Fact-checked by Grok 2 weeks ago

Surface reconstruction

Surface reconstruction refers to the process by which the atoms at the surface of a rearrange into a structure different from the ideal truncation of the bulk crystal , primarily to minimize the surface arising from the reduced coordination and altered bonding environment at the surface. This phenomenon is distinct from surface relaxation, which involves smaller atomic displacements that preserve the surface's two-dimensional periodicity; , in contrast, changes this periodicity and often the of the surface . Surface reconstruction is a fundamental aspect of , influencing the physical, chemical, and electronic properties of materials, with implications for , adsorption, , and resistance. It is prevalent on clean surfaces of metals (e.g., the missing-row structure on face-centered cubic (110) surfaces like Au(110)) and semiconductors (e.g., the (2×1) dimer reconstruction on Si(100)). Adsorbate-induced reconstructions also occur, altering surface structure upon interaction with foreign atoms or molecules. The study of surface reconstruction relies on experimental techniques such as for determining periodicity and for atomic-scale imaging, complemented by theoretical methods like . Detailed mechanisms, notation systems, and specific applications in semiconductors, metals, and oxides are explored in later sections.

Fundamentals

Definition and basic principles

Surface reconstruction in crystal surfaces refers to the spontaneous rearrangement of atoms in the topmost layers into a periodic structure that deviates from the ideal truncation, characterized by a modified two-dimensional and often reduced compared to the underlying lattice. This structural change occurs to minimize the surface , which arises from the incomplete coordination of surface atoms. Unlike surface relaxation, where atomic layers shift rigidly—typically to the surface—without altering the in-plane periodicity, reconstruction involves lateral displacements and the formation of new bonds or motifs that fundamentally alter the surface periodicity. At the atomic scale, surface reconstruction is driven by the reduced of surface atoms relative to the , where atoms in the interior of a are surrounded by their full complement of nearest neighbors in a three-dimensional periodic . Cleaving a to create a surface breaks bonds, leaving unsaturated "dangling bonds" that increase the local energy due to altered interatomic forces and electronic states. To lower this energy, surface atoms reposition themselves, often forming dimers or other bonding configurations; for instance, on clean surfaces like , adjacent atoms pair into dimers to saturate one per atom, reducing the overall through sigma and pi bonding. The phenomenon was first observed in the late 1950s through (LEED) experiments on clean surfaces, where Schlier and Farnsworth reported non-primitive unit cells, such as the 2×1 on Si(100), indicating deviations from periodicity. A landmark advancement came in 1985 with the structural elucidation of the complex Si(111)-7×7 by Takayanagi et al., using to propose the dimer-adatom-stacking-fault model, which highlighted the intricate atomic rearrangements possible on clean surfaces. Understanding surface requires recognizing the distinction between structures, which minimize energy through full coordination, and surface structures, where the surface —defined as the excess per unit area due to bond breaking—drives these atomic-scale adaptations.

Driving forces and thermodynamics

Surface reconstruction is driven by the fundamental thermodynamic imperative to minimize the surface , which arises from the imbalance of atomic bonds at the between a and (or another ). This process allows surface atoms to rearrange laterally, forming new periodic structures that reduce the overall cost of the surface compared to an ideal, bulk-truncated configuration. The surface free energy \gamma is quantitatively expressed as \gamma = \frac{E_{\text{surface}} - N \cdot E_{\text{bulk}}}{2A}, where E_{\text{surface}} is the total energy of a slab model with two symmetric surfaces, N is the total number of atoms in the slab, E_{\text{bulk}} is the energy per atom in the bulk, and A is the area of one surface (the factor of 2 accounts for the two surfaces). Calculations show that reconstructed surfaces can lower \gamma by substantial amounts relative to unreconstructed ones, often by forming additional bonds or optimizing electronic structure. At elevated temperatures, entropic contributions become crucial, influencing the stability of reconstructed phases through the G = H - TS, where H is , T is , and S is (primarily vibrational). This can lead to temperature-dependent reconstructions or transitions, as higher-entropy configurations may become favored despite higher enthalpic costs; for instance, vibrational entropy stabilizes complex reconstructions like the Si(111)-7×7 up to high temperatures. Such effects are incorporated into thermodynamic models to predict diagrams under varying conditions. Surface stress, a second-order tensor \mathbf{f} defined as f_{\alpha\beta} = \frac{\partial \gamma}{\partial \varepsilon_{\alpha\beta}} (with \varepsilon denoting components), further drives by promoting lateral displacements to relieve anisotropic stresses inherent to . Tensile surface stresses, in many clean metal and surfaces, induce or rippling that lowers the total , as observed in systems like Au(111). This stress-relief mechanism is distinct from bulk but couples strongly with reconstruction energetics. In comparison to surface relaxation—which primarily involves atomic shifts and achieves modest energy reductions of ~0.01–0.1 per surface atom—reconstruction yields significantly larger gains, typically 0.1–1 per surface atom, due to more extensive bond reformation and stress alleviation. These differences highlight as a more profound response to surface instability.

Mechanisms of reconstruction

Intrinsic reconstruction

Intrinsic reconstruction refers to the spontaneous reorganization of atoms on clean surfaces, devoid of adsorbates, driven by the instability arising from the truncation of bulk bonds at the surface. This process involves atomic rearrangements such as , where surface atoms tilt to adjust bond angles, or dimerization, where pairs of atoms form new bonds to saturate dangling valences and reduce the number of unsatisfied bonds. These changes occur to minimize the by reforming bonds disrupted during surface creation, often leading to a lower than the bulk termination. Common patterns in intrinsic reconstruction include missing row structures, where alternate rows of atoms are absent, creating a furrowed surface; added row configurations, involving extra atoms protruding in ridges; and rotated domains, where surface layers twist relative to the bulk. These patterns, often denoted using Wood's notation such as (2×1) for doubled periodicity, are thermodynamically favored as they lower the surface energy compared to bulklike terminations and remain stable under conditions to prevent contamination. Achieving equilibrium structures requires overcoming kinetic barriers through , which is thermally activated and facilitated by annealing at elevated temperatures to enable atomic mobility without bulk diffusion. In contrast, most covalent semiconductors undergo significant due to the high energy of dangling bonds and directional bonding, whereas clean metal surfaces predominantly exhibit relaxation—small perpendicular shifts in interlayer spacings—owing to their delocalized electrons that screen charges and stabilize minor adjustments without lateral rearrangements.

