. These parameters define the symmetry constraints that govern lattice formation.
Crystal SystemEdge LengthsAngles
Triclinica \neq b \neq c\alpha \neq \beta \neq \gamma \neq 90^\circ
Monoclinica \neq b \neq c\alpha = \gamma = 90^\circ, \beta \neq 90^\circ
Orthorhombica \neq b \neq c\alpha = \beta = \gamma = 90^\circ
Tetragonala = b \neq c\alpha = \beta = \gamma = 90^\circ
Trigonala = b = c\alpha = \beta = \gamma \neq 90^\circ
Hexagonala = b \neq c\alpha = \beta = 90^\circ, \gamma = 120^\circ
Cubica = b = c\alpha = \beta = \gamma = 90^\circ
Within these systems, there are 14 distinct Bravais lattices, representing all possible unique three-dimensional translations that maintain invariance. These include (P) lattices with points only at cell corners; body-centered (I) with an additional point at the center; face-centered (F) with points at the centers of all faces; and base-centered (C, A, or B) with a point at the center of one pair of faces. The triclinic system has only a primitive lattice. Monoclinic includes primitive and base-centered. Orthorhombic features primitive, base-centered, body-centered, and face-centered. Tetragonal has primitive and body-centered. Trigonal is primitive (often described in rhombohedral setting). Hexagonal is primitive. Cubic encompasses primitive, body-centered, and face-centered. Representative examples include the face-centered cubic (FCC) lattice in metals like aluminum (Al) and (Cu), which achieves high packing efficiency, and the body-centered cubic (BCC) in (Fe) at .
Crystal SystemBravais LatticesExample Material (if applicable)
TriclinicPrimitive (P)
MonoclinicPrimitive (P), Base-centered (C) (C)
OrthorhombicPrimitive (P), Base-centered (C), Body-centered (I), Face-centered (F) (P), La₂CuO₄ (C), U (I), PtN (F)
TetragonalPrimitive (P), Body-centered (I)TiO₂ (P, ), β-Sn (I)
TrigonalPrimitive (R, rhombohedral)
HexagonalPrimitive (P), Zn
CubicPrimitive (P), Body-centered (I), Face-centered (F) (P), α-Fe (I), (F)
Bravais lattices establish the real-space periodicity of crystals, where the structure repeats identically under translations by integer multiples of the basis vectors \mathbf{a}, \mathbf{b}, \mathbf{c}, ensuring long-range order. This periodicity underpins properties like , calculated as the per unit , where the unit cell V for a general triclinic is given by V = abc \sqrt{1 - \cos^2\alpha - \cos^2\beta - \cos^2\gamma + 2\cos\alpha\cos\beta\cos\gamma}, with theoretical density \rho = \frac{n M}{N_A V}, n being atoms per cell, M the , and N_A Avogadro's number. For instance, the FCC yields a packing fraction of \frac{\pi \sqrt{2}}{6} \approx 0.74 for , higher than BCC's \frac{\pi \sqrt{3}}{8} \approx 0.68. parameters also influence elasticity, as elastic moduli (e.g., K) scale with interatomic distances and bonding strength; stiffer lattices with shorter bonds exhibit higher moduli, as seen in diamond's cubic structure versus softer halides. The complements this by providing a geometric in reciprocal space, where vectors are defined as \mathbf{a}^* = \frac{\mathbf{b} \times \mathbf{c}}{V}, \mathbf{b}^* = \frac{\mathbf{c} \times \mathbf{a}}{V}, \mathbf{c}^* = \frac{\mathbf{a} \times \mathbf{b}}{V}, with volume V^* = \frac{1}{V}. It serves as the basis for the Fourier transform of the real-space electron density, facilitating diffraction analysis without delving into derivations here. Point groups may influence which Bravais lattice is selected for a given crystal, as detailed in symmetry discussions.

Symmetry in Crystals

Point Group Symmetry

Point group symmetry in crystallography describes the finite set of symmetry operations—such as rotations, reflections, and inversions—that leave at least one point in space fixed while mapping a crystal's local structure onto itself. These operations characterize the external and internal atomic arrangement at specific sites, without involving translations that extend symmetry across the entire ./02%3A_Rotational_Symmetry/2.04%3A_Crystallographic_Point_Groups) The limits the possible rotational symmetries in periodic crystal structures to 1-, 2-, 3-, 4-, or 6-fold axes, excluding others like 5-fold rotations because higher-order rotations would disrupt the 's translational periodicity. This theorem arises from the requirement that operations must preserve the discrete points, ensuring that rotated positions coincide with equivalent lattice sites. Due to this restriction, there are exactly 32 crystallographic point groups, classified into seven crystal systems: triclinic (2 groups), monoclinic (3), orthorhombic (3), tetragonal (7), trigonal (5), hexagonal (7), and cubic (5). For instance, the triclinic system includes the C_1 group with no symmetry beyond identity and the C_i group with inversion only, while the cubic system features the highly symmetric O_h group incorporating octahedral rotations and reflections. Point groups are denoted using two primary notations: the Schoenflies system, common in molecular (e.g., D_{3d} for a group with a 3-fold axis, perpendicular 2-fold axes, and inversion), and the (Hermann-Mauguin) system, preferred in crystallography (e.g., \bar{3}m for the same group). (CaCO_3), a trigonal mineral, exemplifies the D_{3d} (\bar{3}m) , where its rhombohedral structure exhibits a 3-fold rotation axis, mirror planes, and an inversion center. These symmetries influence physical properties, particularly in non-centrosymmetric point groups (21 of the 32), which lack an inversion and enable phenomena like , where mechanical stress induces electric . Of these, 20 groups exhibit , excluding the cubic 432 group due to its specific rotational constraints that forbid the necessary tensor components.

Space Group Symmetry

Space groups represent the complete symmetry operations of a three-dimensional , encompassing both the rotational and reflectional symmetries of point groups and the infinite translational symmetries of the underlying , augmented by non-symmorphic elements such as screw axes and glide planes. These groups account for the periodic repetition of motifs in crystals, distinguishing them from point groups by incorporating translations that maintain the periodicity. The 230 distinct s in three dimensions arise from systematically combining the 32 crystallographic point groups with the 14 Bravais lattices, while including fractional translations from rotations and glide reflections that are compatible with the lattice. For example, the P21/c (No. 14) is prevalent in monoclinic compounds, featuring a twofold and a that impose specific constraints on molecular arrangements. Each is uniquely identified by its Hermann-Mauguin symbol, which encodes the lattice type (e.g., P for primitive) and the elements present. Within a space group, symmetry operations are formally denoted as \{R \mid \mathbf{t}\}, where R is an from the point group (such as a or ) and \mathbf{t} is a , often fractional relative to the lattice vectors. The full set of these operations generates all equivalent positions from a starting point. classify these equivalent sites, grouping points that share the same site-symmetry subgroup—a that fixes the point under symmetry operations—and multiplicity, which indicates how many such sites occupy the unit cell (e.g., general positions have the highest multiplicity equal to the order of the point group). A consists of all points whose site-symmetry groups are conjugate subgroups of the . The asymmetric unit is the minimal portion of the unit cell that, when subjected to all operations, reproduces the entire structure without redundancy; its content determines the Z, with Z' denoting the number of such units if multiple independent molecules are present. Detailed tabulations of s, including generators, , and diagrams of symmetry elements, are compiled in the International Tables for Crystallography, serving as the standard reference for lookup and application. In determination, symmetry influences patterns through reflection multiplicity—where equivalent reflections arise from symmetric positions—and systematic absences, which are forbidden reflections caused by screw axes or glide planes (e.g., odd h00 reflections absent in P21 due to the screw axis). These features facilitate identification and structure prediction by constraining possible atomic arrangements and reducing the search space during refinement.

