Fact-checked by Grok 2 weeks ago

Low-energy electron diffraction

Low-energy electron diffraction (LEED) is a surface-sensitive for determining the atomic structure of crystalline surfaces by measuring the diffraction patterns formed by low-energy electrons (typically 20–200 ) incident normally on the sample. These electrons have de Broglie wavelengths on the order of 0.9–2.5 , comparable to typical interatomic spacings, enabling the observation of spots that reflect the two-dimensional periodicity of the surface . Due to the short inelastic mean free path of low-energy electrons in solids (approximately 5–10 ), LEED probes only the outermost atomic layers, making it ideal for studying surface reconstructions, adsorbate geometries, and interface structures. The foundational discovery of electron diffraction occurred in 1927, when and Lester Germer observed diffraction patterns from electrons scattered by a , confirming the wave nature of electrons predicted by . However, early experiments were limited by surface contamination in inadequate vacuum conditions, and the technique lay dormant until the 1960s, when advancements in (UHV) technology enabled the preparation and maintenance of atomically clean surfaces. This revival, coupled with theoretical developments in multiple scattering calculations by researchers like , transformed into a quantitative tool for surface by the early 1970s. In a standard LEED setup, a collimated beam of monoenergetic electrons from an is directed at the sample in a UHV chamber (pressures below 10^{-9} ), and the elastically backscattered electrons are accelerated onto a hemispherical fluorescent screen, producing a visible diffraction pattern of glowing spots. Retarding field analyzers or grids filter out inelastically scattered electrons to ensure only elastic diffraction contributes to the pattern. Qualitatively, the spot positions and symmetries reveal the surface unit cell and rotational domains, while quantitative analysis involves measuring intensity versus incident energy (I-V) curves for each diffracted beam and comparing them to simulations based on dynamical diffraction theory, often using reliability factors like the Pendry R-factor to optimize structural models. LEED has been instrumental in , enabling the determination of hundreds, if not thousands, of unique surface structures since the 1970s, including clean metal surfaces, reconstructions, and adsorbate-induced phases such as the c(2×2) CO structure on Ni(100). Its strengths include high sensitivity to atomic positions (accuracy ~0.05 ) and compatibility with complementary techniques like for composition analysis, though limitations arise from the need for long-range order and challenges in modeling complex multiple scattering at low energies. Modern variants, such as spot-profile-analysis LEED (SPA-LEED), extend its capabilities to quantitative domain size measurements and rough surfaces.

Introduction

Basic principles

Low-energy electron diffraction (LEED) is a surface-sensitive diffraction technique that utilizes electrons with kinetic energies typically in the range of 20–200 to probe the atomic structure and crystallography of solid surfaces, particularly single-crystal samples. By directing a of these low-energy electrons onto a surface under conditions, LEED reveals the periodic arrangement of surface atoms through the resulting patterns. The underlying mechanism relies on the wave-particle duality of electrons, where their de Broglie wavelength is comparable to interatomic distances on the order of angstroms, enabling diffraction similar to X-ray or neutron scattering but with enhanced surface selectivity. The de Broglie wavelength \lambda is given by \lambda = \frac{h}{\sqrt{2mE}}, where h is Planck's constant, m is the electron mass, and E is the kinetic energy of the electron; for E \approx 100 eV, \lambda \approx 1.22 Å. At these low energies, elastic scattering predominates, as electrons primarily interact with the electrostatic potentials of surface atoms without significant energy loss, leading to coherent interference that forms observable diffraction patterns on a fluorescent screen or detector. The surface specificity of LEED arises from the short of low-energy electrons in solids, typically 5–10 , which confines the probing depth to the topmost few layers and ensures that only electrons scattered from the near-surface region contribute to the detected signal. This \Lambda(E) can be approximated as \Lambda(E) = a E^{-2} + b E^{1/2}, with empirical constants a \approx 143 and b \approx 0.054 (in appropriate units of and ), highlighting its energy dependence and minimum value around tens of . The resulting patterns consist of discrete spots on a hemispherical screen, each corresponding to vectors of the surface , with spot positions and intensities governed by the Laue conditions for momentum conservation parallel to the surface.

Applications and significance

Low-energy electron diffraction (LEED) serves as a primary tool for determining surface reconstruction, adsorption sites, epitaxial growth, and phase transitions on metals, semiconductors, and oxides, providing atomic-scale insights into surface geometries. For instance, LEED patterns reveal superstructures such as the p(√3 × √3) R30° arrangement of CO molecules adsorbed on Ni{111} surfaces, enabling precise identification of hollow adsorption sites with positional accuracy of ±1-20 pm through intensity-voltage (I/V) analysis. In epitaxial growth studies, LEED spot profile analysis measures terrace widths and step heights, as demonstrated in Pd{100} films where domains of approximately 9.8 nm were resolved, while phase transitions are tracked via changes in spot sharpness indicating order-disorder dynamics on surfaces like those of semiconductors and oxides. The technique holds significant value in catalysis by elucidating structures on catalyst surfaces, where ordered adsorbate phases directly influence reaction pathways, as seen in early observations of chemisorbed species on metal surfaces that informed models of surface phase formation. In processing, evaluates interface quality in thin films by confirming epitaxial alignment and detecting defects, ensuring optimal performance in device fabrication. Within , it characterizes two-dimensional materials like , revealing moiré superlattices and decoupling effects on substrates such as Ni{111}, which is crucial for tailoring electronic properties in van der Waals heterostructures. LEED's advantages stem from its high surface , with electrons at 20-200 penetrating only 0.1-1 nm, allowing exclusive probing of the topmost atomic layers under (UHV) conditions below 10^{-10} mbar for in-situ analysis. This enables real-time monitoring of surface changes, such as during adsorption or annealing, providing qualitative periodicity via patterns and quantitative atomic positions through dynamical simulations. However, limitations include restriction to conductive samples to avoid charging effects, high to that degrades patterns even in UHV, and inability to probe subsurface layers beyond approximately 10 due to rapid . In modern contexts as of , continues to advance studies of clean energy materials, such as oxide perovskites like Sr₂RuO₄, where it maps octahedral rotations and surface terminations under to investigate charge transport at interfaces. Recent advances include ultrafast techniques for time-resolved studies and improved for precise modeling. It also plays a key role in research, facilitating the design of novel by resolving surface electronic structures and dynamics in systems like epitaxial layers, underscoring its enduring impact on .

History

Discovery of electron diffraction

The discovery of electron diffraction marked a pivotal confirmation of Louis de Broglie's 1924 hypothesis that particles possess wave-like properties, with wavelength given by \lambda = h / p, where h is Planck's constant and p is momentum. In 1927, Clinton Davisson and Lester Germer at Bell Laboratories performed scattering experiments using electrons incident on a polycrystalline nickel target in a vacuum apparatus. Initially, their observations showed irregular scattering patterns due to the polycrystalline nature of the target, which they had inadvertently recrystallized into larger single-crystal domains after accidental heating during a vacuum repair. A breakthrough occurred when they detected a pronounced intensity maximum in the scattered electron beam at a polar angle of 50° for incident electrons with of approximately 54 . This energy corresponded to a de Broglie of about 1.65 , calculated as \lambda = h / \sqrt{2 m E}, where m is the and E is the . The peak aligned precisely with Bragg from the (111) planes of the face-centered cubic nickel , which have an interplanar spacing d \approx 2.15 , satisfying the condition n\lambda = 2d \sin\theta for n=2 and \theta = 50^\circ. The observed intensity maxima matched predictions from patterns on the same , providing direct evidence of . The initial irregular results posed interpretation challenges, as Davisson and Germer initially viewed electrons as particles without wave properties, but consultation with wave mechanics principles—prompted by de Broglie's ideas and Schrödinger's equation—resolved this by framing the scattering as coherent wave from the atomic lattice. Concurrently, in 1927, at the independently demonstrated through transmission experiments, directing a beam of ~13,000 eV electrons through thin films such as (~3 × 10^{-6} cm thick) and later metal foils like and . Thomson captured diffraction patterns on photographic plates, revealing central spots surrounded by concentric rings analogous to optical halos, with ring diameters scaling inversely with electron wavelength as per de Broglie's relation. These parallel findings by Davisson-Germer (reflection) and Thomson (transmission) universally established the wave nature of electrons, earning them the 1937 .

