Fact-checked by Grok 2 weeks ago

Crystal structure

A crystal structure is a highly ordered, periodic arrangement of atoms, ions, or molecules in three-dimensional space that defines the microscopic organization of a crystalline solid. This repeating pattern arises from the minimization of energy during solidification, where particles pack as densely as possible while respecting electrostatic and bonding interactions. The fundamental components include a Bravais lattice, which provides the geometric framework of translationally repeating points, and a basis or motif—a group of atoms attached to each lattice point—that specifies the atomic content. Together, these elements form the unit cell, the smallest volume that, when translated via lattice vectors, reproduces the entire structure. Crystal structures are classified into seven crystal systems based on the symmetry of their parameters (edge lengths a, b, c and angles α, β, γ): triclinic (no symmetry constraints), monoclinic (one 2-fold axis), orthorhombic (three perpendicular axes), tetragonal (two equal axes perpendicular to the third), trigonal (a = b = c, α = β = γ ≠ 90°), hexagonal (two equal axes at 120° with a third perpendicular), and cubic (all edges equal and angles 90°). These systems encompass 14 distinct Bravais lattices, accounting for variations like primitive, body-centered, face-centered, and base-centered arrangements that maintain without altering the overall . Common examples include the face-centered cubic lattice in metals like aluminum and the hexagonal close-packed structure in magnesium, both achieving high packing efficiency near 74%. The arrangement in a crystal structure profoundly influences macroscopic properties, such as mechanical strength, electrical conductivity, , and optical anisotropy, making it central to fields like , , and . Techniques like X-ray exploit the periodic nature of to determine structures at , revealing and intermolecular interactions essential for understanding phase transitions and defects. While perfect are idealized, real materials often feature imperfections like vacancies or dislocations that modify behavior without disrupting the underlying .

Basic Elements

Unit cell

In crystallography, the unit cell is defined as the smallest volume element of a crystal lattice that contains all the structural information necessary to describe the entire crystal, such that repeating this volume by pure translations fills the space without gaps or overlaps. This parallelepiped-shaped building block serves as the fundamental repeating unit, encapsulating the positions of atoms, ions, or molecules relative to points. Unit cells are classified as primitive or non-primitive based on the number of lattice points they contain. A unit cell, also known as a simple unit cell, includes exactly one lattice point and has the minimal volume required to represent the translations. In contrast, non-primitive unit cells, such as body-centered (with two lattice points) or face-centered (with four lattice points), contain additional lattice points at internal positions like the body or face centers, resulting in larger volumes but often higher symmetry for practical description. For example, the body-centered cubic structure features a lattice point at each corner and one at the cube's , while the face-centered cubic adds points at the centers of each face. The geometry of a is characterized by three constants—a, b, and c, representing the lengths of the edges along the three crystallographic axes—and three interaxial angles—\alpha (between edges b and c), \beta (between a and c), and \gamma (between a and b). These parameters fully define the shape and size of the , varying across crystal systems; for instance, in the cubic system, a = b = c and \alpha = \beta = \gamma = 90^\circ, forming a with equal edges and right angles. In the hexagonal system, a = b \neq c, with \alpha = \beta = 90^\circ and \gamma = 120^\circ, resulting in a . Visualizations of these often depict the cubic as a symmetric with atoms at corners (and possibly centers for non-primitive types), and the hexagonal as a taller prism with three equivalent basal edges forming 120° angles. Through , identical unit cells are repeated infinitely in three dimensions along the vectors, generating the complete crystal as an extended periodic . This repetition ensures that every point in the crystal can be reached by combinations of the unit cell's defining vectors, preserving the structural integrity across the material.

Crystal lattice

A crystal lattice is defined as an infinite, periodic array of discrete points in , where each point represents the position of a that repeats translationally to fill the entire volume without gaps or overlaps. This arrangement captures the long-range order inherent to crystalline solids, distinguishing them from amorphous materials by their repeating . The periodicity of the crystal lattice is mathematically described by three primitive vectors, conventionally denoted as \mathbf{a}, \mathbf{b}, and \mathbf{c}, which are non-coplanar and connect a lattice point to its nearest neighbors along the three independent directions. Any lattice point can then be reached by linear combinations of these vectors: \mathbf{R} = m\mathbf{a} + n\mathbf{b} + p\mathbf{c}, where m, n, and p are . These vectors define the fundamental translations that preserve the , ensuring that the environment around every lattice point is identical. The provides a dual representation in momentum space, constructed from basis vectors \mathbf{b}_1, \mathbf{b}_2, and \mathbf{b}_3 that satisfy \mathbf{a} \cdot \mathbf{b}_i = 2\pi \delta_{i,j} (with \delta as the ), ensuring to pairs of direct lattice vectors. points correspond to wavevectors where plane waves exhibit the same periodicity as the direct lattice, making it indispensable for analyzing phenomena, as scattering intensities peak at these points in experiments like X-ray . In qualitative terms, direct space describes the real-space positions and arrangements of atoms within the crystal, while reciprocal space captures the Fourier transform of this density, relating spatial frequencies to scattering angles and enabling the interpretation of diffraction patterns as a map of the lattice's periodicity. For instance, in a simple cubic lattice with lattice constant a, the direct lattice vectors are \mathbf{a} = a\hat{x}, \mathbf{b} = a\hat{y}, \mathbf{c} = a\hat{z}, yielding lattice points at coordinates (ma, na, pa) for integers m, n, p. The corresponding reciprocal lattice is also simple cubic but scaled by $2\pi/a, with points at (2\pi h/a, 2\pi k/a, 2\pi l/a) for integers h, k, l.

Indexing and Geometry

Miller indices

Miller indices are a symbolic notation system used in to designate the orientation of planes and directions within a crystal relative to the unit cell axes. This system was introduced in 1839 by the British mineralogist and crystallographer William Hallowes Miller in his work A Treatise on Crystallography, providing a concise way to describe lattice features using small integers derived from geometric intercepts. The notation facilitates the analysis of crystal symmetry and structure without requiring detailed coordinate descriptions, making it essential for identifying specific atomic arrangements in materials. For crystal planes, the are denoted as (hkl), where h, k, and l are integers representing the reciprocals of the fractional intercepts that the plane makes with the crystallographic a, b, and c, respectively, scaled to the smallest integers by clearing fractions. To determine the indices, one identifies the intercepts of the on the (in units of the parameters); takes the reciprocals; and multiplies through by the of the denominators to obtain , with the lowest values preferred. If a is parallel to an , the intercept is , resulting in a zero index for that component (e.g., a parallel to the b- and c- has k = 0 and l = 0). Planes with negative intercepts are indicated by placing a bar over the index (e.g., (\bar{1}00)). The notation {hkl} refers to a family of equivalent planes related by the crystal's operations. For instance, in a cubic , the (100) plane corresponds to a face of the unit cell parallel to the yz-, intersecting the a- at one unit length while being parallel to the others. Directions in the crystal lattice are specified using in the form [uvw], where u, v, and w are the smallest integers proportional to the components of the direction vector along the a, b, and c axes, respectively. Unlike planes, direction indices are not based on reciprocals but directly on the lattice vector coordinates, often reduced to the lowest terms. Negative directions are denoted with bars (e.g., [\bar{1}10]). The notation denotes a family of equivalent directions under . These indices relate directly to the primitive lattice vectors, allowing precise specification of atomic bonds or growth directions in crystals.

Crystal planes and directions

Crystal planes in a crystal structure are defined as families of parallel planes that pass through the lattice points, representing sets of atomic layers stacked in a repeating manner. These planes are fundamental to understanding the geometric arrangement of atoms within the lattice, as they delineate the layers where atoms are densely packed or exhibit specific bonding characteristics. In face-centered cubic (FCC) lattices, for instance, the {111} family of planes consists of close-packed atomic layers that form equilateral triangular arrangements, contributing to the high density and stability of these structures. Crystal directions, in contrast, refer to straight lines that connect lattice points along specific vectors within the crystal , defining pathways for alignment or movement. These directions often coincide with the shortest lattice vectors or high-symmetry axes, influencing processes such as diffusion or motion. In plastic deformation, slip directions are particular crystallographic directions along which dislocations glide, typically the close-packed directions like <110> in FCC crystals, enabling without bond breaking in other orientations. Due to the symmetry of the crystal lattice, multiple planes and directions that are equivalent under rotational or reflection operations form families, denoted by curly braces {} for planes and angle brackets <> for directions. In cubic crystals, the <100> family includes all directions equivalent to , such as and , which point along the principal axes and exhibit identical physical properties due to the lattice's isotropic in these orientations. Crystal planes and directions play a critical role in determining material properties, such as , where crystals fracture preferentially along planes of weak atomic bonding, like the {100} planes in some ionic crystals, resulting in smooth, flat surfaces. Similarly, during , facets often develop perpendicular to low-index directions or along specific planes with energy, influencing the overall of the crystal. In hexagonal close-packed (HCP) structures, the basal plane, denoted as (0001), serves as a prominent example of a close-packed layer that governs anisotropic growth and deformation behaviors in materials like magnesium or .

Interplanar spacing

Interplanar spacing, denoted as d_{hkl}, represents the perpendicular distance between successive parallel crystal planes characterized by (hkl). These planes are defined by their intercepts on the axes, and the spacing provides a key geometric parameter for understanding crystal periodicity and behavior. The concept arises from the arrangement of atoms in the , where parallel planes of atoms scatter waves constructively under specific conditions. In (where positions are expressed relative to the lattice vectors), the equation for planes is h x + k y + l z = p (with p integer for lattice planes). The general formula for interplanar spacing is d_{hkl} = \frac{1}{|\vec{G}_{hkl}|}, where \vec{G}_{hkl} = h \vec{b_1} + k \vec{b_2} + l \vec{b_3} is the reciprocal lattice vector, with reciprocal basis vectors \vec{b_i} defined such that \vec{a_i} \cdot \vec{b_j} = 2\pi \delta_{ij} (or 1 in some conventions). The magnitude |\vec{G}_{hkl}| is computed using the of the reciprocal lattice, which accounts for the angles between direct lattice vectors. This reciprocal formulation is universal and simplifies calculations for , as \vec{G}_{hkl} is to the (hkl) planes by . For crystal systems with orthogonal axes (cubic, tetragonal, orthorhombic), where all angles are 90° but edge lengths may differ, the formula simplifies to d_{hkl} = \left( \frac{h^2}{a^2} + \frac{k^2}{b^2} + \frac{l^2}{c^2} \right)^{-1/2}. In the cubic system, where a = b = c, it further simplifies to d_{hkl} = \frac{a}{\sqrt{h^2 + k^2 + l^2}}. In the hexagonal system, accounting for the sixfold and c/a ratio with \gamma = 120^\circ, the expression is d_{hkl} = \left[ \frac{4(h^2 + hk + k^2)}{3a^2} + \frac{l^2}{c^2} \right]^{-1/2}. These formulas enable precise computation of spacings from known lattice dimensions. A primary application of interplanar spacing lies in diffraction, where relates it to observable diffraction angles: n\lambda = 2d_{hkl} \sin\theta, with n as the diffraction order, \lambda the , and \theta the incidence angle. This equation allows experimental determination of d_{hkl} from measured peak positions, facilitating crystal structure analysis without deriving the full interference conditions here. Interplanar spacing is influenced by external factors that modify lattice parameters. Lattice strain, arising from mechanical deformation or defects, can compress or expand planes, shifting d_{hkl} and thus diffraction peaks. Temperature effects occur via , where increasing heat causes lattice parameters to grow, enlarging d_{hkl} proportionally; for instance, coefficients of thermal expansion quantify this change per degree . As an example, consider (NaCl), which adopts a face-centered cubic structure with lattice parameter a = 5.64 . For the (111) planes, d_{111} = \frac{a}{\sqrt{1^2 + 1^2 + 1^2}} = \frac{5.64}{\sqrt{3}} \approx 3.26 , illustrating how the formula yields atomic-scale distances relevant to and diffraction studies.