Historical Development

Origins and Early Discoveries

The earliest recognition of crystals dates back to ancient times, with Roman naturalist documenting various crystalline minerals, such as rock crystal (), in his (circa 77 CE), describing their formation in rocky terrains and distinguishing them from amorphous stones based on their geometric shapes and transparency. In the medieval period, Islamic scholar (known as Alhazen, 965–1040 CE) advanced the understanding of through his , where he analyzed and in transparent media, laying groundwork for later studies of in crystalline substances. The 17th century marked a pivotal shift toward systematic observation of crystal geometry, beginning with Danish physician Nicolaus Steno's 1669 publication De solido intra solidum naturaliter contento dissertationis prodromus, in which he formulated the law of constancy of interfacial angles, observing that the angles between corresponding faces of crystals remain fixed regardless of size or origin. This principle was expanded in the by French mineralogist Louis Romé de l'Isle, who in his 1772 Essai de cristallographie identified over 110 primitive crystal forms and confirmed Steno's law across diverse minerals, emphasizing that crystal shapes derive from underlying geometric primitives. Building on these ideas, René Just Haüy introduced theory in his 1784 Essai d'une théorie sur la structure des cristaux, proposing that macroscopic crystal habits arise from stacked polyhedral "integrant molecules" arranged in a regular , linking external to internal atomic structure for the first time. Instrumental innovations in the early 19th century enabled precise measurements, as exemplified by British chemist William Hyde Wollaston's 1809 invention of the reflective , a device using light reflection off crystal faces to accurately determine interfacial angles with errors reduced to less than 1 minute of arc. This tool facilitated advancements in crystal classification, including British mineralogist William Hallowes Miller's 1839 introduction of in A Treatise on Crystallography, a using integers to denote plane orientations relative to crystal axes, standardizing descriptions of facets and cleavages. Concurrently, French physicist Auguste Bravais formalized the concept of space lattices in his 1848 memoir Mémoire sur les systèmes formés par des points distribués régulièrement sur un plan ou dans l'espace, identifying 14 unique Bravais lattices as the fundamental arrangements possible in three dimensions, providing a mathematical framework for crystal . Optical studies revealed crystals' anisotropic properties, with French physicist discovering in 1815 that quartz crystals rotate the plane of polarized light, a phenomenon he termed rotary polarization, demonstrating directional dependence in light transmission through solids. extended this in the 1820s by explaining birefringence in calcite and other uniaxial crystals via wave theory, showing that light splits into ordinary and extraordinary rays with perpendicular polarizations due to the medium's anisotropy, thus challenging the prevailing corpuscular model. These findings fueled early 19th-century debates on light's nature, pitting Isaac Newton's corpuscular theory—favoring particle-like propagation to explain straight-line travel and refraction in crystals—against ' wave hypothesis, which better accounted for interference and polarization effects observed in anisotropic media. Such morphological and optical foundations set the stage for later atomic-scale investigations through diffraction methods.

20th-Century Advancements and Key Milestones

The discovery of X-ray diffraction by crystals in 1912 marked the beginning of modern crystallography, when , along with Walter Friedrich and Paul Knipping, demonstrated that X-rays could be diffracted by a , confirming the periodic atomic arrangement in . This experiment provided the first direct evidence of atomic lattices, shifting crystallography from macroscopic morphology to atomic-scale analysis. In 1913, and William Lawrence Bragg formulated , which quantitatively relates the wavelength of X-rays (\lambda), the interplanar spacing (d), the diffraction angle (\theta), and an integer order (n) via the equation: n \lambda = 2 d \sin \theta This law enabled the measurement of atomic distances and became foundational for structure determination. During the 1920s and 1930s, theoretical and experimental advancements deepened the understanding of diffraction. Paul Ewald developed the dynamical theory of X-ray diffraction, accounting for multiple scattering events within crystals and extending beyond the simpler kinematic approximation. Experimentally, the rotating crystal method, developed in the 1910s and advanced by J.D. Bernal in 1926 for data interpretation, allowed for the collection of complete diffraction data sets by rotating the crystal relative to the X-ray beam, facilitating three-dimensional structure analysis. A landmark application came in 1934 when John Desmond Bernal and Dorothy Crowfoot obtained the first X-ray diffraction pattern of a protein crystal, pepsin, preserved in its mother liquor to maintain native structure, opening the door to macromolecular crystallography. Following World War II, new diffraction probes expanded the field. Neutron diffraction emerged around 1946, with initial experiments at nuclear reactors like the Argonne pile, enabling the study of light atoms and magnetic structures that X-rays could not resolve effectively. Electron diffraction, discovered in 1927 by Clinton Davisson and Lester Germer, was advanced in the 1930s and 1940s using transmission electron microscopes, providing high-resolution data for thin crystals and surfaces, complementing X-ray methods for small samples. In 1953, James Watson and Francis Crick elucidated the double-helix structure of DNA, building on fiber diffraction patterns from Rosalind Franklin and Maurice Wilkins, which revealed the molecule's helical parameters and base-pairing geometry. The 1970s and 1990s saw technological revolutions in and phase determination. Synchrotron radiation sources, first utilized for crystallography at facilities like in around 1972, delivered intense, tunable X-ray beams, dramatically reducing exposure times and enabling studies of weakly diffracting samples. Area detectors, such as image plates and charge-coupled devices introduced in the 1980s and 1990s, replaced film by capturing full diffraction patterns in seconds, accelerating high-throughput . For phase solving, the direct methods developed by Herbert Hauptman and Jerome Karle in the 1950s—using probabilistic relations between structure factors—were refined and awarded the 1985 , making structure determination routine for small molecules and influencing macromolecular phasing. Extending into the 21st century, serial femtosecond crystallography (SFX) emerged in the 2010s with X-ray free-electron lasers (XFELs), such as the Linac Coherent Light Source, allowing diffraction data from microcrystals before , enabling time-resolved studies of dynamic processes like enzyme reactions. More recently, in the 2020s, tools like have integrated with crystallography by providing initial models for phase improvement and validation, as seen in hybrid approaches that refine AI predictions against experimental diffraction data to enhance accuracy for challenging structures. In 2024, the was awarded to David Baker, , and Jumper for computational and structure prediction using , which has transformed experimental methods like crystallography by providing accurate initial models for refinement.

