Fact-checked by Grok 2 weeks ago

Lattice plane

In , a lattice plane refers to any plane within a three-dimensional that passes through at least three noncollinear lattice points, forming part of the symmetrical arrangement of atoms or molecules in a solid . These planes intersect in sets of parallel orientations, defining the boundaries of the unit cell and influencing the overall geometry and of the . The orientation and spacing of lattice planes are crucial for understanding crystal properties, as they determine how atoms are organized and interact within the . Lattice planes are systematically described using Miller indices, a notation system denoted as (hkl), where h, k, and l are small integers representing the reciprocals of the intercepts of the plane with the crystallographic axes, scaled to the smallest integers. For example, the (100) plane is parallel to the b- and c-axes and intersects the a-axis at one unit length, while planes like (111) in face-centered cubic lattices represent densely packed atomic layers. This indexing convention allows for precise identification of plane families, denoted by {hkl}, which account for symmetry-equivalent orientations in the crystal. The significance of lattice planes extends to key applications in and physics, particularly in X-ray diffraction, where the interplanar spacing (d) governs the diffraction patterns via : n\lambda = 2d \sin\theta, enabling the determination of crystal structures and atomic arrangements. In crystal growth and , planes that intersect more lattice points tend to form more prominent faces, as per Bravais' law, affecting the external shape and stability of crystals. Furthermore, lattice planes play a role in defect analysis, such as dislocations, and in engineering materials with tailored properties, like semiconductors and metals.

Fundamentals

Definition

In crystallography, a crystal lattice is defined as a regular, repeating array of points in , where each point represents the position of one or more atoms in a crystalline . This lattice arises from the periodic arrangement of atoms, ions, or molecules that forms the basis of a crystal's long-range order. A lattice plane within this lattice is an infinite plane that passes through at least three non-collinear lattice points, thereby intersecting multiple points in the periodic array. Unlike arbitrary planes in space, lattice planes are inherently periodic, with their orientation aligned to the repeating of the , ensuring that the plane repeats the lattice's and contains an infinite series of parallel planes spaced at regular intervals. These planes are fundamental to describing the geometry and properties of crystals, as they reflect the internal atomic arrangement that governs phenomena such as and . Lattice planes are conventionally labeled using , a that specifies their orientation relative to the crystal axes. In a simple cubic , common examples of lattice planes include the (100), (110), and (111) . The (100) is parallel to two of the principal axes (say, y and z), forming a square array of atoms where lattice points lie at the corners of squares with side length equal to the lattice parameter a. The (110) plane cuts diagonally across the cube, intersecting atoms in a rectangular pattern that includes points along the face diagonals./31%3A_Solids_and_Surface_Chemistry/31.02%3A_The_Orientation_of_a_Lattice_Plane_is_Described_by_its_Miller_Indices) Meanwhile, the (111) plane passes through the body diagonal, arranging atoms in a triangular or hexagonal close-packed configuration within the plane, maximizing atomic density among low-index planes in cubic systems. These examples illustrate how lattice planes vary in atomic packing and orientation, influencing the crystal's surface and bulk properties.

Historical Development

The concept of lattice planes in emerged in the early through the foundational work of René Just Haüy, a mineralogist who proposed that crystals are built from repeating geometric units aligned along planar arrangements. In his 1801 treatise Traité de Minéralogie, Haüy described how the external faces of crystals arise from parallel planes of integral molecules, laying the groundwork for understanding internal symmetry and in minerals. This geometric approach shifted from descriptive to a systematic theory of , emphasizing planes as fundamental building blocks. The notation for specifying lattice planes advanced in the mid-19th century, with Carl Friedrich Naumann introducing an early symbolic system in 1826 to denote crystal faces relative to axes, though it was cumbersome for practical use. Building on this and prior efforts by figures like and Auguste Lévy, William Hallowes Miller refined the system in 1839 with his Treatise on Crystallography, proposing the integer-based indices now known as to concisely label planes by their intercepts on crystal axes. Miller's innovation standardized crystallographic description, enabling precise mapping of plane orientations and facilitating comparative studies across mineral species. The marked a pivotal shift with the discovery of X-ray diffraction, where lattice planes were directly linked to atomic-scale structure. In 1912, demonstrated that X-rays diffract off crystal lattices like a three-dimensional , confirming the planar periodicity hypothesized earlier and proving the wave nature of X-rays. Shortly after, and developed the reflection model in 1913, deriving to explain how X-rays constructively interfere at specific angles from parallel lattice planes, revolutionizing structure determination. By the late , the understanding of lattice planes evolved from purely geometric and diffraction-based descriptions to quantum mechanical frameworks in . Felix Bloch's 1928 theorem described electron wavefunctions in crystals as plane waves modulated by the periodic lattice potential, incorporating lattice planes into band theory to explain electronic properties like . This quantum perspective integrated atomic planes with wave mechanics, paving the way for modern applications in semiconductors and materials design.

