Fact-checked by Grok 2 weeks ago

Surface diffusion

Surface diffusion is the thermally activated migration of adsorbed atoms (adatoms), molecules, or atomic clusters across of a , typically occurring at rates significantly higher than bulk diffusion due to lower energy barriers in the two-dimensional surface layer. The was first systematically studied in the early , with initial observations reported in 1918. This process is driven by gradients in that vary along , leading to confined strictly to , as described by Fick's adapted for : J = -D_s \nabla \mu, where J is the , D_s is the , and \mu is the . Surface diffusion plays a pivotal role in numerous and applications, including epitaxial , deposition, , , and processes. More recently, it has gained prominence in the study and manipulation of two-dimensional materials, such as and hexagonal , enabling advances in and nanofluidics as of 2025. In nanocrystal , for example, the relative rates of atom deposition and surface diffusion dictate the final : a high deposition-to-diffusion promotes localized at high-curvature sites like corners, yielding branched structures such as octapods, while a low enables atoms to redistribute evenly, forming more isotropic shapes like cuboctahedrons. This mechanism is essential for controlling the properties of used in , , and , and it underpins nanoscale surface evolution in modern techniques. At the atomic level, surface diffusion proceeds via distinct mechanisms, primarily adatom hopping—where an isolated adatom jumps between adjacent adsorption sites on the surface lattice—and exchange diffusion, in which the adatom swaps positions with an underlying atom. Both are thermally activated processes, with the diffusion coefficient exhibiting Arrhenius behavior: D_s = D_0 \exp(-E_a / kT), where E_a is the barrier (typically 0.5–1 eV for metals), k is Boltzmann's , and T is , allowing diffusion rates to increase exponentially with . Additional pathways, such as step-edge crossing or cluster-mediated transport, can influence long-range diffusion, particularly on stepped or defective surfaces, affecting overall mass transport in practical systems.

Fundamentals

Definition and Scope

Surface diffusion refers to the thermally activated motion of adsorbed atoms (known as adatoms), molecules, or atomic clusters across the surface of a solid material at the atomic scale. This process enables the rearrangement and transport of surface species, which is fundamental to various surface phenomena in and . Adatoms are singly bound atoms on the surface, distinct from those embedded in the , and they typically occupy specific adsorption sites such as atop (directly above a atom), bridge (positioned between two adjacent atoms), or hollow (located in a site surrounded by three or more atoms). The scope of surface diffusion is confined to processes occurring on the two-dimensional surface layer, at temperatures below the of the substrate, where the material remains solid. Unlike bulk diffusion, which involves three-dimensional movement within the interior of the material, surface diffusion is restricted to the outermost atomic layers and excludes transport in gas-phase or liquid environments. This distinction highlights its unique role in surface-specific behaviors, such as epitaxial growth and thin-film formation. Early observations of surface diffusion date back to the 1950s, when field emission microscopy was employed to study the migration of adsorbed species on metal surfaces, notably by Robert Gomer at the . A pivotal advancement occurred in 1966, with the first direct imaging of individual adatom hops using field ion microscopy, as demonstrated by G. Ehrlich and F. G. Hudda on surfaces, providing atomic-resolution evidence of diffusive motion. These developments laid the groundwork for modern understanding of surface dynamics.

Importance in Materials Science and Chemistry

Surface diffusion plays a pivotal role in by facilitating atomic rearrangements that drive , enabling the formation of stable low-energy configurations on solid surfaces during processes like annealing or exposure to reactive environments. It is essential for , where adatoms migrate across terraces to incorporate into growing islands or steps, influencing the and quality of crystalline structures. In , surface diffusion underpins reaction facilitation by allowing adsorbed species to explore the surface and reach reactive sites, thereby controlling the and selectivity of surface-mediated processes. In heterogeneous catalysis, surface diffusion is critical for transporting reactants and intermediates to active sites on catalyst surfaces, thereby dictating reaction pathways and overall efficiency. For instance, rapid diffusion of adsorbates like on catalysts enhances the performance of electrodes by preventing site blocking and promoting continuous turnover. This transport mechanism is particularly vital in supported metal catalysts, where diffusion rates can limit or accelerate the supply of reagents under operating conditions, impacting industrial processes such as synthesis or hydrocarbon reforming. Within , surface diffusion governs epitaxial growth by determining how deposited atoms integrate into lattice sites, which is crucial for fabricating high-quality thin films in devices. It also influences , where atomic migration along particle surfaces promotes neck formation and densification, enabling the production of dense ceramics and metals with enhanced mechanical properties. At interfaces, surface diffusion contributes to processes by allowing selective and repassivation, as seen in dealloying where atoms diffuse to form porous structures that affect material durability. Beyond these core areas, surface diffusion underpins advancements in nanotechnology through self-assembly, where controlled migration of nanoparticles on substrates leads to ordered structures like chains or lattices for applications in sensors and electronics. In energy technologies, such as fuel cells, managing surface diffusion in catalyst layers stabilizes active phases and improves longevity by mitigating atomic rearrangements under electrochemical stress. These roles highlight surface diffusion's interdisciplinary significance in enabling precise control over nanoscale phenomena for technological innovation.