Experimental Techniques

Diffraction Methods

Diffraction methods in crystallography exploit the of waves—such as X-rays, neutrons, or electrons—by the periodic atomic of a to reveal its structure. When a impinges on the , the atoms act as scattering centers, producing waves that interfere constructively under specific conditions dictated by the , which ensure phase differences match integer multiples of 2π. This constructive interference occurs for scattering vectors connecting points, visualized through the Ewald sphere construction: the incident wavevector \mathbf{k}_i originates from a point on a of radius $1/\lambda (where \lambda is the wavelength) centered at the crystal origin, and the scattered wavevector \mathbf{k}_s (also of length $1/\lambda) reaches a point when the difference \mathbf{k}_s - \mathbf{k}_i = \mathbf{G} (a vector). This geometric tool in reciprocal space predicts observable diffraction spots and underscores the periodicity of the . X-ray crystallography remains the cornerstone of these methods, employing with wavelengths comparable to atomic spacings (~0.5–2 Å) to map . Laboratory sources, such as rotating anode generators, provide sufficient flux for routine single-crystal studies, while sources deliver orders-of-magnitude higher brilliance, tunable energies, and smaller beam sizes, enabling data collection from microcrystals or time-resolved experiments with resolutions down to ~0.5 Å or better. X-ray free-electron lasers (XFELs) have further revolutionized the field through serial femtosecond crystallography (SFX), utilizing ultrashort pulses to capture patterns from streams of microcrystals before occurs, facilitating room-temperature structures of biomolecules and ultrafast dynamics studies. As of 2025, advances in sample delivery methods, including fixed-target systems (e.g., - or polymer-based chips) and high-viscosity extruders, have minimized sample consumption to nanoliters while supporting high-throughput data collection at XFELs and . Detectors have evolved from photographic films to charge-coupled devices (CCDs) and pixel array detectors, capturing two-dimensional patterns with and speed. Key techniques include the Laue method, which uses polychromatic "white" to simultaneously record multiple reflections from a stationary , ideal for rapid orientation or snapshot crystallography; the method, where a is rotated around one axis in a monochromatic beam to sweep through reciprocal space and index reflections systematically; and the powder method (Debye-Scherrer), applied to polycrystalline samples where random orientations produce conical beams intersecting detectors as rings, allowing identification without . These approaches measure integrated intensities corrected for geometric factors, yielding for refinement. Neutron diffraction complements X-rays by leveraging neutrons from nuclear reactors or spallation sources, which scatter via nuclear interactions rather than electrons, providing scattering lengths that do not monotonically increase with atomic number. This enables precise localization of light atoms (e.g., carbon, nitrogen, oxygen) and isotopic discrimination, such as between (low scattering) and (high scattering), crucial for hydrogen-bonding studies in organics or hydrides. Unlike X-rays, neutrons cause no and penetrate deeply into bulk samples. A unique advantage is the determination of magnetic structures: the neutron's interacts with atomic moments, producing additional diffraction peaks or modulations that reveal arrangements, antiferromagnetic ordering, or structures in materials like alloys or oxides. Data collection mirrors X-ray methods but requires larger samples (~1 cm³) due to lower fluxes, with resolutions typically 1–2 . Electron diffraction suits nanoscale or surface investigations, where the short de Broglie wavelength (~0.02–0.05 Å at 100–300 keV) allows atomic resolution but demands thin samples to mitigate multiple scattering effects that distort patterns. In transmission electron microscopy (TEM), selected-area electron diffraction probes nanocrystals or defects in thin foils (~100 nm), yielding spot patterns from volumes as small as 200 nm, ideal for beam-sensitive materials like organics or inorganics where X-ray methods fail due to sample size. Low-energy electron diffraction (LEED), using 20–200 eV electrons, characterizes surface reconstructions on single crystals, with beams penetrating only the top 10–20 atomic layers; multiple scattering is prominent and modeled dynamically to refine atomic positions. These techniques often combine with imaging for hybrid real- and reciprocal-space analysis. In all diffraction methods, focuses on measuring diffraction intensities I(\mathbf{h}), where \mathbf{h} = (hkl) indexes the , proportional to the squared modulus of the after corrections for Lorentz, , and multiplicity effects. The F(\mathbf{h}) encodes the atomic arrangement: F(hkl) = \sum_j f_j \exp \left[ 2\pi i (h x_j + k y_j + l z_j) \right] Here, f_j is the atomic scattering factor (dependent on atom type and scattering angle), and (x_j, y_j, z_j) are the of the j-th atom in the unit cell. The phase information in F(hkl) is lost in intensities, necessitating phasing techniques for reconstruction, but |F(hkl)|^2 directly relates to observable electron or nuclear density Fourier transforms.

Imaging and Scattering Techniques

Scanning tunneling microscopy (STM) enables direct visualization of surface atomic lattices in crystalline materials by measuring tunneling currents between a sharp probe tip and the sample surface, achieving resolutions down to the atomic scale under conditions. This technique has been pivotal in resolving reconstructions like the 7×7 structure on Si(111) surfaces, providing real-space insights into surface periodicity and electronic states that complement reciprocal-space methods such as . Atomic force microscopy (AFM), an extension applicable to both conductive and insulating crystals, images surface topography by detecting forces between the tip and sample, often revealing arrangements and defects at sub-nanometer . For instance, AFM has been used to study ionic crystal surfaces, where electrostatic interactions allow atomic-scale contrast without electrical conductivity requirements. Transmission electron microscopy (TEM) provides high-resolution imaging of crystal interiors, particularly for visualizing defects such as dislocations and twins that disrupt perfection. In aberration-corrected TEM, atomic-scale reveals the core structures of dislocations in materials like , showing how they interact with twin boundaries during deformation. This real-space approach is essential for understanding defect dynamics in thin foils, where contrast arises from local strain and differences. Small-angle X-ray scattering (SAXS) and (WAXS) probe nanoscale structures in crystals, such as nanostructures and polymer crystallites, by analyzing scattering at low angles to determine particle sizes, shapes, and interfaces. SAXS is particularly useful for disordered or semi-crystalline systems, where the scattering intensity follows Porod's law, I(q) \propto q^{-4} at high q, indicating smooth interfaces and allowing quantification of surface area per volume. This behavior, derived from the of gradients, helps characterize and aggregation in without requiring long-range order. Inelastic scattering techniques reveal vibrational and local structural dynamics in crystals. measures modes by detecting shifts in inelastically scattered light, providing insights into lattice vibrations and in materials like dichalcogenides. Brillouin scattering, involving interactions with acoustic s, probes hypersonic waves to determine elastic moduli and sound velocities in crystals, with frequency shifts given by \Delta \nu = \frac{2n v \sin(\theta/2)}{\lambda}, where n is the refractive index, v the sound speed, \theta the scattering angle, and \lambda the wavelength. (EXAFS) analyzes oscillations in X-ray absorption above the edge to deduce local coordination environments around absorbing atoms, typically up to 6 Å, in crystalline and amorphous solids. For example, EXAFS quantifies bond lengths and coordination numbers in metal oxides, revealing distortions not evident in average structures. Time-resolved pump-probe methods using ultrafast lasers capture dynamic processes like phase transitions in crystals by exciting the sample with a pump pulse and probing structural changes with delayed pulses. These techniques, often combining optical or X-ray probes, observe lattice responses on picosecond timescales, such as coherent phonon generation during ferroelectric switching or photoinduced melting in semiconductors. In ferroelectrics, they reveal ultrafast polarization dynamics and anharmonic phonon softening, linking electronic excitation to structural reconfiguration.