Adsorption-induced reconstruction

Adsorption-induced reconstruction occurs when adsorbates interact with the substrate surface, leading to structural changes that minimize the total of the system. These interactions often involve the formation of chemical s between the adsorbate and surface atoms, which can induce local lattice strain due to differences in atomic size or electronic structure. For instance, in the case of hydrogen adsorption on the W(001) surface, the H-W formation causes a switching of tungsten dimer orientations, resulting in a reconstructed phase that alleviates strain through atomic rearrangements. Similarly, oxygen adsorption on Cu(100) involves charge transfer from to oxygen, generating compressive surface that drives the formation of overlayer structures, such as the c(2×2) phase at moderate coverages. These bond formations and charge transfers contrast with intrinsic reconstructions by being externally triggered, lowering the surface through adsorbate stabilization without altering bulk fundamentally. The extent of reconstruction depends critically on adsorbate coverage (θ), with distinct behaviors at low and high coverages. At low coverages (θ < 0.25), adsorbates typically occupy isolated sites, causing minimal global changes and localized distortions around adsorption sites, as seen in initial hydrogen binding on where isolated H atoms do not yet trigger widespread dimer flipping. As coverage increases (θ > 0.5), interactions between adsorbates lead to ordered overlayers, either commensurate (e.g., c(2×2) oxygen on , matching the substrate lattice) or incommensurate (mismatched lattices forming moiré patterns). Phase diagrams plotted against coverage and reveal transitions between these states; for H/, Monte Carlo simulations show a critical for , delineating disordered to ordered reconstructed phases. These coverage thresholds highlight how adsorbate-adsorbate repulsion or attraction amplifies strain, promoting collective reconstructions over isolated effects. Specific mechanisms for stress relief in adsorption-induced reconstructions include atomic rippling, where surface layers undulate to accommodate , and , where the surface develops tilted planes to reduce . On (100) with oxygen, compressive from O-Cu bonding is relieved via a missing-row reconstruction in the (√2×√2)R45° phase, where every other Cu row is removed, allowing lateral relaxation and a measured reduction of -0.6 N/m. , observed in organic molecule adsorption on vicinal metal surfaces, creates pyramid-like structures that distribute across facets, enhancing adsorbate binding stability. Reconstructions can be reversible or irreversible: reversible cases, like CO adsorption on Cu electrodes during CO2 reduction, involve dynamic nanoscale clustering that reforms upon desorption, driven by undercoordinated sites. Irreversible reconstructions occur when high-temperature annealing or strong bonding traps the surface in a new configuration, such as persistent missing rows after oxygen exposure. Temperature and pressure significantly influence these processes, often through desorption or vapor-phase interactions. Elevated temperatures promote desorption, which can reverse reconstructions by removing strain-inducing adsorbates; for example, heating H/Mo(100) leads to H desorption and restoration of the unreconstructed surface, with activation energies modulated by prior reconstruction. At ambient pressures, vapor-assisted methods enable controlled reconstructions, as demonstrated in 2025 studies on halide perovskites where vapor-deposited ligands isolate defective surface octahedra, suppressing ion migration and stabilizing the structure against environmental degradation. These effects underscore the dynamic nature of adsorbate-driven changes, where pressure variations in vapor exposure fine-tune coverage and thus the reconstructed morphology.

Description and notation

Standard notation systems

The standard notation systems for describing reconstructed surface structures primarily rely on conventions that capture the periodicity, symmetry, and orientation of the surface unit cell relative to the . These notations emerged to standardize the reporting of patterns and imaging data, enabling concise communication of structural changes without specifying atomic arrangements. Among them, Wood's notation remains the most widely adopted for commensurate reconstructions due to its simplicity and visual intuitiveness. Wood's notation, introduced by Elizabeth A. Wood in 1964, denotes a surface structure as \text{X}(hkl) \, M \times N - R \phi^\circ (\alpha \beta \gamma), where X(hkl) specifies the bulk-terminated surface plane. The parameters M and N represent the integer multiples of the unit cell vectors \vec{a_1} and \vec{a_2} that define the lengths of the reconstructed vectors \vec{b_1} and \vec{b_2}, respectively (i.e., |\vec{b_1}| = M |\vec{a_1}| and |\vec{b_2}| = N |\vec{a_2}|). The optional -R \phi^\circ indicates a of the reconstructed cell by \phi relative to the substrate cell, with R denoting the rotation direction (clockwise if specified). The parenthetical (\alpha \beta \gamma) provides additional details on directions, such as wood directions or basis vectors for centered lattices. Derivationally, it stems from the relating overlayer and substrate lattices, simplified for cases where the angle between \vec{b_1} and \vec{b_2} matches that between \vec{a_1} and \vec{a_2}, assuming commensurate structures with compatible symmetries. Rules include aligning \vec{b_1} parallel to \vec{a_1} when possible, selecting \vec{b_2} anticlockwise from \vec{b_1}, and omitting rotation if \phi = 0^\circ. For instance, a simple doubling in both directions is written as (2×2). Alternative notations address specific cases beyond basic Wood's format. For and centered cells, prefixes like p() denote a with periodicity, while c() indicates a centered rectangular or cell equivalent to two , often used for higher symmetry structures on square or hexagonal substrates. These derive from two-dimensional classifications and are integrated into Wood's notation when additional symmetry is evident from spots. For incommensurate structures, where vectors are not multiples, vector notation employs a \begin{pmatrix} [m](/page/M) & [p](/page/P′′) \\ [n](/page/M+) & [q](/page/M+) \end{pmatrix}, specifying \vec{b_1} = [m](/page/M) \vec{a_1} + [n](/page/M+) \vec{a_2} and \vec{b_2} = [p](/page/P′′) \vec{a_1} + [q](/page/M+) \vec{a_2}, with non- values allowed to capture slight mismatches. The historical evolution of these notations traces back to the 1960s, coinciding with the advent of (LEED) instruments that revealed patterns from clean and adsorbate-covered surfaces, necessitating a systematic vocabulary for non-bulk terminations. Early LEED studies in the mid-1960s prompted Wood's development to describe observed spot arrays efficiently, evolving from labels to formalized rules by the 1970s. Modern refinements incorporate scanning tunneling microscopy (STM) data since the 1980s, which provides real-space confirmation of periodicities, allowing notations to specify domain boundaries or subtle rotations more precisely without altering core formats. Despite their utility, standard notations like Wood's exhibit limitations in complex scenarios, such as quasicrystalline surfaces lacking true periodicity, where integer multiples and rotations fail to capture aperiodic tilings, leading to ambiguity in describing patterns with forbidden symmetries. In such cases, notations resort to approximate commensurate descriptions or extended matrix forms, but they cannot fully represent the non-repeating order inherent to quasicrystals.

Structural models

Structural models in surface reconstruction propose specific atomic arrangements that explain observed superlattice periodicity and symmetry. These models are typically constructed by interpreting diffraction patterns, such as low-energy electron diffraction (LEED) intensities or transmission electron microscopy (TEM) images, to infer bond lengths, layer relaxations, and rearrangements like adatom placement or faulting. For instance, the process begins with identifying the unit cell size and symmetry from reciprocal space data, then iteratively refining atomic coordinates to match experimental structure factors, often using kinematic or dynamical scattering theories. A seminal example is the dimer-adatom-stacking fault (DAS) model for the Si(111)-7×7 reconstruction, proposed based on ultra-high vacuum transmission electron diffraction (UHV-TED) that revealed intensity distributions consistent with stacking faults and adatom densities in a large unit cell containing 49 silicon atoms per layer. The complexity of these models varies from straightforward rearrangements to intricate multilayer configurations. Simple models, such as the missing-row reconstruction on fcc(110) surfaces like Au(110)-(1×2), involve the removal of every other row of surface atoms, reducing the density while doubling the periodicity along one direction; this was confirmed through intensity analysis. In contrast, multifaceted models incorporate multiple structural elements, such as the honeycomb-like in certain adlayer systems, where atoms alternate vertically in a to minimize , though such features are often part of larger reconstructions like those on Si(001). The exemplifies high complexity, featuring 12 adatoms, 18 dimers, and a central within the 7×7 , addressing both electronic passivation and strain relief. Validation of proposed models relies on cross-consistency with diverse experimental techniques, including agreement between diffraction-derived geometries and real-space imaging or spectroscopic signatures. Models must reproduce not only average structure factors but also Pendry R-factors below 0.5 for or phase shifts in TEM, while evolving through refinements as new data emerge; for the , initial TED-based proposal in 1985 was corroborated by high-resolution (STM) in 1986, which visualized protrusions at adatom sites and depressions at rest atoms, confirming the model's faulted half and dimer chains with atomic precision. Post-1980s advancements, such as quantitative and surface , further refined bond angles and relaxations in the DAS structure, reducing discrepancies in interlayer spacings by up to 10%. Modern validations include (DFT) calculations and, more recently, machine-learning force fields that confirm the energetic stability of the DAS model. Thermodynamic stability assessments, like comparing surface free energies, support , favoring those with minimal dangling bonds. Incommensurate reconstructions, where the surface superlattice does not align perfectly with the bulk , are modeled using networks of and to accommodate mismatch. These structures feature partial that relieve through localized , forming walls that separate commensurate , as seen in hexagonal reconstructions on fcc(100) metals like Pt(100)-hex, where a rotated overlayer is pinned by misfit every few unit cells. models, derived from satellite spots, predict diffuse scattering patterns arising from wall spacing variations, with typical densities of 10-20% mismatch leading to arrays spaced 5-10 bulk periods apart. Such models highlight the role of elasticity in stabilizing incommensurate phases over fully commensurate ones.

