Fact-checked by Grok 2 weeks ago

Surface science

Surface science is the interdisciplinary field dedicated to the study of physical and chemical phenomena that occur at the interfaces between different phases, including solid–gas, solid–liquid, and solid–vacuum boundaries. This discipline examines how materials behave at their surfaces, where properties often differ markedly from the bulk due to factors such as arrangement, , and interactions with adsorbates. Key concepts include adsorption (the binding of molecules to surfaces), (the excess energy associated with surface atoms), and (spontaneous rearrangement of surface atoms to minimize energy). These phenomena underpin processes like , , and , making surface science essential for understanding and engineering material interfaces. Emerging in the as a confluence of advances in physics, , and , surface science was propelled by the of ultra-high vacuum (UHV) systems, which maintain pressures below 10⁻⁹ to enable clean surface studies. Early milestones included the recognition of surface-sensitive electron spectroscopies in 1968, such as (LEED) for surface crystallography and for elemental composition. The field's evolution accelerated in the 1980s with computational tools like for modeling surface interactions, and in the 1980s with the invention of scanning tunneling (STM) in 1982, which saw widespread adoption in the 1990s. Techniques such as (XPS) for chemical state analysis, (FTIR) for molecular vibrations, and scanning electron (SEM) for topography, developed primarily in the mid-20th century, are often conducted under UHV conditions to minimize . These methods allow precise of surface (e.g., average roughness Ra < 0.5 μm for smooth interfaces) and oxide layers, such as TiO₂ films on titanium. Surface science has profound applications across multiple domains, driving innovations in microelectronics, enabling the scaling of semiconductor devices to feature sizes below 10 nm as of the late 2010s, in an industry that exceeded $200 billion in value by 2000 and has since grown significantly. In catalysis, it elucidates reaction mechanisms on heterogeneous catalysts, enhancing efficiency in chemical production and environmental remediation. Biomedical applications include designing biocompatible implants, such as dental and orthopedic prosthetics with over 95% success rates after five years, through surface modifications like acid etching and plasma spraying to improve tissue integration. Additionally, it informs corrosion resistance in materials engineering and electrochemistry in energy storage devices, while ongoing research extends to complex systems like liquid-solid interfaces and biological surfaces.

Fundamentals

Definition and Scope

Surface science is the interdisciplinary study of the physical and chemical phenomena occurring at the interfaces between different phases, such as solid-gas, solid-liquid, solid-solid, and liquid-gas boundaries. At these interfaces, material properties deviate significantly from those in the bulk due to reduced atomic coordination, leading to unique behaviors like enhanced reactivity and altered electronic structures that are highly sensitive to external conditions such as temperature, pressure, and adsorbate presence. The scope of surface science encompasses atomic-scale interactions, including adsorption, diffusion, and reaction dynamics, as well as the thermodynamics of surfaces that distinguish them from bulk materials science. For instance, surface energy arises from the imbalance of interatomic forces at the interface, influencing phenomena like wetting and crystal growth, while the quantifies the relationship between surface excess concentration and interfacial tension: \Gamma = -\frac{1}{RT} \left( \frac{\partial \gamma}{\partial \mu} \right)_T where \Gamma is the surface excess, \gamma is the surface tension, \mu is the chemical potential, R is the gas constant, and T is temperature. This equation highlights how solute accumulation at interfaces lowers surface tension, a core contrast to bulk properties where such effects are negligible. Surface science bridges physics, chemistry, materials science, and engineering by integrating tools from condensed-matter physics and physical chemistry to probe and manipulate interfaces. A key example is how clean surfaces exhibit reactivity orders of magnitude higher than contaminated ones; for instance, trace contaminants like carbon monoxide can poison catalytic sites on metal surfaces, drastically reducing reaction rates in processes such as hydrocarbon conversion. This field emerged as a distinct discipline in the mid-20th century, enabled by advancements in vacuum technology that allowed controlled studies of well-defined, contamination-free interfaces.

Key Concepts and Interfaces

Surface science encompasses the study of interfaces between different phases, where properties diverge significantly from the bulk due to reduced coordination and symmetry breaking. Key interfaces include solid-vacuum, solid-liquid, gas-liquid, and solid-solid types, each exhibiting distinct behaviors governed by interfacial energies and interactions. The solid-vacuum interface represents an ideal case for ultra-high vacuum (UHV) studies, minimizing contamination and enabling precise probing of clean surface structures without ambient interference. In contrast, solid-liquid interfaces involve wetting phenomena, quantified by Young's equation, which relates the contact angle \theta to interfacial tensions: \gamma_{sv} = \gamma_{sl} + \gamma_{lv} \cos \theta, where \gamma_{sv}, \gamma_{sl}, and \gamma_{lv} are the solid-vapor, solid-liquid, and liquid-vapor surface tensions, respectively; this equation predicts partial wetting for \theta between 0° and 180°. Gas-liquid interfaces, often stabilized by surfactants, reduce surface tension through adsorption, facilitating processes like emulsification and foam formation by lowering the energy barrier for phase separation. Solid-solid interfaces, exemplified by epitaxial growth, involve oriented atomic alignment across the boundary, driving layer-by-layer deposition in heteroepitaxial systems like semiconductor heterostructures. Surface thermodynamics at interfaces is dominated by surface free energy \gamma, the excess energy per unit area due to incomplete bonding, which dictates stability and morphology. For clean surfaces, atoms rearrange to minimize \gamma: relaxation involves subtle bond length adjustments in the topmost layers, while reconstruction entails larger-scale atomic repositioning, as seen in the Si(111) 7×7 structure, where a dimer-adatom-stacking-fault model reduces dangling bonds from 49 to 19 per unit cell, lowering \gamma by approximately 1.2 J/m². The equilibrium shape of crystals minimizes total surface energy, described by the Wulff construction: planes perpendicular to radii proportional to \gamma in each direction form the bounding polyhedron, favoring low-\gamma facets like {111} in face-centered cubic metals. This principle explains faceted morphologies in nanoparticles and explains why high-\gamma orientations are absent in equilibrium crystals. Kinetic processes on surfaces, such as diffusion and nucleation, control non-equilibrium evolution and are thermally activated. Surface diffusion follows an Arrhenius form, D = D_0 \exp(-E_a / kT), where D_0 is the pre-exponential factor (typically 10^{-4} to 10^{-3} cm²/s for adatoms), E_a is the activation barrier (0.5–1.5 eV for metals), k is , and T is temperature; this governs adatom hopping across terraces, influencing growth uniformity. For thin film nucleation, classical theory for 2D islands posits a free energy barrier \Delta G^* = \pi \gamma^2 / \Delta g, where \Delta g is the free energy gain per unit area (\Delta g = \Delta \mu / \Omega, \Delta \mu the chemical potential supersaturation per atom, \Omega the area per atom), with critical radius r^* = \gamma / \Delta g; nucleation rates scale as \propto \exp(-\Delta G^* / kT), favoring 2D island formation on substrates at low coverage. Defects play a pivotal role as active sites for kinetics, with steps, kinks, and vacancies lowering energy barriers for attachment and detachment. In the terrace-ledge-kink (TLK) model, terraces are flat regions, ledges (steps) are one-atom-high lines separating them, and kinks are interruptions along ledges; adatoms incorporate preferentially at kinks due to higher coordination, reducing the attachment energy by up to 50% compared to flat sites. Vacancies and steps act as sinks or sources, enhancing diffusion paths in the Burton-Cabrera-Frank (BCF) framework, where step flow dominates growth on vicinal surfaces, with kink density dictating overall rates. These defects, often at concentrations of 10^{-3} to 10^{-1} per site, amplify reactivity by orders of magnitude in processes like catalysis.

