Fact-checked by Grok 2 weeks ago

Electron diffraction

Electron diffraction is a quantum mechanical phenomenon in which a beam of electrons interacts with the atomic lattice of a crystalline material, producing interference patterns that reveal the wave-like nature of electrons, as predicted by Louis de Broglie's hypothesis. This effect was first experimentally confirmed in 1927 by and Lester Germer, who observed diffraction peaks when low-energy electrons (around 54 eV) were scattered from a surface at specific angles, such as 50 degrees, demonstrating constructive interference in accordance with , with independent confirmation by G. P. Thomson using transmission through thin polycrystalline films. Their apparatus involved a with an directing a focused beam onto a rotatable target, with scattered electrons detected using a to measure intensity variations. The underlying principles of electron diffraction stem from the de Broglie relation, which states that the λ of an is λ = h / p, where h is Planck's constant and p is the 's , typically yielding wavelengths on the order of 0.002 to 0.01 for accelerated electrons in common experiments. In practice, electrons accelerated by voltages of 2000–4000 V pass through a polycrystalline target, such as or aluminum with known interplanar spacing d (e.g., 2.34 Å for aluminum), forming concentric rings on a fluorescent screen due to multiple events satisfying the Bragg condition nλ = 2d sinθ, where n is an and θ is the angle. These patterns, observable at distances of about 0.14–0.20 m from the target, directly verify wave-particle duality and allow measurement of λ or d by analyzing ring diameters. Beyond fundamental demonstrations, electron diffraction serves as a cornerstone technique in , particularly through (TEM), where it enables atomic-scale characterization of crystal structures, defects, and phases in materials like alloys and oxides. Key applications include identifying quasicrystals, as in David Shechtman's 1984 discovery of icosahedral phases via selected-area electron diffraction patterns showing forbidden fivefold . Advanced variants, such as convergent beam electron diffraction (CBED), provide and thickness information, while modern aberration-corrected TEMs achieve resolutions down to approximately 0.5 Å (50 pm) at 300 kV, facilitating studies of phase transitions and nanostructures. This method's shorter electron wavelengths compared to X-rays offer superior resolution for thin samples, though it requires high and careful sample preparation to minimize multiple scattering.

Introduction

Wave-particle duality of electrons

The wave-particle duality of electrons embodies the quantum mechanical concept that electrons behave both as discrete particles with definite mass and momentum and as waves with associated wavelength and frequency. This duality was hypothesized by Louis de Broglie, who extended the wave-particle nature observed in electromagnetic radiation to matter particles, proposing that every electron is accompanied by a wave whose wavelength \lambda is given by \lambda = h / p, where h is Planck's constant and p = mv is the electron's momentum, with m the electron mass and v its velocity. The associated frequency \nu satisfies W = h\nu, linking the electron's total energy W to the wave properties. For electrons in typical experimental setups, such as those accelerated by an difference V, the non-relativistic is E = eV = p^2 / (2m), yielding p = \sqrt{2meV} and thus \lambda = h / \sqrt{2meV}. Substituting fundamental constants gives the approximate relation \lambda \approx 1.23 / \sqrt{V} , where V is the accelerating voltage in volts; for example, at V = 100 V, \lambda \approx 0.123 , comparable to interatomic distances in . This charged nature of electrons allows precise control of their de Broglie via electrostatic , unlike particles, enabling wavelengths tunable to match the scale of atomic lattices. Electron diffraction arises from the wave-like of these de Broglie waves, analogous to optical where waves constructively through periodic apertures, but here the waves from ordered arrangements of atoms. Observing clear patterns requires a coherent electron beam, in which the phase relationships among are maintained to produce sharp , and a periodic scattering potential, such as the regular array of atoms in a crystal lattice, to enforce constructive at discrete angles.

Basic diffraction phenomena

Diffraction refers to the patterns arising from the of waves by periodic structures, where scattered waves from regularly spaced scatterers reinforce constructively under specific conditions. In the context of electron diffraction, this occurs when an electron beam interacts with a , producing intensity maxima due to the coherent superposition of waves scattered by atoms arranged in a repeating . A fundamental condition for constructive interference in diffraction from crystal planes is given by , expressed as n\lambda = 2d \sin\theta, where n is an (the of diffraction), \lambda is the of the incident , d is the interplanar spacing, and \theta is the between the incident and the planes. For electrons, which exhibit wave-particle duality with de Broglie wavelengths typically on the order of angstroms or smaller depending on their , this implies small diffraction \theta because \lambda is much shorter than in optical or diffraction, allowing probing of atomic-scale structures. For three-dimensional crystals, the more general diffraction conditions are described by the in vector form: \mathbf{k}' - \mathbf{k} = \mathbf{G}, where \mathbf{k} and \mathbf{k}' are the incident and scattered wave vectors, respectively, and \mathbf{G} is a vector. This equation ensures that the change in of the scattered wave matches the periodicity of the crystal lattice, leading to allowed diffraction spots only when the vector difference intersects the reciprocal lattice points. The Ewald sphere construction provides a geometric visualization of these Laue conditions, representing the incident wave vector as originating from a point on a of radius |\mathbf{k}| centered at the origin of the , with diffraction occurring when the sphere intersects points. This method illustrates the selection rules for observable diffracted beams based on the crystal orientation and electron energy. In electron diffraction experiments, is distinguished as , where the retains its and only changes direction (contributing to sharp Bragg peaks), or inelastic, involving energy transfer to the sample (such as excitations or phonons), which broadens patterns but provides additional structural information. dominates the coherent signals used for , while inelastic processes are often minimized or filtered to enhance resolution.

Applications overview

Electron diffraction serves as a fundamental technique in for determining structures, identifying phases within polycrystalline samples, analyzing defects such as dislocations and vacancies through diffuse patterns, and probing surface reconstructions in crystalline materials. These applications leverage the wave nature of electrons to reveal atomic arrangements at scales inaccessible to many other methods, enabling precise characterization of material properties. In practical examples, electron diffraction is employed to measure parameters in , where it provides insights into size-dependent structural variations in nanoparticles and clusters. For semiconductors, it facilitates the study of thin films by elucidating epitaxial relationships and strain effects at interfaces. Additionally, in gas-phase studies, electron diffraction determines the of volatile compounds, yielding bond lengths and angles for free molecules without the need for . Compared to diffraction, electron diffraction offers higher owing to the shorter de Broglie of electrons (typically around 0.002 nm at 200 keV accelerating voltage versus 0.1 nm for common s), allowing analysis of sub-micrometer regions. It also provides stronger cross-sections, enhancing sensitivity to light elements like and carbon that are weakly detected by s. However, electron diffraction is constrained by radiation damage, which can alter or destroy sensitive samples, particularly biological or organic materials, necessitating low-dose strategies or cryogenic conditions. Furthermore, the technique requires environments to prevent by air molecules, limiting its use to compatible sample types. In modern contexts, electron diffraction integrates with (cryo-EM) techniques like MicroED to resolve biomolecular structures in frozen-hydrated states, advancing for proteins and small molecules. It also supports in-situ studies of dynamic processes, such as phase transformations and reaction kinetics during material synthesis, by capturing real-time structural changes under controlled stimuli.

Historical development

Early electron experiments in vacuum

The discovery of the is credited to J. J. Thomson in 1897, who identified as streams of negatively charged particles emanating from the cathode in low-pressure gas discharge tubes. Thomson's experiments involved deflecting these rays using electric and magnetic fields within modified cathode ray tubes, measuring their charge-to-mass and demonstrating that the particles carried negative electricity independently of the gas used in the tube. This work established the as a fundamental constituent of atoms, with a charge-to-mass approximately 1,800 times larger than that of a , laying the groundwork for controlled beams in environments. Subsequent confirmation of the electron's charge came from Robert A. Millikan's oil-drop experiment in 1909, which precisely measured the elementary as e = 1.602 × 10^{-19} C by balancing the gravitational and electric forces on charged oil droplets in air. Millikan ionized the droplets using X-rays and observed their terminal velocities under varying electric fields, revealing that all charges were integer multiples of this fundamental unit, thus quantifying the as a discrete particle with a fixed charge. This measurement was crucial for verifying electron properties essential to beam manipulation in setups. The manipulation of electrons as particle beams required advancements in vacuum technology during the early 20th century, particularly through the development of vacuum tubes and rudimentary guns. John Ambrose Fleming's invention of the two-electrode vacuum diode in 1904 enabled controlled electron emission from a heated in high vacuum, while Arthur Wehnelt's 1904 work on oxide-coated cathodes improved emission stability for beam generation. These devices, housed in evacuated envelopes, allowed electrons to travel without significant , forming the basis for electron guns that accelerated and focused beams using electrostatic fields. By the 1910s, such tubes were integral to cathode-ray oscilloscopes and early particle experiments, demonstrating electrons' utility as controllable particles. Early electron beam experiments faced significant challenges in partial vacuums, including space charge effects where mutual repulsion among electrons caused beam expansion and divergence, limiting focus and intensity. In less-than-ideal vacuums, residual gas molecules exacerbated , further broadening beams and reducing over short distances. To mitigate these issues and enable electrons to travel s greater than 1 mm without collisions, on the order of 10^{-6} were necessary, as higher pressures shortened the to sub-millimeter scales, rendering beams unusable for precise studies. These technical hurdles underscored the need for improved pumping systems and techniques to achieve the required levels for particle-like electron control.