Mathematical Description

Miller Indices

Miller indices provide a standardized notation for specifying lattice planes in a crystal structure, denoted as (hkl), where h, k, and l are small integers representing the reciprocals of the fractional intercepts of the plane with the crystallographic axes a, b, and c, respectively, reduced to the lowest terms by clearing fractions and using a common multiplier. This system, introduced by William Hallowes Miller in the , allows precise identification of planes without reference to absolute coordinates, facilitating comparisons across different crystal orientations. To determine the Miller indices for a given plane, first identify the intercepts of the plane with the a, b, and c axes, expressed as fractions of the unit cell lengths (Weiss parameters). Take the reciprocals of these intercepts to obtain 1/p, 1/q, 1/r, where p, q, r are the intercept values; if an intercept is at infinity (parallel to an axis), the reciprocal is 0. Multiply through by the least common multiple to yield integers h, k, l in their simplest ratio, ensuring no common divisor greater than 1. For example, a plane intercepting the a-axis at 1, and parallel to b and c (intercepts ∞), has reciprocals 1, 0, 0, yielding (100). Similarly, intercepts at 1/2, 1, ∞ give reciprocals 2, 1, 0, so (210). Negative intercepts result in negative indices. Notation conventions distinguish specific planes, directions, and equivalent sets: parentheses (hkl) denote a specific plane, such as (100); square brackets [hkl] specify a particular , like along the a-; angle brackets indicate a family of equivalent directions related by , e.g., <100> includes all permutations and sign changes; and curly braces {hkl} represent a family of equivalent planes, such as {100} encompassing (100), (010), (001), and their negatives. Negative indices are denoted with an overbar, written as (\bar{h}kl), to indicate the direction opposite to the positive axis. In cubic lattices, the three-index system suffices due to the high , with examples like (111) for the plane cutting all axes at unit length, or (110) for the face-centered plane parallel to c. For hexagonal lattices, the four-index Miller-Bravais notation (hkil) is used to reflect the three-fold basal , where i = -(h + k) to maintain equivalence among the a1, a2, a3 axes, and l corresponds to the c-axis; for instance, the basal plane is (0001), and a prism plane is (10\bar{1}0). This extension ensures consistent labeling in low-symmetry directions. The geometric interpretation links to the : the normal to the (hkl) plane is given by \vec{n} = h \vec{a}^* + k \vec{b}^* + l \vec{c}^*, where \vec{a}^*, \vec{b}^*, \vec{c}^* are the basis vectors. The equation of the plane then takes the form \vec{r} \cdot \vec{n} = 1, with the spacing related to the of \vec{n}. This form underscores how the indices directly correspond to components in reciprocal space, providing a foundation for further crystallographic analysis.

