Surface diffusion

Last updated
Figure 1. Model of a single adatom diffusing across a square surface lattice. Note the frequency of vibration of the adatom is greater than the jump rate to nearby sites. Also, the model displays examples of both nearest-neighbor jumps (straight) and next-nearest-neighbor jumps (diagonal). Not to scale on a spatial or temporal basis. Surface diffusion hopping.gif
Figure 1. Model of a single adatom diffusing across a square surface lattice. Note the frequency of vibration of the adatom is greater than the jump rate to nearby sites. Also, the model displays examples of both nearest-neighbor jumps (straight) and next-nearest-neighbor jumps (diagonal). Not to scale on a spatial or temporal basis.

Surface diffusion is a general process involving the motion of adatoms, molecules, and atomic clusters (adparticles) at solid material surfaces. [1] The process can generally be thought of in terms of particles jumping between adjacent adsorption sites on a surface, as in figure 1. Just as in bulk diffusion, this motion is typically a thermally promoted process with rates increasing with increasing temperature. Many systems display diffusion behavior that deviates from the conventional model of nearest-neighbor jumps. [2] Tunneling diffusion is a particularly interesting example of an unconventional mechanism wherein hydrogen has been shown to diffuse on clean metal surfaces via the quantum tunneling effect.

Contents

Various analytical tools may be used to elucidate surface diffusion mechanisms and rates, the most important of which are field ion microscopy and scanning tunneling microscopy. [3] While in principle the process can occur on a variety of materials, most experiments are performed on crystalline metal surfaces. Due to experimental constraints most studies of surface diffusion are limited to well below the melting point of the substrate, and much has yet to be discovered regarding how these processes take place at higher temperatures. [4]

Surface diffusion rates and mechanisms are affected by a variety of factors including the strength of the surface-adparticle bond, orientation of the surface lattice, attraction and repulsion between surface species and chemical potential gradients. It is an important concept in surface phase formation, epitaxial growth, heterogeneous catalysis, and other topics in surface science. [5] As such, the principles of surface diffusion are critical for the chemical production and semiconductor industries. Real-world applications relying heavily on these phenomena include catalytic converters, integrated circuits used in electronic devices, and silver halide salts used in photographic film. [5]

Kinetics

Figure 2. Diagram of the energy landscape for diffusion in one dimension. x is displacement; E(x) is energy; Q is the heat of adsorption or binding energy; a is the spacing between adjacent adsorption sites; Ediff is the barrier to diffusion. Diffusion energy landscape.png
Figure 2. Diagram of the energy landscape for diffusion in one dimension. x is displacement; E(x) is energy; Q is the heat of adsorption or binding energy; a is the spacing between adjacent adsorption sites; Ediff is the barrier to diffusion.

Surface diffusion kinetics can be thought of in terms of adatoms residing at adsorption sites on a 2D lattice, moving between adjacent (nearest-neighbor) adsorption sites by a jumping process. [1] [6] The jump rate is characterized by an attempt frequency and a thermodynamic factor that dictates the probability of an attempt resulting in a successful jump. The attempt frequency ν is typically taken to be simply the vibrational frequency of the adatom, while the thermodynamic factor is a Boltzmann factor dependent on temperature and Ediff, the potential energy barrier to diffusion. Equation 1 describes the relationship:

Where ν and Ediff are as described above, Γ is the jump or hopping rate, T is temperature, and kB is the Boltzmann constant. Ediff must be smaller than the energy of desorption for diffusion to occur, otherwise desorption processes would dominate. Importantly, equation 1 tells us how strongly the jump rate varies with temperature. The manner in which diffusion takes place is dependent on the relationship between Ediff and kBT as is given in the thermodynamic factor: when Ediff < kBT the thermodynamic factor approaches unity and Ediff ceases to be a meaningful barrier to diffusion. This case, known as mobile diffusion, is relatively uncommon and has only been observed in a few systems. [7] For the phenomena described throughout this article, it is assumed that Ediff >> kBT and therefore Γ << ν. In the case of Fickian diffusion it is possible to extract both the ν and Ediff from an Arrhenius plot of the logarithm of the diffusion coefficient, D, versus 1/T. For cases where more than one diffusion mechanism is present (see below), there may be more than one Ediff such that the relative distribution between the different processes would change with temperature.