Historical Development

Early Foundations

The foundations of surface science trace back to the late 19th century, when thermodynamic principles began to address interfacial phenomena. In 1878, Josiah Willard Gibbs published seminal work on the equilibrium of heterogeneous substances, introducing concepts central to surface tension and adsorption isotherms. Gibbs defined the surface excess concentration and derived the , which relates changes in surface tension to variations in solute concentration at interfaces, providing a thermodynamic framework for understanding how molecules accumulate at boundaries between phases. Early 20th-century advancements built on these ideas through experimental and theoretical models of adsorption. Irving Langmuir, a key pioneer, developed the monolayer adsorption model between 1916 and 1918, proposing that gas molecules adsorb onto solid surfaces in a single layer with uniform binding sites. This led to the , expressed as \theta = \frac{KP}{1 + KP} where \theta is the fractional surface coverage, P is the gas pressure, and K is the equilibrium constant reflecting adsorption affinity. Langmuir's model assumed no interactions between adsorbed molecules and reversible binding, revolutionizing the quantitative description of surface coverage. His earlier experiments with oil films on water, detailed in 1917, demonstrated the existence of monomolecular layers by measuring the area covered by known quantities of oleic acid, revealing oriented molecular arrangements at the air-water interface with thicknesses around 2 nm. For these contributions to surface chemistry, Langmuir received the 1932 . Pioneering experiments in the 1910s and 1920s further explored surface processes, particularly in catalysis and crystal growth. Paul Sabatier, awarded the 1912 Nobel Prize in Chemistry jointly with Victor Grignard, demonstrated the hydrogenation of organic compounds on finely divided nickel surfaces, showing how metal catalysts facilitate reactions by adsorbing reactants in specific configurations. In the 1920s, Max Volmer conducted influential studies on crystal growth from vapor phases, observing step-wise advancement on crystal faces and identifying kink sites as preferential incorporation points for adatoms, which laid groundwork for kinetic theories of surface morphology. These early investigations, however, were hampered by the limitations of pre-vacuum technology, where contamination from residual gases and impurities in glass apparatus led to inconsistent and irreproducible results. Surfaces exposed to atmospheric pressures rapidly adsorbed unintended species, obscuring intrinsic behaviors and delaying precise measurements until the development of ultra-high vacuum systems in the 1950s.

20th-Century Advances

The development of ultra-high vacuum (UHV) technology in the 1950s marked a pivotal advancement in surface science, enabling the study of atomically clean surfaces free from contamination. Pioneered by Daniel Alpert at the University of Illinois, the Bayard-Alpert ionization gauge, introduced in 1950, allowed measurement and maintenance of pressures below 10^{-9} Torr, overcoming previous limitations in vacuum gauging that had restricted experiments to higher pressures around 10^{-6} Torr. This breakthrough facilitated reproducible surface experiments by minimizing gas adsorption, which had previously obscured intrinsic surface properties, and spurred the creation of all-metal vacuum systems essential for modern surface studies. Key experimental inventions in the mid-20th century further revolutionized surface imaging and analysis. In 1951, Erwin W. Müller at the Fritz-Haber-Institut in Berlin invented the field ion microscope (FIM), the first instrument to achieve atomic resolution imaging of metal surfaces by ionizing atoms in a high electric field and projecting their images onto a phosphor screen, revealing individual atomic positions with sub-angstrom precision. Complementing this, low-energy electron diffraction (LEED) emerged as a standard tool for surface crystallography in the early 1960s, building on the 1927 electron diffraction experiments of Clinton Davisson and Lester Germer but adapted for surface sensitivity through the use of single-crystal samples and commercial instrumentation that displayed diffraction patterns from the topmost atomic layers. These techniques provided direct visual and structural insights into surfaces, transforming qualitative observations into quantitative data on atomic arrangements./07%3A_Molecular_and_Solid_State_Structure/7.04%3A_Low_Energy_Electron_Diffraction) Theoretical progress in the 1970s advanced the understanding of surface electronic structure through extensions of electron gas models. The , an idealized representation of metal surfaces as a uniform positive background with delocalized electrons akin to , was applied using by and in their seminal 1970 calculations, which predicted surface energy and charge density profiles in the . This work extended the bulk free electron model to surfaces by accounting for the asymmetry at the interface, revealing in electron density and surface states localized near the boundary, providing a foundational framework for later ab initio computations of adsorption and reactivity. Major milestones in the late 20th century solidified surface science's experimental toolkit. The invention of the in 1981 by Gerd Binnig and Heinrich Rohrer at IBM Zurich enabled real-space atomic resolution imaging and manipulation of surfaces under UHV conditions, earning them the 1986 Nobel Prize in Physics for opening nanoscale surface studies to a wide range of materials beyond metals. Concurrently, the proliferation of synchrotron radiation facilities in the 1980s, such as the operational from 1982, enhanced for surfaces by providing tunable, high-brilliance X-rays that improved depth profiling and elemental sensitivity, revealing core-level shifts indicative of surface bonding. The institutionalization of surface science during this era fostered collaboration and standardization. The American Vacuum Society (AVS) was founded in 1953 as the Committee on Vacuum Techniques by a group of 56 scientists in New York, evolving into a key organization for advancing vacuum-based surface research through symposia and the Journal of Vacuum Science & Technology. The first international conferences dedicated to surface science, such as the International Conference on Solid Surfaces organized by the AVS in 1969, built on earlier vacuum congresses and promoted global exchange of techniques like LEED and early PES, establishing surface science as an interdisciplinary field.

Surface Physics

Adsorption and Desorption

Adsorption refers to the process by which molecules or atoms (adsorbates) bind to a solid surface, a fundamental phenomenon in surface physics that governs interactions at interfaces. Desorption is the reverse process, where adsorbates leave the surface. These processes are central to understanding surface coverage, reactivity, and energy landscapes, with applications in catalysis and materials design. The distinction between physisorption and chemisorption arises from the nature of the bonding: physisorption involves weak, non-specific van der Waals forces, typically with binding energies on the order of 10 kJ/mol, allowing reversible attachment without significant alteration of the adsorbate's electronic structure. In contrast, chemisorption entails stronger chemical bonds, often covalent or ionic, with energies exceeding 100 kJ/mol, leading to dissociation or reconfiguration of the adsorbate and potentially activating it for further reactions. This difference influences kinetics, where chemisorption may involve an activation barrier, while physisorption is generally barrierless. A key kinetic feature in both is the precursor state, an intermediate physisorbed configuration that adsorbates occupy before transitioning to chemisorbed sites, facilitating direct or indirect adsorption pathways as observed in molecular beam experiments on metal surfaces. Adsorption behavior is quantitatively described by isotherms, which relate surface coverage to gas pressure or concentration at constant temperature. The Langmuir isotherm models monolayer adsorption on a homogeneous surface with non-interacting sites, assuming each site accommodates one adsorbate molecule and adsorption is reversible. Derived from kinetic principles, it takes the form \theta = \frac{K P}{1 + K P} where \theta is the fractional coverage, P is the pressure, and K is the equilibrium constant related to the adsorption energy. For multilayer adsorption, particularly on porous materials, the Brunauer-Emmett-Teller (BET) isotherm extends the Langmuir model to multiple layers, treating the first layer differently from subsequent uniform layers. The BET equation for adsorbed volume V is V = \frac{V_m C P}{(P_0 - P) \left[1 + (C - 1) \frac{P}{P_0}\right]} where V_m is the monolayer capacity, P_0 is the saturation pressure, and C reflects the energy difference between the first and subsequent layers; this model is widely used to determine surface areas from nitrogen adsorption data. On heterogeneous surfaces, where site energies vary, the Temkin isotherm accounts for adsorbate-adsorbate interactions, assuming uniform coverage distribution and a linear decrease in adsorption heat with coverage, expressed as \theta = \frac{RT}{f} \ln(K P), with f as the heterogeneity factor. This captures scenarios like dissociative adsorption on metals. Desorption kinetics follow Arrhenius-type behavior, modeled by the Polanyi-Wigner equation, which describes the rate of coverage decrease as -\frac{d\theta}{dt} = \nu \theta^n \exp\left(-\frac{E_d}{RT}\right) where \nu is the pre-exponential factor (typically $10^{13} s^{-1}), n is the reaction order (1 for non-dissociative, 2 for recombinative desorption), E_d is the desorption activation energy (often equal to the adsorption energy for non-activated processes), R is the gas constant, and T is temperature. Thermal desorption spectroscopy (TDS), also known as temperature-programmed desorption, probes these kinetics by linearly ramping the surface temperature while monitoring desorbed species via mass spectrometry; resulting spectra show peaks whose position, shape, and intensity reveal binding energies, order, and coverage, with peak shifts indicating precursor-mediated or coverage-dependent processes. For instance, first-order desorption yields asymmetric peaks with a trailing edge, while second-order peaks are symmetric and shift to lower temperatures with increasing coverage. Coverage effects during adsorption influence morphology through adsorbate mobility and interactions. At low coverage, adsorbates diffuse across the surface, governed by Arrhenius-activated hops over barriers, with diffusion coefficient D = D_0 \exp(-E_a / RT), where E_a is typically 0.1–1 eV for physisorbed species on metals, enabling site exploration before binding. As coverage increases, growth modes diverge: the occurs when adsorbate-adsorbate bonds are stronger than adsorbate-substrate bonds, leading to three-dimensional island formation to minimize energy, common for metals on oxides. Conversely, the features layer-by-layer growth when substrate-adsorbate interactions dominate, promoting wetting and smooth films, as in alkali metals on metals. These modes are predicted by comparing surface and interface energies, with transitions possible via strain or alloying.