Interplanar Spacing

The interplanar spacing d_{hkl}, also known as the d-spacing, is defined as the perpendicular distance between two adjacent parallel lattice planes indexed by (hkl) in a . This metric is fundamental in for characterizing the geometry of atomic arrangements and is directly applicable in techniques like X-ray diffraction, where it determines the positions of diffraction peaks via . For cubic lattices, the interplanar spacing is derived from the geometry of the unit cell. Consider a cubic lattice with lattice parameter a. The (hkl) plane intersects the axes at distances a/h, a/k, and a/l from the origin. The equation of the plane is \frac{x}{a/h} + \frac{y}{a/k} + \frac{z}{a/l} = 1, or h x / a + k y / a + l z / a = 1. The normal vector to the plane is \vec{n} = (h/a, k/a, l/a), with magnitude |\vec{n}| = \sqrt{h^2 + k^2 + l^2}/a. The perpendicular distance from the origin to the plane (assuming it passes through the origin for the adjacent plane) gives the spacing between parallel planes as d_{hkl} = \frac{a}{\sqrt{h^2 + k^2 + l^2}}. This formula arises because the reciprocal of the spacing is the magnitude of the vector normal to the planes, scaled by the lattice parameter. In orthorhombic lattices, where the axes are perpendicular but of unequal lengths a, b, and c, the formula generalizes to account for the different dimensions: \frac{1}{d_{hkl}^2} = \frac{h^2}{a^2} + \frac{k^2}{b^2} + \frac{l^2}{c^2}. This expression is obtained by extending the cubic derivation, replacing the uniform scaling with axis-specific terms, as the normal vector components are now weighted by the inverse squares of the parameters. For more general crystal systems, such as triclinic lattices with oblique angles, the formula is extended using the reciprocal g^{ij}, where \frac{1}{d_{hkl}^2} = h_i g^{ij} h_j, with h_i = (h, k, l) and g^{ij} incorporating both lengths and angles between axes. The elements are derived from the direct vectors, ensuring the spacing reflects the full of the structure. To illustrate, consider face-centered cubic (FCC) gold with experimental lattice parameter a = 4.08 Å. For the (100) planes, h=1, k=0, l=0, so d_{100} = \frac{4.08}{\sqrt{1^2 + 0^2 + 0^2}} = 4.08 Å. For the (111) planes, h=1, k=1, l=1, so d_{111} = \frac{4.08}{\sqrt{1^2 + 1^2 + 1^2}} = \frac{4.08}{\sqrt{3}} \approx 2.36 Å. These values highlight how higher-index planes have smaller spacings, influencing patterns in materials analysis. The interplanar spacing is intrinsically linked to the , where d_{hkl} = \frac{1}{|\vec{G}_{hkl}|} and \vec{G}_{hkl} is the reciprocal lattice vector corresponding to the (hkl) planes. This relation stems from the nature of the reciprocal lattice, with \vec{G}_{hkl} perpendicular to the planes and its magnitude inversely proportional to the spacing.

Properties

Atomic Density

The planar atomic density, often denoted as \rho_{hkl} for a plane with Miller indices (hkl), is defined as the number of atoms whose centers lie on that crystallographic plane divided by the area of the plane. This measure quantifies the surface packing of atoms and is calculated by determining the equivalent number of atoms contributed to the plane from the unit cell, accounting for sharing among adjacent cells (e.g., corner atoms contribute 1/4, face-center atoms contribute 1/2), and dividing by the corresponding plane area within the unit cell. For cubic lattices, the planar atomic density is given by \rho_{hkl} = \frac{\text{number of atoms per plane}}{\text{area of plane}}, where the plane area is derived from the lattice parameter a. In body-centered cubic (BCC) structures, for example, the (100) plane intersects four corner atoms, each contributing 1/4 to the plane, for a total of 1 atom over an area of a^2, yielding \rho_{100} = \frac{1}{a^2}. This density reflects the sparser packing on such planes compared to more densely populated orientations in the same lattice. In face-centered cubic (FCC) lattices, the (100) plane includes contributions from four corner atoms (1/4 each, totaling 1) and one face-center atom (full contribution), for 2 atoms over an area of a^2, so \rho_{100} = \frac{2}{a^2}. By contrast, the (111) plane, which is close-packed, intersects three corner atoms (1/6 each, totaling 0.5) and three face-center atoms (1/2 each, totaling 1.5), for 2 atoms over an area of \frac{\sqrt{3}}{2} a^2, resulting in the higher density \rho_{111} = \frac{2}{\frac{\sqrt{3}}{2} a^2} = \frac{4}{\sqrt{3} a^2} \approx \frac{2.31}{a^2}. This comparison highlights how the (111) plane achieves approximately 15% greater atomic density than the (100) plane, influencing surface energetics and reactivity. The general method for computing planar density involves identifying atom positions intersected by the using contributions and calculating the plane area from , sometimes incorporating interplanar spacing to relate and surface metrics. Factors such as type significantly affect density; for instance, close-packed planes like {111} in FCC exhibit the highest densities due to maximal coordination, whereas BCC structures generally show lower densities on equivalent indices because of their less efficient packing. Planar atomic density plays a key role in materials science, particularly for surface properties, as higher-density planes like FCC {111} facilitate greater adsorption sites for catalysis and exhibit preferred slip behavior during deformation due to their dense atomic arrangement, impacting ductility and strength.