Experimental characterization

Diffraction techniques

Diffraction techniques probe the average long-range order of reconstructed surfaces by analyzing the periodic of waves from surface atoms, providing insights into dimensions and symmetry without direct visualization of individual atoms. These methods rely on to produce patterns in reciprocal space, which reflect the two-dimensional periodicity of the surface . Low-energy electron diffraction (LEED) is a primary technique for characterizing surface reconstructions, utilizing electrons with energies typically between 20 and 200 eV that undergo primarily from the top few atomic layers due to their short in solids. The resulting patterns, displayed on a hemispherical screen, consist of spots whose positions and intensities reveal the surface ; for instance, additional spots beyond the bulk lattice pattern indicate reconstructions such as a 2×1 periodicity on Si(100). Quantitative LEED refines structural models by measuring intensity-voltage (I-V) curves, where electron beam energy is varied to record spot intensities, enabling comparison with theoretical simulations for atomic position determination with Ångstrom precision./07%3A_Molecular_and_Solid_State_Structure/7.04%3A_Low_Energy_Electron_Diffraction) Surface X-ray diffraction (SXRD), often performed using sources, employs X-rays with energies above 5 keV, offering greater to access buried interfaces while maintaining surface sensitivity through grazing-incidence , where the beam skims the surface to enhance scattering from the top layers via total external reflection. This configuration minimizes bulk contributions and allows measurement of in-plane and out-of-plane atomic coordinates in reconstructed overlayers or substrate distortions. SXRD excels in determining precise atomic positions, including relaxations and rumplings, through analysis of rod scans—intensity profiles along rods perpendicular to the surface. LEED provides rapid screening of surface order with sub-Ångstrom resolution for determination, making it ideal for in-situ monitoring during preparation, whereas SXRD offers higher accuracy for quantitative structural refinement, particularly for complex or reconstructions, though it requires more sophisticated setups. Historically, LEED confirmed the first surface reconstructions in the late 1950s and 1960s, notably the Si(111)-(7×7) structure observed in , marking a milestone in recognizing deviations from bulk termination./07%3A_Molecular_and_Solid_State_Structure/7.04%3A_Low_Energy_Electron_Diffraction)

Real-space imaging

Real-space imaging techniques provide direct visualization of atomic-scale surface structures, offering complementary local information to the averaged data from methods. Scanning tunneling microscopy (STM) and (NC-AFM) are the primary tools for this purpose, enabling the resolution of reconstructed surface features such as adatom arrays and dimer rows on both conducting and insulating materials. STM operates by measuring the quantum tunneling current between a sharp metallic tip and the sample surface, which depends exponentially on their separation and reflects both topographic and electronic structure information. In constant-current mode, the tip height is adjusted to maintain a fixed tunneling current, yielding topographic maps of the surface; constant-height mode, conversely, records current variations at fixed tip position for faster imaging but requires atomically flat surfaces to avoid crashes. Lateral resolution reaches approximately 0.1 nm, sufficient to image individual atoms in reconstructions like the Si(111)-(7×7) dimer-adatom-stacking-fault model, first resolved in real space by in 1983. These images often reveal not just geometry but also local , distinguishing protrusions due to electronic effects in metallic or reconstructions. For insulating surfaces where is inapplicable due to lack of conductivity, NC-AFM detects short-range force interactions between the tip and sample without contact, using a oscillated near its frequency. detection measures shifts in the oscillation frequency caused by tip-sample force gradients, achieving atomic on wide-bandgap materials like oxides. Pioneered in 1995, NC-AFM resolved the Si(111)-(7×7) with 6 lateral and 0.1 vertical , confirming adatom positions via attractive van der Waals and repulsive Pauli forces. On insulators such as TiO₂(110), it images missing-row reconstructions and oxygen vacancies, providing structural details inaccessible to . Both techniques excel in characterizing defects within reconstructions, such as domain boundaries and step edges, which influence surface stability and reactivity. images of Ge(001)-(2×1) reveal migrating domain boundaries driven by vacancy diffusion, appearing as linear disruptions in the buckled-dimer array. On Si(111)-(7×7), step edges exhibit phase boundaries between reconstructed domains, with protrusions marking stacking faults or adatom clusters. NC-AFM similarly visualizes step-edge reconstructions on surfaces, like the (1×4) pattern on LaAlO₃(100), highlighting atomic rearrangements at edges. These local views confirm and refine structural models derived from by revealing heterogeneity. Post-2010 advancements in functionalization have enabled subatomic , enhancing for in reconstructions. Functionalizing the with a , as demonstrated in 2015 simulations and experiments, sharpens the interaction potential, allowing visualization of intramolecular features on surfaces like over reconstructed metals. This approach, building on earlier work, resolves submolecular details in adsorbates on Cu(111), distinguishing C-C via short-range chemical forces without altering the surface. Such enhancements have been applied to defect studies, distortions at boundaries in reconstructions.

Spectroscopic methods

Spectroscopic methods play a crucial role in characterizing the electronic and chemical alterations that accompany surface reconstruction, providing indirect evidence of structural changes through shifts in energy levels and vibrational signatures. Photoemission spectroscopy (PES), including X-ray photoemission spectroscopy (XPS) and angle-resolved photoemission spectroscopy (ARPES), detects core-level shifts that arise from modifications in the local potential and bonding environment at reconstructed surfaces. These shifts, typically ranging from 0.5 to 2 eV, reflect changes in atomic coordination and charge redistribution; for instance, on the reconstructed Si(001) c(4×2) surface, the Si 2p core-level components exhibit shifts of about -0.5 eV for up-dimers and +0.1 eV for down-dimers relative to bulk, indicating the asymmetric dimer geometry. ARPES further reveals reconstruction-induced band structure alterations, such as the opening of band gaps or dispersion changes, by mapping the momentum-resolved electronic states near the surface. Auger electron spectroscopy (AES) complements PES by probing surface composition and detecting reconstruction-related peak shifts in Auger transitions, which are sensitive to the valence electron density and local chemistry. In semiconductor surfaces like Si and GaAs, AES identifies Fermi-level shifts of 0.2–0.5 eV following cleaning procedures that stabilize reconstructions, such as HF etching, where the shifts correlate with changes in surface band bending and adatom arrangements. These shifts, often smaller than core-level ones in PES, arise from alterations in the final-state screening during the Auger process and can quantify the degree of surface passivation or reconstruction. AES is particularly useful for monitoring adsorbate-induced reconstructions, as peak intensity ratios and positions change with coverage and structural ordering. Vibrational spectroscopies, notably high-resolution (HREELS), elucidate reconstruction effects by resolving adsorbate vibrational modes and surface phonons that are influenced by the altered lattice dynamics. In adsorption-induced reconstructions, such as the Ni(100)/O system forming a c(2×2)-p4g structure, HREELS identifies low-frequency modes around 30–50 meV associated with oxygen bridge-bonding and displacements, providing direct vibrational evidence for the reconstructed geometry. Similarly, reflection absorption (RAIRS) can detect frustrated translation or rotation modes of adsorbates on reconstructed metal surfaces, with frequency shifts indicating bond strengthening or weakening due to periodic distortions. in these techniques relies on or frequency differences—core-level shifts of 0.5–2 eV in PES/AES and vibrational mode displacements of 5–20 meV in HREELS—as reliable signatures of reconstruction, often cross-verified with structural data from or .