Quantum mechanics and wave nature confirmation

In 1924, Louis de Broglie proposed the hypothesis that particles, including electrons, possess wave-like properties, extending the wave-particle duality observed in light to matter. He suggested that the \lambda associated with a particle of p is given by \lambda = h / p, where h is Planck's constant, predicting that electrons could exhibit and phenomena similar to light waves. This idea laid the foundation for wave mechanics, implying that electron beams could produce observable diffraction patterns when interacting with crystalline structures. Building on de Broglie's hypothesis, Erwin Schrödinger developed the wave equation in 1926, providing a mathematical framework to describe the behavior of electrons as waves within atomic systems. The time-independent Schrödinger equation, -\frac{\hbar^2}{2m} \nabla^2 \psi + V \psi = E \psi, where \psi is the wave function, V is the potential, E is the energy, m is the electron mass, and \hbar = h / 2\pi, enabled the modeling of electron probability distributions and wave propagation, confirming the wavelike nature of electrons in quantum systems. Concurrently, Werner Heisenberg's uncertainty principle, formulated in 1927, further underscored the wave-particle duality by stating that the position \Delta x and momentum \Delta p of an electron satisfy \Delta x \Delta p \geq \hbar / 2, a direct consequence of the Fourier relationship between a wave's spatial extent and its momentum spread. Experimental confirmation of electron waves came swiftly in 1927 through two independent studies. Clinton Davisson and Lester Germer at Bell Laboratories directed a of electrons onto a surface and observed intensity maxima in the scattered electrons at angles matching the Bragg condition for , with measured wavelengths aligning precisely with de Broglie's \lambda = h / p for electrons accelerated to 54 volts, yielding \lambda \approx 0.165 . In parallel, at the passed electron beams through thin polycrystalline gold foils and recorded rings on photographic plates, demonstrating patterns consistent with wave from atomic planes, again verifying the de Broglie relation for electrons up to 40 keV. These and experiments provided direct evidence of electron wave , solidifying the quantum mechanical description of matter. The groundbreaking discoveries by Davisson, Germer, and Thomson earned them the 1937 , awarded jointly to Davisson and Thomson "for their experimental discovery of the diffraction of s by crystals," marking a pivotal validation of quantum wave mechanics.

Development of electron microscopy and initial diffraction

In 1931, and Max Knoll constructed the first prototype of a transmission at the , utilizing two magnetic es to achieve a of approximately 400 times, surpassing the capabilities of contemporary microscopes for certain tasks. This device demonstrated the feasibility of for forming enlarged images, building on earlier demonstrations of the wave nature of s, though its resolution was still limited by lens aberrations and electron beam instability. By 1932, Ruska, in collaboration with , refined the design into a more practical transmission electron microscope (TEM) incorporating optimized magnetic lenses with pole pieces to better focus the beam, enabling higher magnifications up to several thousand times and improved image contrast through effects in thin specimens. These advancements laid the groundwork for as a tool for visualizing structures at scales unattainable by optical methods, with the fundamentally governed by the Abbe , approximately \lambda / (2 \mathrm{NA}), where \lambda is the (on the of picometers at typical accelerating voltages) and NA is the of the objective lens, allowing potential atomic-scale imaging far beyond light 's half-micrometer . The integration of diffraction capabilities in these early TEMs advanced rapidly; in 1937, Hans Boersch reported the first electron diffraction patterns obtained directly within a transmission electron microscope, using polycrystalline metal films as specimens to produce ring patterns that confirmed the crystalline structure and orientation of the materials. These patterns highlighted the microscope's dual role in imaging and , with the selected area enabling localized diffraction from micrometer-scale regions. By the 1940s, these innovations enabled initial applications in biological and materials sciences, including the visualization of viruses such as poxviruses in 1938 and bacteriophages in 1940, which confirmed their particulate nature and sub-micrometer dimensions. In materials analysis, diffraction patterns from thin crystal films allowed early insights into parameters and defects, paving the way for structural studies of metals and biological tissues.

Post-1930s advancements in theory and instrumentation

Following the foundational dynamical theory established by in 1928, which accounted for multiple effects in electron diffraction by , post-1930s theoretical advancements focused on computational methods to simulate these complex interactions more accurately. In the , J.M. Cowley and A.F. Moodie introduced the multislice algorithm, a numerical approach that divides the crystal into thin slices to iteratively compute wave propagation, enabling simulations of dynamical for thicker specimens and reducing computational demands compared to earlier methods. This was extended in 1965 by J. Gjønnes and A.F. Moodie, who derived extinction rules for dynamical effects, such as Gjønnes-Moodie lines in convergent-beam electron diffraction patterns, providing analytical insights into symmetry-related intensity variations and aiding pattern interpretation. Instrumentation progressed significantly in the mid-20th century to enhance beam quality and data capture. In the 1960s, field-emission guns (FEGs), pioneered by , replaced thermionic sources in transmission electron microscopes, delivering brighter, more coherent electron beams with reduced energy spread, which improved signal-to-noise ratios in diffraction experiments and enabled higher-resolution studies of crystal structures. By the 1980s, (CCD) detectors were integrated into transmission electron microscopes, allowing digital recording of diffraction patterns with quantitative intensity measurements, supplanting and facilitating automated data processing for large-scale analyses.90187-0) The 1990s saw the introduction of precession electron diffraction (PED) by R. Vincent and P.A. Midgley, a technique that rocks the incident beam in a hollow cone to average dynamical scattering effects, yielding quasi-kinematical patterns for more reliable structure determination in nanocrystals.90154-9) This method reduced multiple scattering artifacts, improving the accuracy of lattice parameter refinement and phase identification. In the 2010s, four-dimensional (4D-STEM) emerged, employing pixelated detectors to record full patterns at each scan position, providing spatially resolved crystallographic information and enabling ptychographic reconstructions with sub-angstrom precision. Recent advancements in the 2020s have incorporated for automated analysis, particularly algorithms for pattern indexing and phase mapping. Convolutional neural networks, trained on simulated and experimental datasets, classify crystal structures and orientations directly from patterns, accelerating phase identification in multiphase materials and reducing manual interpretation time by orders of magnitude. These AI-driven tools, such as those for selective area electron diffraction, achieve over 95% accuracy in indexing complex patterns, enhancing applications in materials discovery.

Theoretical principles

Plane waves, wavevectors, and reciprocal lattice

In electron diffraction, the wave nature of electrons is described by , where the wave function for a propagating in direction \hat{n} is given by \psi(\mathbf{r}) = A \exp(i \mathbf{k} \cdot \mathbf{r}), with the wavevector \mathbf{k} = \frac{2\pi}{\lambda} \hat{n} and de Broglie wavelength \lambda = \frac{h}{p} relating the electron's p to Planck's constant h. This representation arises from the de Broglie hypothesis, which posits that particles exhibit -like properties, and was experimentally confirmed through patterns from crystalline targets. The interaction of these plane waves with a crystal requires consideration of the periodic atomic arrangement, leading to the concept of the reciprocal lattice. The reciprocal lattice is defined by basis vectors \mathbf{a}^* = 2\pi \frac{\mathbf{b} \times \mathbf{c}}{\mathbf{a} \cdot (\mathbf{b} \times \mathbf{c})}, and cyclically for \mathbf{b}^* and \mathbf{c}^*, such that the reciprocal lattice vectors are \mathbf{G}_{hkl} = 2\pi (h \mathbf{a}^* + k \mathbf{b}^* + l \mathbf{c}^*) for integers h, k, l. These vectors correspond to the Fourier components of the crystal's periodic structure and satisfy \exp(i \mathbf{G} \cdot \mathbf{R}) = 1 for any direct lattice vector \mathbf{R}, ensuring the periodicity of plane waves scattered by the lattice. Diffraction conditions are geometrically interpreted using the first , which is the Wigner-Seitz primitive cell in centered at the and bounded by planes to the vectors at their midpoints. This zone delineates regions where electron wavevectors \mathbf{k} lead to Bragg reflections upon reaching zone boundaries, satisfying the Laue condition \mathbf{k}' = \mathbf{k} + \mathbf{G} with |\mathbf{k}'| = |\mathbf{k}|. The Ewald construction extends this framework to electrons by constructing a sphere in reciprocal space with radius |\mathbf{k}| = \frac{2\pi}{\lambda}, centered such that the incident wavevector terminates at the origin of the . Diffraction occurs when a point \mathbf{G} lies on this , corresponding to a diffracted wavevector \mathbf{k} + \mathbf{G}; for electrons, the short de Broglie (typically 0.001–0.004 at accelerating voltages of 100–1000 kV) results in a nearly flat Ewald over small angular ranges, facilitating pattern analysis in . The relates directly to the real-space crystal via the of the periodic potential V(\mathbf{r}), expressed as V(\mathbf{r}) = \sum_{\mathbf{G}} V_{\mathbf{G}} \exp(i \mathbf{G} \cdot \mathbf{r}), where V_{\mathbf{G}} are the Fourier coefficients encoding atomic scattering factors. This decomposition underpins for electron states in crystals, \psi(\mathbf{r}) = u(\mathbf{r}) \exp(i \mathbf{k} \cdot \mathbf{r}) with u(\mathbf{r}) periodic, and highlights how the periodic potential modulates free-electron plane waves to produce the observed .