Orientation and Symmetry

Lattice planes are oriented relative to the crystal axes through their normal vector \vec{n} = (h, k, l), where the Miller indices h, k, and l specify the direction perpendicular to the plane in direct space. This normal vector is parallel to the reciprocal lattice vector \vec{G}_{hkl} = h \vec{a}^* + k \vec{b}^* + l \vec{c}^*, where \vec{a}^*, \vec{b}^*, and \vec{c}^* are the reciprocal basis vectors, confirming that \vec{G}_{hkl} is perpendicular to the (hkl) plane. In cubic systems, this orientation simplifies such that the [hkl] direction is directly normal to the (hkl) plane. The symmetry of lattice planes arises from the point group operations of the , which generate families of equivalent planes denoted by curly braces, such as {hkl}. In cubic crystals, the {100} family comprises six equivalent planes—(100), (010), (001), (\bar{1}00), (0\bar{1}0), and (00\bar{1})—related by rotations and reflections of the cubic . These symmetries ensure that properties like or reactivity are identical for planes within the same family. To visualize plane orientations, pole figures employ stereographic , mapping the normals (poles) of lattice planes onto a two-dimensional from a hemispherical projection, often using equal-area methods to represent distributions accurately. Certain low-index lattice planes exhibit special roles due to their high atomic and weak interlayer bonding. Cleavage planes, typically low-index like {100} or {110}, are those along which crystals fracture preferentially under stress because of reduced bonding strength between planes. In contrast, slip planes, such as {111} in face-centered cubic crystals, facilitate deformation by allowing layers of atoms to glide relative to one another during mechanical loading, with the highest atomic minimizing resistance to shear. The inclination between two lattice planes (h₁k₁l₁) and (h₂k₂l₂) is quantified by the angle \theta, where \cos \theta = \frac{h_1 h_2 + k_1 k_2 + l_1 l_2}{\sqrt{(h_1^2 + k_1^2 + l_1^2)(h_2^2 + k_2^2 + l_2^2)}} This formula, derived from the of their normal vectors, applies directly in cubic systems but requires adjustment via the for lower-symmetry s.

Applications

Diffraction Analysis

Lattice planes play a central role in techniques, where they act as periodic arrays that scatter s or electrons constructively under specific conditions, enabling the determination of crystal structures. In diffraction, incident waves interfere with scattered waves from atoms in these planes, producing diffraction patterns that reveal atomic arrangements. The interplanar spacing d_{hkl} directly influences the scattering . Bragg's law describes the condition for constructive interference from a set of planes: n\lambda = 2 d_{hkl} \sin \theta, where n is an , \lambda is the , d_{hkl} is the spacing between planes with (hkl), and \theta is the angle between the incident beam and the planes. This equation arises because the planes function as reflective gratings for the waves; the path difference between waves scattered from adjacent planes must equal an multiple of the for reinforcement. Derivation follows from considering the extra path length $2 d_{hkl} \sin \theta for the reflected ray, set equal to n\lambda. The provide a formulation of conditions, equivalent to but expressed in : \vec{k}_f - \vec{k}_i = \vec{G}_{hkl}, where \vec{k}_i and \vec{k}_f are the incident and scattered wave (with |\vec{k}_i| = |\vec{k}_f| = 2\pi / \lambda), and \vec{G}_{hkl} = 2\pi (h \vec{a}^* + k \vec{b}^* + l \vec{c}^*) is the corresponding to the (hkl) planes. These three scalar equations (one per ) ensure that occurs only when the scattering matches a , specifying the planes involved./03:_X-rays/3.18:_Laue_equations) In , a polycrystalline sample produces cones of diffracted beams that intersect a detector as rings, with each ring's position corresponding to a specific (hkl) plane via ; indexing assigns to peaks by matching observed $2\theta angles to calculated d_{hkl} values from possible lattice parameters. Single-crystal methods, such as the rotating crystal technique, involve rotating a in a monochromatic to bring different planes into the Bragg sequentially, recording spots on film or a detector to map the . For example, in (NaCl), which has a face-centered cubic structure with lattice parameter a = 5.64 , the (200) planes have spacing d_{200} = a/2 = 2.82 . Using Cu Kα radiation (\lambda = 1.54 ) and n=1, gives \sin \theta = \lambda / (2 d_{200}) = 1.54 / 5.64 \approx 0.273, so \theta \approx 15.8^\circ (or $2\theta \approx 31.6^\circ), corresponding to a prominent peak in the diffraction pattern. Diffraction intensity from (hkl) planes is modulated by the F_{hkl}, which accounts for atomic positions and amplitudes; if F_{hkl} = 0, the is forbidden and absent from the pattern. In face-centered cubic lattices like NaCl or metals such as , reflections like (100) are forbidden because the basis atoms cause destructive (F_{100} = 0), while (200) is allowed (F_{200} = 4(f_\text{Na} + f_\text{Cl})). This systematic absence aids in structure identification.