Atomic-Scale Mechanisms

Adatom Diffusion

Adatom diffusion refers to the thermally activated or quantum-assisted movement of individual atoms adsorbed on a solid surface, distinct from bulk diffusion due to the reduced dimensionality and weaker bonding in the surface layer. This process is fundamental to phenomena such as epitaxial growth, catalysis, and surface reconstruction, where isolated adatoms—atoms bound to the surface but not incorporated into the lattice—migrate across lattice sites. The primary pathways involve overcoming potential energy barriers associated with lattice configurations, with the diffusion barrier typically denoted as E_{\text{diff}}. On face-centered cubic (fcc)(111) surfaces, these barriers for adatom hopping range from approximately 0.01 to 0.2 eV, depending on the metal substrate and adatom species. The most prevalent mechanism is hopping, in which an adatom detaches from its adsorption , such as a or position, and jumps to an adjacent , surmounting an barrier E_{\text{diff}} at the , often a or configuration between sites. This process dominates on open or stepped surfaces and is characterized by nearest-neighbor displacements, though longer jumps can occur under specific conditions. In contrast, the exchange mechanism involves the adatom swapping positions with a neighboring substrate atom, effectively embedding the adatom into the lattice while ejecting the substrate atom to the surface; this is particularly common on close-packed surfaces like fcc(100) or (111), where the barrier is lowered due to concerted atomic motion. For light adatoms like , quantum tunneling provides a temperature-independent pathway, allowing the atom to penetrate the energy barrier via wavefunction overlap rather than classical activation, especially at cryogenic temperatures. This effect is prominent for on surfaces below 50 K, where classical hopping is suppressed, enabling observation of via techniques like scanning tunneling microscopy. Vacancy-mediated , though possible, is rare on surfaces compared to bulk materials, as it requires the presence of surface vacancies that facilitate adatom motion by filling and reforming defects; this mechanism is more relevant in systems or under high vacancy concentrations but contributes minimally to isolated adatom transport on clean metal surfaces.

Cluster Diffusion

Cluster diffusion refers to the mobility of groups of atoms or molecules adsorbed on a surface, distinct from single adatom movement, and typically involves cooperative processes that enable the displacement of the 's . One primary mechanism is the detachment and reattachment of peripheral atoms at the , where atoms temporarily leave the , migrate across adjacent sites, and rejoin, collectively shifting the entire without full . This periphery diffusion process is prevalent for compact s and has been observed in systems like Ag s on Ag(100) surfaces, where atom hops dominate the overall motion. Concerted mechanisms, in contrast, involve the simultaneous or of the entire without , often occurring via terrace across flat surface planes. In such cases, the as a whole overcomes a collective barrier to glide or pivot, which can be energetically competitive for small, symmetric like those with 4-6 atoms on Pt(111). For example, on Pt(111), compact Pt exhibit island through concerted motion alongside peripheral processes. These mechanisms highlight the role of geometry in facilitating non-dissociative transport. The diffusion coefficient of clusters exhibits a strong size dependence, generally decreasing as the number of atoms N increases, with traditional mean-field models predicting D \sim 1/N for periphery-dominated motion on supports. This scaling arises because larger clusters have proportionally fewer mobile edge atoms relative to the total mass, slowing center-of-mass displacement; such behavior has been confirmed for supported metal like faceted fcc metals on oxide-like substrates. Specific examples include _2 dimers on (111), which diffuse primarily via an exchange mechanism where the dimer inserts into the substrate lattice, facilitating rapid mobility through substitutional swaps. Recent studies since 2010 have revealed anomalous behaviors in small clusters, including superdiffusion characterized by , where intermittent long jumps lead to faster-than-normal spreading. For instance, C_{60} molecules on Au(111) surfaces display superdiffusive trajectories due to Lévy-like flight patterns in their motion, transitioning from ballistic to diffusive regimes over time. This superdiffusion underscores how edge adatom precursors can contribute to enhanced cluster mobility in low-coordination environments.

Kinetic Description

Jump Rates and Diffusion Coefficients

Surface diffusion is quantitatively described through probabilistic models that capture the kinetics of atomic motion on a lattice. In the random walk model, adatoms perform uncorrelated jumps between adjacent sites on the surface , providing a foundational framework for understanding diffusion dynamics at low coverage. This approach assumes Markovian processes where each jump is independent, allowing the overall to emerge from the statistics of individual hops. The fundamental quantity in this model is the jump rate \Gamma, which represents the frequency of successful hops from one lattice site to a neighboring site. According to transition state theory, \Gamma = \nu \exp(-E_{\text{diff}} / k_B T), where \nu is the attempt frequency, typically ranging from $10^{12} to $10^{13} s^{-1} and related to surface vibrational modes, E_{\text{diff}} is the energy barrier for diffusion (often arising from hopping mechanisms), k_B is the Boltzmann constant, and T is the absolute temperature. For example, on metal surfaces like Cu(001), \nu incorporates the number of possible saddle points (e.g., 4 for a square lattice), aligning with phonon frequencies around $10^{13} s^{-1}. The mean squared displacement \langle \Delta r^2 \rangle quantifies the extent of adatom spreading over time in this random walk framework: \langle \Delta r^2 \rangle = 2 d \Gamma a^2 t, where d is the dimensionality (typically d=2 for isotropic surface diffusion), a is the lattice spacing, and t is the elapsed time. This expression derives from the statistics of successive uncorrelated jumps, with the factor $2d arising from the equipartition in each dimension. The diffusion coefficient D, which characterizes the macroscopic rate of spreading, connects directly to the microscopic jump kinetics via D = \Gamma a^2 / z, where z is a dimensionality factor (z=1 for 1D diffusion along channels or 2D surface planes). In 2D, this yields D = \Gamma a^2, linking the probabilistic jump process to Fickian diffusion behavior observed in experiments like quasi-elastic helium scattering on systems such as /(001). These relations enable predictions of adatom mobility essential for processes like epitaxial growth.

Temperature and Coverage Dependence

Surface diffusion exhibits a strong temperature dependence, typically following Arrhenius behavior where the diffusion coefficient D is expressed as D = D_0 \exp\left(-\frac{E_\mathrm{diff}}{kT}\right), with D_0 the pre-exponential factor, E_\mathrm{diff} the activation energy for diffusion, k Boltzmann's constant, and T the temperature. Plotting \log D versus $1/T yields a straight line with slope -E_\mathrm{diff}/(k \ln 10), allowing extraction of E_\mathrm{diff} from experimental data; this linear regime holds over wide temperature ranges for thermally activated hopping mechanisms dominant in most systems. At low temperatures, deviations from Arrhenius linearity occur, often indicating quantum tunneling through the diffusion barrier, as observed for oxygen atoms on cold water ice surfaces below 25 K where tunneling enables diffusion rates higher than classical predictions. Coverage \theta, defined as the fraction of surface sites occupied by adatoms, significantly influences through lateral interactions. In systems with repulsive adatom-adatom interactions, D generally decreases with increasing \theta because the effective barrier for hops rises due to electrostatic or repulsions, reducing the availability of low-energy pathways. This coverage-induced slowdown can lead to a critical \theta where becomes favorable, as mobile adatoms aggregate rather than diffuse freely, marking a in surface during adsorption processes. The activation energy E_\mathrm{diff} varies with coverage in surfaces exhibiting repulsive interactions, typically increasing as \theta rises due to the higher energy cost of distorting the adlayer during hops. For example, on the W(110) surface, oxygen adatom diffusion shows E_\mathrm{diff} rising from approximately 0.6 eV at low \theta \approx 0 to 1.2 eV at high \theta \approx 1, reflecting stronger repulsive effects and substrate stiffening at higher coverages. Experimental observations in many systems, such as indium on Si(111), indicate that the pre-exponential factor \nu (related to D_0 via the lattice constant and dimensionality) remains largely coverage-independent, suggesting that the attempt frequency for hops is governed primarily by vibrational modes rather than adlayer density. These dependencies modify the base diffusion coefficient defined in kinetic models, emphasizing the need to incorporate interaction potentials for accurate predictions at varying conditions.