Kinematical diffraction theory

The kinematical diffraction theory, also known as the single-scattering approximation, describes under conditions where multiple scattering events are negligible, treating the crystal as a weak to the incident . This approximation is grounded in the first of quantum scattering theory, where the scattered wave amplitude for a reflection indexed by the vector \mathbf{G} is given by f_{\mathbf{G}} = -\frac{m}{2\pi \hbar^2} \int V(\mathbf{r}) \exp\left[i (\mathbf{k} - \mathbf{k}_0 - \mathbf{G}) \cdot \mathbf{r}\right] d\mathbf{r}, with m the electron mass, \hbar the reduced Planck's constant, V(\mathbf{r}) the crystal potential, \mathbf{k}_0 the incident wavevector, and \mathbf{k} the scattered wavevector (with |\mathbf{k}| = |\mathbf{k}_0| for elastic scattering). This integral represents the Fourier transform of the potential at the scattering vector \mathbf{s} = \mathbf{k} - \mathbf{k}_0 - \mathbf{G}, which vanishes at the exact Bragg condition (\mathbf{s} = 0). The theory assumes the incident wave is unattenuated and ignores back-scattering or higher-order interactions, making it analogous to the kinematic theory in X-ray diffraction but adapted for the stronger electron-matter interaction. For crystalline materials, the scattered amplitude simplifies through the crystal's periodicity, leading to the F_{\mathbf{G}}, which encapsulates the atomic arrangement within the unit cell: F_{\mathbf{G}} = \sum_j f_j \exp\left[2\pi i (h x_j + k y_j + l z_j)\right], where the sum is over all atoms j in the unit cell, f_j is the atomic factor (dependent on the angle and electron energy), and (h, k, l) are the defining \mathbf{G}. The atomic factor f_j itself arises from the of the around atom j, approximated as f_j \approx Z_j () for small angles but decreasing with angle due to the finite size of the atomic potential. The diffracted intensity for the \mathbf{G} reflection is then I_{\mathbf{G}} \propto |F_{\mathbf{G}}|^2, modulated by s specific to electrons; the accounts for the time the crystal planes are in the condition during rotation (for or precessing beams). The kinematical approximation holds under specific conditions: the crystal must act as a weak object, where the mean inner potential (typically 5–20 V) is much smaller than the kinetic (often >100 keV), ensuring the phase shift per atom is small (< \pi); additionally, the specimen thickness t must be less than the extinction distance \xi_g (typically 50–500 nm, depending on material and reflection), beyond which multiple scattering dominates and dynamical effects emerge. These conditions are often met in very thin samples (<10 nm) or light-element materials, allowing direct structure factor determination from intensities. A key prediction is the absence of intensity for reflections where F_{\mathbf{G}} = 0, known as systematically absent or forbidden reflections; for example, in the diamond cubic structure (space group Fd\bar{3}m), reflections like (200) have F_{200} = 0 due to the symmetric placement of carbon atoms at tetrahedral sites, resulting in zero kinematical intensity, though weak dynamical contributions may appear in thicker samples.

Dynamical diffraction and multiple scattering

Dynamical diffraction theory accounts for the strong scattering of electrons in crystalline materials, where the single-scattering approximation of kinematical theory breaks down due to multiple interactions between the electron wave and the periodic lattice potential. This regime is prevalent in electron diffraction because the mean free path for elastic scattering is on the order of tens of nanometers, much shorter than for or neutrons, leading to significant deviations such as anomalous intensities and beam coupling. The theory solves the time-independent for the electron wave function ψ(r) in a periodic crystal potential V(r): i \frac{\partial \psi}{\partial z} = \left[ -\frac{1}{2k_z} \nabla_\perp^2 + V(r) \right] \psi, where z is the beam direction, k_z the z-component of the wavevector, and ∇_⊥ the transverse gradient. The solutions take the form of Bloch waves, ψ = u(r) exp(i k · r), where u(r) is a periodic function with the lattice periodicity, reflecting the wave's modulation by the crystal lattice. These waves propagate as superpositions of plane waves with wavevectors deviated from the incident direction, enabling the description of multiple scattering paths. In the two-beam approximation, which simplifies the full multi-beam dynamical theory by considering only the incident beam and one strongly excited diffracted beam, the electron wave is expressed as a linear combination of these two plane waves. The dispersion surface, a graphical representation in k-space, illustrates the allowed wavevectors as two hyperbolic branches separated near the Bragg condition, with the splitting governed by the excitation error s_g, defined as the deviation from the exact Bragg angle. The parameter ξ_g characterizing this coupling is given by ξ_g = V_g / (2 k (cos γ - cos θ)), where V_g is the Fourier component of the lattice potential for the diffracted beam g, k the incident wavevector magnitude, γ the angle between the beam and the surface normal, and θ the Bragg angle. This approximation captures essential effects like beam interference while being computationally tractable for thin crystals. Key observables in two-beam dynamical diffraction include rocking curves, which plot diffracted intensity as a function of incidence angle, showing oscillatory behavior due to the excitation of different Bloch wave branches. Similarly, Pendellösung fringes arise from the periodic energy transfer between the coupled beams as the electron propagates through the crystal thickness, manifesting as thickness-dependent intensity oscillations in both transmitted and diffracted beams. The characteristic length scale for these oscillations is the extinction distance ξ = 2π / |C V_g|, where C is a relativistic correction factor approximately equal to 1 for non-relativistic electrons but adjusted for high voltages; for typical materials and 100-300 kV electrons, ξ ranges from 10 to 100 nm, reflecting the strong scattering potential. For more complex scenarios involving many beams or thick specimens, the multislice algorithm provides an efficient numerical method to simulate dynamical diffraction by dividing the crystal into thin slices and iteratively applying phase shifts and propagation operators. Originally formulated in the 1950s, its practical computational implementation advanced in the 1970s with the rise of digital computing, enabling accurate modeling of multiple scattering in realistic crystal structures.

Kikuchi lines and pattern formation

Kikuchi lines in electron diffraction patterns emerge as a hallmark of dynamical scattering processes, where multiple scattering events within the crystal lead to characteristic intensity modulations beyond simple kinematical approximations. Within the dynamical diffraction framework, which accounts for Bloch wave propagation, these lines result from the interplay of elastic and inelastic scattering, producing observable signatures in both transmission and reflection geometries. The origin of Kikuchi lines lies in an initial incoherent scattering event—often inelastic, such as thermal excitation—that generates a nearly isotropic distribution of diffuse electrons within the crystal volume. These diffuse electrons then undergo subsequent coherent Bragg diffraction when their trajectories satisfy the Bragg condition relative to lattice planes, creating cones of scattered intensity. The intersection of these cones with the Ewald sphere in reciprocal space produces the Kikuchi pattern, manifesting as excess lines (regions of enhanced intensity) and deficiency lines (regions of reduced intensity) on the diffraction screen. Kikuchi lines typically appear in pairs, with each pair aligned parallel to the projection of specific crystallographic (hkl) in the diffraction pattern. The separation between the two lines in a pair corresponds to twice the Bragg angle, hkl, reflecting the angular deviation required for diffraction from both sides of the . This geometric pairing allows direct indexing of the lines to lattice , facilitating pattern interpretation. An notable feature of Kikuchi line pairs is their intensity asymmetry, where the excess and deficiency lines exhibit unequal brightness due to dynamical absorption effects. Absorption preferentially attenuates certain Bloch wave components, altering the relative contributions from forward- and back-scattered electrons and thus modulating line contrast based on crystal orientation and thickness. In thicker crystals, where multiple scattering is pronounced, Kikuchi pattern formation is dominated by thermal diffuse scattering (TDS) mechanisms. TDS arises from phonon-induced atomic displacements, producing broad cones of diffuse intensity centered on reciprocal lattice points; these large-diameter cones (~180° aperture) intersect the Ewald sphere nearly tangentially, resulting in straight-line appearances rather than curved arcs in the observed pattern. This TDS contribution enhances the visibility of Kikuchi bands in convergent beam electron diffraction (CBED) setups, providing a robust signature for dynamical effects. Kikuchi patterns enable precise applications in , such as orientation mapping, where line intersections identify zone axes and by comparing measured interplanar angles to simulated projections. Additionally, they support determination through the analysis of line pairings, widths, and systematic absences, revealing characteristics without reliance on spot positions alone.

Experimental techniques in transmission electron microscopy

Diffraction pattern generation and selected area methods

In (TEM), diffraction patterns are generated by directing a parallel beam of electrons through a thin sample, where the electrons interact with the atomic lattice and scatter at specific angles according to the Bragg condition. The diffracted electrons are then focused by the objective lens onto its back focal plane, forming a diffraction pattern that represents the of the sample in two dimensions. This ray diagram involves the incident electron beam illuminating the sample uniformly, the sample acting as a , and the back focal plane serving as the imaging plane for the pattern, which can be projected onto a fluorescent screen or detector for observation. Selected area electron diffraction (SAED) enables the acquisition of patterns from localized regions of the sample by inserting a small aperture in the of the objective , which selects electrons originating from a specific area typically ranging from 100 to 500 in diameter. This technique isolates the from the desired region while excluding contributions from surrounding areas, allowing for high analysis of within heterogeneous materials. In SAED patterns from crystalline samples, the central beam corresponds to the undiffracted electrons, surrounded by discrete spots arising from kinematical at Bragg angles, while amorphous regions produce a broad diffuse halo due to short-range atomic order. Defects in the crystal , such as twins or stacking faults, can manifest as spot splitting in SAED patterns, where individual spots divide into pairs or multiples due to overlapping domains with slightly misoriented lattices along the . Accurate measurement of inter-spot distances requires of the camera length L, defined by the relation L = R / \tan \theta, where R is the radial distance from the pattern center to a spot on the recording screen and \theta is the corresponding ; this is typically achieved using a standard sample like polycrystalline with known spacing. Since the , the adoption of pixelated direct detectors has enhanced pattern imaging by enabling high-frame-rate, low-noise recordings that capture subtle features like spot splitting with greater fidelity than traditional screens.