Development of LEED as a surface analysis tool

In , early attempts to observe low-energy diffraction patterns from surfaces were pioneered by researchers like W. Ehrenberg, who developed post-acceleration techniques to amplify the weak diffracted signals and make them visible on fluorescent screens, overcoming the limitations of direct detection in low-vacuum environments. During the 1950s, further refinements by H. E. Farnsworth and others improved these systems, incorporating better and vacuum conditions to produce clearer surface diffraction images, laying the groundwork for LEED as a qualitative tool for face identification. The 1960s marked a breakthrough in LEED's evolution into a practical surface technique, driven by efforts from , which introduced ion-pumped (UHV) chambers integrated with post-acceleration LEED optics around 1965, making the method accessible to laboratories worldwide. Researchers like Gabor A. Somorjai at the , played a pivotal role by combining these commercial UHV-LEED systems with single-crystal samples, enabling the first systematic studies of clean, atomically ordered surfaces free from contamination, which revealed how adsorbates alter surface periodicity. A key milestone came in 1959 when R. E. Schlier and H. E. Farnsworth reported the first LEED observation of on clean (111) and (100) faces, showing deviations from bulk lattice due to atomic rearrangements, such as the 2×1 and 7×7 patterns. By the 1970s, LEED transitioned from qualitative pattern imaging to quantitative structural determination, fueled by advances in dynamical scattering theory and computing power that allowed simulation of multiple electron scattering events for precise atomic position modeling. John B. Pendry's development of efficient computational algorithms for intensity-voltage (I-V) curve analysis during this decade enabled reliable comparison of experimental data with theoretical predictions, establishing LEED as a standard for resolving surface geometries with sub-angstrom accuracy in systems like metal-adsorbate overlayers. From the 1980s onward, LEED instrumentation advanced with the adoption of digital detectors, such as channel-plate electron multipliers and video acquisition systems, which automated pattern recording and intensity measurements, reducing manual intervention and improving data reproducibility for high-throughput experiments. These enhancements, coupled with software for real-time analysis, extended into the and , while integrations with sources in the and beyond combined LEED with techniques like for correlative studies, achieving enhanced spatial resolution and chemical sensitivity on complex surfaces like oxides and organics.

Experimental Setup

Instrumentation components

Low-energy electron diffraction (LEED) systems rely on specialized hardware to generate, direct, and detect low-energy electrons interacting with a sample surface. The primary components include the , , , sample manipulator, and calibration mechanisms, each designed to ensure precise control and high sensitivity in surface analysis. The serves as the source of the incident electron beam, typically utilizing thermionic cathodes, such as or (LaB₆), or field emission sources to achieve monochromatic beams with energies ranging from 20 to 200 eV. These guns incorporate electrostatic lenses and apertures to focus the beam, with typical currents of 1-10 μA to provide sufficient intensity for diffraction without damaging the sample. The design emphasizes stability and low energy spread, often achieving transfer widths exceeding 300 Å at 80 eV for high-resolution patterns. LEED optics consist of a retarding field analyzer equipped with multiple grids (typically three or four) to select electrons by energy and retard the beam, enabling the visualization of diffracted patterns on a hemispherical fluorescent screen or detection via channeltron multipliers for intensity measurements. The fluorescent screen, often phosphor-coated, provides visual feedback of the diffraction pattern, while channeltrons offer quantitative data with gains up to 10⁷ and spatial resolutions around 0.014° angularly. These are arranged in a hemispherical configuration to capture backscattered electrons over a wide angular range, minimizing distortions in the observed pattern. The entire setup operates within an (UHV) chamber to preserve surface cleanliness, with base pressures below 10⁻¹⁰ achieved using ion pumps, titanium sublimation pumps, and liquid-nitrogen cryoshrouds. Such low pressures, often reaching the low 10⁻¹¹ range, prevent adsorption of residual gases that could alter the surface structure during measurements. The sample manipulator enables precise positioning and environmental control of the specimen, featuring x-y-z translational stages, rotational (up to ±180° in polar and azimuthal angles), and heating/cooling capabilities to temperatures from ~77 K to 1500 K. Heating is typically provided by bombardment or radiation, while cooling uses , allowing studies of temperature-dependent surface phases. The manipulator ensures the sample surface remains perpendicular or at controlled angles to the incident beam, with tolerances better than 0.2 mrad for alignment. Safety and calibration procedures are integral to reliable operation, involving regular alignment of the and using test patterns from known surfaces, and achieving energy resolutions of approximately 0.1 through voltage stabilization and grid biasing. These steps minimize and ensure reproducible spot positions, with interlocks preventing high-voltage exposure during maintenance.

Sample preparation and handling

Samples suitable for low-energy electron diffraction (LEED) experiments are typically single crystals of metals such as (Pt), (Au), nickel (Ni), or palladium (Pd), and semiconductors like (Si) or germanium (Ge), which provide well-ordered surfaces for diffraction analysis. Thin films of these materials deposited on conductive substrates can also be used, ensuring electrical conductivity to prevent surface charging under bombardment. Insulating samples pose challenges and require modifications, but conductive samples are preferred for standard LEED setups. Preparation of these samples begins with cutting or cleaving to expose a fresh surface, often followed by mechanical polishing to achieve flatness, as seen in the orientation of NaAu₂(111) crystals using . For semiconductors like Ge(100), initial ultrasonic in solvents such as and removes gross contaminants before (UHV) transfer. To introduce adsorbates for overlayer studies, controlled dosing is performed after cleaning, but clean surface preparation prioritizes removal of oxides and impurities. Mounting involves securing the sample to a manipulator for precise positioning and temperature control, typically using spot-welded (Ta) clips or (W) wires to ensure electrical grounding and resistive heating capability. For example, Au(100) single crystals are mounted at the base of the manipulator with W wires, allowing rotation and alignment for normal incidence. Samples like SmB₆(001) are fixed to a substrate holder using heating for stability during experiments. This setup integrates with UHV systems to maintain pressures below 10⁻⁹ mbar, preventing recontamination. In-situ cleaning is essential and commonly involves cycles of (Ar) followed by annealing in UHV to desorb contaminants and restore surface order. uses Ar⁺ ions at energies of 0.5–2 keV for 15–30 minutes to remove oxides and impurities. Subsequent annealing at temperatures of 500–1000°C, such as 800–850°C for Ge or 1030°C for SmB₆, promotes atomic and defect healing, with cycles repeated several times (typically 2–10) for optimal cleanliness monitored by residual gas analyzers and pressure gauges. For metals like , heating in or reduces surface oxides effectively. Key challenges in sample handling include minimizing surface defects such as steps or facets, which can distort diffraction patterns by introducing mosaicity or , as observed in NaAu₂ surfaces with spreads of 0.3°–1.4°. High-temperature annealing risks elemental segregation or loss, like sodium depletion in NaAu₂ above 450 K, necessitating careful temperature control. For non-conductive samples, charging effects from are mitigated by low beam currents (e.g., 1 nA) or electron flooding, while beam damage must be avoided through brief exposures. Maintaining UHV conditions during transfer and manipulation is critical to prevent adsorption of residual gases.