Diffusion Regimes

Coverage-Dependent Regimes

Surface diffusion regimes are classified based on adatom coverage θ, which influences the nature of atomic motion from isolated hops to collective transport and interaction-dominated behavior. At low coverages, typically θ < 0.01 monolayers (ML), adatoms move independently with minimal interactions, leading to uncorrelated random walks on the surface lattice. Tracer diffusion characterizes this dilute limit, quantifying the mean-squared displacement of individual adatoms through the relation D_t = \frac{1}{4} \Gamma a^2, where Γ is the jump rate and a is the lattice spacing; it reflects single-particle mobility without significant adatom-adatom correlations. Intrinsic diffusion, often synonymous with or closely related to in uniform terraces lacking defects or traps, describes the bare surface mobility of adatoms driven solely by thermal activation over the potential energy barrier, independent of concentration gradients or external sources/sinks. For example, on Si surfaces, the intrinsic diffusion coefficient for Sn and Ga adatoms remains concentration-independent during cluster growth, highlighting its relevance to terrace-limited transport. As coverage increases beyond the dilute regime, typically θ > 0.1 , correlated diffusion emerges, where adatom interactions introduce memory effects and non-random sequences, reducing the effective compared to the low-coverage limit; this is evident in systems like O on W(110), where velocity autocorrelations decay as a . Chemical , prominent at higher coverages, governs collective in response to adatom gradients, incorporating a thermodynamic factor S such that D_c = S D_{CM}, where D_{CM} is the center-of-mass ; it deviates from tracer values due to coverage-induced fluctuations and interactions. Coverage effects on rates, as detailed elsewhere, further modulate these regimes by altering barriers through lateral adatom repulsions or attractions. In the presence of island formation at moderate to high coverages, enhanced perimeter diffusion occurs along edges, where lower barriers (e.g., 0.25 for Ag on Ag(001) steps versus 0.40 on terraces) facilitate faster adatom motion due to partial coordination; this periphery diffusion contributes to overall island shape evolution and scales as D \propto N^{-3/2} for large islands of size N.

Environmental and Interaction Effects

Adsorbate-adsorbate interactions play a crucial role in modifying the surface diffusion coefficient D by altering the effective energy barriers for atomic jumps. Repulsive interactions, such as long-ranged dipole-dipole forces arising from the polar nature of adsorption bonds, typically enhance at higher coverages by repelling adatoms from occupied sites and promoting mobility. For instance, the dipole-dipole interaction scales as r^{-3} and is generally repulsive, leading to shifts in concentration profiles and increased D(\theta) that can be modeled as D(\theta) = D^* [1 + \alpha \theta (1 - \theta)] e^{\alpha \theta}, where \alpha = \beta V_0 and V_0 is the interaction . An example is lithium adatoms on a Dy-Mo(112) surface, where repulsive interactions result in maximum at a coverage \theta_{\max} \approx 0.33, with V_0 \approx 0.16 at 600 K. Attractive interactions, often mediated by substrate strain, conversely reduce D by creating binding potentials that cluster adatoms and raise jump barriers. These forces stem from elastic relaxation of the substrate lattice around adsorbed particles, generating indirect attractions via strain fields. First-principles calculations on systems like alkali atoms on metal surfaces demonstrate that such substrate-mediated attractions dominate at longer ranges, suppressing diffusion and favoring ordered structures over random walks. Gas or liquid environments introduce additional physisorbed layers that impede surface diffusion by site blocking and increased scattering. In ultra-high vacuum (UHV), clean surfaces enable measurement of intrinsic diffusion rates, but ambient conditions lead to adsorption of residual gases like water or oxygen, forming weakly bound layers that occupy diffusion pathways and lower effective D by factors of 10 or more. For oxide surfaces, ambient exposure results in hydroxylated or oxidized layers that alter adatom binding and slow transport compared to reduced UHV states. Surface strain and defects further modulate , often accelerating it along specific features while creating barriers elsewhere. Strained regions, such as those near lattice mismatches, lower coordination and facilitate faster adatom hops due to weakened binding. is typically enhanced parallel to step edges on vicinal surfaces, but perpendicular crossing incurs higher energies, with step barriers approximately 0.5 greater than on flat terraces. For adatoms on (110), the barrier for step-edge reaches 1.5 , compared to 0.5 lower on terraces, highlighting defect-induced in . At high temperatures exceeding the bulk , surface diffusion transitions to a liquid-like regime, characterized by continuous viscous flow rather than discrete activated jumps, enabling ultra-high mobilities. Recent experiments employing heating on polycrystalline films reveal rapid emergence of phases, where diffusion coefficients exceed solid-state values by orders of magnitude due to the fluid nature of the surface layer. Such studies, post-2015, underscore how localized heating induces premelting-like behavior, with adatom transport governed by hydrodynamic effects in the nascent film.