Materials Science and Engineering

In and , lattice planes play a critical role in texture analysis, which quantifies the preferred of crystallographic planes in polycrystalline materials to understand and predict anisotropic . In metals such as rolled aluminum alloys, texture development leads to non-random distributions of orientations, resulting in directional variations in strength, , and formability. This preferred is commonly measured using pole figures, which map the distribution of plane normals relative to sample coordinates, revealing intensity concentrations that indicate anisotropy; for instance, strong {111} fiber textures in face-centered cubic (FCC) metals enhance deep-drawing by aligning close-packed planes favorably for deformation. Lattice planes also define slip systems, the fundamental units of deformation where dislocations glide on specific planes under applied . In FCC crystals like copper and aluminum, the close-packed {111} planes serve as primary slip planes due to their high atomic , which minimizes the energy barrier for dislocation motion along <110> directions, enabling 12 possible slip systems that accommodate general shape changes without cracking. The selection of these planes is influenced by atomic , as higher planar facilitates easier glide by reducing interatomic resistance. This mechanism governs and yield strength; for example, during tensile deformation, initial slip on {111} planes leads to stage I easy glide, followed by multi-slip interactions that increase . In , lattice planes determine behavior and kinetics, directly impacting and processing outcomes in engineering materials. preferentially occurs along low-index planes with weaker interplanar bonds, such as {110} in body-centered cubic (BCC) intermetallics like NiAl, where is lowest (approximately 4.5 MPa√m) compared to {100} planes (8.3 MPa√m), leading to brittle failure under tensile loads even when stresses are not maximal on those planes. rates vary inversely with plane density; in , KOH etches {110} planes hundreds of times faster than {111} planes due to higher density on the former, enabling anisotropic micromachining for device fabrication. These plane-specific properties influence overall material reliability, as along high-energy planes can deflect cracks and enhance in ceramics like . A key application is in semiconductor engineering, where wafer orientation affects device performance through differences in surface bonding and quality. Silicon (100) wafers are preferred for metal-oxide-semiconductor field-effect transistors (MOSFETs) over (111) orientations because the (100) surface exhibits fewer dangling bonds and lower trap densities at the Si/SiO₂ , reducing instability and significantly improving carrier mobility in n-channel devices. In contrast, (111) surfaces have denser atomic arrangements with more strained bonds, leading to higher defect states that degrade electrical performance in high-frequency applications. In , lattice planes control faceting and stability, allowing tailored shapes for enhanced properties. Nanoparticles of FCC metals like preferentially expose low-energy {111} facets during , as these planes minimize ; for example, branched Au nanostructures grown on graphene oxide scaffolds exhibit dominant {111} facets confirmed by , improving catalytic activity for CO oxidation by exposing more active sites. This faceting strategy extends to other , where thermodynamic control selects planes with the lowest to stabilize high-surface-area structures for applications in and .