Random walk statistics describe the mean squared displacement of diffusing species in terms of the number of jumps N and the distance per jump a. The number of successful jumps is simply Γ multiplied by the time allowed for diffusion, t. In the most basic model only nearest-neighbor jumps are considered and a corresponds to the spacing between nearest-neighbor adsorption sites. The root mean squared displacement goes as:

The diffusion coefficient is given as:

where for 1D diffusion as would be the case for in-channel diffusion, for 2D diffusion, and for 3D diffusion. [8]

Regimes

Figure 3. Model of six adatoms diffusing across a square surface lattice. The adatoms block each other from moving to adjacent sites. As per Fick's law, flux is in the opposite direction of the concentration gradient, a purely statistical effect. The model is not intended to show repulsion or attraction, and is not to scale on a spatial or temporal basis. Chemical surface diffusion slow.gif
Figure 3. Model of six adatoms diffusing across a square surface lattice. The adatoms block each other from moving to adjacent sites. As per Fick’s law, flux is in the opposite direction of the concentration gradient, a purely statistical effect. The model is not intended to show repulsion or attraction, and is not to scale on a spatial or temporal basis.

There are four different general schemes in which diffusion may take place. [9] Tracer diffusion and chemical diffusion differ in the level of adsorbate coverage at the surface, while intrinsic diffusion and mass transfer diffusion differ in the nature of the diffusion environment. Tracer diffusion and intrinsic diffusion both refer to systems where adparticles experience a relatively homogeneous environment, whereas in chemical and mass transfer diffusion adparticles are more strongly affected by their surroundings.

Anisotropy

Orientational anisotropy takes the form of a difference in both diffusion rates and mechanisms at the various surface orientations of a given material. For a given crystalline material each Miller Index plane may display unique diffusion phenomena. Close packed surfaces such as the fcc (111) tend to have higher diffusion rates than the correspondingly more "open" faces of the same material such as fcc (100). [10] [11]

Directional anisotropy refers to a difference in diffusion mechanism or rate in a particular direction on a given crystallographic plane. These differences may be a result of either anisotropy in the surface lattice (e.g. a rectangular lattice) or the presence of steps on a surface. One of the more dramatic examples of directional anisotropy is the diffusion of adatoms on channeled surfaces such as fcc (110), where diffusion along the channel is much faster than diffusion across the channel.

Mechanisms

Figure 4. Model of an atomic exchange mechanism occurring between an adatom (pink) and surface atom (silver) at a square surface lattice (blue). The surface atom becomes an adatom. Not to scale on a spatial or temporal basis. Atomic exchange diffusion 2.gif
Figure 4. Model of an atomic exchange mechanism occurring between an adatom (pink) and surface atom (silver) at a square surface lattice (blue). The surface atom becomes an adatom. Not to scale on a spatial or temporal basis.
Figure 5. Model of surface diffusion occurring via the vacancy mechanism. When surface coverage is nearly complete the vacancy mechanism dominates. Not to scale on a spatial or temporal basis. Vacancy diffusion.gif
Figure 5. Model of surface diffusion occurring via the vacancy mechanism. When surface coverage is nearly complete the vacancy mechanism dominates. Not to scale on a spatial or temporal basis.

Adatom diffusion

Diffusion of adatoms may occur by a variety of mechanisms. The manner in which they diffuse is important as it may dictate the kinetics of movement, temperature dependence, and overall mobility of surface species, among other parameters. The following is a summary of the most important of these processes: [12]