Data acquisition techniques

In low-energy electron (LEED), patterns are recorded using detectors that capture the backscattered electrons from the sample surface. Traditional methods relied on visual observation of patterns displayed on fluorescent screens or recording on photographic plates, but modern setups predominantly employ () cameras or video cameras integrated with image processing systems for higher precision and . These detectors allow for times typically ranging from 1 to 10 seconds per , enabling the capture of clear patterns while minimizing dose to the sample. For , -voltage (I-V) curves are acquired by systematically scanning the incident energy, often in steps of 1-2 over ranges such as 30-350 , while continuously monitoring and recording the of selected spots. This provides quantitative data on how spot brightness varies with energy, revealing information about surface atomic positions through subsequent comparison with theoretical models. Acquisition of full I-V datasets can be completed in under 20 minutes using low beam currents around 1 μA to preserve sample integrity. Quantitative measurements of spot intensities require analyzers to select electrons by energy and angle, commonly achieved with retarding field analyzers (RFAs) integrated into optics, which offer angular resolutions of approximately 1°. These devices use hemispherical or grid-based configurations to filter elastically scattered electrons before detection, ensuring high signal-to-noise ratios. Hemispherical analyzers may be employed in advanced setups for enhanced resolution, though RFAs remain standard due to their compatibility with display-type systems. Initial data processing involves software tools for automated analysis, including peak detection and fitting to extract spot profiles, as well as background subtraction to remove noise from or stray electrons. Programs such as those in the ViPErLEED package facilitate this by integrating over spot areas, applying Gaussian or fits, and normalizing intensities against incident beam current, streamlining the transition from raw images to usable I-V data. Common error sources in data acquisition include electron beam-induced damage, which can reduce spot intensities by 50-70% at higher energies due to desorption or structural alterations, and thermal drift causing pattern misalignment over time. Mitigation strategies encompass low-dose operation with reduced beam currents and short exposure times, sample cooling to around 100 K to stabilize the surface, and periodic flashing of the sample to ~500 K to remove contaminants just prior to measurement. These approaches ensure reliable data with minimal artifacts.

Theoretical Basis

Electron-surface interactions and sensitivity

Low-energy electron diffraction (LEED) achieves its surface specificity primarily through the limited penetration depth of the incident electrons, governed by their short (IMFP) in solids. In the typical energy range of 20–200 used for LEED, electrons undergo frequent events, losing approximately 90% of their energy within 5–10 of the surface, thereby confining the probing volume to the top 1–2 atomic layers. This IMFP follows a universal curve that applies across a wide range of materials, as described empirically by Seah and Dench, reaching a maximum of around 10–15 near 100 before slowly increasing at higher energies. The interactions between low-energy electrons and the surface can be categorized into and inelastic processes. preserves the electron's and is responsible for the coherent patterns observed in LEED, arising from of scattered by surface atoms. In contrast, inelastic scattering involves energy loss through mechanisms such as excitations (vibrational modes of the lattice), excitations (collective oscillations of conduction electrons), and the generation of with energies below 50 . These inelastic events contribute to a diffuse background in diffraction patterns and limit the escape of electrons from deeper layers, enhancing the technique's surface selectivity. Compared to bulk diffraction methods like X-ray diffraction, LEED exhibits enhanced backscattering from surface atoms due to their reduced , which alters the local scattering potential and increases the backscattered intensity relative to deeper, fully coordinated bulk atoms. In the backscattering geometry typical of LEED (near-normal incidence and detection), this effect contrasts with bulk probing techniques, where signals average over many layers without such surface enhancement. Several factors influence the surface sensitivity of LEED. The energy dependence is particularly notable: at lower electron energies (e.g., below 50 ), the IMFP decreases, weighting the signal more heavily toward the outermost atomic layer and improving resolution of changes. Additionally, the angle of incidence affects the effective path length through the surface; oblique angles increase penetration slightly but are less common in standard setups, where near-normal incidence maximizes backscattered intensity from the top layers.

Kinematic theory of single scattering

The kinematic theory of single scattering offers a foundational, idealized description of low-energy electron diffraction (LEED) patterns by assuming that incident electrons interact with the surface through a single scattering event only, neglecting any subsequent multiple scatterings within the crystal lattice. This model treats the incident and diffracted electron waves as plane waves and approximates the electron-atom interactions using muffin-tin potentials, where the potential is constant within spherical regions around each atom and zero elsewhere. Such assumptions simplify the problem to a first-order perturbation, enabling qualitative insights into surface periodicity without accounting for complex effects. The fundamental diffraction condition in this framework is governed by the surface-adapted Laue equations: \mathbf{k}_f - \mathbf{k}_i = \mathbf{G} where \mathbf{k}_i is the incident wavevector, \mathbf{k}_f is the final (scattered) wavevector with |\mathbf{k}_f| = |\mathbf{k}_i| due to , and \mathbf{G} is a two-dimensional vector of the surface . Unlike bulk , the surface geometry relaxes the strict three-dimensional momentum perpendicular to the surface, resulting in extended rods rather than discrete points. This allows for a range of out-of-plane momenta, with the in-plane components strictly conserved. The Laue condition is geometrically interpreted through the Ewald sphere construction in reciprocal space, where a sphere of radius $1/\lambda = |\mathbf{k}_i| (with \lambda the de Broglie wavelength) is drawn originating from the end of -\mathbf{k}_i. Diffraction spots become visible when this sphere intersects the reciprocal lattice rods, which extend infinitely perpendicular to the surface plane; each intersection satisfies the momentum transfer requirement and determines the direction of the diffracted beam. As electron energy increases, the sphere radius grows, potentially intersecting more rods and revealing additional spots in the LEED pattern, though the low energies typical of LEED (20–200 eV) limit the number of observable beams. Intensities in the kinematic model are estimated via the first Born approximation, which yields a scattering amplitude f(\theta) proportional to the Fourier transform of the scattering potential, leading to diffracted intensity I(\mathbf{G}) \propto |F(\mathbf{G})|^2, where the structure factor F(\mathbf{G}) is the sum over atomic positions \mathbf{r}_j in the unit cell: F(\mathbf{G}) = \sum_j f_j(\theta) e^{i \mathbf{G} \cdot \mathbf{r}_j}, with f_j(\theta) the atomic scattering factor for atom j. This approximation captures the modulation of spot brightness by the surface's atomic arrangement but ignores phase shifts from multiple interactions. The kinematic theory holds reasonably well for higher electron energies (above ~200 eV), where the lengthens and single dominates, or for simple surface structures with low elements exhibiting weak backscattering. However, it inadequately predicts intensity-energy profiles for typical LEED conditions involving complex surfaces, as multiple distorts the patterns significantly.

Dynamical theory of multiple scattering

The kinematic theory provides a useful starting point for understanding basic patterns in low-energy (LEED), but it inadequately describes intensity variations due to the neglect of multiple events. At low energies (typically 20–200 ), have a short of only a few atomic layers, resulting in strong backscattering and repeated interactions within the crystal lattice that significantly distort observed intensities. This necessitates a dynamical approach to accurately model the effects from these multiple s, which dominate over single-scattering contributions. The dynamical theory addresses this by solving the time-independent for the electron wavefunction in the periodic potential, treating the surface as a on a semi-infinite medium. Formulations often employ a multiple expansion, where the total wave is expressed as a of terms accounting for successive scatterings between sites, or the Korringa-Kohn-Rostoker (KKR) method, which uses techniques to propagate waves layer by layer across the . These approaches incorporate the full three-dimensional of electron propagation, enabling quantitative predictions of intensity-energy (I-V) curves that match experimental data far better than kinematic approximations. Key effects captured in the dynamical framework include thermal diffuse scattering (TDS), which arises from due to atomic vibrations and contributes to background intensity between Bragg peaks, and the Debye-Waller factor, e^{-2M}, which exponentially damps the coherent Bragg scattering amplitude, where M is the mean-square atomic displacement proportional to temperature. These vibrational influences are essential for interpreting temperature-dependent LEED patterns, as TDS can enhance diffuse features while the Debye-Waller factor reduces peak sharpness at higher temperatures. The theory also accounts for band structure influences by approximating the crystal potential with muffin-tin models, where the potential is spherically symmetric within spheres and constant in the interstitial regions, allowing derivation of phase shifts and accurate wavefunctions for scattering calculations. This approximation facilitates the inclusion of electronic band effects, such as surface resonances, which modulate diffraction intensities through constructive or destructive . Despite its precision, the dynamical theory remains computationally intensive, as the multiple scattering series or KKR matrix inversions scale unfavorably with the number of beams and layers, limiting applications to large unit cells without approximations like renormalized forward , which simplifies inter-layer propagation by treating forward-scattered waves perturbatively.