Polycrystalline and composite material analysis

In transmission electron microscopy (TEM), selected area electron diffraction (SAED) applied to polycrystalline specimens produces concentric Debye-Scherrer rings due to the random orientations of numerous small crystallites within the selected area. These rings arise from Bragg reflections averaged over many grain orientations, analogous to powder X-ray diffraction patterns but achieved with a parallel electron beam illuminating a thin foil specimen. The radius r of each ring in the diffraction pattern corresponds to the scattering angle and is given by r = L \tan 2\theta, where L is the effective camera length and \theta is the Bragg angle; for small angles typical in TEM, this approximates to r \approx L \lambda / d, with \lambda the electron wavelength and d the interplanar spacing. Indexing of the rings involves measuring r and computing d = \lambda L / r, then matching these spacings to known crystal structures for phase identification. Preferred orientation, or , in polycrystalline materials manifests as variations in ring intensity or the formation of arcs rather than complete circles, reflecting non-random alignment of crystallites. For instance, in thin films or deformed metals, fiber textures can cause enhanced intensity along specific azimuthal directions, providing insights into processing history and mechanical properties. In composite materials and multilayers, double diffraction occurs when electrons undergo sequential scattering events, such as initial diffraction in one layer followed by further diffraction in an adjacent layer, producing extraneous spots or superlattice-like satellites superimposed on the primary pattern. These artifacts are particularly evident in epitaxial multilayers or phase-separated composites, where the extra reflections can mimic structural ordering but are distinguished by their sensitivity to specimen tilt or thickness. For example, in semiconductor superlattices, such satellites arise from the periodic stacking, aiding analysis of layer periodicity and interface quality. Analysis of polycrystalline and composite diffraction patterns often involves radial of ring intensities to generate one-dimensional profiles comparable to data, enabling quantitative phase fraction determination and texture quantification via peak asymmetry or March-Dollase modeling. In ensembles, ring broadening due to finite size is quantified using the , \Delta \theta = K \lambda / (D \cos \theta), where \Delta \theta is the full width at half maximum, K is a (typically 0.9), and D is the average ; this method has been applied to estimate sizes in metal oxide as small as 5 nm from TEM patterns.

Advanced methods: CBED, PED, and 4D-STEM

Convergent beam electron diffraction (CBED) employs a finely focused probe to illuminate a small crystalline region, typically on the order of nanometers, producing diffraction patterns consisting of overlapping disks rather than discrete spots. These disk patterns arise from the convergence angle of the beam, which excites a range of incident directions, and they reveal detailed intensity distributions within each disk due to dynamical effects. A key feature of CBED patterns is the presence of higher-order Laue zones (HOLZ), which appear as concentric rings of fine lines or rings outside the central zero-order Laue zone (ZOLZ), providing information on parameters and along the beam direction. HOLZ lines intersect the bright-field disk and exhibit deficiency or excess contrasts that are sensitive to crystal orientation and thickness, enabling precise measurements of and tilt with sub-degree accuracy. CBED excels in local symmetry analysis, where the overall pattern , including both ZOLZ and HOLZ, determines the crystal's . The presence of —manifested as inversion across the pattern center—distinguishes centrosymmetric s (e.g., those with an inversion center) from non-centrosymmetric ones, as non-centrosymmetric groups lack this mirror-like inversion in intensity distributions. This analysis classifies the pattern into one of 10 two-dimensional s based on rotational, mirror, and glide symmetries observed in the disks, often confirmed by examining fine details like Kikuchi lines within HOLZ rings. Such has been applied to thin films, such as cadmium arsenide, to verify assignments like \bar{4}2m. Dynamical effects, while complicating intensity quantification, enhance the visibility of these elements compared to kinematical approximations. Precession electron (PED) addresses limitations of standard CBED by rocking the incident beam in a hollow-cone trajectory around the optic axis, effectively averaging over multiple orientations to mitigate dynamical . This reduces intensity transfer between diffracted beams, yielding patterns that more closely approximate kinematical conditions and improving the reliability of measurements. PED patterns exhibit enhanced spot intensities for weak reflections while suppressing multiple artifacts, making it particularly useful for phase identification in nanocrystalline materials. The technique, pioneered in the and refined through multislice simulations, integrates well with scanning modes for orientation mapping at the nanoscale. Four-dimensional scanning transmission electron microscopy (4D-STEM) extends these methods by combining a scanning focused probe with a pixelated detector to record a full two-dimensional pattern at each spatial , generating a four-dimensional of and . This allows simultaneous mapping of local orientation, , and variations across a sample, with patterns providing momentum-resolved information per probe , typically at resolutions down to scales. In the , 4D-STEM has enabled advanced ptychographic reconstructions, where iterative algorithms retrieve the sample's complex transmission function—including shifts—from overlapping data, achieving sub-angstrom resolution for beam-sensitive materials like biomolecules. These techniques, such as single-sideband or difference map methods, surpass traditional imaging by quantifying electromagnetic fields and lattice distortions without prior models.

Scattering from aperiodic, diffuse, and superstructure features

In (TEM), electron diffraction patterns from aperiodic, diffuse, and features provide insights into deviations from ideal periodic lattices, such as modulations, disorders, and defects in crystalline materials. These scattering phenomena arise when electrons interact with structural irregularities that lack long-range , leading to additional reflections or intensity distributions beyond standard Bragg peaks. Unlike kinematical approximations for perfect crystals, these patterns require consideration of dynamical effects and local atomic displacements to interpret the observed intensities accurately. Superstructures in materials manifest as satellite reflections flanking main Bragg peaks in electron diffraction patterns, originating from periodic modulations like waves (s). For instance, in layered dichalcogenides such as 1T-TaSe₂, satellite spots appear at positions corresponding to the wavevector, with intensities modulated by the degree of distortion and beam orientation. These satellites indicate a period that is incommensurate with the underlying crystal , enabling mapping of the amplitude and through pattern analysis. In systems, the satellite reflections arise from the coherent superposition of scattered waves from density modulations, often observed at low temperatures where the phase stabilizes. Aperiodic crystals, exemplified by quasicrystals, produce diffraction patterns with forbidden rotational symmetries, such as five- or ten-fold axes, due to their quasiperiodic arrangements following sequences like the chain. The seminal observation in an Al-14 at.% Mn revealed sharp diffraction spots arranged in icosahedral with fivefold rotational axes, inconsistent with any but exhibiting long-range orientational order without translational periodicity. These patterns feature dense, non-repeating spot arrays that reflect the of positions, allowing identification of quasicrystalline phases through analysis and comparison with simulated from Penrose or tilings. Diffuse scattering encompasses Huang scattering from dilute defects and thermal diffuse scattering (TDS) streaks, both contributing to intensity distributions away from Bragg positions. Huang scattering, prominent near reciprocal lattice points, stems from long-range elastic strain fields around point defects like vacancies or interstitials, as demonstrated in electron-irradiated aluminum where asymmetric intensity lobes reveal defect type and concentration. TDS arises from phonon-induced atomic vibrations, producing streaky features along reciprocal space directions that trace phonon dispersion relations; for example, first- and second-order TDS in simple metals yield diffuse streaks with intensity proportional to the phonon density of states. These phenomena often overlay Kikuchi patterns, requiring background subtraction for accurate quantification. Analysis of these features frequently employs pair distribution functions (PDFs) derived from modulated intensities to probe local atomic . Electron pair distribution function (ePDF) analysis transforms radial intensity profiles from patterns into real-space pair functions, revealing short-range in aperiodic or defective structures without assuming periodicity. For modulated systems, ePDF extracts parameters from or diffuse intensities, enabling refinement of models or defect configurations in . This method is particularly effective for nanoscale samples, where it distinguishes aperiodic motifs from random disorder. Representative examples include phonon dispersions mapped from TDS streaks, where intensity variations along streaks in high-energy electron diffraction correlate with acoustic phonon branches in materials like , providing temperature-dependent dynamics. Similarly, dislocation fields generate Huang-like diffuse tails in patterns, with intensity asymmetry indicating screw or edge character; in high-angle annular dark-field TEM, this enhances from core distortions, facilitating field mapping in deformed crystals. These applications underscore the role of diffuse and modulated in elucidating dynamic and defective structures at scales.

Surface and reflection-based techniques

Low-energy electron diffraction (LEED)

Low-energy electron diffraction (LEED) is a surface-sensitive technique used to determine the atomic structure of crystalline surfaces by analyzing the diffraction patterns formed by low-energy electrons (typically 20-200 eV) incident normally on the sample. The experimental setup consists of a collimated, monoenergetic electron gun that produces a beam with currents around 1 μA and a diameter of 0.5-1 mm, directed at the sample surface. Backscattered electrons are then filtered for elastic scattering using a retarding field analyzer, typically comprising four concentric hemispherical grids to suppress inelastically scattered electrons, before impinging on a fluorescent screen biased at 5-6 kV to visualize the diffraction pattern. The resulting diffraction patterns arise from the interference of electrons scattered by the surface lattice, governed by the Laue equations in reciprocal space, with large scattering angles θ due to the short de Broglie wavelength λ ≈ 0.1 nm at these energies. This short penetration depth, limited by the inelastic mean free path of approximately 5-10 Å, ensures high surface sensitivity, probing primarily the topmost atomic layers. Spot positions and symmetries in the pattern reveal the two-dimensional surface unit cell, while variations in spot shape or splitting indicate reconstructions or adsorbate ordering. Quantitative structural analysis in LEED relies on measuring intensity-voltage (I-V) curves, where the intensity of individual diffraction spots is recorded as a function of incident energy, often over ranges of 100-300 for optimal overlap in structural refinement. These curves are compared to simulated spectra using dynamical models to extract structure factors, achieving precisions of ±1-2 pm for metal surfaces. At low energies, multiple scattering events dominate due to strong electron-atom interactions, necessitating full quantum mechanical treatments rather than kinematical approximations, as briefly outlined in dynamical theory principles. To minimize beam-induced damage to sensitive surfaces, modern systems employ low currents (down to 1 nA) and microchannel plate enhancements for brighter patterns without increasing exposure. A key application of LEED is in characterizing adsorbate-induced surface reconstructions, such as the 2×1 dimer reconstruction on Si(100), where LEED patterns from the clean surface exhibit split spots indicative of the buckled silicon dimer array, often modified by adsorbates like or .