Anisotropy and Surface Structure

Orientational Anisotropy

Surface diffusion exhibits orientational arising from the dependence of adatom mobility on the specific of the substrate, primarily influenced by variations in atomic coordination and surface openness. Close-packed planes, such as the fcc(111) surface, provide higher coordination for adatoms in three-fold hollow sites, resulting in lower activation energies for , typically around 0.3 eV. In contrast, more open planes like fcc(100), with four-fold hollow sites offering less coordination, exhibit higher barriers of approximately 0.6 eV. This orientational dependence leads to markedly faster diffusion on close-packed surfaces. For instance, self-diffusion of Cu adatoms on Cu(111) occurs with an of about 0.04 via simple hopping between adjacent hollow sites, whereas on Cu(100), the barrier rises to roughly 0.5–0.6 , often involving more complex exchange mechanisms; at elevated temperatures, this results in diffusion rates on (111) that are 10–100 times faster than on (100), depending on the exact conditions. The enhanced stability in higher-coordination sites on (111) reduces the penalty for adatom during hops, favoring rapid surface on these orientations. Similar effects manifest on body-centered cubic (bcc) surfaces, where the (110) plane—characterized by zigzag rows of close-packed atoms—shows pronounced orientational . Adatom along these rows proceeds with lower barriers due to favorable nearest-neighbor interactions, achieving rates up to 5 times higher than to the rows, where crossing the troughs between rows incurs higher costs. This row-directed preference underscores how surface geometry dictates overall kinetics across different orientations.

Directional Anisotropy

Directional anisotropy refers to the variation in adatom diffusion rates along different in-plane directions on a surface, arising primarily from the intrinsic geometry of the surface lattice. This phenomenon is particularly pronounced on surfaces with linear structural features, such as atomic rows or troughs, which create preferential pathways for adatom by lowering energy barriers in specific directions compared to others. The resulting differences in diffusion coefficients can span orders of magnitude, influencing processes like island nucleation and growth morphology. Channeled diffusion exemplifies this on face-centered cubic (fcc) (110) surfaces, where adatoms move much faster along the close-packed atomic rows (parallel to the [1̄10] direction) than across the open channels (perpendicular, along ). This leads to quasi-one-dimensional diffusion behavior, with the ratio of parallel to perpendicular diffusion coefficients (D_parallel / D_perp) exceeding 100 in many metal systems, such as Cu(110) and Pt(110). The channeled structure confines adatoms to low-energy paths along the rows, while cross-channel hops require overcoming higher barriers due to the increased coordination changes and electrostatic repulsion. Corrugation effects in the surface landscape further enhance directional preferences by forming elongated troughs or valleys that guide adatom motion. On the reconstructed Si(111)-7×7 surface, for instance, the complex arrangement of adatom rest atoms and dimers creates a with lower diffusion barriers along the troughs between structural faults, facilitating preferential in those directions over random 2D hopping. This stems from the varying bonding environments, where adatoms experience reduced activation energies (around 1.1 eV) for hops along the troughs compared to perpendicular motions exceeding 1.5 eV. Surface reconstructions can also impose strong directional control, as seen on the Si(100)-2×1 surface, where buckled dimer rows align in parallel lines and channel adatom diffusion predominantly along their length. Adatoms, such as or H, exhibit activation barriers of approximately 0.5–1.0 parallel to the rows but over 1.5 eV perpendicular, resulting in highly elongated diffusion trajectories and string-like island formation. This row-directed motion arises from the π-bonding within dimers, which stabilizes intra-row hops while hindering inter-row crossing.

Applications

Heterogeneous Catalysis

In heterogeneous catalysis, surface diffusion often limits the reaction rate, as the migration of adsorbed atoms or molecules (adatoms) to active sites becomes the rather than intrinsic reaction . This diffusion-limited arises when adatom transport is slower than adsorption or desorption processes, leading to coverage gradients across the surface that influence selectivity and turnover frequency. For instance, in like metal-organic frameworks, the critical diffusion length—typically on the order of hundreds of nanometers (e.g., 120–400 nm in UiO-66 thin films)—balances accessibility to active sites while maximizing product yield, as shorter paths enhance rates but may reduce specificity. Site hopping, a key mechanism in surface diffusion, facilitates the movement of adatoms between adsorption sites, enabling spillover to undercoordinated positions that exhibit higher reactivity. On Pt(111) surfaces, (CO) adatoms preferentially hop from low-coordination terrace sites to step edges, where binding is stronger and oxidation proceeds more efficiently, contributing to overall catalytic performance in CO oxidation. This process is governed by statistical rate theory, with migration rates determined by equilibrium exchange and differences across the surface. Diffusion also drives island and ensemble effects, where mobile intermediates aggregate into clusters that form active catalytic ensembles. On transition metal surfaces like Cu(111), adatom ejection from defects followed by surface diffusion leads to subnanometer cluster formation under reaction conditions, dramatically boosting activity—for example, Cu₃ clusters accelerate oxidation by factors up to 2 × 10⁶ compared to extended terraces. These ensembles optimize binding energies for multiple reactants, enhancing reaction pathways that require specific geometric arrangements. A critical example occurs in ammonia synthesis on iron catalysts, where nitrogen (N) adatoms must diffuse across the Fe(111) surface to meet hydrogen adatoms at hydrogenation sites. At typical operating temperatures of 700 K, N adatoms exhibit high mobility due to surface disordering, diffusing approximately 10 nm in 1 ms, which is essential for sustaining the reaction rate under industrial conditions of 600–800 K and high pressure. This rapid transport prevents site blocking and ensures efficient recombination steps in the Haber-Bosch process.