Symmetry Classification

Bravais lattices

A is defined as an infinite array of discrete points in where each point has an identical environment, generated solely by . These lattices represent the distinct ways to arrange points such that no two are equivalent except through pure translations, ensuring the lattice cannot be reduced to a simpler form by redefining the unit cell. In 1850, and crystallographer Auguste Bravais systematically enumerated these unique arrangements, identifying exactly 14 Bravais lattices in three dimensions. The criteria for uniqueness among Bravais lattices emphasize that additional lattice points within the conventional must arise only from translations of the vectors; any extraneous points would imply either a smaller cell or a different type, violating the minimal description. This leads to four primary centering types: (P), where lattice points are only at the corners; base-centered (C), with additional points at the centers of two opposite faces; body-centered (I), with a point at the body center; and face-centered (F), with points at the centers of all six faces. These centering variations, combined with the geometric constraints of the systems, yield the 14 distinct types. The 14 Bravais lattices are classified within seven crystal systems, each defined by specific relationships among the unit cell parameters (lattice constants a, b, c and angles α, β, γ). For instance, the cubic system features equal lengths and right angles (a = b = c, α = β = γ = 90°), while the tetragonal system has a = b ≠ c and α = β = γ = 90°. Representative examples include the cubic lattice in the cubic system and the body-centered tetragonal lattice in the tetragonal system. The full classification is summarized below:
Crystal SystemBravais Lattice TypesDescription
TriclinicNo symmetry constraints; a ≠ b ≠ c, α ≠ β ≠ γ.
Monoclinic, Base-centered (C)One ; a ≠ b ≠ c, α = γ = 90°, β ≠ 90°.
Orthorhombic, Base-centered (C), Body-centered (I), Face-centered (F)Three s; a ≠ b ≠ c, α = β = γ = 90°.
Tetragonal, Body-centered (I)Two equal lengths with s; a = b ≠ c, α = β = γ = 90°.
Trigonal (Rhombohedral)a = b = c, α = β = γ ≠ 90°.
Hexagonala = b ≠ c, α = β = 90°, γ = 120°.
Cubic, Body-centered (I), Face-centered (F)a = b = c, α = β = γ = 90°.
This structure ensures all possible translational symmetries are captured without redundancy.

Lattice systems

Lattice systems classify the possible geometries of crystal lattices into seven distinct categories, determined by the constraints on the unit cell's edge lengths a, b, c and the angles between them \alpha (between b and c), \beta (between a and c), and \gamma (between a and b). This classification arises from the requirement that the lattice must be periodic and translationally symmetric, with the systems reflecting increasing levels of metric symmetry from the lowest in triclinic to the highest in cubic. The grouping enables systematic analysis of crystal structures without considering full rotational symmetries, focusing solely on the metric relations that define distances and angles within the lattice./07%3A_Molecular_and_Solid_State_Structure/7.01%3A_Crystal_Structure) The seven lattice systems, along with their parameter constraints, are summarized in the following table:
Lattice Systemabc\alpha\beta\gammaNumber of Bravais Lattices
Triclinic\neq b\neq carbitrary\neq 90^\circ\neq 90^\circ\neq 90^\circ1 (primitive)
Monoclinic\neq b\neq carbitrary$90^\circ\neq 90^\circ$90^\circ2 (primitive, base-centered)
Orthorhombic\neq b\neq carbitrary$90^\circ$90^\circ$90^\circ4 (primitive, base-centered, body-centered, face-centered)
Tetragonal= b\neq carbitrary$90^\circ$90^\circ$90^\circ2 (primitive, body-centered)
Trigonal (Rhombohedral)= b = c\neq 90^\circ\neq 90^\circ\neq 90^\circ1 (rhombohedral)
Hexagonal= b\neq carbitrary$90^\circ$90^\circ$120^\circ1 (primitive)
Cubic= b = c$90^\circ$90^\circ$90^\circ3 (primitive, body-centered, face-centered)
These constraints define the unique metric properties of each system; for instance, the hexagonal system features a = b \neq c, \alpha = \beta = 90^\circ, \gamma = 120^\circ, which accommodates layered structures with sixfold in the basal plane./07%3A_Molecular_and_Solid_State_Structure/7.01%3A_Crystal_Structure) Each system encompasses one or more of Bravais lattices, which are the distinct translational types compatible with the system's metric constraints; for example, the cubic system hosts three Bravais lattices due to its high allowing , body-centered, and face-centered variants. This partitioning ensures that all possible three-dimensional are covered without in geometric description. To compute distances and angles within these systems, the \mathbf{G} is employed, a symmetric 3×3 whose elements are quadratic forms of the lattice parameters, such that the squared ds^2 = \mathbf{u}^T \mathbf{G} \mathbf{u} for a vector \mathbf{u}. In orthogonal systems like cubic, \mathbf{G} is diagonal with entries a^2, a^2, a^2, simplifying calculations, while in triclinic, all off-diagonal terms are nonzero.

Crystal systems

Crystal systems represent a fundamental classification in crystallography, grouping the 32 point groups into seven categories based on the overall symmetry compatible with the underlying lattice geometry. These systems—triclinic, monoclinic, orthorhombic, tetragonal, trigonal (or rhombohedral), hexagonal, and cubic—define the possible macroscopic symmetries of crystals by integrating rotational and reflection symmetries from point groups with the metric constraints of the lattice. Unlike lattice systems, which focus solely on geometric parameters like axis lengths and angles, crystal systems emphasize the full symmetry repertoire, ensuring that only point groups whose operations preserve the lattice are assigned to each category. Each crystal system corresponds directly to one of the seven lattice systems, but restricts inclusion to point groups that align with the lattice's metric symmetry, creating a mapping that excludes incompatible symmetries. For instance, the orthorhombic lattice system (with three mutually perpendicular axes of unequal length) maps to the orthorhombic crystal system, which accommodates point groups like 222, mm2, and mmm, all of which respect the 90° angles and distinct axis lengths. Similarly, the cubic lattice system aligns with the cubic crystal system, incorporating high-symmetry point groups such as 23, m3, 432, \bar{4}3m, and m\bar{3}m, where operations like threefold and fourfold rotations are feasible due to equal axes and right angles. This mapping ensures that the symmetry elements do not distort the lattice, with lower-symmetry point groups fitting into higher-symmetry systems if their operations are subgroups. Holohedry refers to the point group exhibiting the maximal symmetry within each crystal system, representing the "complete" form that includes all possible symmetry operations allowed by the lattice. For the cubic system, the holohedral group is m\bar{3}m (also denoted 4/m \bar{3} 2/m), featuring inversion, mirror planes, and multiple rotation axes, as seen in structures like halite (NaCl). In the triclinic system, the holohedry is simply \bar{1}, limited to inversion without rotations or mirrors, reflecting the absence of higher symmetries. These holohedral forms serve as benchmarks, with other point groups in the system being hemihedral or merohedral subgroups that omit certain operations. Illustrative examples highlight the diversity: the isometric (cubic) crystal system demonstrates maximal isotropy, with equal lattice parameters (a = b = c) and α = β = γ = 90°, enabling highly symmetric minerals like diamond (point group 4/m \bar{3} 2/m) or pyrite (point group \bar{4} 3 m). Conversely, the anorthic (triclinic) system lacks any symmetry constraints beyond the lattice, with arbitrary parameters (a ≠ b ≠ c, α ≠ β ≠ γ ≠ 90°), as in turquoise (point group 1) or microcline (point group \bar{1}), where even basic rotations are absent. These extremes underscore how crystal systems encapsulate both geometric and symmetric aspects./07%3A_Molecular_and_Solid_State_Structure/7.01%3A_Crystal_Structure) Transitions between crystal systems arise from subtle variations in lattice parameters or symmetry-breaking during phase changes, reclassifying structures when parameters deviate from defining thresholds. For example, a cubic system (a = b = c, all angles 90°) may shift to tetragonal if one axis elongates slightly (a = b ≠ c, angles 90°), reducing the point group from m\bar{3}m to 4/mmm, as observed in some perovskites under . Similarly, orthorhombic can distort to monoclinic by tilting one angle away from 90°, lowering the holohedry from mmm to 2/m; such changes often occur in temperature-driven transitions, where or atomic displacements break higher symmetries while preserving lower ones. These transitions illustrate the continuum of symmetry in , governed by energetic .