Applications

Materials Science and Engineering

Crystallography is fundamental to and , enabling the design and optimization of materials by revealing how atomic-scale arrangements dictate macroscopic properties such as mechanical strength, electrical conductivity, and thermal stability. By analyzing crystal structures, engineers can predict and control behaviors like in alloys or mobility in semiconductors, often using techniques like to map parameters and symmetries. This understanding underpins advancements in high-performance materials, from components to devices. In polycrystalline materials, crystallographic —the preferred of grains—profoundly influences anisotropic , such as the formability of rolled metal sheets or the barrier of polymer films. Pole figures provide a of the distribution of specific crystallographic planes relative to the sample , while orientation distribution functions (ODF) offer a quantitative, three-dimensional representation of the full spectrum, essential for modeling deformation textures in metals like or polymers like . For instance, in face-centered cubic metals, recrystallization textures can be engineered to enhance deep-drawing performance by aligning {111} planes parallel to the sheet surface. Texture analysis via further refines these controls, linking local misorientations to global mechanical response. Crystallographic principles are integral to interpreting phase diagrams and transformations, guiding the manipulation of material phases for desired properties. In steels, the martensitic transformation—a diffusionless, shear-dominated process—converts face-centered cubic to body-centered tetragonal upon rapid cooling, as depicted in Fe-C phase diagrams where the martensite start temperature (M_s) decreases with carbon content. This transformation introduces high hardness but brittleness, critical for tool steels. In pharmaceuticals, controlling polymorphs—distinct crystal forms of the same compound—is vital for solubility and bioavailability; for example, the orthorhombic versus monoclinic polymorphs of acetaminophen exhibit different dissolution rates, with crystallography via powder X-ray diffraction enabling selective nucleation through solvent-mediated processes. Defects in crystal lattices, such as dislocations and grain boundaries, govern plasticity and failure mechanisms, with crystallography providing the tools to characterize and mitigate them. Dislocations are line defects quantified by the , which measures the closure failure in a around the defect , determining slip systems in materials like aluminum where edge dislocations facilitate glide on {111} planes. Grain boundaries, as interfaces between misoriented crystals, impede dislocation motion, strengthening the material per the Hall-Petch relation: \sigma_y = \sigma_0 + k d^{-1/2} where \sigma_y is the yield strength, \sigma_0 is a friction stress, k is the strengthening coefficient, and d is the average grain size; this inverse square-root dependence highlights how nanoscale grain refinement boosts strength in nanocrystalline nickel. Transmission electron microscopy combined with selected-area diffraction visualizes these defects, informing alloy designs for enhanced toughness. In semiconductors and nanomaterials, lattice symmetry directly shapes electronic band structures, enabling tailored optoelectronic performance. Gallium arsenide (GaAs), with its zincblende structure (space group F\bar{4}3m), exhibits a direct bandgap of 1.424 eV at 300 K, where the conduction band minimum at the \Gamma point aligns with the valence band maximum, facilitating efficient radiative recombination for lasers and solar cells. This tetrahedral coordination minimizes dangling bonds, contributing to high electron mobility of over 8500 cm²/V·s. Quantum dots, as zero-dimensional nanocrystals like CdSe, leverage quantum confinement within their wurtzite or zincblende lattices to tune bandgaps from bulk values (e.g., 1.74 eV for CdSe) down to higher energies with decreasing size below 10 nm, enabling applications in displays and biomedical imaging. High-resolution transmission electron microscopy elucidates these size-dependent structures. Additive manufacturing benefits from in-situ crystallography to monitor and control evolving microstructures during layer-by-layer deposition, mitigating defects like or residual stresses. tracks real-time texture development and changes in alloys such as , where epitaxial growth across layers can be tuned by parameters to achieve fine \alpha- laths for improved resistance. For nickel-based superalloys, in-situ observations reveal columnar rotation under thermal gradients, allowing process adjustments to promote equiaxed microstructures that enhance performance at high temperatures. This approach integrates crystallographic data with finite element modeling for predictive design.

Structural Biology and Chemistry

Crystallography has profoundly impacted by enabling the determination of atomic-level structures of biological macromolecules, which is essential for understanding molecular interactions and functions. In protein crystallography, obtaining suitable crystals remains a major challenge due to the flexibility and heterogeneity of proteins, often requiring optimization of conditions to achieve ordered lattices. The hanging drop vapor diffusion method is one of the most widely used techniques for protein crystallization, where a small droplet containing the protein solution is suspended over a reservoir of precipitant, allowing vapor-mediated equilibration to promote . These efforts have populated the (PDB), established in 1971 as the first open-access digital archive for macromolecular structures, which, as of 2025, holds over 230,000 entries and serves as a global resource for structural data. High-resolution structures, typically better than 2.0 Å, reveal detailed side-chain conformations and hydrogen bonding networks, providing insights into and dynamics. In chemistry and , crystallographic structures of enzymes have elucidated catalytic mechanisms by visualizing geometries and binding. For instance, the crystal structures of determined in the late revealed its homodimeric architecture and conserved residues critical for , facilitating the rational design of inhibitors that became cornerstone antiretroviral therapies. These structures demonstrated how inhibitors mimic the , occupying the cleft and preventing viral polyprotein cleavage. Similarly, crystallographic studies of nucleic acids confirmed foundational models of biomolecular architecture; the 1953 double structure of DNA, derived from fiber patterns, established base pairing and helical parameters that underpin genetic replication and transcription. In the 2000s, high-resolution structures, achieved through of bacterial and eukaryotic ribosomes, mapped the center and tRNA binding sites, revealing the ribosomal RNA's catalytic role in formation and earning the 2009 for Venkatraman Ramakrishnan, , and . For small molecules in chemistry, crystallography determines absolute configurations, crucial for in and synthesis. The anomalous dispersion method exploits the phase shift in scattering from atoms like or to distinguish enantiomers, as demonstrated in early applications to and peptides in the 1950s, where Bijvoet pairs of reflections provided the Flack parameter to quantify . This technique has extended to supramolecular assemblies, such as host-guest complexes and metal-organic frameworks, where crystal structures reveal non-covalent interactions like hydrogen bonding and π-stacking that dictate assembly and reactivity. In recent advancements, integration with cryo-electron microscopy (cryo-EM) addresses limitations of crystallography for large complexes exceeding 1 MDa, using hybrid approaches where crystallographic models of components are fitted into cryo-EM density maps to resolve flexible regions and conformational states in assemblies like capsids or supercomplexes.