Thin Film Growth and Nanotechnology

Surface diffusion is pivotal in thin film growth, particularly in dictating epitaxial modes by enabling adatoms to migrate across the before incorporating into the film structure. In the Frank-van der Merwe mode, layer-by-layer predominates when surface diffusion is sufficiently rapid, allowing adatoms to reach step edges and complete monolayers without excessive of new islands. Slower diffusion, conversely, promotes island , shifting the process toward three-dimensional in Volmer-Weber or Stranski-Krastanov modes, where adatoms aggregate into isolated clusters due to limited mobility. Ostwald ripening represents a key diffusion-mediated coarsening mechanism during thin film evolution, wherein smaller islands supply material to larger ones via adatom detachment and surface transport, thereby reducing overall . This process is especially pronounced in the later stages of , influencing island size and film uniformity by favoring thermodynamically stable configurations. effects can modulate ripening rates, with compressive accelerating diffusion pathways and enhancing material redistribution among islands. In applications, surface diffusion governs the of s, enabling precise control over nanoscale structures critical for optoelectronic devices. For example, in the epitaxial growth of InAs quantum dots on GaAs substrates, such as reduce diffusion barriers, increasing adatom mobility to promote larger dot sizes and higher densities while suppressing unwanted . Similarly, in GaAs-on-Si systems, tuned surface diffusion via like or lead adjusts migration lengths, facilitating uniform quantum dot arrays and mitigating mismatch-induced defects. Recent advances since 2015 highlight surface diffusion's role in fabricating van der Waals heterostructures using materials like MoS₂, where controlled adatom diffusion on the basal plane enables high-quality epitaxial stacking and interface formation without covalent bonding. In MoS₂-graphene heterostructures, enhanced surface mobility of precursors during van der Waals epitaxy yields atomically sharp interfaces, improving charge transfer and device performance in . Template-assisted growth on MoS₂ further leverages diffusion to direct lateral expansion of overlayers, such as Sb₂S₃, achieving large-area, defect-free assemblies.

Experimental and Theoretical Methods

Observational Techniques

Observational techniques for surface diffusion encompass both direct visualization of atomic movements and indirect methods that infer diffusion parameters from ensemble behaviors. Direct methods, such as field ion microscopy (FIM), provide atomic-scale resolution of individual adatom hops on metal surfaces under (UHV) conditions and at temperatures typically below 300 K. Pioneered in the mid-20th century, FIM images the surface by ionizing gas atoms (e.g., or ) in a high at the sample tip, allowing real-time tracking of single-atom diffusion events and the determination of hop rates and mechanisms. This technique has been essential for quantifying elementary diffusion processes, including activated hops over barriers on low-index planes like fcc(110) and fcc(111). Scanning tunneling microscopy (STM) complements FIM by enabling the tracking of adatom trajectories on a broader range of substrates, including non-metallic surfaces, through measurement of tunneling currents between a sharp tip and the sample. Operating in UHV, resolves atomic positions and has revealed diffusion pathways, such as nearest-neighbor hops or long-range jumps influenced by surface corrugation. Advancements in the , including video-rate with frame rates up to 80 Hz, have facilitated real-time observation of dynamic processes like adatom and island coarsening, overcoming previous limitations in . Indirect techniques offer averaged insights into diffusion over larger ensembles, avoiding the single-event focus of direct imaging. Quasi-elastic helium atom scattering (QHAS) probes surface dynamics by analyzing the energy broadening of scattered helium beams from adsorbed layers, yielding the diffusion coefficient D via the Debye-Waller factor or linewidth analysis in the quasi-elastic peak. This non-destructive method excels for incommensurate overlayers and has quantified D for systems like CO on Ni(110), revealing coverage-dependent variations due to adsorbate interactions. Thermal desorption spectroscopy (TDS), another indirect approach, infers surface diffusion barriers from the temperature-programmed evolution of desorbed species, modeling the competition between diffusion-limited supply to step edges and desorption kinetics. By fitting TDS peaks to reaction-diffusion equations, energies for can be extracted, often revealing barriers 40-60% of the desorption for weakly bound adatoms. This technique is particularly useful for reactive systems where direct is challenging, such as on metals, and provides complementary data to direct methods for validating kinetic models.

Computational Approaches

Density functional theory (DFT) serves as a cornerstone for computing atomic-scale properties in surface diffusion, particularly the diffusion energy barriers E_\text{diff} and associated migration pathways. By solving the Kohn-Sham equations within the framework of quantum mechanics, DFT enables the determination of stable adsorption sites, transition states, and energy landscapes for adatoms or molecules on crystal surfaces. For instance, calculations on Ag adatoms diffusing on the Ag(111) surface reveal low-energy hopping mechanisms with barriers around 0.05–0.1 eV, highlighting the method's accuracy in capturing electronic interactions. To identify minimum energy paths, the nudged elastic band (NEB) method is routinely employed alongside DFT; this technique optimizes a chain of intermediate images between initial and final states, minimizing forces perpendicular to the path while applying springs to maintain spacing, thus converging to the saddle point efficiently. The climbing image variant of NEB further accelerates convergence by promoting the highest-energy image toward the transition state, making it indispensable for complex surfaces. Kinetic Monte Carlo (KMC) simulations extend DFT-derived parameters to model large-scale, temporally extended diffusion processes on surfaces. In lattice KMC, the system evolves stochastically by selecting diffusion events based on their rates \Gamma_i = \nu_i \exp(-E_{\text{diff},i}/kT), where \nu_i is the attempt frequency from DFT or experiment, allowing prediction of collective behaviors like island nucleation and coarsening over timescales inaccessible to direct dynamics. This approach has been pivotal in simulating epitaxial growth morphologies, such as the formation of stepped or faceted structures in metal-on-metal systems, where adatom diffusion dictates the overall film quality. By incorporating coverage-dependent barriers from DFT, KMC captures realistic spatial correlations, enabling forecasts of diffusion-limited regimes in catalysis and nanotechnology. Molecular dynamics (MD) simulations provide insights into the time-dependent aspects of surface diffusion, revealing dynamic correlations and collective motions at finite temperatures. Classical MD integrates Newton's equations using empirical potentials to track trajectories, but for precise quantum mechanical treatment, ab initio MD (AIMD) couples on-the-fly DFT evaluations with nuclear dynamics, incorporating anharmonic effects and temperature-driven fluctuations. Recent AIMD studies have elucidated quantum nuclear effects, such as zero-point motion influencing light adatom hopping on metal surfaces, leading to enhanced diffusivities beyond classical predictions. These methods are particularly valuable for probing short-time diffusion mechanisms, like precursor-mediated jumps, where vibrational modes couple to translation. Post-2020 advances in potentials (MLPs) have revolutionized computational efficiency by emulating DFT accuracy for extended systems, facilitating large-scale simulations of surface on alloys. Trained on DFT datasets via neural networks or Gaussian approximations, MLPs predict energies and forces rapidly, enabling MD or KMC runs on disordered surfaces like Ni-Mn alloys, where barriers vary with local composition. For example, MLPs have accelerated predictions of vacancy-mediated in , revealing composition-tuned pathways with barriers differing by up to 0.2 from homogeneous metals. These potentials bridge the gap between quantum fidelity and mesoscale modeling, often validated against experimental diffusivities to refine surface process predictions.