Pattern Analysis

Interpretation of diffraction patterns

In low-energy electron diffraction (LEED), the observed diffraction pattern provides direct insight into the surface periodicity and structure. For an unreconstructed surface, the pattern typically exhibits sharp integer-order spots corresponding to the vectors of the bulk-terminated , as seen in the clean Ag(110) surface where these spots align with the rectangular at energies around 60 eV. The positions of these spots, mapped on the Ewald sphere construction, encode the surface dimensions and orientations; specifically, the distances between spots in reciprocal space inversely relate to the real-space vectors via a relationship, allowing immediate determination of the two-dimensional parameters from a single image. Spot shapes in LEED patterns offer qualitative information about surface order and perfection. Sharp, well-defined spots indicate large, coherent ordered domains, typically exceeding the instrument's of 10-20 , as in high-quality epitaxial films where low background confirms minimal defects. In contrast, broadened or diffuse spots signal atomic-scale disorder, such as random adsorbate distributions or thermal vibrations, where the diffuse intensity distribution between spots can be analyzed to quantify local structural irregularities without resolving individual defects. Kikuchi lines, appearing as paired bands in the pattern due to double of inelastically scattered electrons, provide additional cues for surface orientation relative to the incident beam, aiding in multi-domain samples. Intensity variations across the pattern, particularly as a function of incident , reveal deeper structural details through I-V spectra measured for individual spots. These spectra exhibit oscillatory behavior arising from between waves scattered from successive layers, with the period of oscillations directly tied to interlayer spacings; for instance, peaks and minima correspond to constructive and destructive conditions that scale with the of , enabling rough estimates of layer distances before full dynamical modeling. Quantitative involves comparing experimental I-V curves to theoretical ones, where minima in the R-factor (a measure of spectral mismatch, ranging from 0 for perfect agreement to ~1 for random data) pinpoint optimal layer spacings and relaxations, as demonstrated in analyses of metal surfaces with R-factors below 0.2 indicating sub-angstrom precision. Surface reconstructions break the symmetry of the unreconstructed lattice, manifesting as additional fractional-order spots in the pattern that denote an enlarged supercell. These are conventionally labeled using Wood's notation, which specifies the substrate plane, supercell dimensions relative to the primitive cell, and any rotation; for example, the Si(111)-(\sqrt{7} \times \sqrt{7}) reconstruction appears as a \sqrt{7} superlattice of spots rotated by 19.1° from the integer array, reflecting dimer or adatom rearrangements that double or triple the unit cell area./06%3A_Overlayer_Structures_and_Surface_Diffraction/6.01%3A_Classification_of_Overlayer_Structures) Experimental artifacts must be distinguished from intrinsic structural features during pattern interpretation. Multiple coexisting , such as rotated or faceted regions on vicinal surfaces, often produce split spots where sub-peaks reflect the slight misalignment of domain-specific lattices, as observed in Te/(001) where splitting widths quantify domain size distributions down to nanometers. Elevated background intensity, appearing as a hazy glow between spots, primarily stems from events where electrons lose to phonons or plasmons without contributing to coherent , complicating spot profile analysis but correctable via energy filtering or background subtraction in quantitative studies.

Identification of surface superstructures and domains

Low-energy electron diffraction (LEED) identifies surface superstructures through the appearance of fractional-order diffraction spots, which arise from periodic arrangements in reconstructed clean surfaces or adsorbate overlayers that deviate from the primitive bulk lattice. These spots occur at positions corresponding to fractions of the vectors, such as half-order positions, reflecting the enlarged of the superstructure. For instance, a c(2×2) superstructure on an fcc(100) surface, often induced by adsorbates like oxygen on , produces additional spots at (1/2, 1/2) positions relative to the integer-order bulk spots, indicating a doubling of the surface area. Surface domains form when multiple orientations of the superstructure coexist, typically due to slight miscuts in the crystal orientation or step edges that favor different rotational variants. This results in rotated or split diffraction patterns, where spots from distinct domains overlap partially, leading to elongated or multiple sub-spots in the LEED image. The size of these domains can be quantified from the width of the diffraction spots using Fourier analysis of the spot profile, as broader spots indicate smaller, more disordered domains limited by defects or boundaries. Spot profile analysis low-energy electron diffraction (SPA-LEED) enhances this measurement by providing higher angular resolution to resolve subtle broadenings. Representative examples illustrate these phenomena: the clean W(100) surface exhibits a c(2×2) reconstruction below , with LEED spots at half-order positions due to a zigzag pairing of surface atoms, while the Au(111) surface shows a herringbone reconstruction characterized by a (22×√3) periodicity, producing satellite spots around integer orders from the uniaxial strain and rotational domains. Adsorbate-induced superstructures, such as CO on Cu(100), form compressed overlayers at high coverage that yield fractional spots consistent with a c(2×2) or distorted hexagonal arrangement, depending on temperature and exposure. Quantitative assessment of superstructures via LEED includes determining adsorbate coverage from the intensity ratios of fractional to integer-order spots, as the relative strengths scale with the ordered fraction of the surface . Phase transitions in superstructures are observed through temperature-dependent changes, such as spot sharpening upon cooling, which indicates increased domain coherence and long-range order, as seen in the reversible (√2×√2)R45° reconstruction of W(100) below approximately 280 K. Challenges in identifying true superstructures include distinguishing them from defect-induced patterns, such as random dislocations that mimic fractional spots through diffuse ; this requires high-resolution techniques like SPA-LEED to analyze spot shapes and backgrounds for of periodic versus aperiodic . Glancing incidence geometries in advanced LEED setups improve resolution for domain boundaries by increasing the effective path length of electrons near , enhancing sensitivity to lateral variations in superstructures.

Computational Methods

Dynamical LEED simulations

Dynamical LEED simulations utilize computational codes that implement full multiple-scattering theory to predict electron diffraction intensities from proposed surface structures. These codes solve the for electron wave propagation, incorporating effects such as refraction, interference, and backscattering within the crystal. A prominent example is the CLEED package, which performs intensity-voltage (I-V) analysis by computing diffraction amplitudes through iterative matrix inversions based on the equation for surface Green's functions. Similarly, the ViPErLEED employs tensor LEED methods to efficiently calculate I(V) curves, leveraging perturbations around reference structures for rapid evaluation of structural variations. Key input parameters for these simulations include the atomic coordinates defining the surface geometry, often initialized from experimental or theoretical models. Thermal vibrations are modeled using the Debye temperature (Θ_D), which determines mean-square displacements of atoms via the Debye-Waller factor; typical values range from 100–400 K for surface atoms, adjusted to fit experimental data. The electron-ion potential is another critical input, commonly derived from muffin-tin approximations or (DFT) calculations that include exchange-correlation functionals like the local density approximation (LDA) or generalized gradient approximation (GGA) to capture realistic scattering behavior. The primary output from dynamical LEED simulations consists of theoretical I-V curves for selected diffraction beams, representing intensity as a function of (typically 20–200 ). These are quantitatively compared to experimental spectra using the Pendry R-factor, defined as R = \frac{\int |y - x| \, dE + \lambda \int |\frac{dy}{dE} - \frac{dx}{dE}| \, dE}{\int y \, dE + \lambda \int |\frac{dy}{dE}| \, dE}, where x and y are the experimental and theoretical intensities, respectively, E is , and λ (≈0.4–0.7) weights the term to emphasize shapes; R-factors below 0.1 signify excellent agreement between theory and experiment. The simulation workflow begins with candidate structure models, often derived from kinematic approximations or complementary techniques like scanning tunneling microscopy. An initial full-dynamical calculation generates I-V curves, which are evaluated against experiment via the R-factor. Parameters are then iteratively refined using optimization algorithms such as least-squares minimization or genetic algorithms, converging on the structure that minimizes the R-factor while accounting for non-structural factors like inner potential and temperature. Software for dynamical LEED has evolved significantly since the 1970s, when slab methods—treating the crystal as a finite of layers with matching to functions—enabled the first quantitative determinations, as pioneered in early multiple-scattering codes. By the , optimized implementations like CLEED reduced computational demands through C-language efficiency and global search strategies. In the 2020s, modern packages such as ViPErLEED incorporate tensor expansions for faster computations on complex systems, with emerging GPU acceleration allowing simulations of large supercells (up to hundreds of atoms) in hours rather than days.