Notation and Representation

Indexing and Coordinate Systems

In crystallography, provide a standardized notation for specifying crystal planes within a . These indices, denoted as (hkl), are determined by taking the reciprocals of the intercepts of the plane with the crystallographic axes a, b, and c, then reducing to the smallest integers. For a plane intersecting the axes at fractions p, q, r (or if parallel), the indices are h = 1/p, k = 1/q, l = 1/r, cleared of fractions and with a common factor. Negative intercepts are indicated by a bar over the index, such as (\bar{h}kl), ensuring the notation distinguishes parallel planes on opposite sides of the origin. Planes with indices that are permutations or sign changes, such as {hkl}, represent a family of equivalent planes related by . Zone axes, denoted [uvw], describe lines of intersection between such plane families. Crystallographic directions are labeled using direction indices [uvw], where u, v, w are the smallest integers proportional to the direction cosines along the , c axes from the origin to a point. These indices define vectors in direct space, and families of symmetrically equivalent directions are enclosed in angle brackets, , accounting for symmetries that make properties isotropic within the set. The [uvw] specifically refers to a parallel to the of two or more families, facilitating analysis of preferred orientations in polycrystalline materials. Atomic positions within the unit cell are expressed using fractional coordinates (x, y, z), where each value ranges from 0 to 1, representing the relative distances along the a, b, c edges from a chosen origin. These coordinates allow precise description of basis atoms in the asymmetric unit, which are then replicated by symmetry operations to fill the cell. Transformations between fractional and Cartesian coordinates involve the cell metric tensor, with the position vector \mathbf{r} = x\mathbf{a} + y\mathbf{b} + z\mathbf{c}, where \mathbf{a}, \mathbf{b}, \mathbf{c} are the lattice vectors. In reciprocal space, the Miller indices (hkl) correspond to points in the reciprocal lattice, which is the Fourier transform of the direct lattice and consists of vectors \mathbf{G}_{hkl} = h\mathbf{a}^* + k\mathbf{b}^* + l\mathbf{c}^*, perpendicular to the (hkl) planes. The interplanar spacing d_{hkl} is given by d_{hkl} = 1 / |\mathbf{G}_{hkl}|, or more generally, d_{hkl} = 1 / \sqrt{\mathbf{G}_{hkl} \cdot \mathbf{G}_{hkl}}, where the dot product incorporates the metric tensor \mathbf{G} of the direct lattice to account for non-orthogonality: d_{hkl} = \frac{1}{\sqrt{h^2 G_{11} + k^2 G_{22} + l^2 G_{33} + 2hk G_{12} + 2hl G_{13} + 2kl G_{23}}} This formula enables calculation of diffraction angles via Bragg's law. Stereographic projection is a geometric method to represent three-dimensional crystal orientations on a two-dimensional plane, projecting poles (normals to planes) from a unit sphere onto the equatorial plane through the south pole. In crystallography, poles for (hkl) planes are plotted by extending the normal from the sphere center to the plane, with the projection preserving angular relationships between directions. This technique visualizes symmetry, zone axes as great circles, and interfacial angles, aiding in the identification of crystal forms and orientations.

Crystallographic Databases and Software

Crystallographic databases serve as essential repositories for storing and disseminating determined crystal structures, enabling researchers to access, analyze, and build upon existing data for advancing , , and . These databases typically include detailed atomic coordinates, parameters, information, and associated metadata, often searchable by , , or other structural features. Key examples include specialized collections for inorganic, , and biomolecular structures, each with validation protocols to ensure data reliability. The Inorganic Crystal Structure Database (ICSD) is a comprehensive repository focused on fully identified inorganic crystal structures, containing over 210,000 entries as of recent updates. It supports searches by , , and other criteria such as dimensions, facilitating the retrieval of structures for compounds, ceramics, and minerals. Maintained collaboratively by institutions like NIST and FIZ , ICSD emphasizes high-quality, peer-reviewed data from literature since , with tools for deposition and standardized output in formats like . For biomolecular structures, the (PDB) archives over 245,000 experimentally determined three-dimensional structures of proteins, nucleic acids, and complexes as of November 2025, primarily from , NMR, and cryo-EM. These entries include validation metrics such as R-free values, which assess the agreement between the model and experimental data (typically below 0.25 for high-quality structures), and Ramachandran plots, which evaluate backbone dihedral angles to identify outliers (ideally fewer than 5% outliers). Managed by the wwPDB consortium, the PDB provides to these data, supporting research through integrated visualization and analysis tools. The Structural Database (), dedicated to organic and metal-organic compounds, holds over 1.3 million entries, enabling detailed studies of molecular packing and interactions. It is particularly valuable for analyzing hydrogen bonding patterns, with statistical tools revealing propensities for donor-acceptor geometries in supramolecular assemblies, as demonstrated in analyses of millions of structures showing common motifs like O-H···O and N-H···O bonds with angles around 160-180°. Curated by the Crystallographic Data Centre (CCDC), the includes software for querying interaction energies and polymorphism, aiding and materials discovery. Complementing these databases are specialized software tools for structure solution, refinement, visualization, and data exchange. SHELX, a widely used program suite, excels in least-squares refinement of small-molecule structures against diffraction data, incorporating restraints for handling disorder and twinning. Olex2 provides an intuitive interface for structure solution via direct methods and subsequent refinement, integrating SHELX while offering automated model building and publication-ready outputs. For visualization, enables 3D rendering of crystal structures, supporting isosurface plots of and bonding analysis across inorganic and molecular systems. The standard for data interchange is the (CIF) format, a flexible, extensible syntax developed by the International Union of Crystallography (IUCr) for archiving atomic coordinates, experimental details, and metadata in a human- and machine-readable way. In the 2020s, tools have emerged to augment traditional methods, such as generative models for prediction and generation from chemical compositions. For instance, workflows combining neural networks for sampling and relaxation have accelerated discovery, achieving high accuracy in predicting stable polymorphs without exhaustive computational searches. These approaches, often trained on database entries like those from or ICSD, represent a shift toward data-driven refinement and hypothetical structure design in crystallography.