References

  1. [1]
    On the role of surface diffusion in determining the shape or ...
    Surface diffusion is a general process that involves the motion of adsorbed atoms (adatoms), molecules, or atomic clusters on the surface of a solid material (1 ...
  2. [2]
    Surface Diffusion: Motion by <IMG ALIGN=BOTTOM ALT="" SRC ...
    In surface diffusion, mass flux is confined to the surface and is is driven by gradients in chemical potential that vary with position along a surface. In this ...
  3. [3]
    Surface Diffusion at Solid Surface: An Atomic View - NASA/ADS
    The diffusion of individual atoms and molecules across the surface of a crystalline solid is a fundamental process that is central to a number of commercially ...
  4. [4]
    [PDF] Surface Diffusion including Adatoms
    Abstract. The aim of this paper is to study continuum models for surface diffusion taking into account free adatoms on the surface, which is of.
  5. [5]
    Diffusion mechanism of Cu adatoms on a Cu(001) surface ...
    The hopping mechanism with a calculated energy barrier of 0.69 eV is found to be favorable over the exchange mechanism with 0.97 eV. We find from the geometry ...
  6. [6]
    [PDF] Measuring surface mass diffusion coefficients by observing step ...
    To show how experiments probing the equilibrium fluctuations of steps can be used to explore the mechanisms of surface mass self-diffusion, ...<|control11|><|separator|>
  7. [7]
    On the role of surface diffusion in determining the shape or ... - PNAS
    Surface diffusion is a general process that involves the motion of adsorbed atoms (adatoms), molecules, or atomic clusters on the surface of a solid material (1 ...
  8. [8]
    Surface Diffusion - an overview | ScienceDirect Topics
    Surface diffusion is defined as the process in which adsorbed molecules move across a solid surface by hopping between active sites, facilitated by sufficient ...
  9. [9]
    Surface Diffusion - Cambridge University Press & Assessment
    Trembulowicz, Artur Ehrlich, Gert and Antczak, Grazyna 2011. Surface diffusion of gold on quasihexagonal-reconstructed Au(100). Physical Review B, Vol. 84 ...
  10. [10]
    Atomic View of Surface Self‐Diffusion: Tungsten on ... - AIP Publishing
    4. History and technique of field‐ion microscopy are summarized by. E. W. ... This material is described by G. Ehrlich, J. Chem. Phys. (to be published) ...
  11. [11]
    Diffusion in surface layers - Taylor & Francis Online
    Studies of surface diffusion can also provide insights into the nature of the bonding at an interface. Finally, in crystal and thin film growth as well as in ...
  12. [12]
    Subsurface diffusion in crystals and effect of surface permeability on ...
    Sep 17, 2019 · Introduction. Diffusion in crystals and on their surfaces is fundamental to control crystal growth, surface structure, and surface morphology.
  13. [13]
    The effect of surface diffusion on surface reaction rates
    We find that surface diffusion can be an important contributing factor to the rate of reaction. In particular, when surface diffusion is rapid relative to the ...
  14. [14]
    The role of dynamics in heterogeneous catalysis: Surface diffusivity ...
    We find that the dynamical behavior of the surface strongly influences how the reaction occurs as a function of temperature.
  15. [15]
    CO surface diffusion on platinum fuel cell catalysts by ...
    Oct 1, 2008 · We report on surface CO diffusion processes in relation to properties of nanoparticle Pt and Pt/Ru fuel cell catalysts. The COad diffusion ...
  16. [16]
    Diffusion on Semiconductor Surfaces | Physics Today - AIP Publishing
    Jul 1, 2001 · When an individual Si atom is deposited onto the surface at room temperature, it diffuses rapidly and quickly finds another atom with which to ...
  17. [17]
    Surface Self-Diffusion Induced Sintering of Nanoparticles | ACS Nano
    Nov 1, 2024 · Surface self-diffusion is thermally activated due to the tendency to minimize surface energy, with a higher rate at higher temperatures, ...
  18. [18]
    A study at the molecular level of the mechanism of surface diffusion ...
    Surface diffusion is a fundamental event in processes as relevant from the economic point of view as dealloying of alloys [1]or stress corrosion cracking [2].
  19. [19]
    The Mechanisms for Nanoparticle Surface Diffusion and Chain Self ...
    Aug 20, 2015 · Self-assembly kinetics are consistent with a diffusion-driven mechanism; we attribute the change in self-assembly pathway to the increased self- ...
  20. [20]
    Highly stable and active catalyst in fuel cells through surface atomic ...
    Oct 18, 2024 · We developed a strategy that limits the atomic diffusion within surface layers, fostering the phase transition and shape retention during thermal treatment.
  21. [21]
    Mechanisms for adatoms diffusing on metal fcc(111) surfaces
    Oct 20, 2000 · Mechanisms for adatoms diffusing on metal fcc(111) surfaces. Author links open overlay panel. Jun Zhuang , Lei Liu , Xijing Ning , Yufen Li.
  22. [22]
    Jump processes in surface diffusion - ScienceDirect.com
    Here we will briefly review what has been learned about how diffusion of single metal atoms takes place on metal surfaces, what sort of atomic jumps occur and ...
  23. [23]
    Adatom-dependent diffusion mechanisms on a surface | Phys. Rev. B
    May 25, 2010 · A Ag adatom diffuses over the surface through an exchange mechanism in which the adatom and the substrate Ag atoms exchange repeatedly in each diffusion step.
  24. [24]
    Hydrogen tunneling on a metal surface: A density-functional study of ...
    Hydrogen tunneling on a metal surface: A density-functional study of H and D ... These observations strongly suggest that H diffusion in the low-temperature ...
  25. [25]
    Hydrogen adsorption and diffusion on Pd(1 1 1) - ScienceDirect.com
    The adsorption, diffusion and ordering of hydrogen on Pd(1 1 1) was studied by scanning tunneling microscopy in the temperature range of 37–90 K. At low ...
  26. [26]