Point groups

Point groups in crystallography refer to the finite collections of symmetry operations—rotations, reflections, and inversions—that map a crystal lattice onto itself when performed around a fixed point, preserving the overall periodicity of the structure. These groups describe the external symmetry of crystals without involving translations, focusing solely on operations that leave a central point invariant. The possible rotation axes are limited to 1-, 2-, 3-, 4-, and 6-fold due to compatibility with in three dimensions./02%3A_Rotational_Symmetry/2.04%3A_Crystallographic_Point_Groups) The fundamental symmetry elements comprising these point groups include the identity operation (which leaves the unchanged), proper s about principal axes (denoted as n-fold, where n = 1, 2, 3, 4, or 6), mirror planes (perpendicular or parallel to axes), the inversion center (which maps each point to its opposite through the origin), and improper rotoinversions (combinations of and inversion). For instance, a 2-fold reverses direction by 180 degrees, while a mirror plane reflects across its surface. These elements combine in specific ways to form closed groups under composition, ensuring all operations are consistent with lattice invariance. There are exactly 32 crystallographic point groups, arising from the permissible combinations of these elements that align with the seven crystal systems. They are denoted using two primary notations: the Schoenflies system (common in molecular , e.g., D_{4h} for a group with a 4-fold , horizontal mirrors, and planes) and the (Hermann-Mauguin) system (standard in , e.g., 4/mmm for the same group, indicating a 4-fold with mirrors and planes). Examples include the 1 (or C_1, no beyond identity) in the triclinic system and the highly symmetric O_h (or m\bar{3}m) in the cubic system, which incorporates 48 operations including 3-fold, 4-fold, and 2-fold axes along multiple directions. These 32 point groups are distributed across the crystal systems as follows: 2 in triclinic (1, \bar{1}), 3 in monoclinic (2, m, 2/m), 3 in orthorhombic (222, mm2, mmm), 7 in tetragonal (4, \bar{4}, 4/m, 422, 4mm, \bar{4}2m, 4/mmm), 5 in trigonal (3, \bar{3}, , 3m, \bar{3}m), 7 in hexagonal (6, \bar{6}, 6/m, 622, 6mm, \bar{6}m2, 6/mmm), and 5 in cubic (23, m\bar{3}, 432, \bar{4}3m, m\bar{3}m). This distribution reflects the increasing constraints from lower (triclinic) to higher (cubic) systems, with cubic hosting the highest symmetry groups. For clarity, the groups can be summarized in the following table, using international notation with representative Schoenflies equivalents:
Crystal SystemNumber of Point GroupsExamples (International / Schoenflies)
Triclinic21 / C_1, \bar{1} / C_i
Monoclinic32 / C_2, m / C_s, 2/m / C_{2h}
Orthorhombic3222 / D_2, mm2 / C_{2v}, mmm / D_{2h}
Tetragonal74 / C_4, 4mm / C_{4v}, 4/mmm / D_{4h}
Trigonal53 / C_3, 3m / C_{3v}, \bar{3}m / D_{3d}
Hexagonal76 / C_6, 6mm / C_{6v}, 6/mmm / D_{6h}
Cubic523 / T, 432 / O, m\bar{3}m / O_h
This classification ensures each point group corresponds uniquely to observable crystal morphologies, such as the cubic forms in the halite structure. Stereographic projections provide a standard method to visualize point group symmetries by mapping the directions of symmetry elements onto a unit sphere. In this representation, rotation axes appear as points (poles) on the sphere's surface, mirror planes as great circles (equators), and the projection from the south pole onto a plane yields a 2D diagram showing the arrangement of these elements. For example, the projection for the cubic group m\bar{3}m (O_h) displays four 3-fold axes along body diagonals, three 4-fold axes along face normals, and six 2-fold axes along edge midpoints, encircled by appropriate mirror planes. These projections aid in identifying and distinguishing the 32 groups without three-dimensional models.

Space groups

Space groups represent the complete set of symmetries for periodic crystal structures in three dimensions, extending the 32 crystallographic s by incorporating translations along with nonsymmorphic operations such as axes and glide planes. These elements allow the symmetry operations to fill space while maintaining the periodic arrangement of atoms. There are exactly 230 distinct space groups, enumerated and classified in the International Tables for Crystallography. Of these, 73 are symmorphic space groups, which combine point group operations with pure translations without fractional shifts, whereas the remaining 157 are nonsymmorphic, featuring axes (rotations combined with partial translations parallel to the axis) or glide planes (reflections combined with partial translations parallel to the plane)./03:_Space_Groups/3.04:_Group_Properties) Space groups are denoted using the Hermann–Mauguin symbol, as standardized in the International Tables for Crystallography, which specifies the type, principal axes, and any nonsymmorphic elements. For instance, the symbol P2₁/c describes a (P) monoclinic with a twofold screw axis (2₁) and a glide plane (c) perpendicular to the b-axis, reflecting combined rotational and translational symmetries. The distribution of the 230 space groups varies by , reflecting the increasing constraints on symmetry as metric parameters become more equal:
Crystal SystemNumber of Space Groups
Triclinic2
Monoclinic13
Orthorhombic59
Tetragonal68
Trigonal25
Hexagonal27
Cubic36
For example, the cubic system hosts 36 space groups, the highest for any system due to its high symmetry. Determination of a crystal's space group typically involves analyzing X-ray or electron diffraction patterns for systematic absences—missing reflections attributable to translational symmetries like screw axes or glide planes—which narrow down possibilities from the full set of 230. Computational tools, such as the Bilbao Crystallographic Server (launched around 2000 with ongoing EU-funded development), facilitate this process by generating symmetry databases, subgroup relations, and visualization aids to match observed data to specific space groups.

Atomic Coordination

Close packing

Close packing represents the most efficient geometric arrangement of identical , maximizing space utilization in crystal structures by minimizing voids between them. This arrangement begins with a two-dimensional hexagonal layer, where each sphere contacts six neighbors, forming a close-packed plane. Subsequent layers are stacked by placing spheres in the tetrahedral depressions—triangular voids formed by three spheres in the underlying layer—resulting in three possible positions labeled A, B, and C, depending on their lateral offset relative to the first layer. The specific stacking sequence of these layers defines distinct three-dimensional structures. Hexagonal close packing (HCP) follows an ABAB... pattern, with the third layer aligning directly above the first (A position), repeating every two layers to yield a . Cubic close packing (CCP), equivalent to the face-centered cubic (FCC) , employs an ABCABC... sequence, where each successive layer occupies a new position, producing after three layers. Both HCP and CCP achieve identical maximum for equal spheres. Many elemental metals adopt these close-packed structures due to their simple . For instance, (Cu) crystallizes in the FCC arrangement, while magnesium (Mg) forms an HCP structure. In these packings, the nestled spheres create inherent voids: octahedral sites, coordinated by six surrounding spheres, and tetrahedral sites, bounded by four. Each close-packed unit contains one octahedral void and two tetrahedral voids per sphere, providing potential spaces within the otherwise dense array. Both HCP and FCC exhibit the highest among elemental crystal structures, serving as a for .

Atomic packing factor and coordination number

The atomic packing factor (APF), also known as packing efficiency, quantifies the fraction of the unit cell volume occupied by the atoms in a crystal structure, assuming hard-sphere atoms that touch along the closest-packed directions. It is calculated as the total volume of atoms within the unit cell divided by the volume of the unit cell itself. The (CN) represents the number of nearest-neighbor atoms surrounding a given atom in the , providing insight into the local atomic environment and stability. These metrics are fundamental for understanding density and bonding in crystalline materials, with higher APF and CN generally indicating denser packing and stronger metallic or ionic interactions. For the simple cubic (SC) structure, the APF is derived as follows: the unit cell contains 1 atom (contributed by 8 corners × 1/8), with lattice parameter a = 2r where r is the , since atoms touch along the edge. The volume of the atom is \frac{4}{3}\pi r^3, and the unit cell volume is a^3 = (2r)^3 = 8r^3. Thus, \text{APF} = \frac{\frac{4}{3}\pi r^3}{8r^3} = \frac{\pi}{6} \approx 0.52. The CN in SC is 6, corresponding to the six nearest neighbors along the three orthogonal axes. In the body-centered cubic (BCC) structure, 2 atoms occupy the unit cell (8 corners × 1/8 + 1 center), and atoms touch along the body diagonal, giving a = \frac{4r}{\sqrt{3}} since the diagonal length is a\sqrt{3} = 4r. The total atomic volume is $2 \times \frac{4}{3}\pi r^3 = \frac{8}{3}\pi r^3, and the unit cell volume is a^3 = \left(\frac{4r}{\sqrt{3}}\right)^3 = \frac{64r^3}{3\sqrt{3}}. Substituting yields \text{APF} = \frac{\frac{8}{3}\pi r^3}{\frac{64r^3}{3\sqrt{3}}} = \frac{\pi \sqrt{3}}{8} \approx 0.68. The CN for BCC is 8, with the eight nearest neighbors equivalently located along the body-diagonal directions. The face-centered cubic (FCC) structure, along with hexagonal close-packed (HCP), achieves the maximum APF for spherical atoms. It has 4 atoms per unit cell (8 corners × 1/8 + 6 faces × 1/2), with atoms touching along the face diagonal, so a = 2\sqrt{2}r as the diagonal is a\sqrt{2} = 4r. The total atomic volume is $4 \times \frac{4}{3}\pi r^3 = \frac{16}{3}\pi r^3, and the unit cell volume is a^3 = (2\sqrt{2}r)^3 = 16\sqrt{2}r^3. Thus, \text{APF} = \frac{\frac{16}{3}\pi r^3}{16\sqrt{2}r^3} = \frac{\pi}{3\sqrt{2}} = \frac{\pi \sqrt{2}}{6} \approx 0.74. The CN in and is 12, reflecting the close-packed arrangement where each atom has twelve nearest neighbors in the plane and adjacent layers. This highest APF value is realized in close-packed structures like and . A correlation exists between CN and APF: structures with higher CN, such as FCC (CN=12, APF=0.74) versus BCC (CN=8, APF=0.68) or SC (CN=6, APF=0.52), exhibit greater packing efficiency due to more optimal spatial filling by nearest neighbors. In ionic crystals, such as (NaCl) adopting the rock salt structure, both Na⁺ and Cl⁻ ions have a CN of 6, forming an FCC-like arrangement of anions with cations in octahedral sites, resulting in a balanced packing suited to their radius ratio.