References

  1. [1]
    Online Dictionary of Crystallography
    Jul 1, 2024 · The branch of science devoted to the study of molecular and crystalline structure and properties, with far-reaching applications in mineralogy, ...
  2. [2]
    Crystal - Online Dictionary of Crystallography
    Mar 23, 2021 · Direct-space definition. A solid is a crystal if its atoms, ions and/or molecules form, on average, a long-range ordered arrangement. In most ...
  3. [3]
    The success story of crystallography - IUCr Journals
    This discovery propelled the mathematical branch of mineralogy to global importance and enabled crystal structure determination.Missing: definition | Show results with:definition
  4. [4]
    Aspects of the history of the International Union of Crystallography
    The scientific foundations of modern crystallography were laid in 1912 by Laue's discovery of X-ray diffraction followed immediately by the invention of crystal ...
  5. [5]
    2 An Introduction to the Scope, Potential and Applications of X-ray ...
    This is the basis of the so-called powder or Debye-. Scherrer method which is probably the most common technique used in. X-ray crystallography. The powder ...
  6. [6]
    Crystallography Basics - Chemical Instrumentation Facility
    A crystal consists of matter that is formed of an ordered three dimensional arrangement of atoms, molecules or ions.
  7. [7]
    [PDF] What is crystallography?
    What is crystallography? • Deals with the symmetry of crystals and crystal structures. • Provides a descriptive method of describing the symmetry of crystals.
  8. [8]
    Impact and influence of crystallography across the sciences - NIH
    Oct 27, 2016 · Crystallography, whether using electrons, neutrons or X-rays, has transformed the way we look at a problem and the level at which we wish to ...
  9. [9]
    The future of crystallography in drug discovery - PMC - NIH
    X-ray crystallography plays an important role in structure-based drug design (SBDD), and accurate analysis of crystal structures of target macromolecules ...
  10. [10]
    Crystallography of Functional Materials (Editorial Comments) | NIST
    Nov 15, 2017 · This issue provides further evidence of the importance of crystallography in understanding and improving various properties of functional ...
  11. [11]
    [PDF] Basic Concepts of Crystallography
    High temperature (the torch flame) Page 4 Crystallography is the experimental science of the arrangement of atoms in solids. The word "crystallography" derives ...
  12. [12]
  13. [13]
    [PDF] Application of metastable curve and crystal anisotropic for ...
    Jun 27, 2022 · Crystal anisotropy is known as ability of single crystal to show different physical and chemical properties for different axes. Crystal shape ...
  14. [14]
    [PDF] Properties of Minerals Crystal Habits Cleavage and Fracture in ...
    Cleavages can often be identified because there are lots of parallel cracks or layers in a crystal. Cleavage surface often are discontinuous, and may give ...
  15. [15]
    Physical Properties of Minerals - Tulane University
    Sep 16, 2013 · Crystal Habit. In nature perfect crystals are rare. The faces that develop on a crystal depend on the space available for the crystals to grow.
  16. [16]
    Crystalline and Amorphous Solids
    Amorphous solids have two characteristic properties. When cleaved or broken, they produce fragments with irregular, often curved surfaces; and they have poorly ...
  17. [17]
    12.7: Types of Crystalline Solids- Molecular, Ionic, and Atomic
    ### Four Types of Crystalline Solids with Examples and Descriptions
  18. [18]
    X-RAY Crystallography - St. Olaf College
    The unit cell is made up of the smallest possible volume that when repeated, is representative of the entire crystal. The dimensions of a unit cell can be ...
  19. [19]
    Unit Cells
    The unit cell is the simplest repeating unit in the crystal. Opposite faces of a unit cell are parallel. The edge of the unit cell connects equivalent points.
  20. [20]
    361: Crystallography and Diffraction
    Points in 3-D space can be arranged such that each has identical surroundings in 14 ways. These are divided into 7 Crystal Systems. Crystal System. Lattice ...<|control11|><|separator|>
  21. [21]
    [PDF] Handout 4 Lattices in 1D, 2D, and 3D Bravais Lattice
    Bravais Lattices in 3D. There are 14 different Bravais lattices in 3D that are classified into 7 different crystal systems (only the unit cells are shown ...
  22. [22]
    [PDF] Summarize : crystal structure = Bravais lattice + basis
    Examples of Bravais lattices: 1. Simple cubic (sc). 1,2,3 a mutually perpendicular; 1. 2. 3 a a a a. = = = a = lattice constant or lattice parameter. Primitive ...
  23. [23]
    [PDF] Chapter 3: The Structure of Crystalline Solids
    • Calculated theoretical density if atomic mass is A and Avergado number is ... Crystal Structure Determination via. X-Ray Diffraction. • Diffraction ...
  24. [24]
    Point group - Online Dictionary of Crystallography
    Nov 30, 2018 · A point group is a group of symmetry operations all of which leave at least one point unmoved. A crystallographic point group is a point group that maps a ...
  25. [25]
    The symmetry of crystals. The crystallographic restriction theorem
    Crystals can only show 2-fold, 3-fold, 4-fold or 6-fold rotation axes. The reason is that the external shape of a crystal is based on a geometric arrangement ...
  26. [26]
    Crystallography Restriction -- from Wolfram MathWorld
    If a discrete group of displacements in the plane has more than one center of rotation, then the only rotations that can occur are by 2, 3, 4, and 6.
  27. [27]
    Crystallographic Point Groups -- from Wolfram MathWorld
    The crystallographic point groups are the point groups in which translational periodicity is required (the so-called crystallography restriction). There are 32 ...
  28. [28]
    Crystallographic Point Groups
    Crystal System, 32 Crystallographic Point Groups. Triclinic, 1, -1. Monoclinic, 2, m, 2/m. Orthorhombic, 222, mm2, mmm. Tetragonal, 4, -4, 4/m, 422, 4mm, -42m ...
  29. [29]
    [PDF] Crystal Systems and Crystallographic Point Groups
    Spectroscopists use Schoenflies notation to describe symmetry (e.g. C2v, D4h). ○. Crystallographers use Hermann-Mauguin notation (International notation).
  30. [30]
    Crystallographic orientation of uniaxial calcite and dolomite ...
    Dec 14, 2018 · Calcite belongs to the point group ¯3m (or D3d Schönflies notation) and space group. R¯3c, and dolomite, with fewer symmetry operations ...<|control11|><|separator|>
  31. [31]
    Piezoelectricity - Online Dictionary of Crystallography
    Nov 17, 2017 · They are the non-centrosymmetric classes, with the exception of 432. The 20 piezoelectric crystal classes are therefore: 1, 2, m, 222, 2mm, 3, ...
  32. [32]
    Piezoelectricity - TU Graz
    Of all 32 point groups, 21 are noncentrosymmetric, and of those 20 point groups show piezoelectricity (only point group 432 is an exception).
  33. [33]
    space-group symmetry - International Union of Crystallography