    Vacancy-Mediated and Exchange Diffusion in a Surface Alloy
    Mar 24, 2003 · On the other hand, the local spreading of the concentration gradient depends on the rapid, short-distance, vacancy-mediated diffusion process.
  27. [27]
    Cluster diffusion and surface morphological transitions on Pt (111 ...
    We show that for compact clusters with 4 to 6 atoms, this mechanism competes energetically with that of island diffusion through concerted motion. However, as ...
  28. [28]
    Silicon Reactivity at the Ag(111) Surface | Phys. Rev. Lett.
    Jul 8, 2015 · This evidence suggests that once an embedded dimer is formed, it will act as a seed for the rapid exchange reaction of the Si adatoms diffusing ...Abstract · Article Text · Supplemental Material
  29. [29]
    Stochastic analysis of movements on surfaces: The case of C 60 on ...
    Jul 16, 2015 · Superdiffusion includes Levy flights (LFs) and Levy walks (LWs) and may display strong fluctuations [23], [24], where LFs are discontinued ...
  30. [30]
    [PDF] Collective and single particle diffusion on surfaces
    Abstract. We review in this article the current theoretical understanding of collective and single particle diffusion on surfaces and how it relates to the ...Missing: 1950s | Show results with:1950s
  31. [31]
    [PDF] Density Functional Calculations of Self-Diffusion and Au ... - MACAU
    ... t = 4 Γ a2. t,. (2.23) where D denotes the surface diffusion coefficient, Γ is the hopping rate, and a is the nearest- neighbor distance in the (001) ...
  32. [32]
  33. [33]
    Diffusion of Interacting Lattice Gases on Heterogeneous Surfaces
    Figure 6 shows the coverage dependence of surface diffusion coefficients and the normalized mean square fluctuations 〈(δN)2〉/〈N〉 for several representative ...
  34. [34]
  35. [35]
    [PDF] Surface Diffusion Studies by Analysis of Cluster Growth Kinetics
    Aug 27, 1990 · (i) For the surface diffusion mechanism the difference of the actual concentration at the cluster surface, c'(r), and the average free adato ...
  36. [36]
    [PDF] Adatom interaction effects in surface diffusion
    It is shown that by analyzing a time dependence of adatom concentration profiles one can estimate the type and strength of interatomic interactions. Key words: ...
  37. [37]
    First-principles study of substrate-mediated interactions on a ...
    Sep 27, 2005 · The similarity of the adsorbate potential that we find here suggests that surface-state electrons also underlie the interactions shown in Fig. 2<|control11|><|separator|>
  38. [38]
    Surface Physics and Its Relation to Vacuum Science - ScienceDirect
    The interaction of gas molecules with a solid surface is the basis for vacuum technology. Vacuum pumps and pumping, gas transport through a physical system, ...
  39. [39]
    Oxide surfaces as environmental interfaces - ScienceDirect.com
    Many oxide surfaces under UHV conditions lose oxygen atoms to lower the free energy which results in a reduced state of the metal cation. However, since ...<|control11|><|separator|>
  40. [40]
    Diffusion of Fe atoms on W surfaces and films and along surface steps
    Further, the activation energy barrier for Fe diffusion on 1 - ML Fe ∕ W ( 110 ) films has been estimated at Δ E = 1.2 eV . As to the Fe ∕ W ( 100 ) system ...
  41. [41]
    (PDF) Calculation of the activation energy for surface self-diffusion of ...
    Aug 5, 2025 · ... Activation energy of surface self-diffusion E. d. ~in eV! for ... Adatom ~111!~100! Sc 0.21 0.12. Ti 0.37 0.67. Co 0.23 0.42. Y 0.25 0.45.
  42. [42]
    [PDF] Thermal dynamics at surfaces - EPFL
    Oct 5, 2009 · For intrinsic diffusion the enthalpy change is the binding energy difference between initial and transition state and often referred to as ...
  43. [43]
    Diffusion rates of Cu adatoms on Cu(111) in the presence of an ...
    Sep 2, 2005 · The surface diffusion of Cu adatoms in the presence of an adisland at fcc or hcp sites on Cu(111) is studied using the embedded atom model ...
  44. [44]
    Diffusion barriers for Ag and Cu adatoms on the terraces and step ...
    Dec 9, 2009 · The energy barrier for diffusion of Cu adatoms on Cu(100) via exchange is, on the other hand, found to be much larger, close to 1 eV, in ...
  45. [45]
    [PDF] Migration energy barriers and diffusion anisotropy of point defects ...
    Point defects on the (110) surface have the highest diffusivities among the three surfaces and keep a fixed anisotropy factor (1/3) independent of temperature.
  46. [46]
    Atomic events in surface diffusion (Chapter 3)
    These are followed in Fig. 3.3 by fcc(100) and (111) and bcc(100), (110), and (111) planes in Fig. 3.4.
  47. [47]
    Long jumps contribution to the adatom diffusion process near the ...
    Dec 27, 2013 · However, it can exceed 5% and reach 12%. This important contribution of long jumps is due to the lowest activation energy (see Table 1) and ...Missing: seminal | Show results with:seminal
  48. [48]
    Effects of surface structure and of embedded-atom pair functionals ...
    ... diffusion will occur more readily on the (111) surface and that self-diffusion on the (110) surface exhibits directional anisotropy. The diffusion rate ...Missing: channeled | Show results with:channeled
  49. [49]
    None
    Nothing is retrieved...<|control11|><|separator|>
  50. [50]
    Diffusion of an adsorbed Si atom on the Si(111)-(7×7) surface
    Aug 7, 2025 · Si surface diffusion on Si(111)-7 × 7 at room temperature is strongly hindered due to the high energy barrier (E b = 1.14 eV) that separates ...
  51. [51]
    None
    Nothing is retrieved...<|separator|>
  52. [52]
    Energetics of atomic hydrogen diffusion on Si(100) - ScienceDirect
    A strong anisotropy is predicted for hydrogen motion: H should diffuse mainly along dimer rows, where activation energies are ∼ 1.3 eV, while the barrier ...
  53. [53]
    Current-Induced One-Dimensional Diffusion of Co Adatoms on ...
    One-dimensional diffusion of Co adatoms on graphene nanoribbons has been induced and investigated by means of scanning tunnelling microscopy (STM).