Quantitative structure determination

Quantitative structure determination in low-energy diffraction (LEED) involves comparing experimental intensity-voltage (I-V) curves, obtained by measuring the of diffracted beams as a function of , with theoretical curves simulated using dynamical models. The goal is to identify the coordinates that minimize the discrepancy between these datasets, typically through automated optimization techniques such as least-squares minimization of reliability factors (R-factors). R-factors include the Zanazzi-Jona, Pendry, and Van Hove-Tong metrics, which quantify the agreement between experimental and calculated intensities while accounting for noise and phase shifts. These optimizations often employ global search algorithms, such as , to explore the multidimensional parameter space of possible surface geometries, including interlayer spacings, lateral displacements, and rumpling (vertical shifts within atomic layers). The process iteratively refines structural parameters until the R-factor reaches a minimum, providing a quantitative measure of fit; values below 0.1-0.2 typically indicate reliable structures. For instance, in the analysis of the Al(110) surface, such methods revealed multilayer relaxations with contractions of up to 10-15% in the top interlayer spacing relative to bulk values. The precision of LEED-derived structures is generally high, with bond length determinations accurate to approximately 0.05 for vertical relaxations and rumpling, and slightly lower (0.1 ) for lateral positions, enabling detailed quantification of phenomena like multilayer contractions in metals or polar distortions in oxides. This level of accuracy has been benchmarked across numerous systems, where uncertainties arise primarily from statistical noise in I-V data and model assumptions, but systematic errors are minimized through multiple-beam analyses. In modern applications, such as oxide surfaces like Fe₃O₄(001), LEED has resolved complex reconstructions involving oxygen vacancies and metal-oxygen bond adjustments with similar precision. Error analysis in quantitative LEED highlights sensitivities to non-structural parameters, notably the inner potential, which represents the mean electron potential in the crystal and typically ranges from 10-15 ; shifts of this magnitude can alter apparent interlayer spacings by 0.05-0.1 if not properly accounted for. Other sources of uncertainty include Debye-Waller factors for thermal vibrations and surface disorder, which are mitigated by low-temperature measurements (e.g., below 100 K) and inclusion of multiple I-V curves from different incident angles. Pendry's , derived from the curvature of the R-factor minimum, provide a statistical estimate of these uncertainties, often yielding ±0.03 for key bond lengths. Recent advances integrate machine learning to accelerate structure searches, using neural networks or active learning schemes to predict promising candidate geometries from initial I-V data, reducing computational time from days to hours for complex systems. Hybrid approaches combining density functional theory (DFT) with LEED further enhance reliability by initializing searches with DFT-relaxed models, as demonstrated in 2020s studies of multielement oxide surfaces where such methods resolved subtle rumpling with R-factors under 0.05. These developments maintain LEED's status as a gold standard for surface crystallography while addressing challenges in high-dimensional searches.

Auger electron spectroscopy integration

Low-energy electron diffraction (LEED) systems are frequently integrated with Auger electron spectroscopy (AES) to enable simultaneous structural and compositional analysis of surfaces, leveraging shared optical components for efficient operation in ultrahigh vacuum environments. The concentric grids in a typical four-grid LEED optics serve as a retarding field analyzer (RFA) for detecting Auger electrons, which are emitted in the energy range of approximately 30–3000 eV following core-hole decay processes induced by the primary electron beam. This shared optics design, including a common electron gun and hemispherical grids, minimizes the need for additional hardware, allowing the same setup to function for both techniques without mechanical repositioning. The synergy between LEED and AES arises from their complementary capabilities: LEED provides diffraction patterns that reveal surface atomic structure and order, while AES maps elemental composition with a surface sensitivity of about 1% of a monolayer, probing the top 2–10 atomic layers. Operationally, switching between modes is achieved by adjusting bias voltages on the grids; in LEED mode, the grids are set to accelerate and transmit low-energy diffracted electrons (typically 20–200 eV) to the fluorescent screen, whereas in AES mode, a retarding potential is applied to the inner grids to filter electrons by energy, often using a small AC modulation (1–10 V) for derivative-mode spectroscopy that enhances peak visibility through dN/dE spectra. This mode-switching enables rapid alternation, with energy resolution typically 0.2–0.5% at low modulation voltages. In applications, the integrated LEED-AES setup is particularly valuable for monitoring adsorbate coverage and surface preparation, such as tracking oxygen adsorption on metal surfaces like or , where quantifies oxygen signals (e.g., the O KVV peak at ~510 ) while observes evolving superstructures like c(2×2) patterns. It also facilitates impurity detection during cleaning cycles, ensuring clean surfaces for structural studies by identifying contaminants like carbon at sub-monolayer levels. However, limitations include peak overlapping, where Auger transitions from different elements can complicate identification, and matrix effects that alter Auger yields due to changes in the local chemical environment, necessitating quantitative corrections via element-specific sensitivity factors.

Advanced LEED variants

Advanced variants of low-energy electron diffraction () have been developed to address limitations in computational efficiency, , and the ability to probe dynamic processes or spin-dependent phenomena. These extensions enhance the technique's applicability to complex surface structures, particularly in materials like nearly-free-electron metals and nanostructured systems. represents a perturbative to dynamical multiple-scattering , significantly reducing the computational demands of full LEED calculations. By employing transfer matrices to propagate wave functions layer by layer, it achieves speedups of up to two orders of magnitude—equivalent to a 90% or greater reduction in processing time—for systems where the potential is close to free-electron like, such as simple metals. This method, introduced in the late , facilitates quantitative structure determination by allowing rapid iteration over trial geometries without sacrificing accuracy in I-V curve comparisons. Spot profile analysis LEED (SPA-LEED) extends standard by providing high-resolution profiling of spots, enabling the characterization of surface sizes, distributions, and defects at micrometer scales. Developed in the 1980s, it uses electrostatic deflection to scan a finely focused (~1 μm ) across the sample, mapping reciprocal space variations that reveal antiphase boundaries and step distributions. Unlike conventional , which averages over millimeter-scale areas and is insensitive to local heterogeneities, SPA-LEED excels in and monitoring epitaxial growth dynamics, with a spanning seven orders of magnitude in intensity. Micro-LEED variants further localize probing by employing microfocused electron beams (down to ~5 μm) for selected-area diffraction, allowing investigation of inhomogeneous surfaces like polycrystalline or nanostructured films without averaging over large regions. This approach has been instrumental in resolving local crystallography in systems such as multilayer graphene on SiC, where it distinguishes domain orientations and stacking faults. Time-resolved LEED, often termed ultrafast LEED (ULEED), incorporates pulsed electron sources to capture surface dynamics on picosecond to femtosecond timescales. Radio-frequency compression of low-energy electron bunches enables few-picosecond resolution, while photoemission-based schemes achieve 100-fs pulses for pump-probe studies of structural transients, such as adsorbate vibrations or phase transitions. These setups, advanced since the , use laser-synchronized sources to probe non-equilibrium processes inaccessible to continuous-wave LEED. Quantitative low-energy electron diffraction (QLEED) automates structure refinement through integrated software packages that perform least-squares fitting of experimental I-V spectra to simulations, incorporating tensor approximations for efficiency. Tools like AQuaLEED streamline the process by handling symmetrized data and models, enabling reliable determination of atomic coordinates with minimized user intervention. Post-2010 developments include tip-enhanced LEED, where (STM) tips serve as field-emission sources for localized diffraction, achieving sub-micrometer resolution in for site-specific surface analysis. Additionally, synchrotron-based spin-polarized LEED (SPLEED) leverages circularly polarized photons to generate spin-oriented electrons, probing structures and spin-orbit effects with enhanced sensitivity at facilities like Elettra. These integrations expand LEED to and real-time nanomaterial studies.