Theoretical approaches

Ab initio calculations

Ab initio calculations, primarily based on (DFT), offer a first-principles approach to predict and elucidate surface reconstructions by computing the ground-state and total of atomic systems without empirical parameters. In this framework, the Kohn-Sham equations are solved self-consistently to approximate the , enabling the determination of stable atomic configurations that minimize the free energy. For surface systems, DFT utilizes periodic slab models, which represent the surface as a finite stack of atomic layers (typically 5–15 layers thick) separated by vacuum regions (around 15–20 Å) to isolate the surface from its periodic images and simulate the asymmetry of a . These models maintain three-dimensional periodicity, with the vacuum ensuring negligible interaction between slabs. Geometry optimization proceeds via total energy minimization, where the Hellmann-Feynman forces guide iterative relaxation of atomic positions until convergence criteria are met, often yielding reconstructed structures with altered bond lengths and angles driven by surface stress relief. Prominent implementations include the (VASP) and , both of which support plane-wave basis sets and projector-augmented wave methods for efficient handling of periodic systems. For reconstructions involving adsorbates, van der Waals corrections—such as DFT-D3 or optB88-vdW functionals—are incorporated to account for dispersion interactions, improving the description of weak bindings beyond standard generalized gradient approximations. DFT predictions typically achieve high fidelity, with optimized geometries accurate to within ~0.05–0.1 of experimental bond lengths and surface formation energies precise to ~0.05–0.1 per atom, as benchmarked on metal and surfaces. Stable structure searches, often employing evolutionary algorithms or basin-hopping methods interfaced with DFT, systematically explore configuration space to identify low-energy reconstructions by evaluating total energies across candidate geometries. Recent advances integrate to accelerate these computations, exemplified by the MAGUS tool, which combines and evolutionary search with surrogate models to reduce the number of expensive DFT evaluations needed for high-throughput prediction of surface reconstructions. As of 2025, further progress includes for simulating larger-scale and dynamic surface reconstructions, as well as data-driven workflows for predicting structures on complex materials like oxides and systems. Thermodynamic energies derived from these calculations, including surface free energies under varying chemical potentials, further inform stability assessments. Such predictions are routinely validated against experimental structures from or techniques.

Phenomenological models

Phenomenological models provide approximate frameworks for understanding surface reconstruction by incorporating empirical or continuum approximations to capture key physical behaviors, such as stress relief and electronic bond adjustments, without the full quantum mechanical detail required for methods. These models are particularly valuable for simulating large-scale systems or predicting trends in reconstruction patterns, where (DFT) becomes computationally prohibitive. They often rely on simplified parameters derived from experiments or lower-level calculations to model the driving forces behind atomic rearrangements on surfaces. Elastic theory treats surface reconstruction as a response to intrinsic surface stress using continuum mechanics, where the surface layer experiences tensile or compressive forces that are relieved through lateral contractions or expansions. In this approach, the strain ε in the surface is approximately related to the surface stress σ and the bulk Young's modulus Y by ε ≈ σ / (Y ⋅ d), where d is the effective thickness of the stressed surface layer (typically ~1–3 Å, on the order of interatomic spacing), allowing estimation of reconstruction-induced distortions propagating into subsurface layers. For instance, on the Si(001) surface, this model explains the formation of stress domains in the 2×1 dimer reconstruction, where alternating orientations minimize long-range elastic interactions, as demonstrated through calculations showing domain sizes on the order of hundreds of angstroms. Similarly, for metal surfaces like Au(111), the herringbone reconstruction arises from anisotropic surface stress relief, with continuum elasticity predicting the periodic dislocation network that stabilizes the 22×√3 pattern by reducing compressive strain in the top layer. These models highlight how surface stress, often on the order of 1-10 N/m, drives reconstructions to lower the total free energy. Tight-binding approximations simplify the electronic structure by modeling surface atoms with localized orbitals and empirical hopping integrals, focusing on bond saturation to predict stable reconstruction geometries. This method captures the tendency of undercoordinated surface atoms to form dimers or chains, reducing dangling bonds and electronic energy penalties. A classic application is the prediction of the buckled dimer structure on the Si(100) 2×1 reconstructed surface, where tight-binding total energy calculations favor asymmetric dimers due to charge transfer between atoms. Such approximations enable efficient exploration of multiple reconstruction candidates, revealing that electronic relaxation energies can dominate over strain costs in covalent semiconductors. Monte Carlo simulations within phenomenological frameworks, often combined with Ising-like models or gas approximations, investigate transitions and in reconstructed surfaces by sampling configurations based on effective Hamiltonians that include elastic and interaction terms. These simulations reveal critical behaviors, such as order- transitions, where disrupt periodic reconstructions above certain temperatures. For the W(001) surface, methods using parameterized potentials have shown a second-order from the reconstructed c(2×2) to a disordered around 550 K, with simulations reproducing experimental proliferation and effects. By incorporating surface stress anisotropies, these approaches also model adsorption-induced reconstructions, providing insights into how adsorbates pin or depin reconstruction domains. Overall, phenomenological models offer scalable tools for rapid prototyping of mechanisms, serving as benchmarks against more precise DFT results for validation.

Applications and examples

Semiconductor surfaces

Semiconductor surfaces, particularly those of group IV elements like and , exhibit pronounced reconstructions driven by the need to minimize high-energy dangling bonds inherent to their covalent bonding. The Si(100) surface, for instance, forms a 2×1 reconstruction consisting of buckled dimers, where surface silicon atoms pair into asymmetric dimers with one atom slightly elevated relative to the other, effectively reducing the surface energy by forming partial π-bonds. This buckled dimer model, proposed by Chadi in 1979, is energetically favored over symmetric dimers or π-bonded chain alternatives, yielding an energy gain of approximately 1.2 eV per dimer compared to the unreconstructed surface. In contrast, the Si(111) surface adopts a more complex 7×7 reconstruction, the largest primitive observed on elemental semiconductors, spanning 2.7 nm × 2.7 nm. The accepted dimer-adatom-stacking-fault (, introduced by Takayanagi et al. in 1985 through diffraction analysis, incorporates 12 adatoms arranged in a 2×2 within the , along with stacking faults and dimer rows that saturate 42 of the original 49 dangling bonds present in the ideal termination. This intricate arrangement not only passivates most unsaturated bonds but also introduces a subtle metallic character to the otherwise semiconducting surface, as validated by subsequent calculations. The Ge(111) surface similarly reconstructs into a c(2×8) pattern at , featuring structural elements akin to the Si(111) but distinguished by honeycomb chains of atoms that form extended networks, reducing density through lateral bonding. Unlike the Si case, this reconstruction displays temperature-dependent behavior, maintaining the c(2×8) order below approximately 300°C before transitioning to a higher-temperature (2×1) with increased disorder. studies confirm the long-range order and atomic positions in this model, highlighting its role in stabilizing the surface against thermal fluctuations. Post-2010 developments have emphasized hydrogen-terminated variants of these reconstructions for , where controlled passivation of remaining dangling bonds on (100) and (111) surfaces enables atomically precise fabrication of nanowires and quantum devices. These H-terminated structures preserve the underlying reconstruction while providing chemical stability, facilitating applications in scalable silicon-based as demonstrated in first-principles thermodynamic models of morphologies.