Interstitial sites

In crystal lattices, interstitial sites are voids or gaps between the primary atoms that can accommodate smaller atoms or ions without significantly distorting the structure. These sites are particularly prominent in close-packed structures such as face-centered cubic (FCC) and hexagonal close-packed (HCP), where they arise from the efficient arrangement of spheres./08%3A_Ionic_and_Covalent_Solids_-_Structures/8.02%3A_Close-packing_and_Interstitial_Sites) The two primary types of interstitial sites in close-packed lattices are and octahedral. A is surrounded by four host atoms arranged at the corners of a , providing 4-fold coordination to any inserted atom. In contrast, an octahedral site is bounded by six host atoms in an octahedral geometry, offering 6-fold coordination./08%3A_Ionic_and_Covalent_Solids_-_Structures/8.02%3A_Close-packing_and_Interstitial_Sites) The size of these sites is characterized by the radius ratio, defined as the maximum radius of an interstitial atom (r_i) relative to the host atom radius (r_h) that can fit without overlap. For tetrahedral sites in FCC and HCP structures, this ratio is 0.225, while for octahedral sites, it is 0.414. These ratios determine stable insertion: smaller interstitial atoms, such as , preferentially occupy tetrahedral sites due to the tighter fit (e.g., radius ratio ≈0.2 for H in many metals), whereas larger ones like carbon favor octahedral sites. In close-packed structures, the number of interstitial sites per host atom is fixed: there are two tetrahedral sites and one octahedral site per atom. For an FCC with four atoms, this corresponds to eight tetrahedral sites and four octahedral sites. Similarly, HCP structures exhibit the same ratio per atom./08%3A_Ionic_and_Covalent_Solids_-_Structures/8.02%3A_Close-packing_and_Interstitial_Sites) Geometrically, in the FCC , octahedral sites are located at the body center (\frac{1}{2}, \frac{1}{2}, \frac{1}{2}) and at the centers of each edge (e.g., \frac{1}{2}, 0, 0), while tetrahedral sites occupy positions along the body diagonals, such as \left( \frac{1}{4}, \frac{1}{4}, \frac{1}{4} \right) and equivalent sites. These coordinates reflect the symmetry of the and ensure the sites are equidistant from surrounding atoms. Practical examples illustrate the role of these sites. In austenitic steel, carbon atoms occupy octahedral interstitial sites in the FCC iron lattice, enabling up to about 2 wt% carbon. For , metals like and certain alloys accommodate in tetrahedral sites, as seen in interstitial hydrides where the small radius (≈0.37 ) fits the 0.225 ratio relative to typical host atoms (e.g., at 1.43 ).

Crystal Defects

Point defects and impurities

Point defects are localized disruptions in the crystal lattice confined to one or a few atomic sites, contrasting with extended defects like lines or planes. These include intrinsic defects such as vacancies and interstitials, as well as extrinsic defects from impurities. They arise due to , , or intentional addition during , influencing mechanical, electrical, and of materials. Vacancies represent the simplest intrinsic point defect, where an atom or ion is absent from its regular position, creating a void. In elemental crystals like metals, a single vacancy suffices, but in ionic compounds, charge balance requires paired cation and anion vacancies to avoid net charge, known as a . Schottky defects predominate in crystals with similar ion sizes, such as NaCl, where both Na⁺ and Cl⁻ vacancies form to maintain electroneutrality. The formation of these defects involves breaking bonds and rearranging neighboring atoms, typically requiring energies on the order of 1-2 per vacancy pair. Frenkel defects, another intrinsic type prevalent in ionic crystals with significant size disparity between cations and anions, involve a cation displacing to an , leaving a vacancy behind. This maintains overall charge neutrality without requiring surface incorporation. Common in compounds like AgBr or ZnS, Frenkel defects enhance ionic conductivity by facilitating ion hopping. Unlike Schottky defects, they do not alter the total number of lattice sites but redistribute ions. Interstitial defects occur when an extra atom occupies a position between regular sites, often in open structures like close-packed metals. These can be self-interstitials from host atoms or foreign interstitial impurities, such as carbon in iron, which occupies octahedral or tetrahedral voids. Interstitials distort the more severely than vacancies due to bond compression, leading to higher formation energies, typically 3-5 . In ionic crystals, interstitials must pair with vacancies to preserve neutrality, as in Frenkel defects. Substitutional impurities replace host atoms at lattice sites, forming solid solutions up to certain limits determined by atomic size mismatch (Hume-Rothery rules) and solubility. For instance, phosphorus atoms substituting silicon in semiconductors create n-type material by donating an extra valence electron to the conduction band, enabling controlled electrical conductivity. Dopants like phosphorus are introduced at concentrations around 10^{15}-10^{18} cm^{-3}, far exceeding thermal defect levels, to tailor device performance. Solid solution limits, often below 1-10 at.% for many systems, prevent phase separation and maintain single-crystal integrity. The equilibrium concentration of point defects follows the , approximately c \approx \exp\left( -\frac{E_f}{k_B T} \right), where E_f is the defect formation energy, k_B is Boltzmann's constant, and T is ; this yields extremely low intrinsic concentrations (e.g., ~10^{-17} at for typical E_f \sim 1 eV), that rise exponentially with . Point defects profoundly affect properties: vacancies and interstitials mediate , while impurities enable doping for semiconductors. Color centers, such as the in NaCl—an trapped in a Cl⁻ vacancy—absorb visible light, imparting yellow coloration to the otherwise transparent crystal upon irradiation. In oxides like TiO₂ or ZrO₂, oxygen vacancies act as electron donors, enhancing n-type and catalytic activity by creating localized states in the bandgap.

Line defects: dislocations

Line defects, known as dislocations, are one-dimensional imperfections in the crystal lattice that extend along lines and play a crucial role in enabling plastic deformation at stresses far below the theoretical of perfect crystals. These defects were first proposed independently by , E. Orowan, and M. Polanyi in 1934 to explain the observed low yield stresses in metals, where slip occurs along specific crystal planes and directions known as slip systems. Dislocations distort the lattice locally, creating long-range elastic strain fields that interact with each other and with applied stresses, facilitating without breaking atomic bonds across the entire plane. Dislocations are classified into three main types based on the orientation of their line relative to the , a that quantifies the and of the lattice : , , and mixed. An dislocation arises from the insertion or omission of an extra half-plane of atoms, resulting in a perpendicular to the dislocation line; this creates compressive above the line and tensile below it in the slip . A dislocation, in contrast, involves a displacement where the is parallel to the dislocation line, producing a helical that allows the dislocation to propagate perpendicular to the . Mixed dislocations combine characteristics of both, with the at an intermediate angle to the line , and are the most common form observed in deformed crystals. The Burgers vector b is formally defined through the Burgers circuit: a closed loop drawn around the dislocation line in the actual distorted lattice fails to close in a perfect reference lattice, and the closure failure vector is b, given by b = ∮ dl, where the integral is taken along the circuit. In crystalline materials, b is typically a lattice translation vector, such as a/2⟨110⟩ in face-centered cubic metals, ensuring the distortion is consistent with the periodicity of the lattice and minimizing the energy of the defect. This vector not only characterizes the type of dislocation but also determines the slip system along which it moves, linking the defect directly to the crystal's symmetry. Dislocation density, denoted ρ and measured in lines per unit area (typically in cm⁻²), varies widely depending on processing and deformation history, ranging from about 10⁶ cm⁻² in well-annealed metals to 10¹² cm⁻² in heavily deformed ones. Annealing treatments reduce ρ by promoting of dislocations with opposite Burgers vectors and rearrangement into lower-energy configurations, such as subgrain boundaries, thereby softening the material. Dislocations move primarily through two mechanisms: glide and climb, both driven by resolved shear stresses but differing in atomic processes. Glide is a conservative motion where the dislocation translates in its slip plane without changing the number of atoms, occurring via of atomic planes and controlled by the Peierls stress—the intrinsic lattice resistance arising from the periodic potential that dislocations must overcome. This stress, first calculated by R. Peierls in using a sinusoidal potential model, is exponentially sensitive to the dislocation core width and typically low in metals (on the order of G/1000, where G is the ), enabling room-temperature . Climb, a non-conservative requiring diffusion of vacancies or interstitials, allows motion perpendicular to the slip plane and is thermally activated, becoming significant at high temperatures. In metals, dislocations are responsible for , where plastic deformation increases strength through multiplication (e.g., via Frank-Read sources) and interactions that create tangles and forests, impeding further glide and raising the proportionally to √ρ. For instance, in face-centered cubic , initial densities around 10⁷ cm⁻² rise to 10¹¹ cm⁻² after , leading to significant hardening before processes dominate during annealing.

Planar defects: grain boundaries

Grain boundaries are two-dimensional interfaces that separate adjacent crystalline regions, or , within a polycrystalline , arising from variations in crystallographic during solidification or deformation. These planar defects play a crucial role in determining the mechanical, electrical, and thermal properties of by influencing motion and atomic transport. In metals and ceramics, grain boundaries typically span thicknesses of a few atomic layers and can exhibit periodic or disordered atomic arrangements depending on the degree of misorientation between the grains. Grain boundaries are classified into low-angle and high-angle types based on the misorientation θ between the adjacent lattices. Low-angle grain boundaries, with θ < °, consist of ordered arrays of dislocations that accommodate the small rotational mismatch, as described by the dislocation model developed by Read and Shockley. High-angle grain boundaries, with θ > °, feature more disordered structures resembling an amorphous interphase, lacking long-range periodicity and often exhibiting higher defect densities. Tilt grain boundaries, which involve rotation about an perpendicular to the boundary plane, are modeled as arrays of edge dislocations, while twist grain boundaries, involving rotation about an normal to the boundary plane, are composed of screw dislocation networks. These models highlight how dislocations serve as building blocks for low-angle boundaries, linking them conceptually to line defects within s. The energy of grain boundaries, a key thermodynamic property, typically ranges from 0.1 to 1 J/m² in metals, reflecting the disruption to bonding at the . For low-angle boundaries, the Read-Shockley quantifies this energy as \gamma = \gamma_0 \theta (A - \ln \theta), where γ is the boundary energy, γ₀ is a constant related to dislocation core energy, θ is the misorientation angle in radians, and A is a material-dependent . This logarithmic dependence predicts lower energies for small θ, aligning with experimental observations in materials like aluminum and . High-angle boundaries generally have higher, more constant energies due to their disordered nature. Grain boundaries significantly affect material behavior by serving as fast diffusion paths for atoms and vacancies, accelerating processes like and compared to bulk . They also act as preferential sites for of second phases, which can enhance or degrade properties depending on the precipitate type. In terms of mechanical strengthening, grain boundaries impede glide, leading to the Hall-Petch relationship, where strength σ_y increases with decreasing d as σ_y = σ_0 + k d^{-1/2}, with σ_0 and k as material constants; this effect is prominent in polycrystalline metals like and . Prominent examples of grain boundaries occur in polycrystalline metals, such as aluminum alloys used in applications, where they form networks that control fatigue resistance. Triple junctions, the points where three grain boundaries meet, introduce additional complexity, often exhibiting dihedral angles governed by energy balance and serving as nucleation sites for recrystallization or in materials like nickel-based superalloys.