Reflection high-energy electron diffraction (RHEED)

Reflection high-energy electron diffraction (RHEED) employs a of electrons with energies typically ranging from 10 to 50 keV, directed at a grazing incidence of 1° to 5° onto the sample surface. This configuration confines the electron penetration depth to a few atomic layers, enhancing surface sensitivity, while the diffracted beams are projected onto a fluorescent screen or CCD detector to form patterns that reveal the two-dimensional surface . The technique is particularly suited for in-situ monitoring during processes like (MBE) in environments. In RHEED patterns, the diffraction spots from a well-ordered surface appear as elongated streaks rather than discrete points, arising from the limited out-of-plane imposed by the grazing incidence geometry. This rod-like extension in reciprocal space projects the surface lattice as streaks on the screen, with the streak length inversely related to the in-plane domain size and the width reflecting or step density. Consequently, RHEED is highly sensitive to monolayer-scale steps and terracing on the surface, allowing detection of subtle morphological changes during growth. During layer-by-layer epitaxial growth in , RHEED intensity exhibits periodic oscillations in the specular beam, where each cycle corresponds to the completion of one atomic . These oscillations stem from the alternating smooth and rough : maximum intensity occurs on flat layers, decreasing as adatoms form a partial layer, and recovering upon completion of the next layer. For instance, in GaAs(001) growth, persistent oscillations indicate two-dimensional mode, enabling precise calibration of deposition rates to within 1% accuracy. Surface reconstruction during epitaxy manifests as additional spots or split streaks in RHEED patterns, reflecting periodic rearrangements of surface atoms. A prominent example is the 2×4 reconstruction on GaAs(001), where arsenic dimers form rows, producing characteristic superlattice spots offset from bulk positions along the azimuth. These features evolve dynamically with growth conditions, such as As/Ga flux ratio, providing insights into stable surface phases. Recent advancements in the 2020s integrate RHEED with in combined in-situ setups, enabling correlative studies of atomic-scale surface dynamics during . This hybrid approach captures real-time RHEED patterns alongside STM imaging of adatom and island , bridging diffraction-based ensemble averages with local atomic for materials like III-V semiconductors.

Applications to surface structures and epitaxy

Low-energy electron diffraction (LEED) has been instrumental in mapping surface reconstructions, where the resulting diffraction patterns are described using Wood's notation to denote the periodicity and symmetry of the overlayer relative to the substrate. For instance, oxygen adsorption on Pt(111) forms a c(2×2) structure at coverages around 0.25 monolayers, characterized by a centered unit cell that is twice the size of the primitive substrate cell in both directions, as confirmed by LEED patterns showing split spots indicative of this reconstruction. This notation allows precise classification of complex overlayer arrangements, such as rotations or distortions, enabling researchers to correlate structural changes with adsorption-induced modifications. In epitaxial growth, reflection high-energy electron diffraction (RHEED) facilitates real-time monitoring of lattice matching, particularly in heterostructures where mismatch leads to observable spot splitting in diffraction patterns. During the growth of InAs on GaAs substrates, which exhibit approximately 7% lattice mismatch, RHEED spots from the GaAs substrate split as InAs layers form, providing direct evidence of relaxation and coherent onset. Such splitting quantifies the degree of pseudomorphic growth before defects like dislocations emerge, guiding the fabrication of high-quality thin films in devices. RHEED and LEED also enable defect detection at surfaces, such as domain boundaries in grown on metal substrates. On Ir(111), rotational domains of , misoriented by 0° or 30° relative to the substrate, produce distinct spot patterns that reveal ridge-like boundaries where domains meet, influencing electronic properties like charge transport. These techniques highlight how substrate interactions induce moiré superlattices and defects, critical for tailoring 's performance in . Quantitative analysis via dynamical refines adsorption site determination by modeling multiple scattering effects in intensity-voltage curves. For NO on Pt(111), dynamical LEED calculations distinguish between threefold hollow and twofold bridge sites, confirming preferential hollow adsorption with bond lengths around 1.3 for N-O, resolving ambiguities from simpler kinematic models. This approach has similarly identified hollow sites for CO on Co(0001), providing atomic-scale insights into catalytic binding geometries. Emerging time-resolved setups, incorporating laser-pump/probe configurations post-2010, capture surface dynamics by generating sub-300 electron pulses to probe transient structural changes. In these experiments, laser excitation induces surface currents or rearrangements, with subsequent monitoring revealing atomic-scale relaxations on picosecond timescales, as demonstrated in studies of nanodevices where electron bunch durations enable resolution of non-equilibrium states. Such advancements complement RHEED oscillations, which track layer-by-layer growth rates during .

Other specialized techniques

Gas electron diffraction for molecular structures

Gas electron diffraction (GED) employs a beam of high-energy electrons, typically accelerated to 20–60 keV, directed through a gaseous sample in a at pressures around 10^{-4} to 10^{-3} mbar. The setup features sector collimation to minimize multiple and ensure primarily single events at small angles (up to about 10–15°), capturing forward-scattered electrons on photographic plates or digital detectors placed 10–50 cm downstream. This configuration allows determination of isolated molecular geometries in the gas phase without the need for , making it particularly suited for studying transient or thermally labile . The raw diffraction data consist of intensity patterns I(s), where s is the scattering parameter defined as s = \frac{4\pi \sin \theta}{\lambda}, with θ the scattering angle and λ the electron wavelength (approximately 0.005–0.007 nm at these energies). After subtracting incoherent and atomic scattering backgrounds, the molecular scattering intensity M(s) is extracted, and the reduced intensity function sM(s) is computed to emphasize structural features. The radial distribution function P(r), which peaks at internuclear distances, is then obtained via Fourier inversion: P(r) = \frac{2}{\pi} \int_0^{s_{\max}} s M(s) \frac{\sin(sr)}{sr} \, ds, providing a direct visualization of bond lengths and torsional arrangements, though damped by vibrational effects. Structural parameters are refined using least-squares minimization to fit theoretical sM(s) curves, calculated from molecular models incorporating atomic factors and vibrational amplitudes, often supplemented by force fields or quantum chemical potentials to account for . This yields vibrationally averaged geometries (r_g parameters), with lengths accurate to approximately 0.002 and to 0.1–0.2°. GED excels for molecules unstable in solid state or requiring high temperatures (>500 K) for vaporization, as it probes free, unperturbed structures unaffected by packing forces. Unlike spectroscopic methods limited to small or symmetric , GED handles complex, heavy-atom-containing molecules up to ~100 atoms. Notable applications include conformational of n-butane, where GED revealed a trans/gauche with ~85% trans conformer at 20°C, torsional barriers ~3.7 kcal/, and C-C lengths of 1.532 (r_g). It also provides vibrationally averaged structures, such as in XeF_6, capturing fluxional distortions averaged over timescales.

Scanning electron microscopy diffraction

Electron backscatter diffraction (EBSD) is a scanning electron microscopy () technique employed to map the crystallographic microstructure of materials by analyzing diffraction patterns formed from backscattered s. In this method, a focused beam with energies typically ranging from 5 to 20 keV interacts with a tilted sample surface, usually at an angle of about 70 degrees, generating high-angle backscattered s that diffract off the crystal . These diffracted s form Kikuchi patterns, which are projected onto a phosphor screen or direct electron detector positioned behind the sample, enabling the visualization of orientations at each scanned point. The Kikuchi patterns in EBSD consist of characteristic lines and bands that reflect the symmetry of the , briefly referencing the underlying theory of where excess and deficit lines arise from dynamical diffraction effects. For automated analysis, these patterns undergo indexing via the , a computational method that detects linear features corresponding to Kikuchi bands by transforming the pattern into a parameter space where peaks indicate band positions. This process facilitates rapid determination of crystal orientations, phase identification, and misorientation mapping across large areas, with software algorithms refining the solutions to achieve angular accuracies of 0.5 to 1 . EBSD finds widespread applications in , particularly for characterizing polycrystalline metals, where it maps grain boundaries, reveals texture evolution during processing, and quantifies deformation microstructures. For instance, in aluminum alloys or steels, EBSD data delineate high-angle grain boundaries and preferred orientations, aiding in the optimization of mechanical properties like strength and , with spatial resolutions typically achieving around 50 under optimized conditions using field-emission guns. This non-destructive surface-sensitive technique covers areas up to several square millimeters, providing statistically robust datasets for microstructure-property correlations. A variant known as transmission EBSD (t-EBSD), also referred to as transmission Kikuchi diffraction (TKD), extends the method to thinner samples by operating in transmission mode within the , where electrons pass through electron-transparent foils typically 50 to 200 nm thick. This configuration enhances pattern quality and spatial resolution to below 10 nm, making it suitable for analyzing , thin films, or deformed regions near surfaces that are challenging for conventional EBSD. Sample preparation mirrors that for , involving ion milling or , and the technique has been instrumental in studying nanoscale grain structures in lightweight alloys. Advancements in the 2020s have integrated EBSD with serial sectioning techniques, such as milling, to produce three-dimensional () reconstructions of material volumes. In 3D-EBSD, successive surface layers are milled and mapped, allowing volumetric analysis of grain morphologies, boundary networks, and texture gradients over depths up to hundreds of micrometers. Automated systems developed around 2022 enable high-throughput acquisition of large datasets, for example, reconstructing 200 × 200 × 400 μm³ volumes in metals to investigate spatially resolved deformation or recrystallization processes. Recent progress as of 2024 includes frameworks like quaternion residual block self-attention networks (Q-RBSA) for generating high-resolution 3D EBSD maps from sparse data, and large-volume studies of additively manufactured alloys revealing unique grain morphologies.