Figure 6. Surface diffusion jump mechanisms. Diagram of various jumps that may take place on a square lattice such as the fcc (100) plane. 1) Pink atom shown making jumps of various length to locations 2-5; 6) Green atom makes diagonal jump to location 7; 8) Grey atom makes rebound jump (atom ends up in same place it started). Non-nearest-neighbor jumps typically take place with greater frequency at higher temperatures. Not to scale. Surface diffusion jump mechanisms new.png(1) start for horizontal jumps(2) a single jump(3) a double jump(4) a triple jump(5) a quadruple jump(6) start for diagonal jump(7) a diagonal jump (down and to the right)(8) a rebound jumpuse button to enlarge or cursor to identify
Figure 6. Surface diffusion jump mechanisms. Diagram of various jumps that may take place on a square lattice such as the fcc (100) plane. 1) Pink atom shown making jumps of various length to locations 2-5; 6) Green atom makes diagonal jump to location 7; 8) Grey atom makes rebound jump (atom ends up in same place it started). Non-nearest-neighbor jumps typically take place with greater frequency at higher temperatures. Not to scale.
Figure 7. Graph showing relative probability distribution for adatom displacement,Dx, upon diffusion in one dimension. Blue: single jumps only; Pink: double jumps occur, with ratio of single:double jumps = 1. Statistical analysis of data may yield information regarding diffusion mechanism. Diffusion distribution jumps.png
Figure 7. Graph showing relative probability distribution for adatom displacement,Δx, upon diffusion in one dimension. Blue: single jumps only; Pink: double jumps occur, with ratio of single:double jumps = 1. Statistical analysis of data may yield information regarding diffusion mechanism.
Figure 8. Cross-channel diffusion involving an adatom (grey) on a channeled surface (such as fcc (110), blue plus highlighted green atom). 1) Initial configuration; 2) "Dumbbell" intermediate configuration. Final displacement may include 3, 4, 5, or even a return to the initial configuration. Not to scale. Cross channel diffusion.png
Figure 8. Cross-channel diffusion involving an adatom (grey) on a channeled surface (such as fcc (110), blue plus highlighted green atom). 1) Initial configuration; 2) "Dumbbell" intermediate configuration. Final displacement may include 3, 4, 5, or even a return to the initial configuration. Not to scale.
Figure 9. Long range atomic exchange mechanism for surface diffusion at a square lattice. Adatom (pink), resting at surface (1), inserts into lattice disturbing neighboring atoms (2), ultimately causing one of the original substrate atoms emerge as an adatom (green) (3). Not to scale. Long range atomic exchange.png
Figure 9. Long range atomic exchange mechanism for surface diffusion at a square lattice. Adatom (pink), resting at surface (1), inserts into lattice disturbing neighboring atoms (2), ultimately causing one of the original substrate atoms emerge as an adatom (green) (3). Not to scale.

Recent theoretical work as well as experimental work performed since the late 1970s has brought to light a remarkable variety of surface diffusion phenomena both with regard to kinetics as well as to mechanisms. Following is a summary of some of the more notable phenomena:

Figure 10. Individual mechanisms for surface diffusion of clusters. (1) Sequential displacement; (2) Edge diffusion; (3) Evaporation-condensation. In this model all three mechanisms lead to the same final cluster displacement. Not to scale. Cluster diffusion individual mechanisms.png
Figure 10. Individual mechanisms for surface diffusion of clusters. (1) Sequential displacement; (2) Edge diffusion; (3) Evaporation-condensation. In this model all three mechanisms lead to the same final cluster displacement. Not to scale.

Cluster diffusion

Cluster diffusion involves motion of atomic clusters ranging in size from dimers to islands containing hundreds of atoms. Motion of the cluster may occur via the displacement of individual atoms, sections of the cluster, or the entire cluster moving at once. [23] All of these processes involve a change in the cluster’s center of mass.

Cluster dislocation diffusion.gif Cluster diffusion glide.gif
(a) Dislocation(b) Glide
Cluster reptation diffusion.gif Cluster diffusion shear.gif
(c) Reptation(d) Shear
Figure 11. Concerted mechanisms for cluster diffusion.

Surface diffusion and heterogeneous catalysis

Surface diffusion is a critically important concept in heterogeneous catalysis, as reaction rates are often dictated by the ability of reactants to "find" each other at a catalyst surface. With increased temperature adsorbed molecules, molecular fragments, atoms, and clusters tend to have much greater mobility (see equation 1). However, with increased temperature the lifetime of adsorption decreases as the factor kBT becomes large enough for the adsorbed species to overcome the barrier to desorption, Q (see figure 2). Reaction thermodynamics aside because of the interplay between increased rates of diffusion and decreased lifetime of adsorption, increased temperature may in some cases decrease the overall rate of the reaction.

Experimental

Surface diffusion may be studied by a variety of techniques, including both direct and indirect observations. Two experimental techniques that have proved very useful in this area of study are field ion microscopy and scanning tunneling microscopy. [3] By visualizing the displacement of atoms or clusters over time, it is possible to extract useful information regarding the manner in which the relevant species diffuse-both mechanistic and rate-related information. In order to study surface diffusion on the atomistic scale it is unfortunately necessary to perform studies on rigorously clean surfaces and in ultra high vacuum (UHV) conditions or in the presence of small amounts of inert gas, as is the case when using He or Ne as imaging gas in field-ion microscopy experiments.

See also

Related Research Articles

Field ion microscope

The Field ion microscope (FIM) was invented by Müller in 1951. It is a type of microscope that can be used to image the arrangement of atoms at the surface of a sharp metal tip.

The Kirkendall effect is the motion of the interface between two metals that occurs as a consequence of the difference in diffusion rates of the metal atoms. The effect can be observed for example by placing insoluble markers at the interface between a pure metal and an alloy containing that metal, and heating to a temperature where atomic diffusion is possible; the boundary will move relative to the markers.