Electronic and Structural Properties

Surface electronic states arise due to the termination of the periodic crystal lattice at a solid-vacuum interface, leading to localized wavefunctions that differ from bulk band structures. Shockley states, predicted in 1939, emerge in regions of the projected bulk band gap where the wavefunction penetrates multiple atomic layers without strong scattering, forming nearly free-electron-like bands on metal surfaces such as Cu(111) and Au(111). In contrast, Tamm states, theorized in 1932, result from a potential barrier at the surface that localizes electrons within one or two atomic layers, often appearing at the band edges of semiconductors like cleaved GaAs(110). These states are crucial for surface conductivity and reactivity, with experimental confirmation via angle-resolved photoemission spectroscopy revealing parabolic dispersions near the Fermi level. Surface plasmons, collective oscillations of electrons confined to the interface, exhibit a dispersion relation given by \omega = \frac{\omega_p}{\sqrt{1 + \epsilon}} where \omega_p is the bulk plasma frequency and \epsilon is the dielectric constant of the adjacent medium, approaching \omega_p / \sqrt{2} for vacuum interfaces in the non-retarded limit. This relation, derived from classical electrodynamics in 1957, describes how plasmons couple with photons to form surface plasmon polaritons, enabling subwavelength light confinement on noble metals like Ag and Au. The work function \phi, defined as the energy to remove an electron from the to vacuum, is expressed as \phi = -\mu - V_0, where \mu is the chemical potential and V_0 is the surface electrostatic potential relative to bulk. Typical values range from 4.0 to 5.5 eV for transition metals, influencing electron emission and interface barriers. Structurally, clean surfaces often undergo relaxation, with the topmost atomic layer shifting inward by 0.1-0.3 Å due to reduced coordination and enhanced bonding to subsurface atoms, as observed on fcc metals like Al(110) and Pt(111). More pronounced reconstructions involve lateral rearrangements to minimize surface energy; for instance, the Au(111) surface forms a herringbone pattern with alternating zigzag rows, compressing the top layer by about 4% and creating a 23×√3 unit cell. Such superstructures are denoted using Wood's notation, which specifies the primitive unit cell dimensions and symmetry; an example is the Ni(100)-(2×2)p4g reconstruction, where the p4g indicates fourfold rotational symmetry with glide mirrors, often stabilized by subsurface impurities. Surface phonon modes, quantized lattice vibrations localized near the interface, include Rayleigh waves propagating along the surface with velocity v_R \approx 0.87 v_s, where v_s is the bulk shear speed for Poisson's ratio near 0.25, decaying exponentially into the bulk. These acoustic modes dominate low-frequency dynamics, while higher-energy surface resonance modes appear above the bulk continuum, identifiable via vibrational spectroscopy like high-resolution electron energy loss spectroscopy (HREELS). In semiconductors, quantum confinement at the surface can quantize bulk bands into a two-dimensional electron gas (2DEG), as seen on Si(111) where surface states form a metallic sheet with mobility exceeding 10^4 cm²/Vs, enabling quantum Hall effects under gating. This 2DEG arises from band bending and charge transfer, distinct from bulk conduction.

Surface Chemistry

Surface Reactions and Bonding

Surface reactions involve the formation and breaking of chemical bonds between adsorbates and surface atoms, leading to irreversible transformations distinct from physical adsorption processes. These reactions are governed by the nature of the bonds formed, which can be classified as dative, covalent, or ionic. Dative bonding occurs when an adsorbate donates a pair of electrons to an empty orbital on the surface atom, enhancing catalytic activity by stabilizing intermediates, as seen in amine interactions with silicon surfaces. Covalent bonding involves the sharing of electron pairs between the adsorbate and surface, resulting in directional interactions that strengthen adhesion and influence reaction pathways on metal oxides. Ionic bonding at surfaces arises from electrostatic attractions between charged adsorbates and oppositely charged surface sites, often predominant in polar environments like oxide interfaces. To quantify partial charges in these bonds, Bader charge analysis partitions the electron density based on zero-flux surfaces, revealing charge transfer in adsorbate-metal systems; for instance, oxygen adsorbates on metals exhibit partial ionic character with transfers typically around 1-1.5 electrons. Key reaction mechanisms in surface chemistry include the Eley-Rideal and Langmuir-Hinshelwood pathways. The Eley-Rideal mechanism proceeds via direct collision between a gas-phase molecule and an adsorbed species on the surface, without requiring co-adsorption, as originally described in studies of hydrogen exchange on tungsten. In contrast, the Langmuir-Hinshelwood mechanism involves both reactants adsorbing onto adjacent surface sites before reacting, facilitating diffusion and interaction, a process foundational to heterogeneous catalysis since its proposal by Irving Langmuir in 1921 and elaboration by Cyril Hinshelwood. Activation barriers for these mechanisms are analyzed using transition state theory, which predicts the reaction rate constant as k = \frac{k_B T}{h} \exp\left(-\frac{\Delta G^\ddagger}{RT}\right), where k_B is Boltzmann's constant, h is Planck's constant, \Delta G^\ddagger is the Gibbs free energy of activation, R is the gas constant, and T is temperature; this Eyring equation applies to surface elementary steps by treating the transition state as a loosely bound complex. Selectivity in surface reactions is influenced by site-blocking from impurities, known as catalyst poisoning, where adsorbates like sulfur occupy active sites and reduce turnover rates by up to orders of magnitude in hydrogenation processes. In alloy catalysts, ensemble effects arise from the requirement for specific multi-atom site configurations; for example, bimetallic Pd-Au surfaces show diminished reactivity for CO oxidation when isolated Pd atoms lack neighboring ensembles, enhancing selectivity toward desired products. Geochemical surface reactions, such as mineral weathering, exemplify these principles through dissolution processes. Silicate minerals dissolve via surface-controlled reactions, as in quartz weathering: \ce{SiO2 + 2H2O -> H4SiO4}, releasing and influenced by and ligands that lower activation barriers. Metal binding to mineral surfaces is modeled using surface complexation approaches, notably the 2-pK model, which assumes two constants for surface sites (\ce{>SiOH2^+} and \ce{>SiO^-}) to describe charge development and inner-sphere complexes on oxides like . This framework predicts adsorption edges for ions like Pb(II) on silicates, integrating ionic and dative bonding motifs.