Structure Determination and Prediction

Experimental methods for structure determination

The experimental determination of crystal structures relies primarily on diffraction techniques, which exploit the wave nature of radiation interacting with the periodic atomic lattice. In 1912, demonstrated that X-rays could be diffracted by , providing the first evidence for both the wave properties of X-rays and the regular arrangement of atoms in a crystal lattice. This breakthrough was soon followed by the development of by and William Lawrence Bragg in 1913, which relates the of X-rays (\lambda) to the interplanar spacing (d_{hkl}) and the diffraction angle (\theta) via n\lambda = 2d_{hkl} \sin\theta, enabling quantitative analysis of crystal planes. These foundational advances established X-ray (XRD) as the cornerstone method for resolving atomic positions in crystalline materials. Key XRD techniques include the Laue method and the rotating crystal method. The Laue method employs a stationary exposed to a polychromatic beam, producing a pattern that reveals the crystal's and orientation but is less suited for full structure determination due to the complexity of variations. In contrast, the rotating crystal method uses a monochromatic beam and rotates the crystal to access multiple reflections, allowing systematic collection of data from various planes; this approach, refined in the early , forms the basis of single-crystal diffractometers. The of diffracted beams is governed by the F_{hkl}, defined as F_{hkl} = \sum_j f_j \exp\left(2\pi i (h x_j + k y_j + l z_j)\right), where f_j is the atomic factor for the j-th atom, and (x_j, y_j, z_j) are its in the unit cell; this complex quantity encodes both the amplitude and of from the (hkl) ./03%3A_X-rays/3.27%3A_Structure_Factor) The process of structure determination via involves several steps: through exposure of the crystal to X-rays and recording of diffraction intensities on detectors; indexing of reflections to assign (hkl); correction for factors like Lorentz-polarization effects; and solving the phase problem, which arises because only intensities (|F_{hkl}|^2) are measured, not phases. The Patterson function, introduced in 1935, addresses this by computing a map of interatomic vectors from intensity data, facilitating location of heavy atoms or molecular fragments./01%3A_Chapters/1.07%3A_New_Page) For small molecules, direct methods probabilistically estimate phases using statistical relationships between structure factors, often yielding initial electron density maps that are refined iteratively via least-squares minimization against observed data. Complementary diffraction techniques include neutron and , each offering distinct advantages over . Neutron diffraction excels in locating light atoms like , which scatter X-rays weakly, and in probing magnetic structures due to the neutron's intrinsic , making it ideal for materials with isotopic variations or disordered systems./07%3A_Molecular_and_Solid_State_Structure/7.05%3A_Neutron_Diffraction) , typically performed in transmission electron microscopes, provides high spatial resolution for nanoscale crystals or thin films, enabling of beam-sensitive samples where larger crystals are unavailable, though it is limited by multiple effects in thicker specimens. Additional methods support detailed characterization of crystal imperfections and surfaces. (TEM) visualizes defects such as dislocations and stacking faults at atomic resolution by combining imaging with selected-area , revealing local structural deviations. (STM) and (AFM) probe surface atomic arrangements and defects on conductive or insulating crystals, respectively, offering real-space views that complement reciprocal-space data for understanding surface reconstructions or adsorbate effects.

Computational prediction of structures

Computational prediction of crystal structures relies on theoretical methods to determine stable atomic arrangements from first principles, typically starting with only the and thermodynamic conditions. (DFT) serves as a cornerstone for these predictions by enabling the minimization of total energy to identify low-energy configurations, often incorporating corrections for van der Waals interactions to improve accuracy for weakly bound systems. Lattice dynamics calculations extend this framework by accounting for vibrational contributions to , allowing predictions at finite temperatures beyond the zero-Kelvin . Algorithms for structure prediction employ global optimization techniques to explore vast configuration spaces, generating candidate structures within possible space groups as symmetry constraints. The USPEX method, an evolutionary algorithm, iteratively evolves populations of structures through variation operators and local relaxation via DFT, reliably identifying stable phases for systems up to dozens of atoms per unit cell. Complementary approaches include random sampling, which generates diverse initial lattices for energy ranking, and basin-hopping methods that escape local minima to converge on global optima. Databases such as the Inorganic Crystal Structure Database (ICSD) provide repositories of experimentally derived inorganic structures for benchmarking predictions, while the Materials Project, established in 2011, offers a computational database of over 200,000 predicted crystal structures computed with DFT, as of 2025. Recent integrations include structures from machine learning models like GNoME, with over 30,000 added in 2025 updates. In the 2020s, machine learning integrations have accelerated these processes, with equivariant neural networks trained on quantum mechanical data achieving near-DFT accuracy in structure generation and ranking for molecular crystals. Despite advances, challenges persist, including the role of kinetic barriers that hinder access to thermodynamically favored structures during formation, complicating predictions of observed s. Polymorphism introduces further ambiguity, as multiple low-energy structures may compete, particularly in organic systems where DFT struggles with dispersion forces and conformational flexibility, often requiring for reliable energy differences. For instance, DFT-based simulations have successfully predicted high-pressure s of , such as the transition from the diamond structure to the β-Sn structure at around 10 GPa, revealing metastable intermediates like Imma-Si that influence phase stability under compression.

Polymorphism and Variations

Polymorphism

Polymorphism refers to the ability of a single to crystallize into two or more distinct structures, known as polymorphs, which differ in the arrangement and/or conformation of their constituent atoms or molecules while sharing the same . These different forms arise from variations in packing efficiency, intermolecular interactions, or molecular flexibility during crystallization. A classic example is carbon, which can form with a cubic structure characterized by strong tetrahedral bonding and with a layered hexagonal structure featuring weaker van der Waals interactions between planes. Polymorphic transitions between forms can be categorized based on the underlying . Conformational transitions involve changes in the molecular or without breaking bonds, often seen in flexible organic molecules. Displacive transitions occur through small, coordinated atomic displacements, typically reversible and diffusionless, resembling martensitic transformations in metals. Reconstructive transitions, in contrast, require the breaking and reformation of chemical bonds, making them higher-energy processes that are often irreversible under ambient conditions. For carbon, the transition from to is reconstructive, demanding extreme and . The thermodynamic stability of polymorphs is governed by the (G = H - TS), where the form with the lowest G at a given and is most stable. Enantiotropic polymorphism describes systems with a reversible transition point, below which one polymorph is stable and above which the other is, as in the α-β pair. Monotropic polymorphism, however, features one form that is always more stable across all accessible conditions, with the metastable form kinetically trapped, such as relative to at standard conditions. Another carbon polymorph, (hexagonal diamond), exemplifies monotropic behavior, forming under high in impacts but converting to cubic over time. In July 2025, Chinese researchers reported the successful synthesis of high-purity lonsdaleite using advanced methods, highlighting its potential as a harder alternative to cubic . In pharmaceuticals, polymorphism significantly influences drug performance. , an antiretroviral medication, exists in multiple forms; Form I was initially marketed but later discovered to convert to the more stable Form II, which has lower and , prompting a costly reformulation. Such cases highlight industrial implications, including challenges in securing patents for specific polymorphs, as often covers the compound regardless of form, and variations can alter rates, efficacy, and shelf-life . Comprehensive polymorph screening during development is thus essential to mitigate risks in and regulatory approval.

Quasicrystals and aperiodic structures

Quasicrystals represent a class of ordered structures that lack the translational periodicity characteristic of traditional s, thereby challenging the classical definition of a crystal as requiring a repeating . In 1982, observed tenfold symmetry in the pattern of a rapidly solidified aluminum-manganese , marking the first experimental evidence of such a phase. This discovery, initially met with skepticism, was published in 1984 and later recognized with the 2011 awarded to Shechtman for revealing the existence of quasicrystals. Unlike Bravais lattices, which underpin periodic crystals, quasicrystals exhibit forbidden rotational symmetries such as five-, ten-, or twelvefold axes that cannot extend periodically in three dimensions. Key characteristics of quasicrystals include their aperiodic long-range order, evidenced by sharp diffraction peaks in or patterns that appear at positions following sequences like the series, rather than simple integer multiples of a . These patterns arise from the projection of a higher-dimensional periodic onto lower-dimensional , producing dense, non-repeating arrangements of atoms. Theoretical models, such as Penrose tilings introduced in 1974, provided early mathematical frameworks for understanding quasicrystalline order; these two-dimensional aperiodic tilings with fivefold symmetry served as analogs for real materials. For three-dimensional cases, icosahedral quasicrystals, like Shechtman's original Al-Mn phase, feature icosahedral symmetry without translational repetition. Aperiodic structures encompass , which maintain perfect order without periodicity, and incommensurate structures, which are modulated versions of periodic with irrational wavelength ratios leading to apparent aperiodicity. are distinguished by their pure rotational incompatible with three-dimensional periodicity, while incommensurate phases often arise from density waves superimposed on a parent . Notable examples include the decagonal Al-Cu-Fe , which exhibits tenfold along one and periodic stacking perpendicular to it, and natural occurrences such as the icosahedral phase Al63Cu24Fe13 found in the Khatyrka . These materials display unique properties, including high due to their dense atomic packing and low friction coefficients from reduced surface adhesion, making them suitable for coatings in tribological applications.

Structure-Property Relationships

Influence on physical properties

The density of a crystal is fundamentally determined by the arrangement of atoms within its unit cell, calculated as the mass of the atoms or formula units per unit volume. Specifically, the theoretical density \rho is given by the formula \rho = \frac{Z M}{N_A V_\text{cell}}, where Z is the number of formula units per unit cell, M is the molar mass of the formula unit, N_A is Avogadro's number, and V_\text{cell} is the volume of the unit cell. This relationship highlights how tighter atomic packing in structures like face-centered cubic lattices leads to higher densities compared to simpler cubic arrangements. Crystal structure also governs elastic properties, particularly through the directional strength of interatomic bonds. In covalent crystals, exhibits significant , with values higher along directions aligned with strong covalent bonds due to the resistance to deformation in those orientations. For instance, in , the reaches a maximum of approximately 1210 GPa along the direction, corresponding to the tetrahedral carbon-carbon bonds, compared to a minimum of 1050 GPa along . This structural rigidity contributes to 's exceptional , arising from its continuous three-dimensional tetrahedral network that distributes evenly and resists . Thermal expansion behavior is likewise influenced by the crystal lattice, with linear coefficients varying by crystallographic direction due to differences in bond flexibility. In quartz (\alpha-SiO_2), the coefficient along the c-axis can become negative at low temperatures, reflecting transverse vibrations of oxygen atoms that effectively contract the lattice despite increasing thermal energy. Such anisotropy arises from the helical arrangement of SiO_4 tetrahedra in the trigonal structure, where rigid units limit expansion in certain axes. In ionic crystals like NaCl, the rock salt structure—featuring alternating Na^+ and Cl^- ions in an face-centered cubic array—facilitates intrinsic ionic conductivity by providing open pathways for ion hopping, enabling measurable electrical transport even in pure crystals at elevated temperatures. For polycrystalline materials, where random orientations average out single-crystal , the Voigt-Reuss bounds offer theoretical limits on effective moduli. The Voigt average assumes uniform and provides an upper bound on the and moduli, while the Reuss average assumes uniform and yields a lower bound; the (Voigt-Reuss-Hill average) approximates the polycrystalline response. These bounds are essential for predicting macroscopic properties from known single-crystal data, as demonstrated in aggregates of cubic materials where the spread between bounds reflects underlying structural disorder.