Metal surfaces

Surface reconstructions on metal surfaces, particularly noble metals like , and , are generally weaker and more delocalized compared to those on semiconductors, owing to efficient electronic screening that reduces the driving forces for structural changes. These reconstructions often arise from surface stress relief or adsorbate interactions, leading to periodic modulations in atomic positions over large domains. On clean surfaces, such patterns manifest as long-range herringbone or hexagonal arrangements, while adsorbates can induce localized distortions like rumpling or missing rows. Experimental techniques, such as (LEED) and scanning tunneling microscopy (STM), have imaged these domains, revealing their role in surface energetics and reactivity. A prominent example is the Au(111) surface, which exhibits a (22 × √3) herringbone reconstruction driven by intrinsic surface . This structure features alternating face-centered cubic (fcc) and hexagonal close-packed (hcp) stacking domains, forming a pattern with elbows and solitons that accommodate . The involves a uniaxial of approximately 4.4% along the [1̅10] direction, compressing the topmost layer from 2.88 Å to 2.76 Å nearest-neighbor spacing, which lowers the surface by about 0.1 per surface atom. Stress domains in this pattern arise from the spontaneous formation of alternating high- and low- regions, as confirmed by elastic continuum models and atomistic simulations. On , the clean Pt(100) surface undergoes a () reconstruction, forming a quasi-hexagonal overlayer that deviates from the bulk . This consists of a rotated top layer, with a small 0.7° misalignment relative to the underlying above 1100 K, stabilized by the missing-row model where every fourth row of atoms is absent, reducing through closer packing. The hexagonal phase is metastable and can be lifted by adsorbates like , transitioning to the (1×1) via rotational and translational . This enhances catalytic activity but is sensitive to environmental conditions. Adsorbate-induced reconstructions are exemplified by oxygen on Cu(100), where a c(2×2) pattern forms at 0.5 coverage through adsorption in four-fold hollow sites. This configuration induces rumpling of the outermost Cu layer, with Cu atoms beneath oxygen atoms displaced outward by up to 0.2 Å to relieve lateral stress and optimize bonding, as determined by calculations. The rumpled structure stabilizes the adsorbate layer and influences subsequent oxidation pathways, contrasting with subsurface incorporation at higher coverages. Recent investigations in 2023 have highlighted dynamic reconstructions on surfaces under electrochemical conditions, such as potential-dependent lifting of the herringbone phase in acidic electrolytes, revealing transient striped phases that impact electrocatalytic performance. These studies, using operando , underscore how applied potentials modulate surface stress and adsorbate binding, leading to reversible structural fluctuations on and electrodes during or reduction reactions.

Oxide and other materials

Surface reconstructions in oxide materials often arise from the need to mitigate polar discontinuities and ionic imbalances at the surface, leading to complex structural rearrangements distinct from those in simple metals. In rutile TiO₂(110), the (1×2) reconstruction is prominently described by the added Ti₂O₃ row model, where protruding rows of Ti₂O₃ units form along the direction atop the unreconstructed (1×1) termination, effectively doubling the unit cell periodicity. This configuration was first proposed based on scanning tunneling microscopy (STM) observations and has been corroborated by density functional theory (DFT) calculations, which assign it as the lowest-energy structure among competing models. Oxygen vacancies significantly influence this reconstruction, as their formation under reducing conditions promotes Ti interstitial diffusion from subsurface layers to the surface, stabilizing the added rows and altering local electronic properties such as band gap narrowing. Perovskite oxides like exhibit reconstructions driven by octahedral tilts of TiO₆ units, which help relieve surface stress and . Common structures include the (2×1) and (√2×√2)R45° phases, where alternating tilts in and subsurface layers create a rumpled termination with TiO₂-rich or SrO-rich facets, as revealed by (LEED) and (TEM). These tilts, inherited from bulk antiferrodistortive transitions but amplified at , modulate electronic conductivity and facilitate formation in heterostructures. In hybrid systems involving 2D materials, moiré patterns induced by lattice mismatch serve as effective surface reconstructions, modifying electronic and catalytic properties. For graphene grown epitaxially on Ru(0001), a large-scale moiré superlattice emerges due to a ~3% mismatch, forming a (25×25) graphene unit cell over a (23×23) Ru substrate, which buckles the graphene sheet and creates periodic potential variations. This structure, often approximated in smaller commensurate models like √3×√3 for simplified analyses, pins the graphene layer strongly to the metal, enabling applications in spintronics and catalysis through site-specific reactivity at high-symmetry points of the moiré. Battery materials, particularly oxide cathodes, undergo Li-induced surface restructuring during electrochemical cycling, forming protective gradient layers that mitigate degradation. In Li-rich layered oxides like Li₂MnO₃-based cathodes, delithiation triggers phase transitions at the surface, evolving from layered to or rock-salt structures over nanometer-thick gradients, which suppress oxygen release and dissolution. Studies from the highlight how these reconstructions, observed via aberration-corrected TEM, enhance cycle life by creating compositionally graded interfaces with higher Li⁺ , as demonstrated in dual-gradient designs achieving over 80% capacity retention after 500 cycles.

Industrial applications

Surface reconstruction plays a pivotal role in by enabling the formation of active sites on surfaces, particularly for oxidation reactions. In (Pt) and gold (Au) s supported on ceria (CeO₂), dynamic restructuring under CO oxidation conditions leads to faceted structures that enhance catalytic activity; for instance, small Au s (∼2 nm) exhibit perimeter reconstruction, creating low-coordination sites that lower the for CO adsorption and oxidation compared to larger particles. Similarly, Pt s on CeO₂ undergo surface restructuring to form a Pt skin layer, achieving an apparent activation barrier of 39 kJ/mol for CO oxidation, which outperforms bulk Pt surfaces due to optimized oxygen vacancy interactions. These reconstructions are crucial for , where stable faceting prevents deactivation and improves selectivity. In electronics manufacturing, controlled surface reconstruction of () is essential for epitaxial growth, ensuring defect-free layers in devices. During () of , the (2×1) reconstructed Si(001) surface facilitates layer-by-layer growth by minimizing dangling bonds, which is critical for producing high-quality films used in integrated circuits and power MOSFETs. Reconstructed templates also enable precise placement of quantum dots; for example, hydrogen-passivated Si(001):H surfaces with controlled reconstruction direct the self-assembly of InAs quantum dots, improving uniformity for optoelectronic applications like lasers. This preparation step is industrially scaled in () processes to achieve superjunction structures for high-voltage transistors. For energy storage, stabilizing surface reconstructions suppresses defects in battery electrodes and perovskite solar cells, enhancing longevity and efficiency. In all-solid-state batteries, surface reconstruction of oxide cathodes, such as LiNi₀.₈Co₀.₁Mn₀.₁O₂, forms a protective layer that mitigates volume changes during cycling, enabling over 1000 cycles at high voltage (>4.3 V) with 90% capacity retention. In perovskites, nano-polishing-induced reconstruction passivates wide-bandgap surfaces (e.g., MAPb(I₀.₆Br₀.₄)₃), reducing non-radiative recombination and achieving certified efficiencies of 23.7% for single-junction cells and 33.1% for four-terminal tandems, with devices retaining 80% efficiency after 1505 hours of illumination. As of 2025, perovskite-silicon tandem solar cells have achieved certified efficiencies up to 34.6%. Recent vapor-assisted surface reconstruction methods have enabled perovskite solar modules stable for 45 days under outdoor conditions. These modifications address ion migration and degradation, supporting scalable production for grid-scale energy systems. In , engineered drives in thin films, enabling patterned nanostructures for advanced devices. Temperature-controlled in block copolymer thin films, such as -block-poly(4-vinylpyridine), adjusts domain sizes from 10-50 nm by altering interfacial energies, facilitating directed for lithographic templates in . This approach is applied in fabricating uniform arrays on reconstructed surfaces, promoting ordered deposition for sensors and devices without top-down . Despite these advances, industrial adoption faces challenges in and in-situ control of surface reconstruction under operational conditions. High-temperature or electrochemical environments often induce uncontrolled , complicating in large-scale reactors, as seen in electrocatalysts where dynamic changes require to maintain activity. Limited in-situ techniques, such as , hinder precise modulation, while anion regulation strategies show promise but demand further optimization for cost-effective, durable implementations in and energy devices.