Advanced Topics in Solid-State Physics

In , lattice planes in the direct crystal lattice play a fundamental role in reciprocal space, where each family of parallel planes corresponds to a specific point in the . The vector \mathbf{G}_{hkl} is to the (hkl) plane and has a magnitude |\mathbf{G}_{hkl}| = 2\pi / d_{hkl}, with d_{hkl} denoting the interplanar spacing, establishing a direct geometric duality between real-space planes and momentum-space points. This correspondence underpins the construction of s, which are the primitive cells of the defined by the Wigner-Seitz method; the boundaries of the first consist of Bragg planes that bisect the lines connecting a point to its nearest neighbors, to the vectors. These zones delineate the unique range of wavevectors \mathbf{k} for describing electronic and vibrational states in periodic potentials, ensuring that wavefunctions remain Bloch-like within the zone. Lattice plane orientations significantly influence the , particularly through their impact on projections and carrier effective masses. In anisotropic crystals, the curvature of energy bands near the varies with direction relative to specific planes, leading to orientation-dependent effective masses m^* that govern carrier transport; for instance, projections of the onto high-symmetry planes like (100) or (111) in cubic lattices reveal ellipsoidal distortions, altering along those orientations. This arises because the periodic potential modulated by lattice planes folds the free-electron bands at boundaries, modifying the dispersion E(\mathbf{k}) and thus the second derivatives that define m^*_{ij} = \hbar^2 / (\partial^2 E / \partial k_i \partial k_j). Such effects are crucial in materials like semiconductors, where plane-specific symmetries dictate the overall band topology and dynamics. Phonons, as quantized lattice vibrations, are described by plane-wave-like normal modes that propagate across lattice planes, contributing to the phonon dispersion relation \omega(\mathbf{q}). In the harmonic approximation, atomic displacements in a mode with wavevector \mathbf{q} involve coherent motion of entire planes of atoms for propagation along high-symmetry directions, such as in cubic crystals, where planes shift in phase to yield longitudinal or transverse branches. The dispersion curves exhibit folding at Brillouin zone boundaries due to the reciprocal lattice periodicity, with gaps arising from interactions between waves reflected by lattice planes, as captured in the Born-von Kármán model. This plane-wave description across planes enables the calculation of thermal and elastic properties, highlighting how vibrational modes couple to electronic states via electron-phonon interactions. A representative example of plane-specific effects occurs in graphene, a two-dimensional hexagonal lattice, where zigzag and armchair edge terminations—arising from distinct cuts of the basal (0001) plane—profoundly affect electronic conductivity. Zigzag edges host localized edge states near the Fermi level due to flat bands in the projected density of states, leading to suppressed conductance from backscattering and magnetic instabilities, whereas armchair edges gap these states, resulting in semiconducting behavior with higher isotropic transport. These differences stem from the symmetry of the hexagonal lattice planes: zigzag terminations preserve sublattice imbalance, enabling zero-energy states, while armchair edges mix sublattices, opening a band gap proportional to the ribbon width. In topological insulators, specific lattice planes host protected that embody nontrivial bulk topology. For Bi_2Se_3, the (001) plane exhibits a single in the surface , arising from spin-momentum locking and time-reversal symmetry, which confines helical electrons to the surface while the bulk remains insulating. These states are robust against perturbations, enabling dissipationless edge transport, and their linear dispersion E = \hbar v_F |\mathbf{k}| (with Fermi velocity v_F \approx 5 \times 10^5 m/s) directly probes the topological of the material.