Anisotropy in crystals

Anisotropy in crystals refers to the directional dependence of physical properties arising from the asymmetric arrangement of atoms in the , unlike isotropic materials where properties are uniform in all directions. For instance, the in (CaCO₃) varies significantly along different crystallographic axes, leading to double refraction where a single light ray splits into two polarized rays with orthogonal polarizations. This variation stems from the crystal's trigonal structure, which lacks full . In optical properties, manifests as , where the differs for rays polarized parallel and perpendicular to the optic axis, quantified by the difference in refractive indices Δn = n_e - n_o ( and indices, respectively). exhibits strong negative with Δn ≈ -0.172. , another optical effect, involves absorption varying with direction, causing crystals to appear in different colors when viewed along different axes under polarized ; this occurs in anisotropic minerals like due to selective absorption of components. These phenomena are described by second-rank tensors, such as the tensor, whose form is constrained by the crystal's . Electrical is evident in materials with layered structures, where is high along planes of strong bonding but low perpendicular to them. In , the basal plane is orders of magnitude greater than along the c-axis due to delocalized π electrons within sp²-bonded carbon layers, with anisotropy ratios reaching 10² to 10⁵. Similarly, hexagonal (h-BN) displays graphite-like , with high in-plane thermal and electrical from its layered hexagonal lattice of alternating and atoms, while interlayer van der Waals bonds limit perpendicular transport. Mechanical anisotropy arises from structural layering, promoting easy cleavage along weak planes. In , such as (KAl₂(AlSi₃O₁₀)(OH)₂), the perfect basal cleavage parallel to the (001) planes results from weak interlayer bonds between silicate sheets, yielding high tensile strength in-plane but facile perpendicularly. This directional weakness facilitates applications in insulators and composites. Certain anisotropic crystals exhibit , where mechanical stress induces electric polarization, requiring a non-centrosymmetric structure. (SiO₂), with its trigonal α-quartz form belonging to 32, demonstrates this effect: compression along the c-axis generates opposite charges on the ends due to asymmetric silicon-oxygen tetrahedra . The anisotropy of physical properties in crystals is determined by their symmetry, which dictates the rank and form of the associated tensors. In contrast, cubic crystals like (NaCl) or exhibit for many properties, including electrical conductivity and , because their high symmetry ( m3m) results in scalar-like behavior independent of direction.

