Skip to main content

Diffusion

  • Chapter
  • First Online:
Fundamentals of Materials Science
  • 2100 Accesses

Abstract

Material transport by migration of atoms or molecular entities is a fundamental process of nature. Its understanding is, for example, of crucial importance for control of the microstructure of materials. Continuum and atomistic approaches to diffusion are discussed. Possible diffusion mechanisms are presented. The diffusion coefficient is a tool to characterize the diffusion process, via the activation energy. A corresponding, broad discussion of experimental results and their interpretation is given for various materials. It is shown to be necessary to distinguish the overall material flow from the intrinsic diffusion processes, as given by the individual atomic jumps. This culminates in extensive discussion of the Kirkendall effect and the Darken treatment. Via the “driving force” for diffusion, a link can be made from the thermodynamics to the kinetics. Especially along the latter route the so-called intrinsic diffusion coefficients, to be distinguished from the chemical or interdiffusion diffusion coefficient, can be determined, as illustrated experimentally. Fundamental and practical problems are touched upon. Many material systems are subjected to a state of stress, e.g., thin films. A treatment of diffusion in a state of (nonhydrostatic) stress is still not possible in a rigorous way on the basis of boundary conditions compatible with reality. Yet, some interesting considerations and results are presented. Diffusion experiments can be very revealing about the defects in the structural (atomic) arrangement of a material. This is demonstrated here for in particular the role of grain boundaries, including moving grain boundaries.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 59.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 79.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 119.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    The progress of science is tributary immensely to the process of “thinking in analogies”. One of the most striking examples is provided by the (mathematical) similarities in the theories for the conduction of heat in solids (see the book by Carslaw and Jaeger 1959) and for the diffusion of mass in solids (see the book by Crank 1975). Fick was led by such thinking in analogies to his proposal of what we now call Fick’s (first and second) laws of diffusion. He remarks that “It was quite natural to suppose that this law for diffusion … must be identical with that, according to which the diffusion of heat in a conducting body takes place”. And he explicitly links his proposals to the earlier “theory of heat” by Fourier and that due to Ohm for the “diffusion of electricity in a conductor” (Fick 1855).

  2. 2.

    Fick tested successfully the validity of Eq. (8.2) for the case of diffusion in a system composed of dissolved rock salt in water in a stationary (steady) state (cf. Sect. 7.3): Fick kept a bottom layer of water in a vessel saturated with dissolved rock salt and kept the top layer as pure water, leading to a stationary state of constant transport of dissolved salt from bottom to top of the vessel, thus establishing a time-independent dissolved salt concentration profile as a function of height in the vessel; the (net; cf. Sect. 8.2) amount of dissolved salt arriving at a certain height, coming from the bottom, equals the (net; cf. Sect. 8.2) amount of dissolved salt leaving this height, moving upward, and this holds at every height in the vessel. The resulting concentration-height profile must be a straight line (integrate Eq. (8.2) for the case J = constant, i.e. does not depend on x (here x = height in the vessel)), as found by Fick (Fick measured the specific mass as function of the height) (cf. Fig. 7.2).

  3. 3.

    In a strict sense use of the notion “law” should be confined to rules, describing the action of forces (as the law of gravity) and the course of processes (as Fick’s first law), which have been found (initially) to be valid on the basis of empirical (i.e. relying on experience and observation alone) work. However, in science one is not puristic: for example, one also speaks of Bragg’s law (Sects. 4.5 and 6.10), which relation was, also initially, derived theoretically in a straightforward manner and in no way was proposed on the basis of empiricism. Adopting a wide interpretation one could say: a “law” expresses “the regularity of nature”, but this approach introduces a broad and ill defined transition region between just a formula/relation and a law.

  4. 4.

    A Matano-like analysis is also possible for a semi-infinite system (Mittemeijer and de Keijser, Scripta Metallurgica, 11(1977), 113–115).

  5. 5.