Emerging methods and integrations

Electron holography integrated with electron diffraction enhances contrast imaging, enabling the visualization of electromagnetic fields and structural defects at the atomic scale in (TEM). This method records interference patterns from electron waves split by a biprism, reconstructing shifts that reveal charge distributions and strain around defects such as grain boundaries and dislocations. Recent advances, including aberration-corrected implementations, have achieved sub-angstrom resolution for quantitative mapping of electric fields in materials like , where contrast distinguishes electrostatic potentials at interfaces. In defect analysis, off-axis holography combined with differential contrast has quantified 2D electron gases at heterointerfaces, such as GaN/AlInN, by mitigating diffraction artifacts through tilt-averaged acquisitions. These developments, post-2015, leverage hybrid detectors to improve signal-to-noise ratios, facilitating real-time for dynamic defect studies. As of 2024, applications have extended to visualizing demagnetization fields in thin-foiled permanent magnets like Nd2Fe14B using electron holography. Correlative approaches integrating TEM-based electron diffraction with synchrotron X-ray techniques enable multimodal in-situ strain mapping, combining nanoscale resolution from electrons with the of X-rays for bulk samples. In irradiated materials like , synchrotron X-ray diffraction provides lattice parameter evolution during deformation, while TEM diffraction refines local strain fields around dislocations, achieving precisions below 0.1% in reconstructions. setups, such as those using soft X-ray alongside TEM, correlate mesoscale strain gradients with atomic-scale defect configurations in energy materials, as demonstrated in lithium-ion battery electrodes under operando conditions. Post-2015 innovations include automated algorithms that align datasets from both modalities, enhancing accuracy in mapping transient strains during phase transformations. As of 2025, correlative workflows have advanced to in-situ nanochip liquid cell TEM combined with synchrotron techniques for electrochemical studies. Cryo-electron , particularly microcrystal electron (MicroED), has emerged as a powerful tool for determining biomolecular structures, closely linked to cryo-EM workflows through shared cryogenic and low-dose in TEM. MicroED collects patterns from vanishingly small nanocrystals (volumes ~10^{-18} cm³), enabling atomic-resolution structures of proteins and peptides that resist for methods. Since 2015, advances include continuous rotation data collection for complete datasets, yielding structures like the α-synuclein NACore at 1.4 Å and membrane proteins such as NaK channels. with cryo-EM extends to hybrid phasing techniques using , as in the 2020 structure of a GPCR-ligand , bridging and for dynamic biomolecular insights. These methods have resolved hydrogen atoms in small biomolecules, complementing single-particle cryo-EM for complexes under 100 . As of November 2024, MicroED has expanded to small molecules and pharmaceuticals, promising broader access to structures previously unsuitable for analysis. Artificial intelligence and machine learning have revolutionized automated pattern recognition in 4D-STEM datasets, enabling real-time phase identification from electron diffraction patterns. Convolutional neural networks trained on simulated diffraction data classify crystal systems with over 95% accuracy across diverse materials, processing terabyte-scale datasets in minutes. In 2023 models, unsupervised clustering via non-negative matrix factorization disentangles overlapping phases in polycrystalline samples, identifying minor constituents like δ-phase in additively manufactured alloys without prior templates. Deep kernel learning further automates strain and orientation mapping by adapting to experimental noise, as shown in operando battery studies where phase transitions are tracked at video rates. These AI-driven tools, integrated into acquisition software, reduce analysis time from days to hours, enhancing throughput for high-impact applications in nanomaterials. As of October 2024, AI models have improved accuracy in 4D-STEM imaging for delicate materials by predicting Euler angles directly from patterns. Ultrafast electron diffraction (UED) employs electron pulses to probe in pump-probe configurations, capturing non-equilibrium states with atomic spatial and temporal resolution. Laser-driven electron sources generate ~10^5 electrons per pulse at 100-200 keV, minimizing space-charge blurring for diffraction from thin films and gases. In the , advancements include MHz repetition rates for improved statistics, as in 2022 studies of photoinduced phase transitions in VO₂, where lattice expansion is resolved within 100 fs of optical excitation. Pump-probe examples encompass in retinal chromophores (2021, revealing torsional motions) and in alloys (2023, tracking domain growth at 50 fs intervals). These experiments, often at facilities like SLAC, highlight UED's role in visualizing concerted electron-phonon couplings, with future potential in regimes via relativistic compression. As of July 2025, super-resolution MeV-UED has revealed coherent atomic motions in space with enhanced .