Chemisorption is a kind of adsorption which involves a chemical reaction between the surface and the adsorbate. New chemical bonds are generated at the adsorbant surface. Examples include macroscopic phenomena that can be very obvious, like corrosion, and subtler effects associated with heterogeneous catalysis, where the catalyst and reactants are in different phases. The strong interaction between the adsorbate and the substrate surface creates new types of electronic bonds.

Sintering Process of forming and bonding material by heat or pressure

Sintering or frittage is the process of compacting and forming a solid mass of material by heat or pressure without melting it to the point of liquefaction.

Adsorption Process resulting from the attraction of atoms, ions, or molecules from a gas, liquid, or solution sticking to a surface

Adsorption is the adhesion of atoms, ions or molecules from a gas, liquid or dissolved solid to a surface. This process creates a film of the adsorbate on the surface of the adsorbent. This process differs from absorption, in which a fluid is dissolved by or permeates a liquid or solid, respectively. Adsorption is a surface phenomenon, while absorption involves the whole volume of the material, although adsorption does often precede absorption. The term sorption encompasses both processes, while desorption is the reverse of it.

Reflection high-energy electron diffraction (RHEED) is a technique used to characterize the surface of crystalline materials. RHEED systems gather information only from the surface layer of the sample, which distinguishes RHEED from other materials characterization methods that also rely on diffraction of high-energy electrons. Transmission electron microscopy, another common electron diffraction method samples the bulk of the sample due to the geometry of the system. Low-energy electron diffraction (LEED) is also surface sensitive, but LEED achieves surface sensitivity through the use of low energy electrons.

Atomic diffusion

Atomic diffusion is a diffusion process whereby the random thermally-activated movement of atoms in a solid results in the net transport of atoms. For example, helium atoms inside a balloon can diffuse through the wall of the balloon and escape, resulting in the balloon slowly deflating. Other air molecules have lower mobilities and thus diffuse more slowly through the balloon wall. There is a concentration gradient in the balloon wall, because the balloon was initially filled with helium, and thus there is plenty of helium on the inside, but there is relatively little helium on the outside. The rate of transport is governed by the diffusivity and the concentration gradient.

Surface reconstruction refers to the process by which atoms at the surface of a crystal assume a different structure than that of the bulk. Surface reconstructions are important in that they help in the understanding of surface chemistry for various materials, especially in the case where another material is adsorbed onto the surface.

Ion beam mixing is the atomic intermixing and alloying that can occur at the interface separating two different materials during ion irradiation. It is applied as a process for adhering two multilayers, especially a substrate and deposited surface layer. The process involves bombarding layered samples with doses of ion radiation in order to promote mixing at the interface, and generally serves as a means of preparing electrical junctions, especially between non-equilibrium or metastable alloys and intermetallic compounds. Ion implantation equipment can be used to achieve ion beam mixing.

Jump diffusion is a stochastic process that involves jumps and diffusion. It has important applications in magnetic reconnection, coronal mass ejections, condensed matter physics, in Pattern theory and computational vision and in option pricing.

In materials science, segregation is the enrichment of atoms, ions, or molecules at a microscopic region in a materials system. While the terms segregation and adsorption are essentially synonymous, in practice, segregation is often used to describe the partitioning of molecular constituents to defects from solid solutions, whereas adsorption is generally used to describe such partitioning from liquids and gases to surfaces. The molecular-level segregation discussed in this article is distinct from other types of materials phenomena that are often called segregation, such as particle segregation in granular materials, and phase separation or precipitation, wherein molecules are segregated in to macroscopic regions of different compositions. Segregation has many practical consequences, ranging from the formation of soap bubbles, to microstructural engineering in materials science, to the stabilization of colloidal suspensions.

In chemistry, the Terrace Ledge Kink model (TLK), which is also referred to as the Terrace Step Kink model (TSK), describes the thermodynamics of crystal surface formation and transformation, as well as the energetics of surface defect formation. It is based upon the idea that the energy of an atom’s position on a crystal surface is determined by its bonding to neighboring atoms and that transitions simply involve the counting of broken and formed bonds. The TLK model can be applied to surface science topics such as crystal growth, surface diffusion, roughening, and vaporization.

