Surface diffusion

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.

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.

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 very 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 (eq. 2). The diffusion coefficient is given as D = Γa2/z (eq. 3), where z = 2 for 1D diffusion as would be the case for in-channel diffusion, z = 4 for 2D diffusion, and z = 6 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.

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.
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]

(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 jump use 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,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.
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

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.

(a) Dislocation (b) Glide
(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

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

This article is issued from Wikipedia - version of the 12/3/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.