    The diffusion coefficient of hydrogen in metals is very large—of the same order of magnitude as found for atomic diffusion in liquids—and, moreover, is characterized by a very low value of the activation energy, except at relatively high temperatures; diffusion of hydrogen in metals is not dominated by the thermally activated atomic jumps over barriers, but rather is governed by quantum mechanical tunneling (Fukai 2005).

  6. 6.

    It is remarkable to observe that the octahedral interstitial sites in the close packed crystal structures (f.c.c. and h.c.p.) are larger than in the less close packed b.c.c. crystal structure (for the same size of the (metal) atom taken as rigid sphere; this approximately holds for iron (see Sect. 9.5.2.1)). Yet, the total relative interstitial site volume is larger in the b.c.c. crystal structure, because per (metal) atom there are three times as much interstices: three octahedral and six tetrahedral interstices per metal atom in b.c.c. and one octahedral and two tetrahedral interstices per metal atom in f.c.c. and h.c.p.

    As long as the reasoning is based on the relative amount of “free space” (see the main text), adopting the rigid (hard, solid) sphere model (cf. Sect. 4.2; for the f.c.c. crystal structure the atoms are in touch along the <110> directions (face diagonals of the unit cell); for the b.c.c. crystal structure the atoms are in touch along the <111> directions (body diagonals of the unit cell), a difference in the absolute value for the size of the atom in the one and the other crystal structure is irrelevant. Adopting the atomic volume (i.e. the volume per atom, for example calculated from the unit cell volume divided by the number of atoms in the unit cell) of the element considered as a measure for atom size, it follows, perhaps surprisingly/counter-intuitively, for many metals which can crystallize in f.c.c./h.c.p. and b.c.c. modifications, that the not close packed b.c.c. modification exhibits the smaller atomic volume (Rudman 1965). An exception is iron, where the reverse holds (also see Table 4.4 in Sect. 4.2.5), thereby lending support for the simple argumentation given above, also if one does not depart from the metal atom as a rigid sphere.

  7. 7.

    The size of the octahedral interstices in b.c.c. iron is smaller than in f.c.c. iron (see footnote 6 in this chapter and see Sect. 9.5.2.1). This may explain the pronouncedly larger solubility of carbon in f.c.c. iron and the very small solubility of carbon in b.c.c. iron. Thus, the conclusion of the deliberations in the above main text and in this footnote is that (1) the diffusion coefficient of carbon in f.c.c. iron is smaller than in b.c.c. iron, because it is more difficult in f.c.c. to establish the displacement of the iron atoms pertaining to the activated state of the carbon atom jumping from one to the next octahedral interstitial site (less “free space” in f.c.c.), while (2) the solubility of carbon in f.c.c. is larger than in b.c.c. iron, because the size of the octahedral interstitial site in f.c.c. is larger.

  8. 8.

    Computer simulations have suggested that the vacancy volume in (bcc) metals depends about linearly on an elastic constant (the Poisson constant, ν; see Sect. 12.2): from about 1 × the atomic volume for ν = 0.25 down to about 0.6 × the atomic volume for ν = 0.4 (Kurita and Numakura 2004).

  9. 9.

    At this stage of the analysis, instead of defining a lattice-fixed frame of reference, it would have been possible just to speak of a frame of reference fixed to a plane oriented perpendicular to the diffusion direction, as in fact done by Darken (1948), thereby avoiding seeming restriction to diffusion in crystalline materials. The lattice-fixed frame of reference is already introduced here to facilitate the interpretation of the material flow with velocity v below Eq. (8.35) as a consequence of the vacancy mechanism of diffusion.

  10. 10.

    Note that this approach to the relative movement of one coordinate system with respect to another coordinate system also occurs in Einstein’s first, special theory of relativity (1905).

  11. 11.