References

  1. [1]
    Davisson-Germer Experiment - HyperPhysics
    Davisson and Germer designed and built a vacuum apparatus for the purpose of measuring the energies of electrons scattered from a metal surface. Electrons from ...
  2. [2]
    [PDF] Electron Diffraction - Physics
    This allows us to sample all possible angles of incidence without changing the direction of the electron beam. If electrons act like a wave, different atomic.
  3. [3]
    Electron Diffraction Using Transmission Electron Microscopy - PMC
    Electron diffraction via the transmission electron microscope is a powerful method for characterizing the structure of materials, including perfect crystals ...
  4. [4]
    [PDF] Louis de Broglie - Nobel Lecture
    This hypothesis is necessary to explain how, in the case of light interferences, the light energy is con- centrated at the points where the wave intensity is ...
  5. [5]
    PHY 133 Lab Manual
    In the homework, you will show that this is λ = (12.26 Ǻ) / , where V is the accelerating voltage. (This comes from making some substitutions into λ = h / p.)Missing: derivation | Show results with:derivation
  6. [6]
    Ultrafast electron diffraction: Visualizing dynamic states of matter
    Dec 6, 2022 · This electron beam can be defined by the sum of isolated electrons correlated in time by periodic emission (the stroboscopic approach) ( Baum, ...
  7. [7]
    Electron Diffraction - an overview | ScienceDirect Topics
    As electrons are interacting actively with the matter, this technique needs a very thin sample to observe a diffraction pattern in transmission. Ever more ...
  8. [8]
    Critical Role of Inelastic Interactions in Quantitative Electron ...
    Jan 17, 2008 · Here both elastic and inelastic scattering of the beam electrons are included. The results of the two experiments were conducted on different a ...
  9. [9]
    [PDF] Electron Diffraction Using Transmission Electron Microscopy
    Electron diffraction via the transmission electron microscope is a powerful method for characterizing the structure of materials, including perfect crystals ...
  10. [10]
    Advances in the electron diffraction characterization of atomic ...
    Novel methods based on electron diffraction have been used to efficiently study individual nanoparticles and clusters and these can overcome the obstacles ...
  11. [11]
    Crystal structure and orientation of organic semiconductor thin films ...
    We use microcrystal electron diffraction (MicroED) to determine structures of three organic semiconductors, and show that these structures can be used along ...
  12. [12]
    Requiem for gas-phase electron diffraction | Structural Chemistry
    Apr 11, 2023 · This molecule has a trigonal bipyramidal shape with the axial P − F bonds longer than the equatorial P − F bonds.
  13. [13]
    Introduction - DoITPoMS
    The small electron wavelength also makes the diffraction angles θ small (1-2°); this can be seen by substituting a wavelength of 2.51 x 10-12 m into the Bragg ...
  14. [14]
    A crystal structure determined with 0.02 Å accuracy by electron ...
    Jul 11, 1996 · ELECTRON crystallography has two important advantages over X-ray crystallography for the determination of atomic positions in crystal ...
  15. [15]
    Review Applications and limitations of electron 3D crystallography
    Nov 2, 2023 · For protein crystals, the measurements should be conducted with liquid-nitrogen cooling to reduce radiation damage and maintain a hydrated ...
  16. [16]
    3D Electron Diffraction for Chemical Analysis: Instrumentation ...
    Electrons are absorbed by air. The flight tube of a TEM is under vacuum. Usually, the pressure is below 10–6 mbar. Many types of crystals will deteriorate under ...<|control11|><|separator|>
  17. [17]
    An Overview of Microcrystal Electron Diffraction (MicroED) - PMC
    This review discusses the accomplishments and future directions of the microcrystal electron diffraction (MicroED) cryo-EM technique. MicroED was developed for ...
  18. [18]
    In Situ Measurements of Thermodynamics and Reaction Kinetics ...
    Oct 29, 2019 · Multiscale spectroscopy combined with electron diffraction as well as TEM/STEM imaging are employed for in situ observations and quantifications of dynamic ...
  19. [19]
    J. J. Thomson 1897 - Cathode Rays - Le Moyne
    This experiment proves that something charged with negative electricity is shot off from the cathode, travelling at right angles to it, and that this something ...
  20. [20]
    August, 1913: Robert Millikan Reports His Oil Drop Results
    Aug 1, 2006 · Millikan's reported value for the elementary charge, 1.592 x 10-19 coulombs, is slightly lower than the currently accepted value of 1.602 x 10- ...Missing: primary | Show results with:primary
  21. [21]
    Historical development and future trends of vacuum electronics
    Sep 7, 2012 · The electron tube production started slowly in 1905 and made use of the existing vacuum technology for mass production of incandescent lamps, ...
  22. [22]
    [PDF] Space Charge Effects
    Space charge forces are responsible for several unwanted phenomena related to beam dynamics, such as energy loss, shift of the synchronous phase and frequency, ...Missing: partial | Show results with:partial
  23. [23]
    Vacuum Basics and Applications | Abbess Instruments
    Thus the mean free path would be 5 cm at 0.001 Torr and 50 meters at 1×10^-6 Torr. The lengthening of mean free path at low pressures is a key enabler for ...
  24. [24]
    The Nobel Prize in Physics 1937 - NobelPrize.org
    ... 1937 was awarded jointly to Clinton Joseph Davisson and George Paget Thomson "for their experimental discovery of the diffraction of electrons by crystals".
  25. [25]
    History of the Electron Microscope - News-Medical
    In 1933, Ernst Ruska developed on the original model further to develop an electron microscope that was capable of producing an image of higher resolution than ...
  26. [26]
    Electron Microscopy in Berlin 1928–1945 - ScienceDirect
    In the spring of 1931 Knoll and Ruska finished the construction of the first two-stage transmission electron microscope with magnetic lenses.
  27. [27]
    Microscope Resolution: Concepts, Factors and Calculation
    In order to increase the resolution, d = λ/(2NA), the specimen must be viewed using either a shorter wavelength (λ) of light or through an imaging medium with ...
  28. [28]
    Electron Microscopy Techniques, Strengths, Limitations and ...
    Jan 2, 2024 · This was called the "Abbe diffraction limit". Abbe deduced that a microscope could not resolve two objects located closer than λ/2NA, where λ is ...
  29. [29]
    Electron Microscopy for Rapid Diagnosis of Emerging Infectious ...
    Mar 3, 2003 · The first electron micrograph of poxvirus was published in 1938. In 1941, immunologic procedures were first used in electron microscopic studies ...Missing: crystal lattice
  30. [30]
    The first phage electron micrographs - PMC - NIH
    The first phage electron micrographs were published in 1940 in Germany and proved the particulate nature of bacteriophages.Missing: crystal lattice
  31. [31]
    Electron backscattering from thin films - AIP Publishing
    Apr 1, 1982 · The basic principles of electron backscattering from atoms and solids in the energy range 10 to 100 keV are reviewed.
  32. [32]
    Bacteriophage imaging: past, present and future - ScienceDirect.com
    The visualization of viral particles only became possible after the advent of the electron microscope. The first bacteriophage images were published in 1940 ...
  33. [33]
    Brillouin zones and Kikuchi lines for crystals under electron ...
    The geometry of the Kikuchi lines in high- and low- energy electron diffraction patterns is defined in terms of intersections of the Brillouin zone boundaries ...
  34. [34]
    Kinematical theory of electron diffraction (Chapter 3)
    In the kinematical theory, we consider the diffraction of a plane wave (of wavelength λ) incident upon a three-dimensional lattice array of identical scattering ...Missing: seminal | Show results with:seminal
  35. [35]
    [PDF] Phase Identification in a Scanning Electron Microscope Using ...
    In the first de- scription, Kikuchi patterns are formed by the elastic scattering (diffraction) of previously inelastically scat- tered electrons [1].
  36. [36]
    Kikuchi Lines and Bands in Electron Diffraction
    Kikuchi lines originate with two electron scatterings: the first scattering is incoherent (sometimes inelastic), and then it is followed by a coherent (elastic ...
  37. [37]
    None
    ### Summary of Kikuchi Lines Formation, Geometry, and Applications in Orientation and Symmetry
  38. [38]
    (IUCr) Dynamical calculation of thermal diffuse electron scattering
    Thermal diffuse scattering calculated across the dif- fraction pattern for ... However, it is not clear whether this suggestion of Kikuchi lines is dependent on ...
  39. [39]
    The Interpretation and Application of Electron-Diffraction `Kikuchi ...
    The interpretation of electron-diffraction Kikuchi-line patterns from single crystals is developed as a powerful independent and general means of determining ...
  40. [40]
    Selected Area Electron Diffraction - an overview | ScienceDirect Topics
    Selected area electron diffraction (SAED) is a crystallographic experimental technique that is performed by using a transmission electron microscope.
  41. [41]
    Thon rings from amorphous ice and implications of beam-induced ...
    Thon rings from amorphous carbon are routinely used to adjust astigmatism and set the defocus of a microscope.
  42. [42]
    Diffraction artefacts from twins and stacking faults, and the mirage of ...
    Mar 15, 2021 · This artefact may come from an overlap along the electron beam axis of some domains that are twin-related. It leads to extra-spots created by ...
  43. [43]
    Fast Pixelated Detectors in Scanning Transmission Electron ...
    The use of fast pixelated detectors and direct electron detection technology is revolutionizing many aspects of scanning transmission electron microscopy (STEM) ...
  44. [44]
    [PDF] The Characterization of Textures of Thin Films by Electron Diffraction
    Two types of preferred orientations (POs) are commonly observed, namely a lamella type and a fibrous type of texture¹¹¹. In lamella texture, the crystalline ...
  45. [45]
    Top-bottom effects in double diffraction - ScienceDirect
    The effect may be useful as a simple method for determining on which side of a TEM specimen a particular epitactic feature lies.
  46. [46]
    Symmetry analysis - Tanaka - 1989 - Wiley Online Library
    The crystal point- and space-group determination method using convergent-beam electron diffraction (CBED) is reviewed. Crystal symmetry elements that can be ...Missing: centrosymmetry | Show results with:centrosymmetry
  47. [47]
    Dynamical diffraction effects on higher-order laue zone lines in ...
    We studied the dynamical diffraction effects on the defect lines of higher-order Laue zone reflections (HOLZ lines) in bright field disks of convergent-beam ...
  48. [48]
    Strain measurements by convergent-beam electron diffraction
    Jul 26, 2004 · A broadening of the high order Laue zone lines in the transmitted disk of CBED patterns is observed when approaching the NiSi ∕ Si interface.
  49. [49]
    Point group symmetry of cadmium arsenide thin films determined by ...
    Aug 20, 2019 · Here, we apply convergent beam electron diffraction (CBED) to determine the point group of thin films grown by molecular-beam epitaxy.
  50. [50]
    Development of ultrafast four-dimensional precession electron ...
    Integrating the excitation error and off-zone axes incidence of the electron beam in PED significantly reduces the uneven diffracted intensities caused by the ...
  51. [51]
    [PDF] Precession Electron Diffraction and its Advantages for Structural ...
    Nov 6, 2008 · This paper outlines the foundations of precession electron diffraction (PED) in order to illustrate its utility for struc- tural fingerprinting ...
  52. [52]
    Four-Dimensional Scanning Transmission Electron Microscopy (4D ...
    In this paper, we review the use of these four-dimensional STEM experiments for virtual diffraction imaging, phase, orientation and strain mapping.
  53. [53]
    Synchronization of scanning probe and pixelated sensor for image ...
    Here we present a solution to the problem of synchronizing the electron probe scan with the diffraction image acquisition, and demonstrate on a fast hybrid- ...
  54. [54]
    Phase Imaging Methods in the Scanning Transmission Electron ...
    Jun 28, 2025 · This review introduces phase imaging methods in the STEM and explores how the most recent innovations are driving progress in nanoscience.
  55. [55]
    Sub-ångström resolution ptychography in a scanning electron ...
    Oct 14, 2025 · It creates a quantitative model of the amplitude and phase shift accrued by the electron wave on passing through the sample using a computation ...<|control11|><|separator|>