Catalysis and Electrochemistry

In , surface reactions occur at the interface between a solid catalyst and gaseous or liquid reactants, enabling efficient conversion at lower temperatures and pressures than homogeneous processes. The governs optimal catalyst performance, stating that the of reactants to the surface should be intermediate—neither too weak to prevent adsorption nor too strong to hinder desorption and product release. This principle, originally proposed by Paul Sabatier, is visualized in volcano plots where catalytic activity peaks at balanced adsorption strengths, as quantified by calculations of binding free energies. Catalytic efficiency is often measured by turnover frequency (TOF), defined as the number of reactant molecules converted to product per per unit time under specified conditions, such as low concentration and reactant . TOF provides insight into intrinsic site activity, typically reported in s⁻¹, and is calculated as the of turnovers with respect to time. Promoters enhance TOF by modifying surface electronic properties; for instance, (K) in iron-based catalysts for synthesis weakens adsorption, increasing activity by up to an through electron donation and site isolation effects. A seminal example is the Haber-Bosch process for ammonia synthesis (N₂ + 3H₂ → 2NH₃) on iron catalysts, where the dissociative mechanism involves N₂ adsorption and stepwise hydrogenation on Fe(111) surfaces under industrial conditions (e.g., 673 K, 20 atm). The rate-determining step is N₂ dissociation, with predicted TOF of approximately 18 s⁻¹ per site, closely matching experimental values and highlighting the role of surface steps in activating the N≡N bond. Another key reaction is CO oxidation on platinum (2CO + O₂ → 2CO₂), following Langmuir-Hinshelwood kinetics where the rate is proportional to the product of coverages, rate = k θ_CO θ_O, with maximum activity when θ_CO ≈ θ_O ≈ 0.5 monolayers due to balanced adsorption sites. In , the electrical double layer at -electrolyte interfaces structures distribution and influences . The Helmholtz model describes a compact layer of solvated s adsorbed at fixed distance from the , akin to a parallel-plate with linear potential drop and C_H = ε_r ε_0 / l, where l is the layer thickness (typically 0.3 nm). The Gouy-Chapman extension accounts for a diffuse layer of thermally agitated s, solving the Poisson-Boltzmann equation to yield the potential distribution \psi(x) = \frac{2kT}{ze} \ln\left( \frac{1 + \gamma e^{-\kappa x}}{1 - \gamma e^{-\kappa x}} \right), where \gamma = \tanh\left( \frac{ze\psi_0}{4kT} \right), \psi_0 is the potential at the surface, \kappa is the inverse Debye length, z is the ion valence, e is the elementary charge, k is Boltzmann's constant, and T is temperature. This formulation captures ion screening over the Debye length, with capacitance C_GC increasing with surface charge as C_GC = ε_r ε_0 κ cosh(zeφ / 2kT), where κ is the inverse Debye length. Underpotential deposition (UPD) involves the reversible formation of a foreign metal on a at potentials positive to the bulk deposition potential, driven by favorable heterometallic bonding energies. For example, UPD on occurs at ~0.2-0.4 V vs. , enabling precise control of surface composition for electrocatalytic tuning. In electrodes, catalyzes the hydrogen oxidation reaction (HOR: H₂ → 2H⁺ + 2e⁻) and evolution reaction (HER: 2H⁺ + 2e⁻ → H₂), proceeding via Volmer (H⁺ + e⁻ → *H), Tafel (*H + *H → H₂), or Heyrovsky (*H + H⁺ + e⁻ → H₂) steps, with Pt's near-zero *H binding (ΔG_H ≈ 0) yielding Tafel slopes of 30-120 mV/dec depending on . Corrosion resistance arises from passivation, where protective films form spontaneously on metals. On aluminum, a thin Al₂O₃ layer (2-5 nm) passivates the surface in neutral aqueous environments, inhibiting further oxidation. Pourbaix diagrams delineate stability: Al₂O₃ prevails in the passivation region ( 4-9, potentials -1.2 to -0.5 V vs. SHE at 25°C), bounding (Al³⁺ dissolution at low ) and immunity (Al metal stability) zones, with solubility limits at 10⁻⁶ M for practical thresholds.

Characterization Techniques

Spectroscopic Methods

Spectroscopic methods in surface science provide energy-resolved insights into the , , and at surfaces and interfaces, enabling the identification of adsorbates, chemical states, and molecular orientations without direct spatial . These techniques exploit the of photons or electrons with surface to probe vibrational modes, valence and levels, and unoccupied states, offering complementary information to structural methods. Vibrational spectroscopies reveal adsorbate dynamics and contributions, while and photoelectron spectroscopies map band structures and elemental distributions with atomic sensitivity. Vibrational spectroscopies are essential for characterizing surface s and adsorbate vibrations. High-resolution (HREELS) measures energy losses of low-energy electrons scattered from surfaces, achieving resolutions of approximately 10 cm⁻¹ to resolve phonon dispersions and adsorbate modes such as C-O stretches in on metals. This technique is particularly sensitive to dipole-allowed excitations and dipole-forbidden modes via impact scattering, providing insights into adsorbate geometry and coverage on single crystals. Sum-frequency generation (SFG) , a nonlinear optical , selectively probes interfaces by generating a sum-frequency signal from overlapping visible and beams, with the second-order susceptibility χ⁽²⁾ proportional to the surface density N_surface, ensuring negligible bulk contribution. SFG is ideal for buried interfaces, such as liquid-solid or -substrate boundaries, where it reveals vibrational spectra of interfacial or polymer chains, highlighting orientational order through phase-sensitive detection. Electronic spectroscopies target and levels to elucidate bonding and composition. photoelectron spectroscopy () uses ultraviolet photons, typically the He I line at 21.2 eV, to eject electrons, mapping the near the and on surfaces like metal oxides or semiconductors. spectra distinguish from bulk bands, as seen in the band structure of clean (111) surfaces, with enhancing selectivity. X-ray photoelectron spectroscopy (XPS) employs X-rays (e.g., Al Kα at 1486.6 eV) to levels for and chemical shifts, with calculated as E_b = h\nu - E_{kinetic} - \phi, where h\nu is the , E_{kinetic} the measured , and \phi the . The limits sensitivity to 5-10 nm depth, enabling non-destructive profiling of oxide layers or adsorbates on catalysts. Auger electron spectroscopy (AES) complements XPS by detecting Auger electrons following core-hole creation, with the modified Auger parameter \alpha' = E_{kinetic}^{Auger} + E_{binding}^{XPS} minimizing relaxation effects to identify chemical environments via shifts of 1-5 eV. AES excels in high (<10 nm) for mapping impurities on semiconductors, though self-convolution of levels broadens peaks for light elements. Synchrotron-based near-edge X-ray absorption fine structure (NEXAFS) spectroscopy probes unoccupied states above core edges (e.g., C K-edge at 285 eV), revealing molecular orbitals and in organic monolayers. Linear dichroism in polarized measurements quantifies orientational order, as the intensity ratio I_parallel / I_perpendicular reflects tilt angles in self-assembled monolayers on . NEXAFS provides element-specific sensitivity to π* and σ* resonances, distinguishing aromatic from aliphatic carbons at surfaces.