References

  1. [1]
  2. [2]
    Lattice Planes and Miller Indices (all content)
    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.
  3. [3]
    Crystal Morphology, Crystal Symmetry, Crystallographic Axes
    Aug 13, 2010 · This arrangement of atoms in crystals is called a lattice. In 2-dimensions a plane lattice consists of an orderly array of points.<|control11|><|separator|>
  4. [4]
    Crystallography III, X-ray Diffraction
    Planes are known as lattice planes if a lattice point is on the plane. So the plane that passes through the origin is a lattice plane because the lattice point ...
  5. [5]
    Crystallography Basics - Chemical Instrumentation Facility
    A unit-cell is the smallest building block of a crystal and is representative unit of the repetative motifs in the crystal structure. Lattice is the geometrical ...
  6. [6]
    Geos 306, Lecture 10, Crystallography II
    Any 3 (not co-linear) lattice points define a lattice plane. All faces of ... When n = 0, then the plane passes through the origin. When n = 1 then ...
  7. [7]
    Examples of lattice planes
    The (100), (010), (001), (100), (010) and (001) planes form the faces of the unit cell. Here, they are shown as the faces of a triclinic (a ≠ b ≠ c, α ≠ β ≠ γ) ...
  8. [8]
    (IUCr) René-Just Haüy and the birth of crystallography
    Dec 4, 2022 · René-Just Haüy, a French abbot living in Paris, at almost 40 years old, began a scientific career focused on the study of crystals.
  9. [9]
    Geos 306, Fall 2012, Lecture 9, Crystallography I
    René-Just Haüy introducted the concept of "molécules intégrantes" (i.e. the modern day unit-cell) in his treatise "Traité de Minéralogie" (1801). He noticed ...<|separator|>
  10. [10]
    [PDF] commented chronology of crystallography and structural chemistry
    Feb 15, 2018 · Mohs, Levy (1825) and Naumann (1826) had previously proposed less practical notations. Miller's indexes are used also to describe lattice planes ...
  11. [11]
  12. [12]
    [PDF] Max von Laue and the discovery of X-ray diffraction in 1912
    Thus, the experiment is evidence for the wave nature of X-rays and the space lattice of crystals at the same time. This is how the discovery entered the ...
  13. [13]
    [PDF] LAUE Laue Back-Reflection of X-Rays - Department of Physics
    In 1912, Max von Laue used X-ray diffraction from crystals to prove that X-rays were very short wavelength electromagnetic radiation, for which he was ...
  14. [14]
    The reflection of X-rays by crystals - Journals
    William Henry Bragg, Proceedings A, 1913. The analysis of crystals by the X-ray spectrometer. William Lawrence Bragg, Proceedings A, 1914. III. On the ...
  15. [15]
    The birth of X-ray crystallography - Nature
    Nov 7, 2012 · Bragg and his son Lawrence worked feverishly on X-ray diffraction all that summer at the University of Leeds, UK, where William was ...
  16. [16]
    [PDF] Bloch's theorem - Rutgers Physics
    A simple justification of the Bloch theorem is given in Introduction to Solid State Physics, by Charles Kittel, seventh edition (Wiley, New York. 1996) in the ...
  17. [17]
    [PDF] Solid State Physics PHYS 40352
    May 22, 2017 · The equation of the lattice plane closest to the origin. (but not passing through it) is Gs · r = 2π. Then, in terms of the xi, this becomes.
  18. [18]
    Quantum Crystallography: the 100-year revolution - PMC - NIH
    Oct 31, 2025 · Evidence for all this comes readily from this anniversary compilation issue, celebrating 100 years of quantum mechanics in crystal structure ...
  19. [19]
    Miller Indices - Crystal Planes in Semiconductors - BYU Cleanroom
    All lattice planes and lattice directions are described by a mathematical description known as a Miller Index. This allows the specification, investigation, ...
  20. [20]
    Axial Ratios, Parameters, Miller Indices - Tulane University
    Sep 7, 2016 · Axial ratios are defined as the relative lengths of the crystallographic axes. They are normally taken as relative to the length of the b crystallographic axis.
  21. [21]
    [PDF] PLANES DIRECTIONS Lattices Crystals - UCSC Physics
    Miller indices are used to specify directions and planes. ❑ These directions and planes could be in lattices or in crystals.
  22. [22]
    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 ...
  23. [23]
    None
    ### Summary of Miller Indices for Hexagonal Lattices and Miller-Bravais Notation
  24. [24]
    [PDF] LATTICE GEOMETRY, LATTICE VECTORS, AND RECIPROCAL ...
    Planes are natural features associated with the crystalline state. When the macroscopic shape of a crystal is considered the bounding surfaces are disturbances ...
  25. [25]
    361: Crystallography and Diffraction
    The distance from the origin to h k l | r → h k l * | = 1 / d h k l where d h k l is the inter-planar spacing. Consider a monoclinic crystal with a = 4 Å , b = ...
  26. [26]
    [PDF] Miller Indices and Interplanar Spacing – 3.091 Introduction to Solid ...
    To find the distance between neighboring parallel planes, we calculate the interplanar spacing, dhkl.Missing: formula | Show results with:formula
  27. [27]
    [PDF] Lecture Outline Crystallography
    Lattice planes are represented by the vector that is normal (perpendicular to them), these are 3D vectors in reciprocal (or dual) space (reciprocal space is ...
  28. [28]
    Prediction of Au lattice constant in SC, FCC and HCP crystal ...
    Feb 27, 2020 · In this post, optimal lattice parameters of gold(Au) are analytically derived using Density Functional Theory(DFT) methods.
  29. [29]
    [PDF] Introduction to Solid State Physics, 8th Edition Charles Kittel