References

  1. [1]
    Crystallography Basics - Chemical Instrumentation Facility
    A crystal structure is a unique arrangement of atoms, molecules or ions in a crystal. It is composed of a motif, which is a set of atoms arranged in a ...
  2. [2]
    [PDF] Experiment 7: Understanding Crystal Structures - Bellevue College
    When a solid crystallizes, the atoms or ions that form it try to pack as tightly as possible. In the case of ions, repulsive forces between ions of the same ...
  3. [3]
    Page not found | Evgeny Tsymbal | Nebraska
    Insufficient relevant content. The requested URL (https://tsymbal.unl.edu/sites/unl.edu.cas.physics.tsymbal/files/media/file/Section%201_Crystal%20Structure.pdf) returns a "Page not found" error, indicating the content is inaccessible or does not exist.
  4. [4]
    [PDF] Basic Concepts of Crystallography
    • Crystal Structure: Basic Definitions - lecture. • Calculation of material ... By definition crystal is a periodic arrangement of repeating “motifs ...
  5. [5]
    [PDF] Crystal Structure and Defects
    Crystal structures are regular 3-dimensional patterns of atoms in space. ➢. The regularity with which atoms are packed in solids arises from geometrical.
  6. [6]
    [PDF] X-Ray Crystallography Laboratory Department of ... - MSU chemistry
    A crystal structure determines the connectivity of the atoms in a compound and the way the molecule (or molecules) pack in three dimensions to form a solid ...<|control11|><|separator|>
  7. [7]
    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.The Simplest Repeating Unit... · Three-Dimensional Graph · NaCl and ZnS
  8. [8]
    Crystal Morphology, Crystal Symmetry, Crystallographic Axes
    Aug 13, 2010 · The crystallographic axes are imaginary lines that we can draw within the crystal lattice. These will define a coordinate system within the crystal.
  9. [9]
    [PDF] Crystal basis:
    Each of these basis units is called a unit cell. The unit cells are repeated over and over again in three dimensions to make up a macroscopic crystal.
  10. [10]
    Crystal Structure (Theory) : Solid State Physics Virtual Lab
    Unit cell can be of primitive as well as non primitive type. A primitive cell is a minimum volume unit cell and has only one lattice point in it and the ...<|control11|><|separator|>
  11. [11]
    Axial Ratios, Parameters, Miller Indices - Tulane University
    Sep 7, 2016 · For example, the mineral quartz is hexagonal, with the following unit cell dimensions as determined by x-ray crystallography: a1 = a2 = a3 ...
  12. [12]
    Crystal structure - DoITPoMS
    The lattice type defines the location of the lattice points within the unit cell. The lattice parameters define the size and shape of the unit cell.Missing: authoritative source
  13. [13]
    [PDF] LATTICE GEOMETRY, LATTICE VECTORS, AND RECIPROCAL ...
    Reciprocal vectors are defined to be perpendicular to two of the three lattice vectors and with length equal to 1/length of the third vector. (. ) (. ) (. ) *.
  14. [14]
    [PDF] Lattices, Reciprocal Lattices and Diffraction - Chemistry
    Crystallographers classify these sets of planes using the intercepts on the unit cell axes cut by the plane adjacent to the plane through the origin. • The ...
  15. [15]
    [PDF] Crystal Structure Analysis - Rutgers Physics
    THE RECIPROCAL LATTICE. • the reciprocal lattice is defined in terms of a Bravais lattice. • the reciprocal lattice is itself one of the 14 Bravais lattices.
  16. [16]
    [PDF] Handout 5: Reciprocal Space
    “Reflection from lattice planes” is just a concept that helps us to visualize what is happening in a diffraction experiment.
  17. [17]
    Miller Indices - Crystal Planes in Semiconductors - BYU Cleanroom
    Miller indices are referenced to the crystallographic axes of a crystal. They therefore do not have to be oriented at right angles, though they correspond to ...
  18. [18]
    Miller Indices
    Miller indices provide a description of crystal Directions and crystal Planes in terms of a group of numbers < u v w > or {h k l} contained in brackets with a ...
  19. [19]
    Lattice Planes and Miller Indices (all content) - 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.Missing: paper | Show results with:paper
  20. [20]
    1.3: Surface Structures- fcc Metals - Chemistry LibreTexts
    Feb 21, 2022 · The diagram below shows the conventional birds-eye view of the (111) surface - emphasizing the hexagonal packing of the surface layer atoms.
  21. [21]
    [PDF] Crystallographic Points, Directions, and Planes. - nanoHUB
    Define basic terms and give examples of each: • Points (atomic positions). • Vectors (defines a particular direction - plane normal).
  22. [22]
    Slip in Single Crystals (all content) - DoITPoMS
    When a single crystal is deformed under a tensile stress, it is observed that plastic deformation occurs by slip on well-defined parallel crystal planes.
  23. [23]
    Crystallographic Slip - an overview | ScienceDirect Topics
    Crystallographic slip is defined as the sliding of crystal blocks over one another along a specific slip plane, typically initiating near the yield point ...
  24. [24]
    [PDF] Chapter 3: Crystallographic directions and planes
    Example a b c. Crystallographic planes (continue). In hexagonal unit cells the same idea is used. Example a1 a2 a3 c. 4. Miller-Bravais Indices. (1011). 1 ...
  25. [25]
    3.5.2: Cleavage, Fracture, and Parting - Geosciences LibreTexts
    Dec 16, 2022 · The whole set of planes, collectively referred to as a cleavage, represents planes of weak bonding in the crystal structure. If cleavage ...
  26. [26]
    [PDF] CHAPTER 3: CRYSTAL STRUCTURES & PROPERTIES
    CRYSTALLOGRAPHIC DIRECTIONS. AND PLANES. • In BCC, the close-packed planes are the {110} planes, and the close-packed directions are the <111>. • In FCC, the ...<|control11|><|separator|>
  27. [27]
    None
    ### Extracted Formulas for Interplanar Spacings
  28. [28]
    [PDF] Cohen's method
    HEXAGONAL CRYSTALS. For a hexagonal system, with no errors we have. A(h2 + hk + k2)i + Cl2 i = sin2 θi. (21). Where A = λ2/3a2 o and C = λ2/4c2 o. As in the ...Missing: formula | Show results with:formula<|separator|>
  29. [29]
    X-ray reflection in accordance with Bragg's Law
    Solving Bragg's Equation gives the d-spacing between the crystal lattice planes of atoms that produce the constructive interference. A given unknown crystal is ...
  30. [30]
    [PDF] Back-to-Basics tutorial: X-ray diffraction of thin films
    Oct 13, 2021 · The Bragg Equation relates the angular position of diffracted X-rays to the lattice spacing and underpins all XRD measurements. the goniometer, ...Missing: law | Show results with:law
  31. [31]
    361: Crystallography and Diffraction
    Figure12.2 illustrates the 012 reflection in the h = 0 layer of a cubic unit cell. In this example incident beam I o is represented by the incident wave ...<|separator|>
  32. [32]
    [PDF] Determination of lattice constant in NaCl
    The obtained lattice constant of NaCl is a= 5.6413(6) Ĺ and it is in agreement with the values found in literature: ·a= 5.640(8) Ĺ [1] ·a= 5.642(2) Ĺ [2] ·a= 5 ...<|control11|><|separator|>
  33. [33]
    The Structure of Metals
    In 1850, Auguste Bravais showed that crystals could be classified in terms of 14 unit cells, which meet the following criteria.Missing: historical | Show results with:historical
  34. [34]
    [PDF] Handout 4 Lattices in 1D, 2D, and 3D Bravais Lattice
    The Wigner-Seitz (WS) primitive cell of a Bravais lattice is a special kind of a primitive cell and consists of region in space around a lattice point that ...Missing: non- | Show results with:non-
  35. [35]
    None
    Error: Could not load webpage.<|control11|><|separator|>
  36. [36]
    Section 1.5.2.2. Metric tensors of direct and reciprocal lattices
    2.2. Metric tensors of direct and reciprocal lattices. is determined by the matrix P, i.e. G is not affected by an origin shift p.
  37. [37]
    [PDF] Crystal Systems and Space Groups - McMaster Chemistry
    There are 7 crystals systems and they are named: Triclinic, Monoclinic, Orthorhombic, Tetragonal, Trigonal, Hexagonal, and Cubic.
  38. [38]
    External Symmetry of Crystals, 32 Crystal Classes - Tulane University
    Aug 20, 2013 · The 32 crystal classes represent the 32 possible combinations of symmetry operations. Each crystal class will have crystal faces that uniquely define the ...
  39. [39]
    11 Crystallography – Mineralogy - OpenGeology
    Crystals belong to one of seven crystal systems. • Unit cells of distinct shape and symmetry characterize each crystal system. • Total crystal symmetry depends ...
  40. [40]
    [PDF] Crystal Systems and Crystallographic Point Groups
    There are 7 crystal systems and they are named: Triclinic, Monoclinic, Orthorhombic,. Tetragonal, Trigonal, Hexagonal, and Cubic. What differentiates one ...
  41. [41]
    Phase Transitions
    Often the symmetry change will involve in going from one crystal system to another (e.g. cubic to tetragonal or orthorhombic to monoclinic).
  42. [42]
    16 Symmetry relations at phase transitions - Oxford Academic
    Dec 19, 2024 · Phase transitions are driven by vibrational modes. According to the Landau theory, a continuous phase transition requires that the space group ...<|control11|><|separator|>
  43. [43]
  44. [44]
    Crystallographic point group notations/description: Hermann ...
    Table 1623 lists the notations of the 32 crystallographic point groups. The Schoenflies notation system has nomally been used to describe the symmetry of ...
  45. [45]
    [PDF] Contents - Physics Courses
    5.11 HM and Schoenflies notation for the 32 crystallographic point groups . ... (international) notation for simple crystallographic point groups. The last ...Missing: distribution | Show results with:distribution
  46. [46]
    [PDF] MAT SCI 361: Crystallography and Diffraction
    Jan 7, 2025 · This course covers elementary crystallography and basic diffraction theory with a reciprocal space perspective.
  47. [47]
    Teaching crystallographic and magnetic point group symmetry using ...
    In section 3, we describe how the 32 crystallographic point groups can be represented graphically by means of computer-generated 3-D drawings. Then we introduce ...Missing: across | Show results with:across
  48. [48]
    (International Tables) Home page
    The diagrams and tables of data for the plane groups (two-dimensional space groups), as listed in Volume A, are useful for teaching students about space-group ...
  49. [49]
    Space-group determination | Crystal Structure Analysis
    This chapter describes the factors involved in deciding which of the 230 space groups is correct for a particular crystal structure.Prior knowledge and... · Systematic absences · The statistical distribution of...
  50. [50]
  51. [51]
    5 Close-Packed Structures - International Union of Crystallography
    1. The close-packing of spheres. common close-packed structures which occur in nature are: (i) the hex- agonal close-packing (hcp)with a layer stacking ABAB.. ...
  52. [52]
    11.7 Lattice Structures in Crystalline Solids – Chemistry Fundamentals
    ... metals crystallize in closest-packed arrays with coordination numbers of 12. Metals that crystallize in an HCP structure include Cd, Co, Li, Mg, Na, and Zn ...
  53. [53]
    5 Close-Packed Structures - International Union of Crystallography
    1. The close-packing of spheres. common close-packed structures which occur in nature are: (i) the hex- agonal close-packing (hcp)with a layer stacking ABAB.. ...
  54. [54]
    [PDF] Lecture Outline Crystallography
    Hexagonal unit cell. Covalently bonded layer. (b) Graphite. Buckminsterfullerene (C. 60. ) molecule (the. "buckyball" molecule). The FCC unit cell of the.
  55. [55]
    [PDF] CHAPTER 3: CRYSTAL STRUCTURES & PROPERTIES
    = 1 atoms/2R. Page 5. • APF for a simple cubic structure = 0.52. Atomic Packing Factor (APF). Page 6. a. R. 8. • APF = 0.68: Body Centered Cubic (BCC) Structure.Missing: formula derivation
  56. [56]
    [PDF] Bravais Lattice and Primitive Vectors - Simple, Body-Centered, and ...
    The conventional unit cell is generally chosen to be bigger than the primitive cell and to have the required symmetry. Thus one frequently describes the body- ...
  57. [57]
    Primary Metallic Crystalline Structures - NDE-Ed.org
    The BCC crystal has twice the number of atoms per unitcell than the BCC crystal structure ... The packing factor is 0.74, which is the same as the fcc unit cell.Missing: derivation | Show results with:derivation
  58. [58]
    [PDF] CHAPTER 3: Crystal structures and properties
    • APF for a simple cubic structure = 0.52. APF = a3. 4. 3 π (0.5a)3. 1 atoms unit ... 3.1(a),. Callister 6e. Atomic Packing Factor: FCC. Page 11. 11. • ABCABC ...Missing: derivation | Show results with:derivation
  59. [59]
    Closest Packed Structures - Chemistry LibreTexts
    Jan 29, 2023 · The hexagonal closest packed (hcp) has a coordination number of 12 and contains 6 atoms per unit cell. The face-centered cubic (fcc) has a ...
  60. [60]
    [PDF] Structures of Ceramics
    Coordination numbers and geometries for various cation-anion radius ratios (Rc/Ra). Page 3. 3. AX-type crystal structure. □ Rock salt structure. • Sodium ...
  61. [61]
    Interstitial Site - an overview | ScienceDirect Topics
    In fcc iron (austenite) the largest interstice or 'hole' is at the centre of the unit cell (coordinates 1 2 , 1 2 , 1 2 ) where there is space for an atom of ...
  62. [62]
    Interstitial Sites: Size, Types, Applications, And Calculations
    Crystal Structure, BCC, FCC ; Octahedral Interstitial Site Radius, 0.155r, 0.414r ; Tetrahedral Interstitial Site Radius, 0.291r, 0.225r ; Cubic Interstitial Site ...
  63. [63]
    Tetrahedral and Octahedral Voids - GeeksforGeeks
    Feb 22, 2022 · Thus, for an atom to occupy a tetrahedral void, its radius must be 0.225 times the radius of the sphere. Octahedral void. The octahedral space ...
  64. [64]
    Tetrahedral Sites - Dallas R. Trinkle
    Again, the tetrahedral site has a defined geometry and offers space for an interstitial atom. ... We have (6 · 4)/2 = 12 positions per unit cell. To index · go to ...
  65. [65]
    Distribution of interstitials in fcc iron-carbon austenite: Monte Carlo ...
    Oct 4, 2002 · A study of the distribution of carbon atoms in the octahedral interstitial sites of the face-centered-cubic (fcc) austenite phase in iron-carbon ...
  66. [66]
    [PDF] Ionic point defects - MIT Open Learning Library
    Schottky Defect. -- a paired set of cation and anion vacancies. • Frenkel Defect. -- a cation vacancy-cation interstitial pair. Material. KCl. TiO.
  67. [67]
    Frenkel Defect - Point Defect in Ionic Crystal - Materials Science