    This conclusion (that the migration of the markers occurs to the side of the diffusion couple where the largest intrinsic diffusion coefficients prevail) is generally accepted (e.g. see Liu et al. 2008), and has been validated also in cases where a concentration dependence of the molar volume has to be accounted for. However, it has been claimed that the direction of the marker movement would be dictated by the sign of the slope of Vegard’s relation (Eq. (4.7)) for the component pair considered, in the way that the markers would move in the direction of increasing lattice parameter (marker-movement direction predictions on this basis can (and very often are) but need not be the same as those obtained by the first reasoning above). This was ascribed to the larger tendency of the larger atoms to be surrounded by a vacancy atmosphere, thereby relaxing the local and global misfit, coherency strain and at the same time enhancing the exchange of the larger atoms with vacancies, i.e. increasing their intrinsic diffusion coefficient … (Kirkaldy and Savva 1997).

  12. 12.

    The history of the discovery of the Kirkendall effect is described by H. Nakajima (JOM, 49 (1997), 15–19), on the basis of personal recollections of Kirkendall.

  13. 13.

    The replacement of the concentration gradient by the chemical potential gradient, as “driving force” for diffusion, indicates the possibility that diffusion can occur in a direction opposite to that of the concentration gradient, i.e. from lower to higher concentration as long as the chemical potential of the component concerned becomes reduced. This can regularly be observed in (but not only) multicomponent systems. Anyway, a (multicomponent) system strives for decreasing its total (Gibbs) energy and this will govern the diffusion directions of the various components and, consequently, considering only one component and its transport by diffusion in a multicomponent system will generally lead to erroneous conclusions.

  14. 14.

    At this place in this book some thermodynamic knowledge is required that has not been presented before (in Chap. 7). This is no obstruction to understand the main message of the treatment in this section. Yet, as clarification for (re)reading at a later stage, the corresponding short derivation is given in this footnote as follows. The chemical potential of the considered component of concentration c is given by

    μ = μ0 + RTlna = μ0 + RTlnγc = μ0 + RTlnγ + RTlnc, with a = γc as the so-called activity of the component considered of concentration c and with γ as the so-called activity coefficient; μ0 is a constant (reference chemical potential). Differentiating the left-hand and right-hand sides of this equation with respect to c, and multiplying thereafter both sides with c, leads to the expression for ∂μ/∂lnc given above (Note that μ can also be written as μ = \( \mu^{\prime}_{0} \) + RTlnγX, with the atomic fraction given as X = c/c0 and where-RTlnc0 has been taken up into \( \mu^{\prime}_{0} \). This is used in footnote 16, where \( \mu^{\prime}_{0} \) is yet written without prime).

  15. 15.

    In the case of self-diffusion there is neither a net flow of vacancies (nor of matter). \(D^{s}_{A} = D_{A} \left( {X_{A} = 1} \right)\) can therefore directly be determined from the diffusion of A* into A by straightforward fitting of a concentration profile for A* as determined on the basis of Eq. (8.6b) with D taken as \(D_{A}^{s}\).

  16. 16.

    At this place, again, some thermodynamic knowledge is required that has not been presented before. Knowledge of and understanding of the background of this footnote (which cannot be expected for a beginning materials scientist) are no prerequisite for understanding the message of this section leading to the final result presented as Eqs. (8.56) and (8.57).

    Using:

    (i) μA = μ0,A + RTlnγAXA (cf. footnote 14) and thus ∂μA = RT[∂lnXA + ∂lnγA] and, similarly,

    ∂μB = RT[∂lnXB + ∂lnγB],

    (ii) XA + XB = 1, and thus ∂XA = −∂XB,

    (iii) XA∂μA + XB∂μB = 0, i.e. a so-called Gibbs–Duhem equation,

    the result presented above as Eq. (8.53) is obtained.

  17. 17.

    It should be recognized that a planar state of stress is characterized by also three principal stress components, albeit one of these equals nil. Hence, the equivalent hydrostatic stress component (given by the sum of the three principal stress components divided by three) for the planar state of stress considered in Fig. 8.16 and pertaining to Eq. (8.60), with two equal principal stress components parallel to the surface, equals 2/3 . σ.

  18. 18.