    CHAPTER 2: WAVE DIFFRACTION AND THE RECIPROCAL LATTICE. Diffraction of ... interplanar spacing is. hkA. ˆn. 1. ˆ /h. ⋅n a . But. , whence . (c) For a simple ...
  30. [30]
    [PDF] chapter 3 fundamentals of crystallography problem solutions
    4.57 (a) Derive planar density expressions for FCC (100) and (111) planes in terms of the atomic radius R. (b) Compute and compare planar density values for ...
  31. [31]
    Linear Density and Planar Density - with Solved Example Problems
    The planar density of a crystal is the density of atoms in a crystal plane. This is defined as the number of atoms per unit area on a crystal plane. This ...
  32. [32]
    [PDF] CHAPTER 3: CRYSTAL STRUCTURES & PROPERTIES
    • In FCC, the close-packed planes are the {111} planes, and the close-packed directions are the <110> direction. • In HCP, the close-packed planes are the (0001) ...
  33. [33]
  34. [34]
  35. [35]
    Pole Figure - an overview | ScienceDirect Topics
    The pole figure is defined as a two-dimensional stereographic projection that illustrates the variation of pole density with pole orientation for a selected set ...
  36. [36]
    Crystals and Their Slip Systems - Stanford Advanced Materials
    Jul 24, 2025 · A slip system consists of a combination of a slip plane and a slip direction. The slip plane is the plane with the highest atomic density ...
  37. [37]
    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 ...
  38. [38]
    Bragg's Law
    This satisfies the conditions of Bragg's law for diffraction of the x-rays from the crystal lattice planes.
  39. [39]
    Powder diffraction/Indexing powder patterns - classe
    Indexing the pattern is a process where each peak is identified by the (hkl) Miller-index of the lattice plane that it results from.
  40. [40]
    nglos324 - nacl
    The crystal lattice parameter is 0.563 nm. The diagram shows both a unit cell with ion locations indicated (a) and a space filling model (b) of ionic hard ...
  41. [41]
    [PDF] X-ray Diffraction (XRD)
    To satisfy Bragg's Law, θ must change as d changes. e.g., θ decreases as d increases. λ = 2dhklsinθhkl. Different planes have different spacings. Page 12 ...
  42. [42]
    Visualization of texture components using MTEX - PMC - NIH
    Feb 18, 2020 · Texture and Anisotropy: Preferred Orientations in Polycrystals and Their Effect on Materials Properties. Cambridge University Press. Law ...
  43. [43]
    Slip in Single Crystals (all content)
    Dislocations glide through the material, resulting in bulk plastic deformation. ... slip occurs along close-packed directions on the close-packed planes.
  44. [44]
    Slip Plane - an overview | ScienceDirect Topics
    A slip plane is a specific plane within a crystal lattice where permanent displacement occurs during plastic deformation, typically with the greatest atom ...
  45. [45]
    [PDF] Generalized Reliability Methodology Applied to Brittle Anisotropic ...
    frequently fracture along specific planes defined by the crystal structure, even when the externally applied stresses are not a maximum on the specific planes.
  46. [46]
    [PDF] Printable Single-Crystal Silicon Micro/Nanoscale Ribbons, Platelets ...
    Aug 28, 2007 · difference in etching rate rests in the lower density of atoms and higher density of dangling bonds on the {110} planes than the {111} planes.
  47. [47]
    [PDF] the effect of crystallographic orientation on ductile material removal ...
    For silicon, the fracture toughness is found to vary with the crystal plane orientation. Chen and Leipold give the values of fracture toughness in certain ...
  48. [48]
    [PDF] Transport Enhancement Techniques for Nanoscale MOSFETs
    Unlike silicon, for which (100) wafer orientation is known to provide the best theoretical electron mobility and at the same time the best interface quality ...
  49. [49]
    Branched Au Nanostructures Enriched with a Uniform Facet - Nature
    Jun 11, 2014 · On the other hand, as it is known, among the facets of the fcc crystals, {111} planes exhibit the lowest surface energy. As a result, the as– ...
  50. [50]
    [PDF] Reciprocal Space and Brillouin Zones in Two and Three Dimensions
    the sets of planes that are perpendicular bisectors to the lattice vectors connecting the origin in k-space to its nearest neighbor reciprocal lattice points.
  51. [51]
    [PDF] MORE RECIPROCAL SPACE - Non-secure http index page
    Oct 12, 2011 · •The Brillouin zone is just the Wigner-Seitz cell in reciprocal space. •The boundaries of the Brillouin zone are Bragg planes. •The ...Missing: corresponding | Show results with:corresponding
  52. [52]
    [PDF] arXiv:1405.6898v1 [cond-mat.mtrl-sci] 27 May 2014
    May 27, 2014 · Thus, although the 2D electronic structure (effective masses, orbital order- ing) depends on the surface orientation, the thickness and carrier ...
  53. [53]
    [PDF] Lecture-7-phonons-normal-modes.pdf - IISc Physics
    Valid for high-symmetry planes like [100], [110], [111] of a cubic lattice – for wave propagating along this direction entire planes of atoms move in phase ...
  54. [54]
    [PDF] 1 Phonon I: lattice waves - bingweb
    Feb 18, 2019 · Abstract. A lecture note on the lattice waves in the solid is presented. In a crystal each atom are coupled with the neighboring atoms by ...
  55. [55]
    The electronic properties of graphene | Rev. Mod. Phys.
    Jan 14, 2009 · The most studied edges, zigzag and armchair, have drastically different electronic properties. Zigzag edges can sustain edge (surface) states ...
  56. [56]
  57. [57]
    Landau Quantization of Topological Surface States in | Phys. Rev. Lett.
    Aug 9, 2010 · We report the direct observation of Landau quantization in B i 2 ⁢ S e 3 thin films by using a low-temperature scanning tunneling microscope.Missing: (001) paper