Stranski–Krastanov growth is one of the three primary modes by which thin films grow epitaxially at a crystal surface or interface. Also known as 'layer-plus-island growth', the SK mode follows a two step process: initially, complete films of adsorbates, up to several monolayers thick, grow in a layer-by-layer fashion on a crystal substrate. Beyond a critical layer thickness, which depends on strain and the chemical potential of the deposited film, growth continues through the nucleation and coalescence of adsorbate 'islands'. This growth mechanism was first noted by Ivan Stranski and Lyubomir Krastanov in 1938. It wasn’t until 1958 however, in a seminal work by Ernst Bauer published in Zeitschrift für Kristallographie, that the SK, Volmer–Weber, and Frank–van der Merwe mechanisms were systematically classified as the primary thin-film growth processes. Since then, SK growth has been the subject of intense investigation, not only to better understand the complex thermodynamics and kinetics at the core of thin-film formation, but also as a route to fabricating novel nanostructures for application in the microelectronics industry.

Surface phonon

In solid state physics, a surface phonon is the quantum of a lattice vibration mode associated with a solid surface. Similar to the ordinary lattice vibrations in a bulk solid, the nature of surface vibrations depends on details of periodicity and symmetry of a crystal structure. Surface vibrations are however distinct from the bulk vibrations, as they arise from the abrupt termination of a crystal structure at the surface of a solid. Knowledge of surface phonon dispersion gives important information related to the amount of surface relaxation, the existence and distance between an adsorbate and the surface, and information regarding presence, quantity, and type of defects existing on the surface.

Dislocation creep is a deformation mechanism in crystalline materials. Dislocation creep involves the movement of dislocations through the crystal lattice of the material, in contrast to diffusion creep, in which diffusion is the dominant creep mechanism. It causes plastic deformation of the individual crystals, and thus the material itself.

Thermocompression bonding describes a wafer bonding technique and is also referred to as diffusion bonding, pressure joining, thermocompression welding or solid-state welding. Two metals, e.g. gold (Au)-gold (Au), are brought into atomic contact applying force and heat simultaneously. The diffusion requires atomic contact between the surfaces due to the atomic motion. The atoms migrate from one crystal lattice to the other one based on crystal lattice vibration. This atomic interaction sticks the interface together. The diffusion process is described by the following three processes:

Field-emission microscopy (FEM) is an analytical technique used in materials science to investigate molecular surface structures and their electronic properties. Invented by Erwin Wilhelm Müller in 1936, the FEM was one of the first surface-analysis instruments that approached near-atomic resolution.

Adsorption is the accumulation and adhesion of molecules, atoms, ions, or larger particles to a surface, but without surface penetration occurring. The adsorption of larger biomolecules such as proteins is of high physiological relevance, and as such they adsorb with different mechanisms than their molecular or atomic analogs. Some of the major driving forces behind protein adsorption include: surface energy, intermolecular forces, hydrophobicity, and ionic or electrostatic interaction. By knowing how these factors affect protein adsorption, they can then be manipulated by machining, alloying, and other engineering techniques to select for the most optimal performance in biomedical or physiological applications.

The strength of metal oxide adhesion effectively determines the wetting of the metal-oxide interface. The strength of this adhesion is important, for instance, in production of light bulbs and fiber-matrix composites that depend on the optimization of wetting to create metal-ceramic interfaces. The strength of adhesion also determines the extent of dispersion on catalytically active metal. Metal oxide adhesion is important for applications such as complementary metal oxide semiconductor devices. These devices make possible the high packing densities of modern integrated circuits.

In mathematics and physics, surface growth refers to models used in the dynamical study of the growth of a surface, usually by means of a stochastic differential equation of a field.

References

  1. 1 2 Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 325
  2. Antczak, Ehrlich 2007, p.39
  3. 1 2 Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 349
  4. Antczak, Ehrlich 2007, p. 50, 59
  5. 1 2 Shustorovich 1991, p. 109
  6. Shustorovich 1991, p. 109-111
  7. Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 327
  8. Structure and Dynamics of Surfaces II (Topics in Current Physics), W. Schommers, P. Von Blanckenhagen, ISBN   0387173382. Chapter 3.2, p. 75
  9. Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 330-333
  10. Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 333
  11. Shustorovich 1991, p. 114-115
  12. Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 336-340
  13. Shustorovich 1991, p. 111
  14. Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 338
  15. Antczak, Ehrlich 2007, p. 48
  16. Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 338-340
  17. Shustorovich 1991, p. 115
  18. Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 340-341
  19. Antczak, Ehrlich 2007, p. 51
  20. Antczak, Ehrlich 2007, p. 58
  21. Antczak, Ehrlich 2007, p. 40-45
  22. Antczak, Ehrlich 2007, p. 48-50
  23. Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 341
  24. Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 343-344
  25. Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 343-345
  26. Oura, Lifshits, Saranin, Zotov, and Katayama 2003, p. 341-343

Cited works