    A higher surface concentration implies the occurrence of a larger concentration gradient and thus (cf. Eq. (8.2)) a larger diffusional flux, which results in a larger penetration depth. Also recognizing that the case considered (cf. Fig.8.16) implies a case of semi-infinite diffusion, Eq. (8.17) (for constant D) in Sect. 8.3 already indicates a larger penetration depth for larger surface concentration, c0.

  19. 19.

    This phenomenon parallels the observation for the grain-boundary mobility, where rate control appears to be governed by groups of atomic jumps (in this case not necessarily only along the grain boundary) as well (cf. Sect. 10.7 and Bos et al. 2007).

  20. 20.

    Upon heating a polycrystalline material, close to the melting point of the bulk material grain boundaries can be “wetted”, i.e. be covered by a liquid film. The condition for this process to occur is that the energy of the solid/liquid interface becomes smaller than half the grain-boundary energy. (cf. the discussion on grain-boundary wetting in Sect. 9.4.5). The occurrence of a grain-boundary diffusivity, at temperatures close to the melting temperature of the bulk material, which approaches the diffusivity in the liquid state, is clear indication of grain-boundary wetting in the system considered (Divinski and Herzig 2008). This is a recent example of the power of diffusion analysis to reveal the microstructure of a material (see Sect. 8.9.1 for a famous, “old” example). The occurrence of “superplasticity” at high strain rate (up to 102/s) in nanostructured materials upon plastic deformation at such elevated temperature has been ascribed to such grain-boundary wetting by a liquid film (see footnote 23 in Sect. 12.16.1).

  21. 21.

    The first observations of (edge) dislocations, made by using a transmission electron microscope, were published in 1956.

  22. 22.

    Upon reading this early note (F.N. Rhines and A.M. Montgomery, Nature, 141 (1938), 413) it is clear that these authors had no idea of the background of their observation of the “disturbed” grain boundary of a Cu bicrystal upon inward diffusion of Zn.

  23. 23.

    After having written the text of this Intermezzo in the first edition of this book, upon preparing the second edition I became aware of a recent, very readable booklet where this disgrace of modern science, and many even more serious ones, have been exposed and illustrated by illuminating examples: see G. Pacchioni, The Overproduction of Truth, Oxford University Press, 2018.

References

General

  • H.S. Carslaw, J.C. Jaeger, Conduction of Heat in Solids (Clarendon Press, Oxford, 1959).

    Google Scholar 

  • J. Crank, The Mathematics of Diffusion, 2nd edn. Oxford University Press (1975).

    Google Scholar 

  • P.F. Green, Kinetics, Transport, and Structure in Hard and Soft Materials (Taylor & Francis, London, 2005).

    Book  Google Scholar 

  • I. Kaur, Y. Mishin, W. Gust, Fundamentals of Grain and Interphase Boundary Diffusion, 3rd revised and enlarged edition. (Wiley & Sons, Chichester, 1995).

    Google Scholar 

  • J.S. Kirkaldy, D.J. Young, Diffusion in the Condensed State (The Institute of Metals, London, 1987).

    Google Scholar 

  • H. Mehrer, Diffusion in Solids (Springer, Heidelberg, 2007).

    Book  Google Scholar 

  • J. Philibert, Atom Movements; Diffusion and Mass Transport in Solids (Les Editions de Physique, Les Ulis, 1991).

    Google Scholar 

  • P. Shewmon, Diffusion in Solids, 2nd edn. (The Minerals, Metals & Materials Society, Warrendale, 1989).

    Google Scholar 

Specific

  • M. Akhlaghi, M. Jung, S.R. Meka, M. Fonović, A. Leineweber, E.J. Mittemeijer, Dependence of the nitriding rate of ferritic and austenitic substrates on the crystallographic orientation of surface grains; gaseous nitriding of Fe-Cr and Ni-Ti Alloys, Philos. Mag. 95, 4143–4160 (2015)

    CAS  Google Scholar 

  • H.K.D.H. Bhadeshia, Carbon–Carbon Interactions in Iron. J. Mater. Sci. 39, 3949–3955 (2004)