References

  1. [1]
    [PDF] Critical Compilation of Surface Structures Determined by Low ...
    Oct 15, 2009 · This review critically compiles all surface structures derived from low-energy electron diffraction (LEED) crystallography reported in the ...
  2. [2]
    [PDF] Phys 446: Solid State Physics / Optical Properties - NJIT
    Low Energy Electron Diffraction (LEED) λ= h/p = h/(2mE)1/2. E = 20 eV → λ ≈ 2.7Å;. 200 eV → 0.87 Å. Small penetration depth (few tens of Å). – surface ...<|control11|><|separator|>
  3. [3]
    Low-Energy Electron Inelastic Mean Free Path of Graphene ... - NIH
    Sep 18, 2021 · The inelastic mean free path (IMFP) can be defined as the mean distance between two subsequent inelastic scattering events (measured along its ...
  4. [4]
    Fundamental Research at Bell Laboratories: The Discovery of ...
    Jan 1, 1981 · M. D. Fagen, ed., A history of engineering and science in the Bell System. ... Germer, "Low energy electron diffraction," Physics today, 77:7 ( ...
  5. [5]
    The birth and evolution of surface science - NIH
    The combination in the 1960s and 1970s of ultra-high-vacuum (i.e., P ... The “low-energy” electron diffraction (LEED) of back-scattered electrons in ...
  6. [6]
    [PDF] The phenomenon of the diffraction of electrons was first postulated ...
    This required the development of Ultra-High-Vacuum technology together ... [4] Low Energy Electron Diffraction by J. B. Pendry (Academic Press, 1974) ...
  7. [7]
    [PDF] First-Principles Computation of Low-Energy Electron Diffraction ...
    Low-energy electron diffraction (LEED) is an experimental tool that measures the intensity of low-energy electrons back-diffracted from crys- talline surfaces.
  8. [8]
    Low Energy Electron Diffraction - Wellesley College
    Electrons from an electron gun diffract off the sample and head back towards the electron gun and the grids surrounding it.
  9. [9]
    [PDF] High-resolution low-energy electron diffraction - TUE Research portal
    Jan 1, 1990 · This chapter is foliowed by a review of the kinematic theory of LEED. ... In this paper we report a surface morphology change on Ga.As(001).
  10. [10]
    [PDF] Ultrafast low-energy electron diffraction at surfaces - eDiss
    chapter provides an introduction to the basic principles of low-energy electron diffraction, as well as the concepts necessary to describe the investigated ...
  11. [11]
    [PDF] Low Energy Electron Diffraction - LEED
    LEED is a surface-sensitive technique using low energy electrons, where only electrons scattered from the near surface can leave. It forms diffraction patterns.Missing: principles | Show results with:principles
  12. [12]
    [PDF] Low-energy electron diffraction crystallography of surfaces and ...
    Two review articles by Heinz et al. [Hein94] and Over [Over98] provide good overviews and dis- cussions of LEED structure determinations of clean and ad-.
  13. [13]
    Discovery of Surface Phases by Low Energy Electron Diffraction ...
    This chapter describes low energy electron diffraction (LEED) patterns found after adsorption. LEED experiments shows that chemisorptions of simple ...
  14. [14]
    Low Energy Electron Diffraction - an overview | ScienceDirect Topics
    LEED is a technique using low-energy electrons to resolve surface structures, providing structural information of crystalline surfaces.
  15. [15]
    Graphene/Ni(111) surface structure probed by low-energy electron ...
    Oct 29, 2014 · The structure and the morphology of graphene on Ni(111) has already been studied by low-energy electron diffraction (LEED), density functional ...
  16. [16]
    Electric-field-driven octahedral rotation in perovskite - Nature
    Jan 8, 2021 · npj Quantum Materials volume 6, Article number: 5 (2021) Cite this article ... 2: Low-energy electron diffraction and angle-resolved ...
  17. [17]
    Designing Quantum Materials for Future Electronics
    Aug 21, 2025 · Designing Quantum Materials for Future Electronics. August 21 ... The beamline is equipped with 3D very-low-energy electron diffraction ...<|control11|><|separator|>
  18. [18]
    Diffraction of Electrons by a Crystal of Nickel
    ### Summary of Davisson-Germer Experiment (Phys. Rev. 30, 705)
  19. [19]
    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 ...Missing: details | Show results with:details
  20. [20]
    Quantum Milestones, 1927: Electrons Act Like Waves
    Feb 3, 2025 · Davisson and Germer showed that electrons scatter from a crystal the way x rays do, proving that particles of matter can act like waves.Missing: details | Show results with:details
  21. [21]
    Diffraction of Cathode Rays by a Thin Film - Nature
    A fine beam of homogeneous cathode rays is sent nearly normally through a thin celluloid film (of the order 3 × 10 6 cm. thick) and then received on a ...
  22. [22]
    Elastic Scattering of Low-Energy Electrons from Surfaces
    The electrons are scattered within a few atomic planes of the surface and their diffraction is intermediate between the kinematic and dynamic limits. ... Low- ...
  23. [23]
    The Evolution of Surface Chemistry. A Personal View of Building the ...
    From the 1960s new techniques permitted atomic and molecular studies of surfaces. The technique developments started in surface physics where applications ...
  24. [24]
    John Pendry: His Contributions to the Development of LEED Surface ...
    Such scaling could never be beaten by advances in computer speed, even at the rate of growth predicted by Moore's Law. Diffuse LEED John's first shot across ...
  25. [25]
    LEED intensity measurements and surface structures - IOP Science
    LEED intensity measurements and surface structures: The dynamical approach (Illustrated by Ni/chemisorbed Na system). S Andersson and J B Pendry.<|separator|>
  26. [26]
    A new pulse counting low‐energy electron diffraction system based ...
    Jan 1, 1992 · A new low‐energy electron diffraction (LEED) system has been constructed with a pulse counting position sensitive detector using channel ...
  27. [27]
    Combined EELS, LEED and SR-XPS study of ultra-thin crystalline ...
    We report the effect of annealing on the electronic energy loss and on the valence bands at different energies using synchrotron radiation XPS and LEED pattern ...
  28. [28]
    LEED Experiment (Chapter 3) - Surface Structure Determination by ...
    Aug 18, 2022 · This chapter describes specialised equipment and techniques used to perform LEED experiments and to measure intensities of diffracted LEED beams.
  29. [29]
    [PDF] Development of a High-Resolution Low Energy Electron Diffraction ...
    Aug 8, 1991 · The electron optics basically consist of a triode electron gun. a lens column of two identical einzel lenses El and E3 and a steering lens E2, ...
  30. [30]
    REVIEW ARTICLE Instrumentation for low-energy electron diffraction
    In this section we give a brief review of diffraction theory, with the purpose of putting into perspective the measure- ments that are required to obtain ...
  31. [31]
    (PDF) Low‐energy‐electron‐diffraction system using a high ...
    The system produces digitized images of LEED patterns as well as high‐resolution spot profiles of individual beams using incident currents in the picoampere ...
  32. [32]
    Low Energy Electron Diffraction (LEED) - SPECS Group
    Low Energy Electron Diffraction is a standard method to determine the long-range order, crystal structure and quality of sample surfaces.
  33. [33]
    [PDF] Bulk Single Crystal Growth and Sample Surface Preparation ... - OSTI
    During the study by Kwolek to determine the boundary conditions to reliably produce a clean surface by Ar ion sputtering and/or annealing in. UHV suitable ...
  34. [34]
    [PDF] LEEM and LEED Analysis of Ge(100) - UC Davis
    The technique, called sputtering and anneal- ing, is commonly used for ultra-high vacuum (UHV) experiments. Firing argon ions at the sample's surface acts as a ...Missing: single | Show results with:single
  35. [35]
    The growth of sulfur adlayers on Au(100) - AIP Publishing
    A Au(100) single crystal (diameter 1 cm) was mounted with W wires at the bottom of a manipulator; the sample was heated with resistive heating of the mounting ...
  36. [36]
    Evidence for in-gap surface states on the single phase SmB6(001 ...