    Frenkel Defect is one type of point defect where an atom (better to say cation) leaves its original lattice site & occupies interstitial position.
  68. [68]
    [PDF] Point and Line Defects: Defects in crystals
    0-dimensional: point defects. - vacancy (missing atom). - self-interstitial (extra atom, off lattice site). - impurity, substitutional or interstitial.
  69. [69]
    [PDF] Chapter 4 Point defects and dislocations
    A point defect in a crystal is (i) the occupancy of a lattice sites by impurity atoms/ions or a voids (i.e. vacancy); or (ii) extra atoms/ions not in regular ...
  70. [70]
    n-type Semiconductor | Toshiba Electronic Devices & Storage ...
    An n-type semiconductor is a group IV intrinsic semiconductor such as silicon (Si) doped with group V elements such as phosphorus (P), arsenic (As), or ...
  71. [71]
    Molecular dynamics study of the point defects in bcc uranium
    May 10, 2021 · The equilibrium concentration c of point defects is governed by their formation energy E f and temperature T [11, 12] as c ∝ e − E f / k B T , ...
  72. [72]
    Production of F-centers - MRSEC Education Group
    The F-centers are point defects and can be readily formed in alkaline halides with the aid of a source of ionizing radiation, such as an X-ray source.
  73. [73]
    Insights into oxygen vacancies from high-throughput first-principles ...
    Dec 27, 2021 · We developed codes for high-throughput point-defect calculations and applied them to characterize oxygen vacancies in 937 oxides.
  74. [74]
    The mechanism of plastic deformation of crystals. Part I.—Theoretical
    The mechanism of plastic deformation of crystals. Part I.—Theoretical. Geoffrey Ingram Taylor.
  75. [75]
    Sixty years of dislocations | Bulletin of Materials Science
    It is well-known that with the appearance of three independent papers by Taylor, Polanyi and Orowan in the year 1934, the concept of crystal dislocations was ...
  76. [76]
    Dislocation Density - an overview | ScienceDirect Topics
    In fully annealed material, the dislocation density is about 103–104 mm−2, and in material that has undergone a large plastic deformation at room temperature, ...
  77. [77]
    Effect of Annealing Treatment on Mechanical Properties of ... - NIH
    Feb 13, 2015 · During annealing, the dislocation density decreases, resulting from annihilation of dislocations with opposite Burger's vectors and dislocation ...
  78. [78]
    Energetics and structure of grain boundary triple junctions in graphene
    Jul 6, 2017 · There are six grains, 12 triple junctions, six symmetric grain boundaries (between two rotated grains), and 12 asymmetric grain boundaries ( ...
  79. [79]
    [PDF] Grain Boundary Properties: Energy
    Apr 2, 2014 · Read-Shockley model describes the energy variation with angle for low-angle boundaries successfully in many experimental cases, based on a ...
  80. [80]
    Tilt and Twist Boundaries | MTEX
    In this chapter we discuss the computation of the misorientation axes at subgrain boundaries and discuss whether they vote for twist or tilt boundaries.
  81. [81]
    [PDF] Theoretical and experimental grain boundary energies ... - DiVA portal
    Jun 8, 2023 · For W, Mo, Nb, and Ta, the energies are 0.79, 0.36, 0.44, and 0.56. J∕m2, respectively, at 1500 ◦C, which compare well with the calculated. 𝛴3( ...
  82. [82]
    Hall-Petch Relationship - an overview | ScienceDirect Topics
    The inverse Hall-Petch phenomenon should be responsible for the grain size weakening effects induced by the deformation mechanism of grain boundary sliding.
  83. [83]
    Grain boundaries in polycrystalline materials for energy applications
    Jan 30, 2024 · In this review, we provide an introduction to key concepts and approaches for investigating grain boundaries using these methods.
  84. [84]
    Triple junction excess energy in polycrystalline metals - ScienceDirect
    Oct 15, 2024 · This paper develops a methodology to assess polycrystalline average grain boundary and triple junction excess energies for pure fcc metals Ni, Cu, Al, Pd, Pt, ...
  85. [85]
    Grain Boundary Triple Junction Dynamics: A Continuum ... - SIAM.org
    The microstructure of polycrystalline materials consists of networks of grain boundaries (GBs) and triple junctions (TJs), along which three GBs meet.
  86. [86]
    Max von Laue – Facts - NobelPrize.org
    In 1912, Max von Laue came upon the idea that X-rays passing through crystals might create similar patterns. That is, that a crystal's structure would ...Missing: primary | Show results with:primary
  87. [87]
    The reflection of X-rays by crystals - Journals
    The reflection of X-rays by crystals. William Henry Bragg. Google Scholar · Find this author on PubMed · Search for more papers by this author. and. William ...
  88. [88]
    The phase problem - IUCr Journals
    The paper will canter through direct methods, isomorphous replacement, anomalous scattering and molecular replacement. As phasing is the most acronymic realm ...The Phase Problem · 2. Recovering The Phases · 4. Phase Improvement
  89. [89]
    Electron Diffraction - an overview | ScienceDirect Topics
    Electron diffraction is a technique that allows determination of the crystal structure of materials. When the electron beam is projected onto a specimen, its ...
  90. [90]
    Crystallographic Defect - an overview | ScienceDirect Topics
    A crystallographic defect is defined as an imperfection in a crystal structure, which can include point defects (such as vacancies and interstitials), ...Missing: STM | Show results with:STM
  91. [91]
    Atomic-level defect modulation and characterization methods in 2D ...
    Oct 21, 2021 · Atomic force microscopy (AFM) analyzes surfaces ... HAADF STEM has been used to visualize the crystal structure and defects in TMDCs.
  92. [92]
    Crystal Structure Prediction from First Principles - PubMed Central
    We demonstrate that it is possible to predict the crystal structures of a biomedical molecule using solely first principles calculations.
  93. [93]
    Crystal structure prediction at finite temperatures - Nature
    Oct 21, 2023 · It is based on a previously developed evolutionary crystal structure prediction method USPEX, which predicts the lowest-energy crystal structure ...
  94. [94]
    USPEX—Evolutionary crystal structure prediction - ScienceDirect.com
    Dec 15, 2006 · USPEX can very reliably find the most stable crystal structure of systems with up to several dozen atoms/cell. Due to local optimization and the ...
  95. [95]
    ICSD: Home
    FIZ Karlsruhe provides the scientific and the industrial community with the world's largest database for completely identified inorganic crystal structures, ...ICSD ProductsThe crystal structure depot
  96. [96]
    Materials Project
    The Materials Project provides open web-based access to computed information on known and predicted materials as well as powerful analysis tools.Machine Learning · About · The Materials Project · Dashboard
  97. [97]
    Data-efficient machine learning for molecular crystal structure ...
    In this contribution, we present tailored Δ-ML models that allow screening a wide range of crystal candidates while adequately describing the subtle interplay.
  98. [98]
    Machine Learning Predicts Structures with Quantum Accuracy
    Sep 12, 2024 · DAC Researchers created and applied a machine learning protocol using equivariant neural networks to precisely predict molecular crystal ...<|control11|><|separator|>
  99. [99]
    Control and prediction of the organic solid state: a challenge to ...
    Sep 19, 2018 · It is challenging to calculate the relative thermodynamic stability of different organic crystal structures, let alone understand the kinetic ...
  100. [100]
    Computational prediction of organic crystal structures and ...
    Source: Crystal Growth & Design. Predicting Lattice Energy of Organic Crystals by Density Functional Theory with Empirically Corrected Dispersion Energy.
  101. [101]
    Metadynamics Sim of High-Pressure Silicon Phases w/ NN Potential
    May 5, 2008 · We study in a systematic way the complex sequence of the high-pressure phases of silicon obtained upon compression by combining an accurate high-dimensional ...
  102. [102]
    Polymorphism, what it is and how to identify it: a systematic review
    This review on polymorphism is a personal, non-comprehensive view on the field of polymorphism – a term which is often misused.<|control11|><|separator|>
  103. [103]
    Open questions in organic crystal polymorphism - Nature
    Oct 19, 2020 · In this opinion piece, the authors review the latest developments in the field of polymorphism and discuss standing open questions.
  104. [104]
    Polymorphic phase transitions: Macroscopic theory and molecular ...
    ... (displacive/martensitic, reconstructive and diffusive), effects of defects ... Crystal polymorphism is defined as the ability of a substance, of a ...
  105. [105]
    Polymorphic Solid Solutions in Molecular Crystals: Tips, Tricks, and ...
    Sep 6, 2023 · Crystal polymorphism has been a topic of much interest for the past 20 years or so, especially since its scientific (and legal) importance ...
  106. [106]
    A robust crystal structure prediction method to support small ... - Nature
    Mar 5, 2025 · Crystal polymorphism, the existence of different crystal structures for the same chemical compound, is a common phenomenon in chemistry.
  107. [107]
    Recent Advances in Polymorph Discovery Methods of Organic ...
    Dec 1, 2022 · This paper reviews the recent advances in the polymorph screening methods of organic crystals, mainly including solution crystallization, melt crystallization,
  108. [108]
    [PDF] The DiSCovery of QuaSiCrySTalS - Nobel Prize
    The notebook of Dan Shechtman dating the discovery of quasicrystals to 8 April 1982. (www.quasi.iastate.edu/discovery.html). Page 6. 4 (11). To make the ...
  109. [109]
    Five-fold symmetry in crystalline quasicrystal lattices - PNAS
    An experimental diffraction pattern can be represented as the 5-fold average of a pentile model pattern that best fits the measured width of the Bragg ...
  110. [110]
    A simpler approach to Penrose tiling with implications for ... - Nature
    Aug 1, 1996 · An influential model of quasicrystal structure has been the Penrose tiling4, in which two types of tile are laid down according to 'matching ...
  111. [111]
    Aperiodic approximants bridging quasicrystals and modulated ...
    Jul 11, 2024 · Aperiodic crystals constitute a class of materials that includes incommensurate (IC) modulated structures and quasicrystals (QCs).
  112. [112]
    Natural quasicrystal with decagonal symmetry | Scientific Reports
    Mar 13, 2015 · Icosahedrite, Al63Cu24Fe13, the first natural quasicrystal to be identified, was found in the same meteorite. The new quasicrystal was found ...
  113. [113]
    Self-lubricating, low-friction, wear-resistant Al-based quasicrystalline ...
    Quasicrystals exhibit intrinsically low friction when sliding against hard steel (Figure 1) [3, 4], but unfortunately they are rather brittle at room ...
  114. [114]
  115. [115]
    The mechanical properties of various chemical vapor deposition ...
    Mar 2, 2012 · While the Young's modulus shows relatively little anisotropy with a minimum of 1050 GPa in the [100] direction and a maximum of 1210 GPa in the ...
  116. [116]
    How can graphite and diamond be so different if they are both ...
    Oct 9, 2007 · Diamond's tetrahedral structure makes it hard and strong, while graphite's layered structure makes it soft, slippery, and a conductor of ...
  117. [117]
    Structural Origin of Negative Thermal Expansion in High ...
    Nov 14, 2005 · 1(c) show that the dominant negative thermal expansion is in direction of the c axis, whereas weak negative thermal expansion is observed ...Abstract · Article Text
  118. [118]
    The Voigt‐Reuss‐Hill Approximation and Elastic Moduli of ...
    The Voigt-Reuss-Hill (VRH) approximation converts anisotropic single-crystal elastic constants into isotropic polycrystalline elastic moduli.
  119. [119]
    [PDF] Anisotropic Elasticity
    Mar 15, 2020 · Anisotropic elasticity means elastic response varies with direction, often seen in composites and textured polycrystals, where stiffness varies ...
  120. [120]
    Calcite - Birefringence - HyperPhysics
    A simple demonstration of the large birefringence of calcite is to put a dot on a piece of paper and put the calcite crystal over it. You see two distinct dots.
  121. [121]
    Birefringence – double refraction, uniaxial, biaxial crystals, induced ...
    For example, it is observed for crystalline quartz, calcite, sapphire and ruby. It is also found in various laser crystal materials such as Nd:YVO4, Nd:YLF and ...
  122. [122]
    Principles of Birefringence | Nikon's MicroscopyU
    Crystals are classified as being either isotropic or anisotropic depending upon their optical behavior and whether or not their crystallographic axes are ...
  123. [123]
    Calcite - Smith College
    Max Birefringence, 0.172 ; Extinction, symmetrical to cleavage traces, when crystal is in an extinct position, birefringent calcite dust formed by grinding is ...
  124. [124]
    Pleochroism in Faceted Gems: An Introduction | Gems & Gemology
    Describes the phenomenon of pleochroism and its complex effect on the appearance of colored stones.
  125. [125]
    Tensors in Materials Science (all content) - DoITPoMS
    The tensors in the circles are those that can be applied and measured in any orientation with respect to the crystal (e.g. stress, electric field) and are known ...Missing: lattice | Show results with:lattice
  126. [126]
    Large Anisotropy of the Electrical Conductivity of Graphite - Nature
    GRAPHITE is a hexagonal crystal with a perfect basal cleavage. The carbon atoms in it are arranged in layers parallel to the basal plane, the atoms in each ...
  127. [127]
    Hexagonal boron nitride: From fundamentals to applications
    Apr 15, 2025 · Its layered structure, similar to graphite, grants h-BN similar physical characteristics, earning it the nickname “white graphite”. BN ...Hexagonal Boron Nitride... · Structure Of H-Bn · Applications Of H-Bn
  128. [128]
    Experimental deformation of muscovite - ScienceDirect.com
    The strength and mechanical anisotropy of muscovite have been investigated by shortening single crystals at 45°, 0° and 90° to the basal plane (001)
  129. [129]
    Piezoelectric Materials (all content) - DoITPoMS
    To do this, the piezoelectric atomic structure must be non-centrosymmetric, i.e. there must be no centre of symmetry.
  130. [130]
    Introduction - DoITPoMS
    Cubic crystals are isotropic for many properties, including thermal and electrical conductivity, but crystals with lower symmetry (such as tetragonal or ...