References

  1. [1]
    A survey on surface reconstruction based on 3D Gaussian splatting
    Aug 5, 2025 · Surface reconstruction is a foundational area in computer graphics and vision, aiming to generate complex and accurate surfaces from sparse ...
  2. [2]
    Surface Reconstruction: Roles in the Field of Computer Vision and ...
    Surface Reconstruction is the most potent aspect of 3D computer vision. It allows recapturing or imitating of the shape of real objects.
  3. [3]
    [PDF] REVIEW OF SURFACE RELAXATION AND RECONSTRUCTION ...
    Surface reconstruction. 6.1. Metal surfaces. The surfaces (001) of the bcc transition metals W, Mo and Cr reconstruct below room temperature. The ...
  4. [4]
    Surface reconstruction of InAs (001) depending on the pressure and ...
    Sep 6, 2017 · Surface reconstruction occurs to minimize the surface energy (γ) of a specific crystallographic plane by altering some of its bonding ...
  5. [5]
    [PDF] Principles of Semiconductor Surface Reconstruction
    Dec 28, 1994 · The term "recon- struction" is used for surface structures which exhibit a lower symmetry parallel to the surface than the bulk structure.
  6. [6]
    Atomic and Electronic Structures of Reconstructed Si(100) Surfaces
    Jul 2, 1979 · New structural models for 2×1 and 4×2 reconstructed (100) surfaces of Si determined from energy-minimization claculations are presented.
  7. [7]
    Structure and Adsorption Characteristics of Clean Surfaces of ...
    R. E. Schlier, H. E. Farnsworth; Structure and Adsorption Characteristics of Clean Surfaces of Germanium and Silicon. J. Chem. Phys. 1 April 1959; 30 (4): 917– ...
  8. [8]
    [PDF] Chapter 4 DFT Calculations for Solid Surfaces - Refubium
    with inversion symmetry), the surface energy, γ at T = 0 K of a clean surface, can then be calculated as, γ = 1. 2A. (Etotal slab − NsEtotal bulk ) . (4.8).
  9. [9]
    Crucial role of vibrational entropy in the surface structure stability
    Jun 15, 2022 · Our study shows that vibrational entropy plays a crucial role in the stabilization of 7 × 7 reconstruction at elevated temperatures, and ...
  10. [10]
    The role of surface stress in reconstruction, epitaxial growth and ...
    Several examples are presented where the surface stress is either the driving force for the reconstruction of the surface or is at least facilitating the ...
  11. [11]
    [PDF] Stress Relief in Reconstruction - UMD Physics
    Jun 2, 1997 · For both surfaces the reconstruction reduces the tensile stress. This is not surprising, since the average bond distances between the surface ...
  12. [12]
    [PDF] Influence of material, surface reconstruction and strain on ... - TIB
    From thermodynamic considerations, the island formation is energetically favourable because the energy, which is gained from a partial strain relaxation by ...
  13. [13]
  14. [14]
    SURFACE RECONSTRUCTION AND ,CATALYSIS - Annual Reviews
    surface reconstruction, as oxidation and reduction will change atomic distances and restructure the surface with the period of the turnover frequency (82).Missing: definition | Show results with:definition<|control11|><|separator|>
  15. [15]
    Reconstruction of Clean and Adsorbate-Covered Metal Surfaces
    Effects of lattice strain on noble metal (110) surface: Missing row reconstruction and adsorption properties. Applied Physics Letters 2021, 119 (14) https ...
  16. [16]
    Surface Relaxation - an overview | ScienceDirect Topics
    Surface relaxation is defined as small interlayer spacing changes in clean metal surfaces, typically involving the top layer of atoms moving inward towards ...<|control11|><|separator|>
  17. [17]
    (PDF) Adsorbate-induced surface reconstruction and surface-stress ...
    Aug 6, 2025 · These results indicate that surface stress reduction plays a role in causing the reconstruction. The discrepancies between theory and experiment ...
  18. [18]
    Adsorbate-induced surface reconstruction and surface-stress ...
    Oct 3, 2006 · Both experiment and theory show that oxygen adsorption leads to a compressive change in the surface stress that is larger for the overlayer.
  19. [19]
    Adsorption-induced surface reconstruction: predictions of a simple ...
    The phase diagram of the real system is reproduced only at low coverage (θ < 0.4). The phase diagram is obtained by the mean-field approximation as well as by ...
  20. [20]
    Correlations in coverage-dependent atomic adsorption energies on ...
    May 18, 2009 · First, the adsorption energy of an adsorbate is often dependent on the coverage due to bonding competition with the surface, or repulsive ...
  21. [21]
    Nanoscale patterning, macroscopic reconstruction, and enhanced ...
    Jan 16, 2017 · In conclusion, the change of surface stress during faceting of a metal surface induced by adsorption of an organic molecule has been measured ...Missing: relief | Show results with:relief
  22. [22]
    Spectroscopic Observation of Reversible Surface Reconstruction of Copper Electrodes under CO2 Reduction
    ### Summary of Reversible Reconstruction Induced by CO Adsorption on Cu Surfaces
  23. [23]
    The effect of surface reconstruction on desorption: H/Mo(100)
    We have measured adsorption isobars for the system H/Mo(100) and from them we have extracted the energy of activation, Ed, and the frequency factor, ν, ...
  24. [24]
    Vapor-assisted surface reconstruction enables outdoor-stable ...
    May 29, 2025 · We addressed this issue by isolating defective octahedra at the perovskite surface using a vapor-deposited polydentate ligand. Surface octahedra ...
  25. [25]
    [PDF] Surface crystallography - physics.muni
    Examples of Wood's Notation. Masel, „Principles of Adsorption… „ p.80-82. Page ... The (7 x 7) reconstruction of the Si(111) surface. DAS (Dimer-Adatom ...
  26. [26]
    [PDF] Crystallography of Quasicrystals
    The part “Methods” discusses experimental techniques for the study of real quasicrystals as well as power and limits of methods for their structural analysis.<|control11|><|separator|>
  27. [27]
    Surface Crystallography - an overview | ScienceDirect Topics
    The noted X-ray crystallographer, Elizabeth A. Wood (1963), developed the first systematic notation for surface crystallography.
  28. [28]
    Structure analysis of Si(111)-7 × 7 reconstructed surface by ...
    The atomic structure of the 7 × 7 reconstructed Si(111) surface has been analysed by ultra-high vacuum (UHV) transmission electron diffraction (TED).
  29. [29]
    A new model for the reconstructed (110)-1×2 surfaces of Ir, Pt and Au
    Starting from a discussion of the “missing row model” for the reconstructed (110)-1×2 surfaces of Ir, Pt and Au, we come to the conclusion that the ...Missing: seminal paper
  30. [30]
    Structural analysis of Si(111)‐7×7 by UHV‐transmission electron ...
    May 1, 1985 · A new structural model of Si(111)‐7×7 is derived. The model basically consists of 12 adatoms arranged locally in the 2×2 structure, nine dimers on the sides of ...Missing: DAS 7x7
  31. [31]
    Surface reconstruction of Au(001): High-resolution real-space and ...
    Jul 28, 2014 · The hexagonal reconstruction of Au(001) has been a model system for the understanding of complex metal surface reconstructions, and it has ...
  32. [32]
    Diffraction from reconstructed surfaces with incommensurate domain ...