    Oct 9, 2017 · Structural and electronic properties of the SmB6(001) single-crystal surface prepared by Ar+ ion sputtering and controlled annealing are
  37. [37]
    An electron spectrometer for LEED fine structure measurements
    LEED ... This spectrometer differs from others used for these type of measurements in that it used a retarding field analyzer to collect the elastically scattered ...
  38. [38]
    Wide-angle display-type retarding field analyzer with high energy ...
    Dec 12, 2017 · Wide-angle display-type retarding field analyzer with ... Auger electron spectroscopy (AES) using the LEED optics is also well established.
  39. [39]
    ViPErLEED documentation — ViPErLEED 0.14.1 documentation
    The ViPErLEED project comprises a set of open-source tools that aims at drastically reducing the effort for quantitative low-energy electron diffraction.
  40. [40]
    A Quantitative low-Energy Electron Diffraction Study | Phys. Rev. Lett.
    ... low-energy electron diffraction. Successful ... To minimize surface contamination, the sample was flashed to ∼ 500 K just prior to data acquisition.
  41. [41]
    Structure determination by low-energy electron diffraction—A roadmap to the future
    ### Summary of LEED Surface Sensitivity, Electron Interactions, and Mean Free Path
  42. [42]
    Quantitative electron spectroscopy of surfaces: A standard data base ...
    Abstract. A compilation is presented of all published measurements of electron inelastic mean free path lengths in solids for energies in the range 0–10 000 eV ...
  43. [43]
    Theory of Low Energy Electron Diffraction. III. Inelastic Scattering
    May 3, 2018 · The effects of diffraction of low energy electron beam on the inelastic scattering are predicted. The diffraction enhances or diminishes the ...
  44. [44]
    Low-energy electron inelastic mean free path and ... - AIP Publishing
    Feb 5, 2021 · This work shows that the classical electron trajectory framework still works for revealing the physics picture of low-energy electron ...
  45. [45]
    Low-energy electron diffraction
    The Ewald sphere illustrates the points of constructive interference formed by the incident and diffracted electron waves. The upper half of the sphere can be ...
  46. [46]
    The Physics of LEED | SpringerLink
    Low-energy electron diffraction is ... This is readily seen in the failure of “kinematic” theory (single-scattering theory) to explain LEED intensity vs.
  47. [47]
    Low-energy electron diffraction for surface structure analysis
    This review aims to give an introduction into the experimental, theoretical and analytical procedures needed for a successful structure determination.
  48. [48]
    Dynamical theory of reflection of very low energy electrons at crystal ...
    It is shown that the surface-state resonances, acting through the associated conditions of destructive interference, can have a dominant effect in very low ...Missing: backscattering | Show results with:backscattering
  49. [49]
    Multiple-Scattering Treatment of Low-Energy Electron-Diffraction ...
    A theory of intensities in the diffraction of low-energy (;5100-eV) electrons by crystals is formulated on the basis of Lax's multiple-scattering equations.
  50. [50]
    Dynamical LEED Theory | SpringerLink
    J. Korringa: Physica 13, 392 (1947). Article ADS MathSciNet Google Scholar. W. Kohn, N. Rostoker: Phys. Rev. 94, 1111 (1954). Article MATH ADS Google Scholar. K ...
  51. [51]
    The determination of surface debye temperatures from low-energy ...
    ... LEED spectra are calculated from dynamical theory (layer-KKR method) for a model of Ag{111} with a given value of θDeff and then the usual kinematic ...
  52. [52]
    (PDF) A Multiple Scattering Theory of Thermal-diffuse Scattering for ...
    This theory takes into account the multiple scattering of the electrons within the crystal, and is so formulated that it may be used in conjunction with present ...<|control11|><|separator|>
  53. [53]
    [PDF] Low-energy electron diffraction study of the thermal expansion of Ag ...
    The fact that the thermal diffuse scattering dominates the LEED pattern at temperatures above 720 K may be due to this. In summary, we have studied the ...
  54. [54]
    Dynamical calculations of LEED intensities at clean copper (001 ...
    The scattering properties of the individual atoms in the lattice are evaluated in the muffin-tin approximation for the periodic potential and inelastic ...
  55. [55]
    Dynamical and approximate LEED theories applied to layers of ...
    The description of the full dynamical theory is presented to demonstrate its limits of practical application. In turn approximative methods are described ...
  56. [56]
    Silicon quantum wires on Ag(1 1 0): Fermi surface and quantum well ...
    Fig. 1a shows the LEED pattern of the unreconstructed Ag(1 1 0) surface taken at 62.5 eV primary energy displaying clearly the characteristic integer spots of ...
  57. [57]
    Low-Energy Electron Diffraction: Experiment, Theory and Surface ...
    Low-Energy Electron Diffraction or LEED has become the prime tech nique used to determine atomic locations at surfaces.
  58. [58]
    LEED Kikuchi pattern: Phonon and plasmon contributions
    LEED Kikuchi pattern: Phonon and plasmon contributions ... We also find effects related to the orientation of the incident beam relative to the surface.
  59. [59]
    Low-energy electron diffraction study of the multilayer relaxation of ...
    Both theoretical studies predict relaxations as deep as the eighth interlayer spacing, although neither predicts relax- ations greater than 4% for these deeper ...
  60. [60]
    Structure determination by low-energy electron diffraction—A ...
    In this perspective article approaches will be discussed for extracting structural information from disordered and rough surfaces.
  61. [61]
    Domain wall structure of Te/Si(001) surface studied by LEED
    The splitting of diffraction spots has been reported in the case of O/Ni(110) [7], Li/Si(001) [8], Ge(111) [9] and so on, and has been explained by the ...
  62. [62]
    (PDF) Electron inelastic mean free path, electron attenuation length ...
    Aug 7, 2025 · The inelastic scattering of the primary electron in the electron gas of the material is introduced into the LEED theory in terms of the electron ...
  63. [63]
    Low-energy electron diffraction structural analysis of the (2 X 2 ...
    This LEED pattern can be caused by two fundamentally different adsorbate superstructures: a p(2X2) array of oxygen atoms (one-quarter monolayer coverage) or ...
  64. [64]
    Determination of domain size distributions from LEED beam profiles
    Aug 5, 2025 · For one-dimensionally disordered structures, the spot profiles for given domain length distributions can be calculated using the domain matrix ...
  65. [65]
    [PDF] A NEW LEED INSTRUMENT FOR QUANTITATIVE SPOT PROFILE ...
    The last condition is usually checked by the more or less qualitative inspection, if the pattern is "good", that means sharp spots and low background. On ...Missing: shape | Show results with:shape
  66. [66]
    LEED intensity analysis of the clean W(100) c(2×2) surface ...
    A LEED structure analysis is performed for the W(100) c(2*2) surface reconstruction. Models with pure vertical displacements of the surface atoms can be ...
  67. [67]
    Origin of the herringbone reconstruction of Au(111) surface at the ...
    Oct 5, 2022 · The herringbone reconstruction on Au(111) surface is first formed by densification into the stripe pattern and then the transformation into the ...
  68. [68]
    On the structure of CO adlayers on Cu(100) and Cu(111)
    Alternative models for the high coverage compressed overlayers of CO on copper (100) and (111) are proposed in which the LEED patterns previously attributed ...Missing: superstructure | Show results with:superstructure
  69. [69]
    (a) LEED patterns (at 53 eV) obtained for several Sb coverages ...
    At a coverage of about 1.0 ML, the intensity ratio has decreased to a very low value (1/10 th the maximum value), where only a (1 £ 1) LEED is observed.
  70. [70]
    [PDF] Growth mechanism and structure of ultrathin Pd films vapor ...
    The “beat” structure observed at 125 K is characterized by diffuse spots; the pattern sharpens only when the Pd layer is annealed above 300 K.
  71. [71]
    Liam-Deacon/CLEED: Computational package for Low ... - GitHub