    Space group G is the set of all symmetry operations of a crystal pattern, including all translations that are symmetry operations.
  34. [34]
    Wyckoff position - Online Dictionary of Crystallography
    Nov 20, 2017 · Definition. A Wyckoff position of a space group G consists of all points X for which the site-symmetry groups are conjugate subgroups of G.
  35. [35]
    (International Tables) Volume A home page
    It includes chapters on groups, crystallographic and space-group symmetry, descriptions of space groups, coordinate-system transformations and methods of space- ...<|control11|><|separator|>
  36. [36]
    Pliny, Natural History, 37 (a) - ATTALUS
    Glass-ware has now come to resemble rock-crystal in a remarkable manner, but the effect has been to flout the laws of Nature and actually to increase the value ...
  37. [37]
    [PDF] THE OPTICS OF IBN AL-HAYTHAM - Monoskop
    Preface to the [whole] book. Inquiry into the properties of sight. Inquiry into the properties of lights and into the manner of radiation o f lights.
  38. [38]
    NICOLAUS STENO (1638 - 1686). De Solido intra Solidum. Florence
    The law of constancy of angles was confirmed and shown to be true for crystals of many other substances by Romé de l'Isle in 1783.
  39. [39]
  40. [40]
    (IUCr) René-Just Haüy and the birth of crystallography
    Dec 4, 2022 · René-Just Haüy, a French abbot living in Paris, at almost 40 years old, began a scientific career focused on the study of crystals.
  41. [41]
    XVI. Description of a reflective goniometer - Journals
    From the advances that have been made of late years in crystallography ... Description of a reflective goniometer. William Hyde Wollaston. Google Scholar.
  42. [42]
    Introduction - DoITPoMS
    Miller Indices are a method of describing the orientation of a plane or set of planes within a lattice in relation to the unit cell. They were developed by ...Missing: 1825 | Show results with:1825
  43. [43]
    (IUCr) Auguste Bravais, from Lapland to Mont Blanc
    Feb 1, 2019 · In essence, Bravais' theory (1848-1850) uses a polyhedral ... As crystallographers you all know Bravais and his famous 14 crystal lattice types.
  44. [44]
    Augustin Fresnel and the wave theory of light - Photoniques
    In 1819, he studied birefringent crystals, and turned his attention to double refraction and the associated properties of polarization. In a memoir and ...
  45. [45]
    Light through the ages: Ancient Greece to Maxwell - MacTutor
    However, the work strongly argued for a corpuscular theory of light, with the most telling argument being that light travels in straight lines yet waves are ...
  46. [46]
    The birth of X-ray crystallography - Nature
    Nov 7, 2012 · Laue's lecture reported the first observation by his colleagues Walter Friedrich and Paul Knipping of the diffraction of X-rays by a crystal — ...
  47. [47]
    [PDF] The discovery of the diffraction of X-rays by crystals
    Walter Friedrich and Paul Knipping, with Max Laue's proposal, observed X-ray diffraction by a crystal on April 21, 1912, in Munich.Missing: milestones | Show results with:milestones
  48. [48]
    Full article: The Nobel Science: One Hundred Years of Crystallography
    Nov 27, 2015 · WHB was more interested in X-rays than crystal structures and he discovered that metals give X-rays with characteristic energies (Ewald 1962), a ...
  49. [49]
    Laue diffraction and time-resolved crystallography: a personal history
    Apr 29, 2019 · A personal, historical view is presented of Laue X-ray diffraction and its application to time-resolved studies of dynamic processes.
  50. [50]
    [PDF] commented chronology of crystallography and structural chemistry
    Feb 15, 2018 · Observations on crystal morphology led him to the conclusion that certain angles between faces of crystals of a natural specimen are constant.
  51. [51]
    X-Ray Photographs of Crystalline Pepsin - NASA ADS
    X-ray photographs taken of the crystals in the usual way showed nothing but a vague blackening. This indicates complete alteration of the crystal and explains ...Missing: crystallography advancements 1920s- 1930s
  52. [52]
    (IUCr) 50 Years of X-ray Diffraction
    This book covers the beginnings of X-ray diffraction, its principles, crystal structure analysis, and its growing power and applications.<|separator|>
  53. [53]
    Electron Diffraction of 3D Molecular Crystals | Chemical Reviews
    Aug 15, 2022 · This Review covers concepts and findings relevant to the practicing crystallographer, with an emphasis on experiments aimed at using electron diffraction.
  54. [54]
    [PDF] Ninety Years of Powder Diffraction: from Birth to Maturity
    During the 1970s dedicated synchrotron laboratories began operation and the PD results from such sources have made a significant contribution to the litera ...
  55. [55]
    Advancements in macromolecular crystallography: from past to ...
    Developments in robotic crystallization techniques, intense synchrotron X-ray sources, faster detectors as well as automated data processing, structure solution ...
  56. [56]
    The Ewald sphere construction for radiation, scattering, and diffraction
    Apr 1, 2017 · In this paper, we show that the Ewald sphere construction, a powerful tool for predicting crystallographic diffraction patterns, can also be ...
  57. [57]
    A new era of synchrotron-enabled macromolecular crystallography
    May 7, 2021 · The future of macromolecular crystallography includes new X-ray sources, enhanced remote-accessible capabilities and time-resolved methods to capture ...
  58. [58]
    [PDF] Laue X-Ray Diffraction - Physik-Institut | UZH
    Sep 26, 2023 · In x-ray crystallography, the Laue method is a simple and fast way of recording a diffraction pattern of a single-crystalline sample.<|separator|>
  59. [59]
    Neutron diffraction: a primer | ORNL
    Jun 1, 2024 · With high neutron flux and sufficient brilliance, neutron diffraction also excels for diffuse scattering, for in situ and operando studies as ...
  60. [60]
    Neutron Diffraction of Magnetic Materials - GeoScienceWorld
    Jul 13, 2017 · Neutron diffraction is a powerful tool for studying magnetic materials. Neutrons have a magnetic moment, and are scattered by the magnetic ...
  61. [61]
    7.4: Low Energy Electron Diffraction - Chemistry LibreTexts
    Aug 28, 2022 · Low energy electron diffraction (LEED) is a very powerful technique that allows for the characterization of the surface of materials.Principles and Diffraction... · LEED Experimental Equipment · LEED Applications
  62. [62]
    3.27: Structure Factor - Chemistry LibreTexts
    Jun 30, 2023 · The structure factor F h ⁡ k ⁢ l is a mathematical function describing the amplitude and phase of a wave diffracted from crystal lattice planes ...
  63. [63]
    Interactions between dislocations and twins in deformed titanium ...
    In this work, we use aberration-corrected scanning transmission electron microscopy (STEM) to resolve the interactions between dislocations and twins at atomic ...
  64. [64]
    Review of the fundamental theories behind small angle X-ray ... - NIH