    Article  CAS  Google Scholar 

  • C. Bos, F. Sommer, E.J. Mittemeijer, Atomistic study on the activation enthalpies for interface mobility and boundary diffusion in an interface-controlled phase transformation. Phil. Mag. 87, 2245–2262 (2007)

    Article  CAS  Google Scholar 

  • F.J.A. den Broeder, Onderzoek naar de Diffusie in het Systeem Chroom-Wolfraam. Ph.D. Dissertation, Delft University of Technology, pp. 90–95 (in Dutch, 1970)

    Google Scholar 

  • F.J.A. den Broeder, Interface reaction and a special form of grain boundary diffusion in the Cr-W system. Acta Metall. 20, 319–332 (1972)

    Article  Google Scholar 

  • L. Dabrowski, A. Andreev, M. Georgiev, Carbon diffusion in α-iron: Evidence for quantum mechanical tunneling. Metallur. Mater. Trans. A 37A, 2079–2084 (2006)

    Article  CAS  Google Scholar 

  • L.S. Darken, Diffusion, mobility and their interrelation through free energy in binary metallic systems. Trans. AIME 175, 184–194 (1948)

    Google Scholar 

  • S. Divinski, C. Herzig, Radiotracer Investigation of diffusion, segregation and wetting phenomena in grain boundaries. J. Mater. Sci. 43, 3900–3907 (2008)

    Article  CAS  Google Scholar 

  • A. Fick, Über Diffusion. Poggendorff’s Annalen, 94, 59–86 (in German, 1855); published in English, in abstracted form, as A. Fick, On liquid diffusion. Philos. Mag. 10, 30–39 (1855)

    Google Scholar 

  • D.H.R. Fors, G. Wahnström, Nature of Boron solution and diffusion in α-Iron. Phys. Rev. B 77, 132102 (2008)

    Article  Google Scholar 

  • Y. Fukai, The Metal-Hydrogen System; Basic Bulk Properties (Springer, Berlin, 2005)

    Google Scholar 

  • G. Grube, A. Jedele, Die Diffusion der Metalle im Festen Zustand. Zeitschrift Für Elektrochemie 38, 799–807 (in German, 1932)

    CAS  Google Scholar 

  • M. Hillert, G.R. Purdy, Chemically induced grain boundary migration. Acta Metall. 26, 333–340 (1978)

    Article  CAS  Google Scholar 

  • A.H. King, Diffusion induced grain boundary migration. Int. Mater. Rev. 32, 173–189 (1987)

    Article  CAS  Google Scholar 

  • J.S. Kirkaldy, G. Savva, Correlation between coherency strain effects and the Kirkendall effect in binary infinite diffusion couples. Acta Mater. 45, 3115–3121 (1997)

    Article  CAS  Google Scholar 

  • L. Klinger, E. Rabkin, Theory of the Kirkendall effect during grain boundary interdiffusion. Acta Mater. 59, 1389–1399 (2011)

    Article  CAS  Google Scholar 

  • P. Klugkist, A.N. Aleshin, W. Lojkowski, L.S. Shvindlerman, W. Gust, E.J. Mittemeijer, Diffusion of Zn along tilt grain boundaries in Al: pressure and orientation dependence. Acta Mater. 49, 2941–2949 (2001)

    Article  CAS  Google Scholar 

  • N. Kurita, H. Numakura, Formation volume of atomic vacancies in body-centred cubic metals. Z. Metallkde. 95, 876–879 (2004)

    CAS  Google Scholar 

  • S.J.B. Kurz, S.R. Meka, N. Schell, W. Ecker, J. Keckes, E.J. Mittemeijer, Residual stress and microstructure depth gradients in nitrided iron-based alloys revealed by dynamical cross-sectional transmission X-ray microdiffraction. Acta Mater. 87, 100–110 (2015)

    Article  CAS  Google Scholar 

  • F.C. Larché, J.W. Cahn, The interactions of composition and stress in crystalline solids. Acta Metall. 33, 331–357 (1985)