Microscopic and Diffraction Techniques

Scanning probe microscopy techniques offer real-space imaging of surfaces at the atomic scale, providing direct visualization of topography and local electronic or mechanical properties. These methods, typically performed in ultrahigh vacuum to minimize contamination, rely on a sharp probe interacting with the sample surface. Scanning tunneling microscopy (STM), invented by Binnig and Rohrer in 1982, images conductive surfaces by measuring the quantum tunneling current between a metallic tip and the sample. The current follows I \propto \exp(-2 \kappa d), where d is the tip-sample separation and \kappa = \sqrt{2m \phi}/\hbar with \phi the average barrier height, enabling atomic resolution under typical conditions of 0.5-1 nm tip-sample distance. Applying a bias voltage allows mapping of electronic density of states near the Fermi level, revealing surface reconstructions and adsorbate configurations. Atomic force microscopy (AFM), developed by Binnig, Quate, and Gerber in 1986, extends imaging to insulating materials by detecting forces between the tip and sample via deflection. In contact mode, the tip physically touches the surface, while non-contact modes oscillate the to sense long-range forces; force-distance curves quantify , elasticity, and short-range interactions. AFM achieves sub-nanometer lateral resolution and is widely used for diverse surfaces including polymers and biological samples. Electron microscopy provides complementary morphological information over larger areas. Scanning electron microscopy () scans the surface with a focused electron beam, detecting to generate topographic images with resolutions approaching in field-emission instruments. SEM excels at revealing surface features like roughness and defects, though non-conductive samples often need coating to prevent charging. Transmission electron microscopy (TEM) on cross-sectioned samples enables high-resolution (~0.1 nm) imaging of buried interfaces and subsurface structures. Preparation involves mechanical polishing and ion milling to thin samples to ~100 nm, allowing transmission to visualize arrangements at surface-adjacent layers, such as in thin films or heterostructures. techniques determine surface crystallography by analyzing scattered waves from periodic arrays. (LEED) directs s (20-200 eV) normal to the surface, producing spot patterns from the top few layers due to the short (~1 nm); intensity-voltage (I-V) curves facilitate quantitative structure fitting via dynamical scattering calculations. LEED is essential for verifying adsorbate-induced reconstructions on single crystals. Reflection high-energy electron diffraction (RHEED) uses grazing-incidence electrons (10-50 keV) to monitor epitaxial growth in real time, with streaked patterns indicating surface smoothness and intensity oscillations reflecting monolayer-by-monolayer deposition rates. RHEED is standard in chambers for semiconductors and oxides. Grazing-incidence diffraction (GIXRD) employs s at shallow angles (<1°) to confine penetration to surface layers (~10 nm), mapping in-plane and out-of-plane lattice parameters of thin films and interfaces. This non-destructive method complements electron techniques for larger samples and provides strain and texture information. Recent advances in non-contact AFM using enhance sensitivity to short-range chemical forces. By oscillating the at its frequency and detecting shifts proportional to force gradients, true atomic imaging of insulators is achieved, distinguishing bond types and molecular orbitals without tip-sample contact. These developments, building on qPlus designs, enable submolecular on complex surfaces like .

Applications

Materials and Nanotechnology

Surface science plays a pivotal role in the design and fabrication of and thin films, where control over surface properties enables precise at the and molecular scales. Techniques leveraging surface interactions allow for the creation of structures with tailored , , and characteristics, essential for applications in , , and . This involves methods for epitaxial growth, , and functionalization that exploit adsorption , surface energies, and interfacial chemistry to achieve uniformity and functionality. Recent advances have extended these principles to two-dimensional (2D) materials, such as and transition metal dichalcogenides (TMDs), where surface-controlled (CVD) enables large-area epitaxial growth on substrates like or , yielding defect densities below 10^{11} cm^{-2} and enabling applications in and . Thin film growth techniques in surface science, such as (), enable layer-by-layer deposition of materials with atomic precision under conditions. Developed in the , directs beams of atoms or molecules from sources toward a heated substrate, promoting epitaxial alignment and high-purity films. The flux of material in is governed by the rate from the source, given by F = \frac{P}{\sqrt{2\pi M k T}}, where P is the , M the , k Boltzmann's constant, and T the source temperature; this formula ensures controlled delivery for growth rates on the order of fractions of a per second. 's ability to monitor surface structure in real-time via reflection high-energy electron diffraction facilitates the fabrication of heterostructures like quantum wells. Atomic layer deposition (ALD) complements by providing conformal coatings through sequential, self-limiting surface reactions in a vapor-phase process. Pioneered in the for thin-film electroluminescent displays, ALD alternates exposure to precursors and purge steps, ensuring uniform thickness even on high-aspect-ratio nanostructures. Each cycle deposits a fraction of a via , with growth rates typically 0.1–1 Å per cycle, enabling precise control over film in multicomponent materials like high-k dielectrics. Self-assembly emerges as a bottom-up approach in surface science, where molecules or nanoparticles spontaneously organize driven by surface interactions. form ordered films on substrates like , with alkanethiols serving as a model ; the reaction HS–R + → Au–S–R + 1/2 H₂ produces densely packed layers up to 2 nm thick, stabilized by van der Waals forces between alkyl chains. This organization yields tilt angles of ~30° and domain sizes exceeding micrometers, providing tunable wettability and reactivity. For nanoparticles, shape control relies on differential surface energies of facets, as predicted by the , where lower-energy planes dominate the equilibrium morphology—e.g., cubic habits for materials with {100} facets of minimal energy. Ligands selectively stabilize high-energy facets, enabling synthesis of rods or tetrahedra from isotropic seeds. At the nanoscale, surface effects dominate, altering bulk properties through confinement and interfacial phenomena. In quantum dots, electron-hole pairs experience quantum confinement, shifting the bandgap by an energy ΔE ∝ 1/r², where r is the radius; this leads to size-tunable emission from infrared to ultraviolet, as modeled for semiconductor particles like CdSe. For metallic nanoparticles, (SPR) arises from collective electron oscillations, with the resonance wavelength λ_max red-shifting by tens of nm as particle diameter increases from 10 to 100 nm due to enhanced . These effects underpin optoelectronic devices, where surface passivation minimizes non-radiative recombination. In biomaterials, surface science guides functionalization to enhance by mitigating adverse interactions like protein fouling. Grafting poly() (PEG) chains onto surfaces creates hydrophilic brushes that resist nonspecific adsorption; for instance, PEG-terminated SAMs on reduce fibrinogen adsorption by over 90% compared to bare surfaces, attributed to steric repulsion and effects. This modification extends implant lifetimes and supports scaffolds with controlled .

Energy and Environmental Systems

Surface science plays a pivotal role in advancing technologies, particularly in , where interfacial phenomena dictate device and stability. In dye-sensitized solar cells (DSSCs), the anchoring of dye molecules to the TiO₂ surface is crucial for efficient electron injection, with groups forming strong bidentate bonds that ensure stable attachment and prevent aggregation. This surface interaction facilitates rapid charge separation at the dye-TiO₂ interface, occurring on timescales, which minimizes recombination losses and enhances generation. The performance of these cells is often evaluated using incident photon-to-current (IPCE), defined as IPCE = \frac{1240}{\lambda} \times EQE, where λ is the in nanometers and EQE is the external , providing a measure of wavelength-dependent charge collection . Perovskite solar cells, another key photovoltaic technology, face stability challenges due to surface defects involving PbI₂ residues, which form at grain boundaries and interfaces during fabrication or degradation, leading to ion migration and in current-voltage characteristics. These defects trap charge carriers, reducing and long-term operational stability under ambient conditions. Mitigating PbI₂-related surface imperfections through passivation strategies, such as incorporating chloride ions, has been shown to improve device lifetimes by stabilizing the lattice. As of 2025, certified power conversion efficiencies for perovskite solar cells have reached 27%, with recent advances in protective coatings, such as fluorinated barriers and alumina nanoparticles, tripling operational stability under continuous illumination. In and conversion, surface processes are essential for solid oxide fuel cells (SOFCs) and lithium-ion batteries. For SOFCs, the surface, typically , governs oxygen incorporation via the reaction O₂ + 4e⁻ → 2O²⁻ at the triple-phase boundary, where oxygen reduction kinetics are influenced by surface oxygen vacancies and catalytic activity of the material. Optimizing these surface properties enhances ionic and reduces losses, enabling efficient high-temperature operation. In lithium-ion batteries, the solid electrolyte interphase (SEI) forms on anodes during initial cycling, primarily from (EC) decomposition, yielding a passivation layer rich in Li₂CO₃ that prevents further reduction while allowing Li⁺ diffusion, with typical thicknesses of 10-50 nm that critically determine cycle life and capacity retention. Environmental applications of surface science leverage catalytic surfaces for pollutant remediation and atmospheric processes. TiO₂-based enables under ultraviolet light via the reaction 2H₂O → 2H₂ + O₂, driven by band edge alignment where the conduction band minimum (~ -0.5 V vs. NHE) exceeds the hydrogen evolution potential and the valence band maximum (~2.7 V) surpasses the potential, generating electron-hole pairs that drive reactions on the surface. This process is enhanced by surface modifications like doping to extend visible-light absorption, achieving quantum yields up to several percent for . In , over polar regions is accelerated on ice surfaces in stratospheric clouds, where heterogeneous reactions such as N₂O₅ + HCl → ClNO₂ + HNO₃ on ice particles release radicals that catalytically destroy O₃, contributing to seasonal ozone holes. Geosurfaces provide natural platforms for , particularly through adsorption mechanisms. Heavy metal ions, such as Cd²⁺ and Pb²⁺, are effectively removed from contaminated soils via adsorption onto clay minerals like , where surface and aluminol groups form complexes, often modeled by the Freundlich isotherm q = K C^{1/n}, indicating multilayer heterogeneous adsorption with n values typically between 1 and 10 for enhanced capacity at low concentrations. Zeolites, with their microporous frameworks, excel in carbon capture by selectively adsorbing CO₂ through interactions at cationic sites, achieving capacities up to 4-5 mmol/g at ambient conditions, which supports post-combustion capture and technologies by exploiting tunable surface acidity and pore sizes. Emerging sorbents, such as metal-organic frameworks (MOFs), have demonstrated capacities exceeding 5 mmol/g under similar conditions as of 2025, offering higher selectivity and regenerability for large-scale CO₂ mitigation.