    In this paper, the fundamental concepts and equations necessary for performing small angle X-ray scattering (SAXS) experiments, molecular dynamics (MD) ...
  65. [65]
    100th Anniversary of Brillouin Scattering: Impact on Materials Science
    Brillouin scattering is a non-contact and non-destructive method to measure sound velocity and attenuation. It is possible to investigate the elastic ...
  66. [66]
    Probing Ultrafast Dynamics of Ferroelectrics by Time‐Resolved ...
    This review offers a clear guidance of ultrafast dynamics of ferroelectric orders, which may promote the rapid development of next‐generation devices.4. Lattice Dynamics Of... · 4.1. Optical Phonon... · 4.2. Acoustic Phonon...
  67. [67]
  68. [68]
    [PDF] A New Approach to Texture Measurements: Orientation ... - OSTI.GOV
    Traditionally, in ODF is determined from pole figures for a relatively small number of reflections. These pole figures are measured with x-rays or neutrons ...
  69. [69]
    [PDF] UF Kocks, CN Tomé and H.-R. Wenk
    This book is about the measurement and analysis of textures, the prediction of poly- crystal properties from measured textures and known single-crystal ...Missing: ODF | Show results with:ODF
  70. [70]
    Crystallography of martensite transformations in steels - ScienceDirect
    This chapter describes the unique features of martensitic transformations in steels. It covers the characteristics that serve to distinguish and identify ...Missing: diagrams | Show results with:diagrams
  71. [71]
    Controlling Polymorphism in Pharmaceutical Compounds Using ...
    Jan 4, 2018 · In this work, two model pharmaceutical compounds, glycine and acetaminophen, were crystallized using a flow coating technique termed ...
  72. [72]
    Hall-Petch Law Revisited in Terms of Collective Dislocation Dynamics
    Aug 15, 2006 · The Hall-Petch (HP) law, that accounts for the effect of grain size on the plastic yield stress of polycrystals, is revisited in terms of ...
  73. [73]
    Band structure parameters of wurtzite and zinc-blende GaAs under ...
    Jul 26, 2011 · Quasiparticle self-consistent GW calculations are used to study the band structure in wurtzite and zinc-blende GaAs. The band-gap change ...
  74. [74]
    Quantum dot nanomaterials: Empowering advances in ...
    Quantum dots (QDs) of colloidal semiconductors are nanoscale crystals floating in a solution. Optoelectronic devices, imaging, and fundamental science find ...
  75. [75]
    Microstructural control of additively manufactured Ti6Al4V via in-situ ...
    Jan 15, 2024 · In this work, we explore the use of layer-by-layer large area surface annealing with a secondary diode laser to control the microstructure of ...
  76. [76]
    In situ observation of crystal rotation in Ni-based superalloy during ...
    May 23, 2023 · The in-depth understanding of the crystal rotation and SG formation mechanisms could be helpful for optimising additive manufacturing approaches ...
  77. [77]
    Crystallization of Membrane Proteins by Vapor Diffusion - PMC - NIH
    Feb 16, 2016 · In the hanging-drop technique, the mixed drop is included on the glass slide itself, inverted and sealed with grease atop the chamber. For ...
  78. [78]
    PDB History
    The PDB was established in 1971 at Brookhaven National Laboratory under the leadership of Walter Hamilton and originally contained 7 structures.
  79. [79]
    Protein X-ray Crystallography: Principles & The Technique
    At medium resolution (3.5-2.5 Å), we start to distinguish side chains and can build the model. When the resolution is better than 2.8 Å, we can clearly ...
  80. [80]
    HIV Protease: Historical Perspective and Current Research - PMC
    May 6, 2021 · This review covers the historical background of studies on the structure and function of HIV protease, the subsequent development of antiviral inhibitors,
  81. [81]
  82. [82]
    The Nobel Prize in Chemistry 2009 - Illustrated Presentation
    Ribosome structures will save lives. All three Nobel Laureates have generated structures showing exactly where different antibiotics attack bacterial ribosomes.
  83. [83]
    Assignment of Absolute Configuration from Anomalous Dispersion ...
    Jan 1, 1970 · There have been several recent determinations of the absolute configuration of molecules using the anomalous scattering of X-rays by oxygen.
  84. [84]
    combining Cryo-TEM, X-ray crystallography, and NMR - PMC - NIH
    The combination of Cryo-TEM and other methods such as X-ray crystallography, nuclear magnetic resonance spectroscopy, and molecular dynamics modeling
  85. [85]
    NIST Inorganic Crystal Structure Database (ICSD) - National Institute ...
    The NIST ICSD is a comprehensive database of inorganic crystal structures with over 210,000 entries, searchable by various criteria, and requires a ...
  86. [86]
  87. [87]
    RCSB PDB: Homepage
    RCSB Protein Data Bank (RCSB PDB) enables breakthroughs in science and education by providing access and tools for exploration, visualization, and analysis.About RCSB PDB · Team Members · PDB Statistics · Protein Data Bank
  88. [88]
    Assessing the Quality of 3D Structures - RCSB PDB
    Oct 27, 2023 · The validation slider for the PDB entry 7s98 is shown here (Figure 2). Bars representing the R-free, clashscore, Ramachandran outliers, ...Assessing The Quality Of 3d... · Documentation · Accessing And Interpreting...
  89. [89]
    wwPDB: Worldwide Protein Data Bank
    Since 1971, the Protein Data Bank archive (PDB) has served as the single repository of information about the 3D structures of proteins, nucleic acids, and ...FAQ · PDB Archive Downloads · PDB deposition FAQs · PDB Versioned Archive
  90. [90]
    1.3M Structures and Counting. Highlights From the Latest CSD Data ...
    Sep 11, 2024 · Image of CSD entry LALNIN91 showing the hydrogen bonding network with refinement from high-resolution View of CSD Entry LALNIN91 showing the ...
  91. [91]
    New insights and innovation from a million crystal structures in ... - NIH
    The Cambridge Structural Database (CSD) is a large collection of crystal structures ... molecules in the CSD have no hydrogen bond donors or acceptors. It also ...
  92. [92]
    Shelx - Universität Göttingen
    No information is available for this page. · Learn why
  93. [93]
    Olex2 | OlexSys
    Olex2 is an easy-to-use program containing everything you need to solve, refine and finish small-molecule crystal structures using an intuitive user interface.For Crystallographers · For Chemists · Open Source
  94. [94]
    (IUCr) Crystallographic Information Framework
    The acronym CIF is used both for the Crystallographic Information File, the data exchange standard file format of Hall, Allen & Brown (1991) (see Documentation) ...Software for CIF and STAR · CIF dictionaries · Mailing list: cif-developers · Cif2
  95. [95]
    Crystal structure prediction of organic molecules by machine ...
    This study underscores the utility of combining machine learning models with efficient structure relaxations to accelerate organic crystal structure discovery.Missing: tools 2020s