    This paper generalizes the kinematic approximation to an (N × 1) reconstructed surface with a distribution of various types of lateral displacements at a step.Missing: dislocations | Show results with:dislocations
  33. [33]
    The birth and evolution of surface science: Child of the union ... - PNAS
    Although the existence of the Si(111)–(7 × 7) structure had been discovered in 1959 by LEED, the determination of its atomic geometry and the origin thereof had ...
  34. [34]
    Introduction | SpringerLink
    Dec 10, 2013 · LEED complemented with dynamical intensity calculation and first principle density function calculation has been used to study surface ...
  35. [35]
    Surface X-ray Diffraction (Chapter 7)
    The surface slab consists of the adsorbate layers or reconstructed layers as well as distorted substrate layers down to a depth where deviations from the ...
  36. [36]
    Surface X-Ray Diffraction - ResearchGate
    Surface X-ray diffraction (SXRD) is a method to observe the atomic configuration or molecular orientation at surfaces or interfaces (Robinson in Handbook on ...
  37. [37]
    7 × 7 Reconstruction on Si(111) Resolved in Real Space
    Jan 10, 1983 · The 7 × 7 reconstruction on Si(111) was observed in real space by scanning tunneling microscopy. The experiment strongly favors a modified adatom model.<|control11|><|separator|>
  38. [38]
    Atomic Resolution of the Silicon (111)-(7×7) Surface by ... - Science
    The reconstructed silicon (111)-(7×7) surface was imaged in a noncontact mode by force microscopy with atomic resolution (6 angstroms lateral, 0.1 angstrom ...
  39. [39]
    NC-AFM Imaging of Surface Reconstructions and Metal Growth on ...
    Abstract. There has been an explosion of interest in metal oxide surfaces in the past decade. This reflects the technological importance of these surfaces ...
  40. [40]
    Vacancy diffusion mediated dynamics of domain boundaries on Ge ...
    Jul 3, 2024 · Domain boundaries often play a pivotal role in surface dynamics and surface phase transitions. In this study we scrutinize the dynamics of ...
  41. [41]
    High-resolution imaging of LaAlO3(100)–(1 × 4) reconstructed ...
    Jul 16, 2018 · Non-Contact Atomic Force Microscopy (NC-AFM), which is capable of imaging insulating surfaces with atomic resolution, can be used to ...
  42. [42]
    CO tip functionalization in subatomic resolution atomic force ...
    Oct 22, 2015 · A clear advance in nc-AFM technology is the use of a CO-functionalized tip, which has produced high resolution images for molecules and graphene ...
  43. [43]
    Unique Determination of “Subatomic” Contrast by Imaging Covalent ...
    Mar 14, 2014 · The origin of so-called “subatomic” resolution in dynamic force microscopy has remained controversial since its first observation in 2000.
  44. [44]
    [PDF] Mechanism of high-resolution STM/AFM imaging with functionalized ...
    Aug 19, 2014 · High-resolution STM/AFM imaging uses functionalized tips, and probe relaxation due to tip-sample interaction causes submolecular resolution. ...
  45. [45]
    [PDF] Surface core-level shifts on clean Si(001) and Ge(001) studied with ...
    Mar 31, 2010 · In this paper calculated surface core-level shifts (SCLS) are used to aid in the decomposition of high-resolution Si 2p and Ge 3d core-level ...Missing: spectroscopic | Show results with:spectroscopic
  46. [46]
    Photoemission studies of surface core-level shifts and their ...
    Surface core-level shifts measured using photoemission with synchrotron radiation are reviewed for a variety of transition metals (Ta, W, Ir), rare earth metals ...Missing: spectroscopic | Show results with:spectroscopic
  47. [47]
    [PDF] Detecting Fermi-level shifts by Auger electron spectroscopy in Si ...
    Dec 25, 2017 · In this paper, changes in surface Fermi-level of Si and GaAs, caused by doping and cleaning, are investigated by Auger electron spectroscopy.
  48. [48]
  49. [49]
    Vibrational evidence of surface reconstruction in the Ni(100)/O ...
    Oct 15, 1990 · Vibrational evidence of surface reconstruction in the Ni(100)/O system obtained using high-resolution electron-energy-loss spectroscopy.
  50. [50]
    (PDF) Surface Reconstruction of W2C (0001) - ResearchGate
    Aug 7, 2025 · In the off-specular HREELS, an atomic vibration of 44.8 meV (361 cm − 1) appeared within the gap energy region of the bulk phonon bands, which ...
  51. [51]
    Density functional theory: Its origins, rise to prominence, and future
    Aug 25, 2015 · This paper reviews the development of density-related methods back to the early years of quantum mechanics and follows the breakthrough in their application ...
  52. [52]
    Advances in Density-Functional Calculations for Materials Modeling
    Jul 1, 2019 · Here, we review the recent advances in DF calculations for materials modeling, giving a classification of modern DF-based methods when viewed ...
  53. [53]
    [PDF] First steps with a periodic DFT code : Quantum Espresso
    − building of a slab model for a surface. − determination of the surface energy. − study of the surface reconstruction. − characterization of adsorption ...
  54. [54]
    Benchmarking DFT Accuracy in Predicting O 1s Binding Energies on ...
    Oct 8, 2025 · Across several common DFT methods, the accuracy of DFT predictions for binding energies of O/metal surfaces remains similar.
  55. [55]
    Advances and challenges in treating van der Waals dispersion ...
    Sep 24, 2012 · An accurate description of dispersion is highly challenging, with the most widely used electronic structure technique, density functional theory (DFT), failing ...
  56. [56]
    Adsorption studies of C6H6 on Cu (111), Ag (111), and Au (111 ...
    Oct 4, 2012 · ... energies that are within 0.1 Å and 0.05 eV, respectively, of experimental values. Similarly, in Tonigold and Groß's QM:QM hybrid DFT-D ...
  57. [57]
    [2212.11549] Prediction of surface reconstructions using MAGUS
    Dec 22, 2022 · Condensed Matter > Materials Science · Title:Prediction of surface reconstructions using MAGUS · Bibliographic and Citation Tools · Code, Data and ...Missing: accelerated DFT 2023
  58. [58]
    Prediction of surface reconstructions using MAGUS - AIP Publishing
    May 3, 2023 · Once given the chemical formula, an EA program can automatically determine the most stable structures under different external conditions.
  59. [59]
    [PDF] Theoretical study of the Si(100) surface reconstruction
    10. We estimate the energy difFerence between the p(2 x 1)s and p(2 x 1)areconstructions, the buckling energy, to be. 0.12+0.01 eV. This value was obtained ...
  60. [60]
    (PDF) Toward a Detailed Understanding of Si(111)-7×7 Surface and ...
    Sep 11, 2025 · (a) The image extends over an area of 30 nm × 30 nm. The amplificatory (7 × 7) unit cell was indicated in the inset.
  61. [61]
    Surface structure and long-range order of the Ge(111)-c(2×8 ...
    Nov 15, 1988 · We have performed an x-ray diffraction study of the Ge(111)-c(2×8) surface. From the analysis of 43 in-plane, symmetry-inequivalent ...Missing: honeycomb chains
  62. [62]
    Surface core-level shifts on Ge (111) c (2× 8): Experiment and theory
    Aug 7, 2025 · A careful analysis of photoelectron spectra, in combination with DFT calculations of surface core-level shifts for silicon atoms in ...