References

  1. [1]
    Surface (Surface Science) - an overview | ScienceDirect Topics
    Surface science is defined as the study of the physical and chemical phenomena occurring at the interfaces between different phases, such as solids, liquids, ...
  2. [2]
    Surface Science (Nix) - Chemistry LibreTexts
    Feb 21, 2022 · Surface science is the study of physical and chemical phenomena that occur at the interface of two phases, including solid–liquid interfaces ...
  3. [3]
    1.2 Basic concepts and terminology in surface science - Fiveable
    Surface science is the study of physical and chemical phenomena that occur at the interface of two phases (solid–liquid, solid–gas, solid–vacuum, liquid-gas) ...
  4. [4]
    Surface and Interface Science | Research Starters - EBSCO
    Surface and interface science is a specialized field that investigates the characteristics of materials at their surfaces and at the interfaces where different ...
  5. [5]
    The birth and evolution of surface science - NIH
    Surface science emanated from the confluence of concepts and tools in physics and chemistry with technological innovations that made it possible to determine ...<|control11|><|separator|>
  6. [6]
    Surface Science – How it all began, from an experimental perspective
    Dec 18, 2017 · In surface science, various phenomena at the intersection of different phases are explored. The phases could be any of vacuum, gas, liquid and ...
  7. [7]
    The birth and evolution of surface science: Child of the union ... - PNAS
    Surface science is the child of the union of science and technology. Essentially all of the concepts and theoretical tools used in the field emanate from ...
  8. [8]
    The Evolution of Surface Chemistry. A Personal View of Building the ...
    Surface chemistry, like other branches of physical chemistry, historically developed through macrosopic studies. These included measurements of ...The Historical Development of... · High-Pressure Surface... · Conclusion
  9. [9]
    Gibbs Adsorption Equation - an overview | ScienceDirect Topics
    The Gibbs adsorption equation refers to the relationship that describes how an increase in the chemical potential of a species at an interface leads to a ...
  10. [10]
    Comparing Contact Angle Measurements and Surface Tension ...
    Sep 3, 2010 · ... surface tension of the solid is given by the Young equation (1)Here, θY is the Young CA, σs and σl are the surface tensions of the solid and ...
  11. [11]
    Modeling Gas–Liquid Interfaces by Dissipative Particle Dynamics
    Nov 20, 2020 · Adsorption of surfactants at gas–liquid interfaces that causes reduction in the surface tension is a classical problem in colloid and interface ...
  12. [12]
    Chapter 1 Surface processes in epitaxial growth - ScienceDirect.com
    This chapter discusses surface processes in epitaxial growth. The solid-vapor interface can be observed by an array of surface science and microscope-based ...
  13. [13]
    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).
  14. [14]
    Wulff shape of crystals - Scholarpedia
    Oct 24, 2013 · The shape of an equilibrium crystal is obtained, according to the Gibbs thermodynamic principle, by minimizing the total surface free energy associated to the ...Thermodynamics of... · Interfaces in statistical... · Wulff shape in statistical...
  15. [15]
    [PDF] Surface energies - Kristin Persson
    Sep 13, 2016 · relaxation of the (110) surface of diamond Si exceeds this value due to its tendency to reconstruct to achieve stability. Code availability.
  16. [16]
    David A. REED and Gert EHRLICH 1. Introduction Surface diffusion ...
    The connection is established between the surface diffusivity. D(S) and QH), the rate of atom jumps in an adsorbed gas at a coverage 0.
  17. [17]
    [PDF] Nucleation Theory and the Early Stages of Thin Film Growth
    A review is given of nucleation and growth models as applied to the earliest stages of thin film growth. Rate equations (RE), kinetic Monte Carlo (KMC) and ...Missing: classical ΔG
  18. [18]
    How does your crystal grow? A commentary on Burton, Cabrera and ...
    (b) Schematic diagram of the terrace–ledge–kink model of a surface of some arbitrary (high-index) or vicinal orientation. Early evidence of a distinct ...Missing: vacancies | Show results with:vacancies
  19. [19]
    Macrosteps dynamics and the growth of crystals and epitaxial layers
    Theoretically, crystal growth by atomic step flow mode was formulated relatively early by Burton, Cabrera and Frank (BCF) [39]. The growth proceeds via parallel ...
  20. [20]
    Gibbsian Surface Thermodynamics - ACS Publications
    Oct 22, 2020 · This work lays out the relationships between important equilibrium equations including the following: the Gibbs–Duhem equation, the Gibbs ...
  21. [21]
    [PDF] Irving Langmuir - Nobel Lecture
    By cooling the bulb in liquid air the adsorption increased. The maximum amount of hydrogen adsorbed on the glass in these experiments was 0.03 mm3 per cm2, ...Missing: monolayer | Show results with:monolayer
  22. [22]
    Paul Sabatier – Nobel Lecture - NobelPrize.org
    During the period 1901 to 1905, together with Senderens, I showed that nickel is very suitable for the direct hydrogenation of nitriles into amines and, no less ...
  23. [23]
    (PDF) Crystal Growth through the Ages - ResearchGate
    chemist Max Volmer (1885–1965) and his student Immanuel Estermann (1900–1973) ... Detailed studies of bulk and surface structures are therefore needed.
  24. [24]
    [PDF] Thoughts on the Past, Present and Future of UHV Surface Chemistry
    In this perspective we briefly discuss how the field developed, identify some challenges, and highlight Single-Atom Alloys (SAAs), a new class of heterogeneous ...Missing: issues early
  25. [25]
    Vacuum science and technology: 1950–2003 - AIP Publishing
    Sep 2, 2003 · The “new vacuum technology” began in 1950 with the development of the Bayard–Alpert gauge which ended the limitation on the lowest ...
  26. [26]
    [PDF] Erwin Müller - National Academy of Sciences
    It is fascinating and somewhat ironic that knowledge gained through research using Müller's FEM was important in developing the present-day atomic resolution ...
  27. [27]
    Effects of density functional semilocality and van der Waals nonlocality
    The work of Lang and Kohn (12–14) in the early 1970s demonstrated the ability of the simple local density approximation (LDA) (11, 15) for the exchange- ...
  28. [28]
    The Nobel Prize in Physics 1986 - NobelPrize.org
    ... microscope", the other half jointly to Gerd Binnig and Heinrich Rohrer "for their design of the scanning tunneling microscope". To cite this section. MLA ...
  29. [29]
    Surface Spectroscopies with Synchrotron Radiation - Science
    A variety of photoelectron spectroscopies using synchrotron radiation have been devised to study solid surfaces.
  30. [30]
    History - AVS
    A Brief History of the AVS Organizational Structure: Original Group - On June 18, 1953, fifty-six people from diverse technical backgrounds, gathered in New ...
  31. [31]
    Review Article History of the International Union for Vacuum Science ...
    The. American Vacuum Society'S Surface Science Division initi- ated the First International Conference on Solid Surfaces in conjunction with this Congress.
  32. [32]
    Chemisorption, Physisorption and Dynamics - Surface Science
    Mar 30, 2012 · This chapter contains sections titled: Types of interactions. Binding sites and diffusion. Physisorption Advanced Topic: Theoretical ...Missing: seminal | Show results with:seminal
  33. [33]
    Following the microscopic pathway to adsorption through ... - Science
    Sep 18, 2020 · The intricate interplay between physisorption and chemisorption states is believed to take place in precursor-mediated adsorption (9–11) and ...Missing: seminal | Show results with:seminal