    Article  Google Scholar 

  • Y. Liu, L. Zhang, D. Yu, Y. Ge, Study of Diffusion and Marker Movement in fcc Ag-Au Alloys. J. Phase Equilibria and Diffusion 29, 405–413 (2008)

    Article  CAS  Google Scholar 

  • C. Matano, On the relation between the diffusion coefficients and concentration of solid metals (the nickel-copper system). Jpn. J. Phys. 8, 109–113 (1933)

    CAS  Google Scholar 

  • E.J. Mittemeijer, A.M. Beers, Recrystallization and interdiffusion in thin bimetallic films. Thin Solid Films 65, 125–135 (1980)

    Article  CAS  Google Scholar 

  • M.A. Noah, D. Flötotto, Z. Wang, M. Reiner, C. Hugenschmidt, E.J. Mittemeijer, Interdiffusion in epitaxial, single-crystalline Au/Ag films studied by auger electron spectroscopy sputter-depth profiling and positron annihilation. Acta Mater. 107, 133–143 (2016a)

    Article  CAS  Google Scholar 

  • M.A. Noah, D. Flötotto, Z. Wang, E.J. Mittemeijer, Interdiffusion and stress development in single-crystalline Pd/Ag bilayers. J. Appl. Phys. 119, 145308 (2016b)

    Article  Google Scholar 

  • B. Okkerse, Zelfdiffusie in Lood, Ph.D. Dissertation, Delft University of Technology, pp. 55–69 (in Dutch, 1954a)

    Google Scholar 

  • B. Okkerse, Self-diffusion in lead. Acta Metall. 2, 551–553 (1954b)

    Article  Google Scholar 

  • J. Philibert, Beyond Fick’s equations, an overview. Int. J. Mater. Res. 100, 744–749 (2009)

    Article  CAS  Google Scholar 

  • P.S. Rudman, The atomic volumes of the metallic elements. Trans. Metallur. Soc. AIME 233, 864–871 (1965)

    CAS  Google Scholar 

  • J. Silva, R.B. McLellan, Diffusion of carbon and nitrogen in bcc Iron. Mater. Sci. Eng. 26, 83–87 (1976)

    Article  Google Scholar 

  • R. Simmons, R.W. Balluffi, Measurements of equilibrium vacancy concentrations in aluminum. Phys. Rev. 117, 52–61 (1960)

    Article  CAS  Google Scholar 

  • A.C. Smigelskas, E.O. Kirkendall, Zinc diffusion in alpha brass. Trans. AIME 171, 130–142 (1947)

    Google Scholar 

  • W.T.M. Straver, H.C.F. Rozendaal, E.J. Mittemeijer, Consequences of the heterogeneous nitriding of α-iron: dislocation production and oriented precipitation. Metallur. Trans. A 15, 627–637 (1984)

    Google Scholar 

  • A. Suzuki, Y. Mishin, Atomic mechanisms of grain boundary diffusion: low versus high temperatures. J. Mater. Sci. 40, 3155–3161 (2005)

    Article  CAS  Google Scholar 

  • D. Turnbull, R.E. Hoffman, The effect of relative crystal and boundary orientations on grain boundary diffusion rates. Acta Metall. 2, 419–426 (1954)

    Article  CAS  Google Scholar 

  • J.Y. Wang, E.J. Mittemeijer, A new method for the determination of the diffusion-induced concentration profile and the interdiffusion coefficient for thin films by auger electron spectroscopical sputter depth profiling. J. Mater. Res. 19, 3389–3397 (2004)

    Article  CAS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Eric J. Mittemeijer .

Appendix: Concentration-Depth Profiles in Thin Layer Systems

Appendix: Concentration-Depth Profiles in Thin Layer Systems

With a view to the great practical importance of thin film systems, as in the microelectronic industry, a number of solutions to Fick’s second law for a variety of thin layer systems is summarized at the end of this chapter in this appendix. Various cases with different initial and boundary conditions can be considered for different diffusion stages. It will be assumed that the diffusion coefficient can be taken as concentration independent.