  54. [54]
    Diffusion-programmed catalysis in nanoporous material - Nature
    Feb 3, 2025 · In the realm of heterogeneous catalysis, the diffusion of reactants into catalytically active sites stands as a pivotal determinant ...
  55. [55]
    A statistical rate theory description of CO diffusion on a stepped Pt ...
    Jul 22, 1999 · The statistical rate theory approach is used to describe far-from-equilibrium diffusion of carbon monoxide on a stepped Pt(111) surface at ...
  56. [56]
    Formation of active sites on transition metals through reaction-driven ...
    Apr 6, 2023 · We used density functional theory calculations to elucidate the conditions that lead to cluster formation and show how adatom formation energies ...
  57. [57]
    [PDF] Kinetics of Epitaxial Growth: Surface Diffusion and Nucleation
    Epitaxial growth involves depositing a mono-crystalline film. Kinetics are determined by surface diffusion and nucleation, where adatoms either form new ...
  58. [58]
    [PDF] Epitaxial growth of thin films | MTI Japan
    Different atomistic processes may occur on the surface during film growth: deposition, diffusion on ... Frank-van der Merwe (FV) growth mode. During FV or layer- ...
  59. [59]
    Influence of strain, surface diffusion and Ostwald ripening on the ...
    Apr 1, 2003 · This study explores the evolution of nanoscale islands and wire structures during deposition and surface ripening.
  60. [60]
    Ostwald ripening of three-dimensional clusters on a surface studied ...
    Jul 18, 2011 · We have studied the Ostwald ripening of three-dimensional islands on a homogeneous surface with an original off-lattice kinetic Monte Carlo ...
  61. [61]
    [PDF] Quantum dot self-assembly driven by a surfactant-induced ... - arXiv
    Here we show that such a morphological phase transition can be induced on- demand using surfactants. We explore Bi as a surfactant in the growth of InAs on GaAs ...
  62. [62]
    Surfactant effect on the surface diffusion length in epitaxial growth
    Sep 15, 1993 · Abstract. It is shown that Te and Pb, which segregate at the surface during the epitaxial growth of GaAs, respectively, decrease and increase ...Missing: Si dots
  63. [63]
    Van der Waals Epitaxy of Two-Dimensional MoS2-Graphene ...
    Aug 6, 2025 · In this work we demonstrate direct van der Waals epitaxy of MoS2-graphene heterostructures on a semiconducting silicon carbide (SiC) ...
  64. [64]
  65. [65]
    "Direct Observations of Surface Diffusion" by Gert Ehrlich
    The field ion microscope, with its ability to depict individual metal atoms on crystals, has made it possible to measure diffusivities of individual atoms.Missing: seminal | Show results with:seminal
  66. [66]
    A historic perspective of FIM and STM studies of surface diffusion
    The earliest direct observations and successful studies of surface diffusion of individual atoms used field ion microscopy (FIM).Missing: seminal papers
  67. [67]
    Resolving atomic diffusion in with spiral high-speed scanning ...
    Transition state theory proposes a migration pathway for the diffusion in the oxygen adlayer. With spiral scan geometries—a new approach to high-speed STM—the ...
  68. [68]
    (PDF) Scanning Probe Microscopy at Video-Rate - ResearchGate
    Aug 6, 2025 · PDF | Recent results have demonstrated the feasibility of video-rate scanning tunneling microscopy and video-rate atomic force microscopy.
  69. [69]
    Quasielastic helium atom scattering measurements of microscopic ...
    Jul 22, 1999 · The diffusion coefficient is 2 ... Quasielastic helium atom scattering measurements of microscopic diffusion of CO on the Ni(110) surface.
  70. [70]
    Quasi-elastic helium-atom scattering from surfaces - IOP Science
    May 31, 2002 · This review provides a topical summary of both the quasi-elastic helium-atom scattering technique and the available data in relation to the ...
  71. [71]
    Surface Diffusion Measured Using Laser Induced Thermal Desorption
    Experiments at low coverages between 260–330 K revealed an activation barrier for surface diffusion of 4 kcal/mole and a preexponential of 6.3 × 10−4 cm 2/sec.
  72. [72]
    Revisited reaction-diffusion model of thermal desorption ...
    Jul 28, 2015 · Desorption phase of thermal desorption spectroscopy (TDS) experiments performed on tungsten samples exposed to flux of hydrogen isotopes in ...
  73. [73]
    Density-functional theory calculations of hopping rates of surface ...
    Nov 15, 1998 · Using density-functional theory we compute the energy barriers and attempt frequencies for surface diffusion of Ag on Ag(111) with different ...
  74. [74]
    Theoretical foundations of dynamical Monte Carlo simulations
    In this paper, we present the theoretical basis for a dynamical Monte Carlo method in terms of the theory of Poisson processes.
  75. [75]
    A Practical Guide to Surface Kinetic Monte Carlo Simulations
    Apr 8, 2019 · This review article is intended as a practical guide for newcomers to the field of kinetic Monte Carlo (KMC) simulations, and specifically to lattice KMC ...Abstract · Introduction · KMC Simulations: From... · Steady-State and Transient...Missing: seminal | Show results with:seminal
  76. [76]
    Toward an ab Initio Description of Adsorbate Surface Dynamics
    Jul 27, 2024 · Detailed analysis of the MLP/MD-calculated diffusivities sheds light on the potential shortcomings of using DFT-based nudged elastic band to ...Missing: seminal | Show results with:seminal
  77. [77]
    Predicting hydrogen diffusion in nickel–manganese random alloys ...
    Aug 23, 2025 · This study develops a machine-learning interatomic potential (MLIP) for the Ni–Mn–H ternary system by efficiently sampling training data through ...
  78. [78]
    Prediction of vacancy defect diffusion paths in high entropy alloys ...
    Aug 20, 2025 · In this work, a workflow was developed using ML and KMC techniques to generate accelerated predicted vacancy defect diffusion paths in HEAs.