  63. [63]
    Predictive Morphology Control of Hydrogen-Terminated Silicon ...
    Jan 7, 2014 · We present a theoretical model based on density functional theory first principles thermodynamics for the optimal shape and surface reconstruction of hydrogen- ...
  64. [64]
    Atomically Precise Manufacturing of Silicon Electronics - PMC
    The review presented here will focus on details related to APM and Hydrogen Desorption Lithography (HDL) from silicon surfaces with atom defined precision in ...
  65. [65]
    (PDF) Surface Reconstruction: Noble Metals - ResearchGate
    Aug 10, 2025 · The paper reviews the most important achievements in the study of the surface-reconstruction phase transition on the low-index single ...
  66. [66]
    Elastic stress domains and the herringbone reconstruction on Au(111)
    Sep 7, 1992 · We suggest that the herringbone reconstruction of the Au(111) surface results from the spontaneous formation of stress domains.Missing: theory | Show results with:theory
  67. [67]
    [PDF] Coexistence of the (23 × x3) Au(111) Reconstruction and a Striped ...
    The study found that the Au(111) reconstruction can coexist with a striped phase monolayer, contrary to the belief that it deconstructs upon monolayer ...Missing: 22x √ stress
  68. [68]
    Structure and local reactivity of the Au(111) surface reconstruction
    Aug 7, 2025 · The close-packed (111) surface of gold is well known to show a 22×√3 reconstruction on single nm lengths with a long-range herringbone pattern on scales of a ...
  69. [69]
    Mechanism and Dynamics of the CO-Induced Lifting of the Pt(100 ...
    Feb 14, 2003 · This reconstruction, normally referred to as the hex phase, which undergoes a slight rotation of 0.7 ° above about 1100 K, restructures to the ...Missing: formation | Show results with:formation
  70. [70]
    CO oxidation on Pt and Pd model catalysts - RSC Publishing
    Jun 7, 2017 · To lower its surface free energy, the cubic Pt(100) surface exhibits a (111)-like, hexagonal reconstruction when it is free from adsorbates. Two ...
  71. [71]
    [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 ...
  72. [72]
    Density functional study of oxygen on Cu(100) and Cu(110) surfaces
    Feb 23, 2010 · ... thermodynamics of surface reconstruction for this technologically important material. The combined oxygen-dependent free energies of the ...
  73. [73]
    A surface extended x-ray absorption fine-structure study - PubMed
    Structure of c(2 x 2) oxygen on Cu(100): A surface extended x-ray absorption fine-structure study. Phys Rev B Condens Matter. 1985 Feb 15;31(4):2532-2534.Missing: checkerboard rumpling
  74. [74]
    Noble Metal-Based Heterogeneous Catalysts for Electrochemical ...
    Feb 8, 2023 · In this review, we comprehensively discussed the research progress on noble metal-based heterogeneous electrocatalysts used in water electrolysis for hydrogen ...
  75. [75]
    Direct in-situ insights into the asymmetric surface reconstruction of ...
    Feb 22, 2024 · The rutile TiO2 (110) surface has been reported to reconstruct into TiO2 (110)-(1×1), TiO2 (110)-(1×2), and pseudo-hexagonal rosette structures, ...
  76. [76]
    Octahedral rotations in strained LaAlO3/SrTiO3 (001) heterostructures
    Feb 11, 2014 · Symmetry-dependent interfacial reconstruction to compensate polar discontinuity at perovskite oxide interfaces (LaAlO3/SrTiO3 and LaAlO3 ...
  77. [77]
    Thickness-Dependent Perovskite Octahedral Distortions at ...
    May 22, 2019 · In this study, we analyze how the octahedral tilts and rotations of thin films of LaNiO3 and LaAlO3 grown on different substrates, ...
  78. [78]
    Graphene on Ru(0001): A Supercell | Phys. Rev. Lett.
    Sep 19, 2008 · Here we show, using surface x-ray diffraction (SXRD), that graphene forms a surprising superstructure when grown on Ru(0001), whereby 25 × 25 ...
  79. [79]
    Growth Mechanism of Graphene on Ru(0001) and O2 Adsorption on ...
    Apr 16, 2009 · There coexist large islands with typical Moiré patterns at monolayer graphene structure on metals (12-14) and nanographene clusters with an ...Missing: 3x | Show results with:3x
  80. [80]
    Surface Reconstruction Enhanced Li-Rich Cathode Materials for ...
    Sep 18, 2024 · Based on the molten salt-assisted sintering method, a lithium rich cathode material with surface rich Ni and Co is designed and prepared.
  81. [81]
    Argonne builds on past success with cathode design for lithium-ion ...
    Oct 29, 2024 · A breakthrough “dual-gradient” design for the battery's cathode substantially increases energy storage capacity, stability and lifetime while reducing costs.
  82. [82]
    Size-dependent dynamic structures of supported gold nanoparticles ...
    Jul 9, 2018 · We report size-dependent structure dynamics of single Au nanoparticles on ceria (CeO 2 ) in CO oxidation reaction condition at room temperature.
  83. [83]
    Surface restructuring and predictive design of heterogeneous catalysts
    Nov 22, 2024 · Surface restructuring is defined as the breakup of surface metal-metal or metal-nonmetal bonds that leads to a reorganization of the surface ...
  84. [84]
    Overcoming activity/stability tradeoffs in CO oxidation catalysis by Pt ...
    Aug 12, 2025 · The use of redox active metal oxides to support noble metals is critical in the design of highly-active CO oxidation catalysts for gas ...
  85. [85]
    Initial stages of silicon molecular beam epitaxy: Effects of surface ...
    Jan 1, 1985 · Initial stages of silicon molecular beam epitaxy: Effects of surface reconstruction. ... High energy ion scattering/channeling and low energy ...
  86. [86]
    Placement Control of Nanomaterial Arrays on the Surface ...
    We present a control strategy for the facile placement of densely packed nanomaterial arrays (ie, nanoparticles and nanorods) on surface reconstructed ...
  87. [87]
  88. [88]
    Surface Reconstruction Enables High‐Voltage, Long‐Life All‐Solid ...
    Aug 25, 2025 · All-solid-state batteries (ASSBs) offer enhanced energy density and improved safety through the utilization of solid electrolytes.
  89. [89]
    Surface reconstruction of wide-bandgap perovskites enables ...
    Dec 4, 2024 · We develop a surface reconstruction method by removing the defect-rich crystal surface by nano-polishing and then passivating the newly exposed high- ...
  90. [90]
    Surface Reconstruction of Perovskites with Organosilanes for High ...
    Sep 23, 2024 · Surface or interface engineering is one of the most effective strategies to improve the device performance and stability of perovskite solar ...Missing: storage batteries<|separator|>
  91. [91]
    Feature size control using surface reconstruction temperature in ...
    Here we present a method to control the size of the openings in hexagonally organized BCP thin films of poly(styrene)-block-poly(4-vinylpyridine) (PS-b-P4VP) ...
  92. [92]
    In-depth understanding and precise modulation of surface ...
    Jan 9, 2025 · Precise modulation of heterogeneous electrocatalyst surface reconstruction can be achieved through three dimensions, namely pre-catalyst ...
  93. [93]
    Progress in Manipulating Dynamic Surface Reconstruction via Anion ...
    Aug 8, 2023 · Anion modulation represents a potential direction for controllable extent of surface reconstruction and enhancement of reconstructed ...