  34. [34]
    THE ADSORPTION OF GASES ON PLANE SURFACES OF GLASS ...
    Replacing the Langmuir Isotherm with the Statistical Thermodynamic Fluctuation Theory. The Journal of Physical Chemistry Letters 2024, 15 (13) , 3683-3689 ...
  35. [35]
    Adsorption of Gases in Multimolecular Layers - ACS Publications
    ... Brunauer–Emmett–Teller Isotherm Model. Industrial & Engineering Chemistry Research 2025, 64 (32) , 15841-15856. https://doi.org/10.1021/acs.iecr.5c00860.
  36. [36]
    Adsorption isotherm modeling for methylene blue removal onto ...
    Jun 2, 2021 · Temkin isotherm. Temkin isotherm model (Temkin and Pyzhev 1940) incorporates the effects of adsorbate–adsorbate interactions on the ...
  37. [37]
    Analysis of Temperature-Programmed Desorption via Equilibrium ...
    Temperature-programmed desorption (TPD) experiments in surface science are usually analyzed using the Polanyi–Wigner equation and/or transition-state theory ...Introduction · Case 2: An Ideal Lattice Gas · Discussion · Conclusions
  38. [38]
    23 Thermal excitation: thermal desorption spectroscopy (TDS)
    Oct 31, 2023 · Abstract. TDS has been used very extensively, but not as a general analytical tool in the same sense as have XPS or AES.
  39. [39]
    Surface Diffusion of Adsorbed Molecules in Porous Media
    This review provides an overview of surface diffusion and capillary condensate flow in porous media. Emphasis has been placed on the distinction between ...Missing: barriers | Show results with:barriers
  40. [40]
    Growth Modes - DoITPoMS
    Island Growth is also known as Volmer-Weber Growth. This occurs when the ... Layer Growth is also known as Frank-van der Merwe Growth. This occurs when ...
  41. [41]
    On the Surface States Associated with a Periodic Potential | Phys. Rev.
    The wave functions and energy levels associated with a finite one-dimensional periodic potential field are investigated.Missing: seminal | Show results with:seminal
  42. [42]
    Surface tamm states at curved surfaces of ionic crystals
    I. E. Tamm, Phys. Zs. Sowjetunion 1, 733 (1932); see also I. E. Tamm, Collection of Scientific Papers (Nauka, Moscow, 1975), Vol. 1, pp.
  43. [43]
    Vibrational Surface Electron-Energy-Loss Spectroscopy Probes ...
    Dec 7, 2017 · The study of surface phonons, collective atomic vibrations localized on the surface of solids, has long been helped by insights from ...
  44. [44]
    Work Function: Fundamentals, Measurement, Calculation ...
    The work function, which is the energy barrier for an electron escaping from the surface of a material, is a fundamental material surface property with many ...BASIC PHYSICS OF THE... · WORK-FUNCTION... · APPLICATIONS OF WORK...
  45. [45]
    Understanding the interplay of surface structure and work function in ...
    Jul 16, 2020 · This work function is, therefore, a function of the specific surface being considered and can be expressed as Φ = Evac − EF = Φ0 + ΔΦdipole.Missing: V0 | Show results with:V0
  46. [46]
    Surface Relaxation Influenced by Surface States | Phys. Rev. Lett.
    A detailed theoretical investigation of the relaxation of the simple metal surface Al(100) is presented. We show the influence of electronic surface states ...
  47. [47]
    (PDF) Structure determination of p4g Ni(100)-(2 × 2)C by LEED
    The surface structure of Ni(lOO)-(2 X 2)C is determined by LEED. The dynamical calculations include reconstruction of the top two layers.
  48. [48]
    Rayleigh Wave Velocity - an overview | ScienceDirect Topics
    Hence, the definition c = ω / ξ yields the wave speed. In principle, Eq. (22) has three double roots. However, for practical situations, only one is real, ...
  49. [49]
    A non-invasive gating method for probing 2D electron systems on ...
    Oct 14, 2020 · Intrinsic Si(111) surfaces passivated with atomic hydrogen are an ideal platform to host two-dimensional electron systems.
  50. [50]
    Silicon Surfaces as Electron Acceptors: Dative Bonding of Amines ...
    These results suggest that the ability to form these stable dative-bonded adducts is strongly connected with the ability of these surfaces to transfer charge ...
  51. [51]
    First-principles definition of ionicity and covalency in molecules and ...
    Apr 10, 2024 · Covalent bonding is directional and results in low-coordination structures with high barriers of rearranging the structure and, consequently, ...
  52. [52]
    Strong H Bonding between NH4+ and Bi-layered V2O5
    Jun 13, 2019 · Inserted K+ ions interact ionically with the oxide. In contrast, Bader charge analysis shows that with one NH4+ inserted 0.17 of an electron is ...
  53. [53]
    [PDF] Measuring charge transfer between adsorbate and metal surfaces
    We use Bader charge analysis to derive the trend on the charge state of surface species. The Bader charges of the chemisorbed O atoms (7.00e) are nearly the ...
  54. [54]
    A first-principles analysis of the charge transfer in magnesium ...
    Sep 14, 2020 · As a reference, the Bader charges for the clean Mg(0001) surface as well as for the adsorbates of interest are illustrated in Fig. S1 ...Results · The Volmer Step · Reaction Kinetics
  55. [55]
    [PDF] Eley–Rideal, the Other Mechanism - ETH Research Collection
    Apr 11, 2018 · The mechanism that Eley and Rideal really studied some 75 years ago is important in heterogeneous as well as homogeneous catalysis in the liquid ...Missing: seminal | Show results with:seminal
  56. [56]
    Practical Considerations for Understanding Surface Reaction ...
    Oct 30, 2024 · Here, we present our perspective on differentiating between various surface reaction mechanisms in light of pioneering studies by leaders in the field.
  57. [57]
    Calculation of the transition state theory rate constant for a general ...
    The transition state theory rate constant can be expressed in terms of the free energy of activation ΔG‡: k TST = 1 β h e - β Δ G ‡ , where β = 1/kBT.Abstract · Introduction · Computational MethodsMissing: barriers | Show results with:barriers<|separator|>
  58. [58]
    Selective Blockage Of The Catalyst Active Sites For The Hydrogena
    “Selective” poisoning involves preferential adsorption of the poison on the most active sites at low concentrations. If sites of lesser activity are blocked ...
  59. [59]
    Atomic Ensemble Effects in Electrocatalysis: The Site-Knockout ...
    Aug 10, 2025 · The poisoning of undesired intermediates or impurities greatly hinders the catalytic performances of noble metal‐based catalysts. Herein ...
  60. [60]
    [PDF] Silica Dissolution and Precipitation - European Commission
    Silica dissolves as silicic acid according to. SiO2 + 2H2O = H4SiO4 , Keq = aH4SiO4. aSiO2. aH2O. 2. (3). For the case of pure silica, the solid silica activity ...
  61. [61]
    Evidence of the Existence of Three Types of Species at the Quartz ...
    The existence of the >SiOH2+ and >SiO- surface species inferred from this approach allows us to propose the 2-pK model and explain the quartz surface charge and ...Introduction · Background · Acknowledgment
  62. [62]
    Surface properties, solubility and dissolution kinetics of bamboo ...
    Surface acid–base titrations allowed generation of a 2-pK surface complexation model (SCM) for larch, elm and horsetail phytoliths.
  63. [63]
    The Sabatier Principle in Electrocatalysis: Basics, Limitations, and ...
    The Sabatier principle, which states that the binding energy between the catalyst and the reactant should be neither too strong nor too weak, has been widely ...Introduction · The Sabatier Principle and its... · Challenges of the Sabatier...Missing: seminal paper
  64. [64]
    “Turning Over” Definitions in Catalytic Cycles | ACS Catalysis