  56. [56]
    Differentiating electron diffuse scattering via 4D-STEM spatial ...
    Specifically, the heterogeneous-domain diffuse-scattering signals are attributed to nanoscale planar defects, while the homogeneous diffuse-scattering of the ...
  57. [57]
    Direct observation of charge-density waves in | Phys. Rev. B
    Apr 30, 2008 · ... electron diffraction and direct observation in real space using ... satellite spots of charge-density wave modulations. The satellite ...
  58. [58]
    Metallic Phase with Long-Range Orientational Order and No Translational Symmetry
    ### Summary of Key Findings on Electron Diffraction Patterns in Quasicrystals
  59. [59]
    Investigation of Interstitials in Electron-Irradiated Aluminum by ...
    Sep 15, 1973 · The diffuse scattering of x rays ( C u ⁢ K 𝛼 ⁢ l ) from single crystals of aluminum has been investigated after low-temperature electron ...
  60. [60]
  61. [61]
    Review Pair Distribution Function Obtained from Electron Diffraction
    Feb 3, 2021 · Pair distribution function from electron diffraction (ePDF) is a remarkable technique capable of elucidating the atomic arrangement of ...
  62. [62]
    Dynamics of thermal diffuse scattering in high-energy electron ...
    The TDS streaks are defined by the qx-qy curves which satisfy δi(q)=O, where δi(q) is the phonon dispersion relationship determined by the two-dimensional ...
  63. [63]
    Dislocation contrast in high-angle hollow-cone dark-field TEM
    the creation of diffuse scattering due to lattice distortion around the dislocation ...
  64. [64]
    [PDF] Low-energy electron diffraction crystallography of surfaces and ...
    The basic principle of a standard LEED experiment is very sim- ple: a collimated mono-energetic beam of electrons is directed towards a single crystal surface ...
  65. [65]
    None
    ### Summary of LEED Observation of Si(100) 2x1 Reconstruction
  66. [66]
    Reflection High-Energy Electron Diffraction - ScienceDirect.com
    RHEED patterns are obtained by the diffraction of a high-energy electron beam (typically in the 10–35 keV range) at a grazing incidence to the substrate.
  67. [67]
    [PDF] Reflection High-Energy Electron Diffraction - UBC Physics
    Nov 22, 2019 · In this paper, we present an overview of the diffraction technique ... DIFFRACTION AND EWALD CONSTRUCTION. It is helpful to describe ...
  68. [68]
    [PDF] Reflection High-Energy Electron Diffraction (Shuji Hasegawa)
    Reflection high-energy electron diffraction (RHEED) uses a finely collimated electron beam with energy of. 10–100 keV. The beam irradiates a sample surface ...
  69. [69]
    RHEED streaks and instrument response
    Reflection high-energy electron diffraction (RHEED) is so sensitive to surface morphology that it is difficult to separate the roles of instrument response and ...
  70. [70]
    Dynamics of film growth of GaAs by MBE from Rheed observations
    Detailed observations have been made of the intensity oscillations in the specularly reflected and various diffracted beams in the RHEED pattern during MBE.
  71. [71]
    Development of in situ characterization techniques in molecular ...
    By combining STM, reflection high energy electron diffraction (RHEED), and angle-resolved photoelectron spectroscopy (ARPES), Song et al. were able to ...
  72. [72]
    Density functional theory study of the initial oxidation of the Pt(111 ...
    Mar 27, 2009 · Consistent with prior studies, our calculations predict that oxygen atoms adsorb on fcc sites and form p ⁢ ( 2 × 2 ) and p ⁢ ( 2 × 1 ) ...Missing: 2x2) notation<|separator|>
  73. [73]
    Real-Time Characterization Using in situ RHEED Transmission ...
    Jan 26, 2018 · This shift was because of the relatively larger lattice mismatch of ca. 7% between GaAs and InAs. The RHEED spots from GaAs split off from ...
  74. [74]
    Defects of graphene on Ir(111): Rotational domains and ridges
    We use low-energy electron microscopy (LEEM), low-energy electron diffraction (LEED), and scanning tunneling microscopy (STM) to study different ...
  75. [75]
    Dynamical LEED analyses of the Pt(111)-p(2× 2)-NO and the Ni(111 ...
    The threefold hollow-site adsorption determined for these systems also contrasts with the lower coordinated bridge and top sites generally found for CO ...
  76. [76]
    LEED investigations on Co(0001): The overlayer | Phys. Rev. B
    Mar 23, 2001 · The local adsorption structure of the coadsorption phase of K and CO on Co(0001) has been determined at 160 K using dynamical low-energy ...
  77. [77]
    Femtosecond electrons probing currents and atomic structure in ...
    Oct 31, 2014 · Here we introduce an alternative approach for the implementation of time-resolved LEED using the potential of our electron gun design to ...
  78. [78]
    Intensity oscillations for electron beams reflected during epitaxial ...
    Apr 15, 1987 · The demonstrated utility of RHEED oscillations for monitoring controlling growth in GaAs is now applicable to metal growth. References (11).
  79. [79]
    None
    Summary of each segment:
  80. [80]
    Gas-Phase Electron Diffraction - SpringerLink
    Hargittai, I. (1988) A Survey: The Gas-Phase Electron Diffraction Technique of Molecular Structure Determination. Chapter 1 in Hargittai, I. and Hargittai, M.
  81. [81]
    Gas-phase electron diffraction | Accurate Molecular Structures
    Oct 31, 2023 · Gas electron diffraction is one of the few experimental techniques able to determine molecular geometry in the vapor phase. The other principal ...
  82. [82]
    Direct Evaluation of Equilibrium Molecular Geometries Using Real ...
    Apr 1, 1994 · Effect of vibronic interactions on molecular structures determined by gas electron diffraction. Structural Chemistry 2015, 26 (5-6) , 1197 ...
  83. [83]
    Gas-Phase Electron Diffraction for Molecular Structure Determination
    This paper reviews advances of modern gas electron diffraction (GED) method combined with high-resolution spectroscopy and quantum-chemical calculations in ...Missing: seminal | Show results with:seminal<|control11|><|separator|>
  84. [84]
    The Molecular Structure and Rotational Isomerization of n-Butane 1,2
    Molecular Structure of Piperazine as Studied by Gas Electron Diffraction. Bulletin of the Chemical Society of Japan 1971, 44 (9) , 2352-2355. https://doi ...
  85. [85]
    Gas-phase structure of nickel dichloride. An electron-diffraction ...
    Apr 15, 2019 · In 1973 we completed a gas-phase electron-diffraction (GED) ... accelerating voltage, 44 kV; exposure times, 2–3 min (LC) and 5–6 ...
  86. [86]
    Electron Backscatter Diffraction - an overview | ScienceDirect Topics
    This sequence includes the original diffraction pattern, Hough transformed pattern, peaks detected in the Hough transformation, corresponding Kikuchi bands ...
  87. [87]
    Electron Backscatter Diffraction (EBSD) Pattern Formation
    Learn about the physics behind electron backscatter diffraction pattern (EBSP) formation, including the calculation of Kikuchi band intensities.
  88. [88]
    Constraints on the effective electron energy spectrum in backscatter ...
    Feb 27, 2019 · Electron backscatter diffraction (EBSD) is a technique which is used to reveal the microstructure of crystalline materials, including metals, ...
  89. [89]
    Basics of Automated Indexing for EBSD - Oxford Instruments
    A Hough transform is used to identify the positions of the Kikuchi bands. The bands are seen as peaks in Hough space. A filter (such as a “butterfly” filter) is ...
  90. [90]
    Polarity Determination in EBSD Patterns Using the Hough ...
    The current article will show how to use the Hough transformed EBSP to determine the orientation with respect to the polar planes and hence provide a step ...Introduction · Experimental Methods · Theory · Results and Discussion
  91. [91]
    High-precision orientation mapping from spherical harmonic ...
    This paper presents a brief overview of EBSD indexing techniques, including conventional Hough-transform indexing, the established high-precision technique ...
  92. [92]
    Recent Advances in EBSD Characterization of Metals - MDPI
    The grain size, crystallographic orientation, texture, and grain boundary character distribution can be obtained by EBSD analysis. Despite the limited ...
  93. [93]
    What is EBSD? | Electron BackScatter Diffraction - Bruker
    Grain Orientation Mapping: EBSD is commonly used to determine the orientation, size and shape distribution of grains in crystalline materials. This ...
  94. [94]
    EBSD spatial resolution for detecting sigma phase in steels
    Two types of spatial resolution are usually defined: the physical resolution, which represents the maximum distance to a high angle grain boundary for which the ...
  95. [95]
    Features of Transmission EBSD and its Application | JOM
    Jul 27, 2013 · The quality of t-EBSD patterns is very much dependent on the thickness of the specimen. This dependence also has a strong effect on the spatial ...
  96. [96]
    Transmission Kikuchi Diffraction - Oxford Instruments
    Learn how transmission Kikuchi diffraction (TKD) or transmission electron backscatter diffraction (t-EBSD) can be used to characterise nanomaterials.
  97. [97]
    Transmission EBSD in the Scanning Electron Microscope | NIST
    May 1, 2013 · The purpose of this paper is to describe this new way to obtain electron diffraction patterns with nanometer resolution from thin samples within ...Missing: thinner | Show results with:thinner
  98. [98]
    Development of a new, fully automated system for electron ...
    We report the development of a fully automatic large-volume 3D electron backscatter diffraction (EBSD) system (ELAVO 3D), consisting of a scanning electron ...
  99. [99]
    [2506.17534] Large volume 'chunk' lift out for 3D tomographic ... - arXiv
    Jun 21, 2025 · In this work, we present an overview of site-specific large volume 'chunk' lift out and 3D serial sectioning of substantive volumes (eg 200 x 200 x 400 um3).
  100. [100]
    Recent Progress of Digital Reconstruction in Polycrystalline Materials
    Mar 1, 2025 · Recent research on the serial sectioning methodologies has focused on large-scale datasets, high efficiency, and high-resolution reconstruction ...
  101. [101]
    Differences between differential phase contrast and electron ...
    In this work two complementary electron microscopic methods, differential phase contrast (DPC) and electron holography (EH), are used for characterization of a ...Missing: advances | Show results with:advances
  102. [102]
    Advances in synchrotron x-ray diffraction and transmission electron ...
    The focus of the present study is to demonstrate how advances in these two techniques allow for a more detailed analysis of irradiated material.B. X-Ray Diffraction · 1. Synchrotron Xrd · Iii. Results<|separator|>
  103. [103]
    Correlative transmission electron and soft x-ray microscopy for ...
    Correlative transmission electron and soft x-ray microscopy yield highly complementary information from atomic to mesoscale.
  104. [104]
    [PDF] Electron Diffraction Imaging of Materials Structural Properties - arXiv
    Zuo JM & Spence JCH (2017) Advanced transmission electron microscopy, imaging and diffraction ... Reimer L ed (1995) Energy-filtering Transmission Electron ...<|control11|><|separator|>
  105. [105]
  106. [106]
    The CryoEM Method MicroED as a Powerful Tool for Small Molecule ...
    Nov 2, 2018 · This method uses a transmission electron cryo-microscope to collect electron diffraction data from extremely small 3-dimensional (3D) crystals.
  107. [107]
    Automated crystal system identification from electron diffraction ...
    4D-STEM allows spatially resolved electron diffraction patterns (DPs) to be collected. The structural information gleaned from such DPs is critical in ...
  108. [108]
    Machine learning for automated experimentation in scanning ...
    Dec 20, 2023 · Machine learning (ML) has become critical for post-acquisition data analysis in (scanning) transmission electron microscopy, (S)TEM, imaging and spectroscopy.
  109. [109]
    High-Throughput Sub-Pixel Electron Diffraction Pattern Recognition
    Jun 4, 2025 · Here, we investigate for the first time the applicability of a fast object detection machine learning architecture in sub-pixel precision peak ...Missing: indexing 2020s
  110. [110]
    Shot-to-shot acquisition ultrafast electron diffraction - arXiv
    Jul 18, 2025 · We demonstrate a novel shot-to-shot acquisition method for optical pump - keV electron energy probe in ultrafast scattering experiments.Missing: examples | Show results with:examples
  111. [111]
    High-repetition-rate ultrafast electron diffraction with ... - AIP Publishing
    Sep 27, 2024 · UED is often employed in pump-probe configuration where an optical (“pump”) pulse photoexcites a sample away from its ground-state structure ...INTRODUCTION · Direct electron detection · Time-overlap between optical...