Case 1 (Fig. 8.23a): A bilayer (AB) or multilayer (ABABAB …) for which the thickness of each sublayer is much larger than the diffusion length, \(\sqrt {Dt}\), corresponding to a very early diffusion stage for the (multi)layer.

Fig. 8.23
figure 23

Concentration-depth profiles in thin film systems. a Bilayer or multilayer with the thickness of each sublayer much larger than the diffusion length. b Trilayer or multilayer with the thickness of each sublayer A much smaller than that of sublayers B and the diffusion length much smaller than the thickness of sublayers B. c Trilayer with thickness of each sublayer of the order of the diffusion length. d Multilayer with the thickness of each sublayer of the order of the diffusion length

The solution for the bilayer AB (possibly as part of the multilayer ABABAB…) with the A/B interface at z = zi and initial conditions C = C0 for z ≤ zi and C = 0 for z > zi is given by (Crank 1975):

$$C(z,t) = \frac{{C_{0} }}{{2\sqrt {\pi Dt} }}\int\limits_{{z - z_{i} }}^{\infty } {\exp \left( { - \frac{{\xi^{2} }}{4Dt}} \right){\text{d}}\xi } = \frac{1}{2}C_{0} {\text{erfc}}\frac{{z - z_{i} }}{{2\sqrt {Dt} }}$$
(8.64)

where erfc (=1 − erf) denotes the error-function (erf) complement (cf. Eq. 8.16). Note that, evidently, C = C0/2 at z = zi for all t > 0.

Case 2 (Fig. 8.23b): A trilayer (BAB) or multilayer (BABABA…) for which the thickness of sublayers A is much smaller than that of the sublayers B, i.e. dA ≪ dB, and \(\sqrt {Dt}\) ≪ dB.

For the trilayer BAB (possibly as part of the multilayer BABABA…) with z = zi at the centre plane of sublayer A and initial conditions C = C0 for zi − dA/2 ≤ z ≤ zi + dA/2 and C = 0 for z < zi − dA/2 and z > zi + dA/2, the concentration profile is obtained as (cf. Eq. 8.64)

$$\begin{aligned} C(z,t) & = \frac{{C_{0} }}{{2\sqrt {\pi Dt} }}\int\limits_{{z - (z_{i} + d_{A} /2)}}^{{z - (z_{i} - d_{A} /2)}} {\exp \left( { - \frac{{\xi^{2} }}{4Dt}} \right){\text{d}}\xi } \\ & = \frac{1}{2}C_{0} \left[ {{\text{erf}}\frac{{d_{A} /2 + (z - z_{i} )}}{{2\sqrt {Dt} }} + {\text{erf}}\frac{{d_{A} /2 - (z - z_{i} )}}{{2\sqrt {Dt} }}} \right] \\ \end{aligned}$$
(8.65)

It is clear that the system can be cut in half by a plane at z = zi without affecting the distribution, which is symmetrical about z = zi. Therefore, Eq. (8.65) also holds for a bilayer system composed of a sublayer A of thickness of dA/2, with the surface or a diffusion barrier at z = zi, on top of the semi-infinite sublayer B (substrate, see Fig. 8.23b). Redefining dA/2 as h and taking zi = 0, it follows for the diffusion-induced concentration profile in this case

$$C(z,t) = \frac{{C_{0} }}{2}\left[ {{\text{erf}}\frac{h + z}{{2\sqrt {Dt} }} + {\text{erf}}\frac{h - z}{{2\sqrt {Dt} }}} \right]$$
(8.66)

Case 3 (Fig. 8.23c): A trilayer (BAB) for which the thickness of each sublayer is not much larger than the diffusion length \(\sqrt {Dt}\), which represents a relatively advanced diffusion stage for the trilayer.