    CLEED is a computational package for Low Energy Electron Diffraction (LEED) IV analysis. It fits experimental IV data curves with those simulated by the ...Missing: dynamical | Show results with:dynamical
  72. [72]
    ViPErLEED package I: Calculation of I⁢(V) curves and structural ...
    Jun 27, 2024 · Hammer, U. Diebold, and M. Riva, ViPErLEED package III: Data acquisition for low-energy electron diffraction, Phys. Rev. Research (2024).
  73. [73]
    Surface Structure of V2O3(0001): A Combined I/V-LEED and STM ...
    Sep 11, 2015 · The vibrational amplitudes of the atoms (and therefore their Debye temperatures) are relevant input parameters for I/V-LEED computations, and ...
  74. [74]
    Reliability factors for LEED calculations - IOPscience
    A new R-factor is defined which is insensitive to intensity discrepancies between theory and experiment but retains a simple functional form.
  75. [75]
  76. [76]
    [PDF] Automated determination of complex surface structures by LEED
    The challenges of structural complexity. 1 .I. Limits of conventional LEED crystallography. 1.2. Automation of LEED crystallography.
  77. [77]
    A simple and effective procedure for the refinement of surface ...
    The procedure requires no changes to existing LEED code, and has therefore a quite general applicability. The results of tests of the procedure are presented ...
  78. [78]
    Global search problem in LEED-IV surface structural analysis
    A simulated annealing algorithm is also one of the global search methods implemented in a LEED code completely written in the C language (CLEED) by Held et al.Abstract · Introduction · Control Parameters...
  79. [79]
    [PDF] Fast generation of quantum dynamics data using a GPU ... - arXiv
    A GPU implementation of the FDTD method simulates the time-dependent Schrodinger equation, achieving a 350x speed increase over a CPU, for electron diffraction.
  80. [80]
  81. [81]
    7.4: Low Energy Electron Diffraction - Chemistry LibreTexts
    Aug 28, 2022 · Low energy electron diffraction (LEED) is a ... ultra-high vacuum technology made possible the commercial availability of LEED instruments.
  82. [82]
    Global optimization in LEED structure determination using genetic ...
    By application of search algorithms, low energy electron diffraction (LEED) structure determination is now capable of solving relatively complicated systems.
  83. [83]
    Multilayer relaxation of the Al(110) surface - IOPscience
    The surface structure of Al(110) is determined via r-factor comparisons of new measurements at 100K of LEED intensity-energy spectra with spectra calculated ...
  84. [84]
    A combined DFT/LEED-approach for complex oxide surface ...
    Aug 5, 2025 · A combination of density functional theory (DFT) calculations and low energy electron diffraction (LEED) analysis is used to determine the ...
  85. [85]
    Low-Energy Electron Diffraction - an overview | ScienceDirect Topics
    Low-energy electron diffraction (LEED) is a technique used to study surface reconstruction by bombarding a sample surface with low-energy electrons ...
  86. [86]
    Some observations on the use of reliability indices in LEED ...
    Included in the latter are the energy dependence of the real part of the inner potential for Cu(111), and the individual variations of the imaginary potential ...
  87. [87]
    ViPErLEED package I: Calculation of $I(V)$ curves and structural ...
    Jan 3, 2025 · The focus is on the computational part of ViPErLEED, which performs highly automated LEED- calculations and structural optimizations.
  88. [88]
    Machine learning predicts atomistic structures of multielement solid ...
    Mar 4, 2024 · An active learning scheme for identifying stable surface structures of multielement solids under working conditions is proposed.
  89. [89]
    LEED 800 - Scienta Omicron
    The LEED 800 with integral shutter has LEED and AES capabilities using a miniature electron gun, set of concentric grids and a conductive, phosphor coated ...
  90. [90]
    LEED / Auger (AES) Spectrometers - OCI Vacuum Microengineering
    The LEED 600 MCP has LEED and AES capabilities using a miniature electron gun, set of concentric grids and a conductive, phosphor coated screen. In addition, ...Missing: shared retarding
  91. [91]
    Auger spectroscopy and low-energy electron diffraction studies of ...
    The Auger transition energies provide no general evidence for 'chemical shifts' but chemisorption leads to a significant modification of some Auger transition ...Missing: integration | Show results with:integration
  92. [92]
    Low energy electron diffraction and auger electron spectroscopy ...
    Low energy electron diffraction (LEED) studies of the structure of adsorbed molecules on crystal surfaces revealed that ordered surface structures ...
  93. [93]
    Chemical Effects in Auger Electron Spectroscopy - AIP Publishing
    Chemical effects in AES include shifts in Auger electron energies, changes in spectrum shape, and changes in relative intensities of Auger lines.
  94. [94]
    [PDF] CHAPTER 10 AUGER ELECTRON SPECTROSCOPY
    Because Auger features are much more pronounced in derivative spectra, such spectra are commonly used in qualitative AES investigations. Each element has a ...
  95. [95]
    Tensor LEED: A Technique for High-Speed Surface-Structure ...
    Dec 8, 1986 · A fundamental limitation of LEED as a surface-structure probe is the complexity of the theory compared with x-ray scattering from solids.
  96. [96]
    Tensor LEED I: A technique for high speed surface structure ...
    We present the theory and implementation of the new Tensor LEED technique for the rapid calculation of LEED I/V spectra. ... Low Energy Electron Diffraction.
  97. [97]
    Components - SPA-LEED
    The spot profile analysis low energy electron diffraction (SPA-LEED) instrument was developed in the 1980s by Prof. Martin Henzler and coworkers at the ...
  98. [98]
  99. [99]
    Stacking of adjacent graphene layers grown on C-face SiC
    Sep 6, 2011 · On one sample, a probing area of 5 μm can be used to collect a 1×1 LEED pattern from a multilayer graphene grain. ARPES is used to determine the ...
  100. [100]
    2}$ by low-energy electron microscopy and microprobe diffraction
    Apr 4, 2014 · However, real-space probing by LEEM in conjunction with -space probing by μ-LEED shows that the quality of CVD-grown can be comparable to that ...Missing: micro- | Show results with:micro-
  101. [101]
    Ultrafast Low-Energy Photoelectron Diffraction for the Study of ...
    A novel configuration for pump-probe low-energy electron diffraction uses photoelectrons from the substrate and achieves 100-fs time resolution.
  102. [102]
    Radio-frequency controlled electron pulses for time-resolved LEED ...
    We demonstrate radio-frequency compression and streaking of low-energy electron pulses for ultrafast diffraction from surfaces with few-ps time resolution.
  103. [103]
    LEED Calculation Home Page - icts - HKBU
    The LEED packages are designed for surface structure determination from experimental LEED IV curves. They are primarily used with ordered, commensurate surface ...Overview · Acknowledgments · Guide To The Selection Of A...Missing: CLEED | Show results with:CLEED
  104. [104]
    Quantitative Multilayer Cu(410) Structure and Relaxation ... - Nature
    Nov 15, 2019 · Starting from the (accurate) QLEED-determined structure (carried at 120 K), we can determine the corresponding electron distribution of the ...Missing: precision | Show results with:precision
  105. [105]
    Development of an advanced low-energy electron diffraction ...
    Aug 6, 2025 · Development of an advanced low-energy electron diffraction technique using field-emitted electrons from scanning tunneling microscope tips.
  106. [106]
    Low-energy spin-polarized electrons: their role in surface physics
    Light from a synchrotron source expands the experimental capabilities of momentum microscopy due to a wide range of a tunable photon energy and a selectable ...Missing: post- | Show results with:post-
  107. [107]
    Development of a spin polarized low energy electron diffraction system
    Feb 18, 2016 · We have designed and constructed a spin polarized low energy electron diffraction system working in the reflected electron pulse counting ...Missing: synchrotron post-