    Nov 8, 2012 · “The standard turnover frequency, TOF°, is the TOF measured at a hypothetical 1 M standard concentration of reactants and products (or 105 Pa in ...Introduction · Typical Misuses of the TOF · “Turning Over” Definitions · Conclusion
  65. [65]
    Potassium promoter in industrial ammonia synthesis catalyst
    The chemical state of potassium in the catalyst changes from a stronger bound ionic, probably in the form of Fe O K, to a weaker, more covalent one during the ...
  66. [66]
    CO oxidation on Pt and Pd model catalysts - RSC Publishing
    Jun 7, 2017 · This reaction is an example of the Langmuir–Hinshelwood (LH) mechanism and a maximum rate is expected if θCO and θO are equal. The co ...
  67. [67]
    [PDF] Classical Models of the Interface between an Electrode ... - COMSOL
    2.2 Gouy-Chapman Model​​ They pictured a diffuse double layer (DDL) consisting of counterions (i.e. ions of opposite charge to the surface), which are attracted ...Missing: formula | Show results with:formula
  68. [68]
    Modeling the Gouy–Chapman Diffuse Capacitance with Attractive ...
    We propose a double-layer model that is capable of reproducing the main experimental findings in a simple and (in parts) analytical way.Introduction · Methods · Results · Discussion and Outlook
  69. [69]
    Underpotential Deposition - an overview | ScienceDirect Topics
    Underpotential deposition refers to the phenomenon where the first monolayer of a depositing metal onto a different material is formed at a potential that ...
  70. [70]
    Some remaining puzzles in hydrogen electrocatalysis mechanisms ...
    Jul 17, 2024 · We summarize some of the most intensively debated puzzles in HER/HOR mechanisms on platinum surfaces and the evolution and current status of our understanding.
  71. [71]
    Aluminum E-pH (Pourbaix) diagram
    Aluminum E-pH (Pourbaix) Diagram. The E-pH diagram of aluminum is one of the simplest E-pH diagrams. It will be used here to demonstrate how such diagrams ...
  72. [72]
    On the fate of high-resolution electron energy loss spectroscopy ...
    From its advent, high-resolution electron energy loss spectroscopy (HREELS) has emerged as one of the most versatile tools in surface science.
  73. [73]
    High Resolution Electron Energy Loss Spectroscopy (HREELS)
    Aug 7, 2025 · We review the principles of high resolution electron energy loss spectroscopy (HREELS) and of its extensions to time, momentum and spin ...Missing: seminal | Show results with:seminal
  74. [74]
    High resolution electron energy loss spectroscopy - ScienceDirect.com
    This paper provides a good example of the use of EELS to investigate the surface structure and stoichiometry of complex oxide materials by monitoring the ...Missing: seminal | Show results with:seminal
  75. [75]
    Sum frequency generation vibrational spectroscopy (SFG-VS) for ...
    This review focused on the SFG-VS spectral lineshape and some controversial issues in SFG-VS that are likely be crucial for the application of quantitative SFG- ...
  76. [76]
    Investigating buried polymer interfaces using sum frequency ... - NIH
    This paper reviews recent progress in the studies of buried polymer interfaces using sum frequency generation (SFG) vibrational spectroscopy.
  77. [77]
    Ultraviolet photoelectron spectroscopy: Practical aspects and best ...
    When helium is used, the most convenient emission line is the He I line ... However, this can sometimes be a disadvantage in that the valence band UPS spectrum ...
  78. [78]
    X-ray Photoelectron Spectroscopy (XPS) - Chemistry LibreTexts
    Aug 21, 2022 · Here, binding energy is the energy of an electron attracted to a nucleus; photon energy is the energy of X-ray photons being used by the ...
  79. [79]
    X-ray Photoelectron Spectroscopy (XPS)
    ... energy via an electron analyzer makes it possible to calculate the binding energy: Eb = hn + Ek + f. Since binding energies of core electrons are ...
  80. [80]
    Auger Electron Spectroscopy - an overview | ScienceDirect Topics
    Auger electron spectroscopy (AES) is a characterization technique that is commonly used for the study of material surface.
  81. [81]
    NEXAFS Spectroscopy
    NEXAFS spectroscopy refers to the absorption fine structure close to an absorption edge, about the first 30eV above the actual edge.
  82. [82]
    Near-edge X-ray absorption fine structure spectroscopy in studies of ...
    Apr 23, 2023 · This article reviews the application of near-edge X-ray absorption fine structure (NEXAFS) spectroscopy to characterization of self-assembled monolayers (SAMs)
  83. [83]
    First-Principles Predictions of Near-Edge X-ray Absorption Fine ...
    Apr 21, 2017 · Near-edge X-ray absorption fine structure (NEXAFS) spectroscopy can reveal molecular orientation and ordering by probing element-specific ...
  84. [84]
    Scanning tunneling microscopy---from birth to adolescence | Rev
    Jul 1, 1987 · This collection marks the 35th anniversary of scanning tunneling microscopy (STM) and the 30th anniversary of atomic force microscopy (AFM).
  85. [85]
    Surface Studies by Scanning Tunneling Microscopy | Phys. Rev. Lett.
    Jul 5, 1982 · The papers included in the collection have been made free to read. References (14). G. Binnig, H. Rohrer, Ch. Gerber, and E.
  86. [86]
    Atomic Force Microscope | Phys. Rev. Lett.
    These papers, all published in the Physical Review journals, highlight the positive impact that STM and AFM have had, and continue to have, on physical science ...
  87. [87]
    Scanning Electron Microscopy - Nanoscience Instruments
    Scanning electron microscopy is a highly versatile technique used to obtain high-resolution images and detailed surface information of samples.
  88. [88]
    Cross-section preparation for transmission electron microscopy of ...
    The cross-section preparation technique based on sandwiched face-to-face layered structures has been developed to obtain side views of interface structures ...
  89. [89]
    Low-energy electron diffraction for surface structure analysis
    Low-energy electron diffraction (LEED) is the most important technique for studying the atomic structures of crystalline surfaces.
  90. [90]
    Reflection high energy electron diffraction studies of epitaxial growth ...
    May 1, 1986 · Reflection high‐energy electron diffraction (RHEED) is so well suited to studies of epitaxial growth that it has become the standard in situ ...
  91. [91]
    X-ray diffraction under grazing incidence conditions
    In this Primer, grazing incidence X-ray diffraction (GIXD) is comprehensively introduced, starting from basic considerations on X-ray diffraction at crystals ...
  92. [92]
    Noncontact atomic force microscopy: Bond imaging and beyond
    The bond-resolved NC-AFM presents chemical bonds as apparent interconnectivity with fundamental information of the bond length, bond angle and even bond order.
  93. [93]
    Molecular Beam Epitaxy - Science
    Molecular beam epitaxy is an ultrahigh vacuum technique for growing very thin epitaxial layers of semiconductor crystals.
  94. [94]
    Molecular beam epitaxy - IOPscience
    This article reviews the major physico-chemical aspects of molecular beam epitaxy (MBE), especially as applied to the deposition of thin epitaxial films.
  95. [95]
    A brief review of atomic layer deposition: from fundamentals to ...
    Atomic layer deposition (ALD) is a vapor phase technique using sequential, self-limiting reactions to produce thin films with exceptional conformality.
  96. [96]
    Kinetic and Thermodynamic Modified Wulff Constructions for ...
    Wulff constructions are a powerful tool to predict the shape of nanoparticles, which strongly influences their performance in catalysis, sensing, and surface- ...Introduction · The Modified Wulff Construction · The Kinetic Modified Wulff...