The outer surfaces of both B sublayers are barriers for mass transport. Hence,

$$\frac{\partial C}{{\partial z}} = 0\quad {\text{at}}\;\;z_{1} = z_{i} - \left( d_{B} + \frac{d_{A}}{2} \right)\;\;{\text{and}}\;\;z_{2} = z_{i} + \left(d_{B} + \frac{d_{A}}{2} \right)$$

The resulting concentration profile can be constructed by the superposition (reflection) principle as follows (cf. Crank 1975). The concentration profile given by Eq. (8.65) is reflected at the plane at boundary z2, and this reflected profile is obtained by replacing zi in Eq. (8.65) by zi + (2 dB + dA). This firstly reflected curve is reflected at the plane at zi, and the secondly reflected profile is obtained by replacing zi in Eq. (8.65) by zi − (2dB + dA). Then, the secondly reflected profile is reflected again at the boundary z2 (zi → zi + 2(2dB + dA)) and at zi(zi → zi − 2(2dB + dA)) and so on. Therefore, the complete solution as the result of such successive reflections is given by

$$\begin{aligned} C(z,t) & = \frac{1}{2}C_{0} \sum\limits_{n = - \infty }^{\infty } {\left[ {{\text{erf}}\frac{{d_{A} /2 - n(2d_{B} + d_{A} ) + z - z_{i} }}{{2\sqrt {Dt} }}} \right.} \\ & \left. {\quad + {\text{erf}}\frac{{d_{A} /2 + n(2d_{B} + d_{A} ) - z + z_{i} }}{{2\sqrt {Dt} }}} \right] \\ \end{aligned}$$
(8.67)

Case 4 (Fig. 8.23d): A multilayer (ABABAB…) for which the thickness of each sublayer is not much larger than the diffusion length \(\sqrt {Dt}\), which represents a relatively advanced diffusion stage for the multilayer.

Considering the initial conditions: C = C0 for zi − dA/2 ≤ z ≤ zi + dA/2 and C = 0 for zi − (dA/2 + dB) < z < zi − dA/2 and zi + dA/2 < z < zi + (dA/2 + dB) with zi denoting the centre plane of sublayer A, and recognizing that for t ≥ 0 ∂C/∂z = 0 at the centre plane of the sublayers B, the total concentration profile for the trilayer BAB in the multilayer is obtained by the superposition (reflection) principle (see also Case 3) as follows

$$\begin{aligned} &C(z,t) \\&= \frac{1}{2}C_{0} \sum\limits_{n = - \infty }^{\infty } {\left[ {{\text{erf}}\frac{{d_{A} /2 - n(d_{B} + d_{A} ) + z - z_{i} }}{{2\sqrt {Dt} }} + {\text{erf}}\frac{{d_{A} /2 + n(d_{B} + d_{A} ) - z + z_{i} }}{{2\sqrt {Dt} }}} \right]} \end{aligned}$$
(8.68)

The diffusion-induced concentration profiles as given by the Eqs. (8.64)–(8.68) have been derived considering volume (bulk) diffusion. In polycrystalline thin films the role of grain-boundary diffusion is often dominant, recognizing the high grain-boundary density and the usually applied relatively low diffusion annealing temperatures (as compared to the melting point of the components). Often columnar microstructures occur in thin films; i.e. the grain boundaries are oriented more or less perpendicular to the film surface and sublayer interfaces. Then, if the volume diffusion length, (Dbt)1/2, is much smaller than the grain-boundary diffusion length, (Dgbt)1/2, with Db and Dgb as the volume and grain-boundary diffusion coefficients, respectively, it can be shown (see Wang and Mittemeijer 2004) that the laterally averaged concentration profile for Cases 1–4, as induced by grain-boundary diffusion, are also given by Eqs. (8.64)–(8.68), provided C is identified with \(\overline{C}\), the laterally averaged concentration, D is identified with the grain-boundary diffusion coefficient Dgb, and C0 is identified with C0δη (δ is grain-boundary width and η is grain-boundary length per unit area for the plane parallel to the surface.

Rights and permissions

Reprints and permissions

Copyright information

© 2021 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Mittemeijer, E.J. (2021). Diffusion. In: Fundamentals of Materials Science. Springer, Cham. https://doi.org/10.1007/978-3-030-60056-3_8

Download citation

Publish with us

Policies and ethics