12.07.2015 Views

Introduction to Diffusion - Department of Materials Science and ...

Introduction to Diffusion - Department of Materials Science and ...

Introduction to Diffusion - Department of Materials Science and ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Materials</strong> <strong>Science</strong> & MetallurgyMaster <strong>of</strong> Philosophy, <strong>Materials</strong> Modelling,Course MP6, Kinetics <strong>and</strong> Microstructure Modelling, H. K. D. H. BhadeshiaLecture 3: <strong>Introduction</strong> <strong>to</strong> <strong>Diffusion</strong>Mass transport in a gas or liquid generally involves the flow <strong>of</strong> fluid(e.g. convection currents) although a<strong>to</strong>ms also diffuse. Solids on theother h<strong>and</strong>, can support shear stresses <strong>and</strong> hence do not flow except bydiffusion involving the jumping <strong>of</strong> a<strong>to</strong>ms on a fixed network <strong>of</strong> sites.Assume that such jumps can somehow be achieved in the solid state,with a frequency ν with each jump over a distance λ.For r<strong>and</strong>om jumps, the root mean square distance isx = λ √ ndiffusion distance ∝ √ twhere n is the number <strong>of</strong> jumps= λ √ νt where t is the time<strong>Diffusion</strong> in a Uniform Concentration GradientλCC+δCCxδCFig. 1: <strong>Diffusion</strong> Gradient


Concentration <strong>of</strong> solute, C, number m −3Each plane has Cλ a<strong>to</strong>ms m −2 (Fig. 1){ } ∂CδC = λ∂xA<strong>to</strong>mic flux, J, a<strong>to</strong>ms m −2 s −1J L→R = 1 6 νCλJ R→L = 1 ν(C + δC)λ6Therefore, the net flux along x is given byJ net = − 1 6 ν δC λ= − 1 { } ∂C6 ν λ2 ∂x{ } ∂C≡ −D∂xThis is Fick’s first law where the constant <strong>of</strong> proportionality is called thediffusion coefficient in m 2 s −1 . Fick’s first law applies <strong>to</strong> steady state fluxin a uniform concentration gradient. Thus, our equation for the me<strong>and</strong>iffusion distance can now be expressed in terms <strong>of</strong> the diffusivity asx = λ √ νt with D = 1 6 νλ2 giving x = √ 6Dt ≃ √ DtNon–Uniform Concentration GradientsSuppose that the concentration gradient is not uniform (Fig. 2).{ } ∂CFlux in = −D∂x1{ } ∂CFlux out = −D∂x2[{ } { ∂C ∂ 2 }]C= −D + δx∂x ∂x 21


unitareaδx1 2CxFig. 2: Non–uniform concentration gradientIn the time interval δt, the concentration changes δCδCδx = (Flux in – Flux out)δt∂C∂t = CD∂2 ∂x 2assuming that the diffusivity is independent <strong>of</strong> the concentration. Thisis Fick’s second law <strong>of</strong> diffusion.This is amenable <strong>to</strong> numerical solutions for the general case butthere are a couple <strong>of</strong> interesting analytical solutions for particular boundaryconditions. For a case where a fixed quantity <strong>of</strong> solute is plated on<strong>to</strong>a semi–infinite bar (Fig. 3),boundary conditions:C{x, t} =∫ ∞0C{x, t}dx = B<strong>and</strong> C{x, t = 0} = 0√ B { −x2expπDt 4DtNow imagine that we create the diffusion couple illustrated in Fig. 4,by stacking an infinite set <strong>of</strong> thin sources on the end <strong>of</strong> one <strong>of</strong> the bars.<strong>Diffusion</strong> can thus be treated by taking a whole set <strong>of</strong> the exponential}


Carea undereach curve isconstantx∞Fig. 3: Exponential solution. Note how the curvaturechanges with time.functions obtained above, each slightly displaced along the x axis, <strong>and</strong>summing (integrating) up their individual effects. The integral is in factthe error functionerf{x} = 2 √ π∫ xso the solution <strong>to</strong> the diffusion equation is0exp{−u 2 }duboundary conditions:C{x = 0, t} = C s<strong>and</strong> C{x, t = 0} = C 0{ } xC{x, t} = C s − (C s − C 0 )erf2 √ DtThis solution can be used in many circumstances where the surfaceconcentration is maintained constant, for example in the carburisationor decarburisation processes (the concentration pr<strong>of</strong>iles would be thesame as in Fig. 4, but with only one half <strong>of</strong> the couple illustrated). Thesolutions described here apply also <strong>to</strong> the diffusion <strong>of</strong> heat.Mechanism <strong>of</strong> <strong>Diffusion</strong>A<strong>to</strong>ms in the solid–state migrate by jumping in<strong>to</strong> vacancies (Fig. 5).The vacancies may be interstitial or in substitutional sites. There is,


C AC sC Bx∞∞Fig. 4: The error function solution. Notice that the“surface” concentration remains fixed.InterstitialSubstitutionalFig. 5: Mechanism <strong>of</strong> interstitial <strong>and</strong> substitutionaldiffusion.nevertheless, a barrier <strong>to</strong> the motion <strong>of</strong> the a<strong>to</strong>ms because the motionis associated with a transient dis<strong>to</strong>rtion <strong>of</strong> the lattice.Assuming that the a<strong>to</strong>m attempts jumps at a frequency ν 0 , thefrequency <strong>of</strong> successful jumps is given by}ν = ν 0 exp{− G∗kT{ }}S∗≡ ν 0 exp ×exp{− H∗k kT} {{ }independent <strong>of</strong> Twhere k <strong>and</strong> T are the Boltzmann constant <strong>and</strong> the absolute tempera-


ture respectively, <strong>and</strong> H ∗ <strong>and</strong> S ∗ the activation enthalpy <strong>and</strong> activationentropy respectively. SinceD ∝ ν we find D = D 0 exp{− H∗A plot <strong>of</strong> the logarithm <strong>of</strong> D versus 1/T should therefore give a straightline (Fig. 6), the slope <strong>of</strong> which is −H ∗ /k. Note that H ∗ is frequentlycalled the activation energy for diffusion <strong>and</strong> is <strong>of</strong>ten designated Q.kT}Fig. 6: Typical self–diffusion coefficients for puremetals <strong>and</strong> for carbon in ferritic iron. The uppermostdiffusivity for each metal is at its melting temperature.The activation enthalpy <strong>of</strong> diffusion can be separated in<strong>to</strong> two components,one the enthalpy <strong>of</strong> migration (due <strong>to</strong> dis<strong>to</strong>rtions) <strong>and</strong> the enthalpy<strong>of</strong> formation <strong>of</strong> a vacancy in an adjacent site. After all, for thea<strong>to</strong>m <strong>to</strong> jump it is necessary <strong>to</strong> have a vacant site; the equilibrium concentration<strong>of</strong> vacancies can be very small in solids. Since there are manymore interstitial vacancies, <strong>and</strong> since most interstitial sites are vacant,interstitial a<strong>to</strong>ms diffuse far more rapidly than substitutional solutes.Kirkendall Effect<strong>Diffusion</strong> is at first sight difficult <strong>to</strong> appreciate for the solid state. Anumber <strong>of</strong> mechanisms have been proposed his<strong>to</strong>rically. This includes avariety <strong>of</strong> ring mechanisms where a<strong>to</strong>ms simply swap positions, but controversyremained because the strain energies associated with such swapsmade the theories uncertain. One possibility is that diffusion occurs by


a<strong>to</strong>ms jumping in<strong>to</strong> vacancies. But the equilibrium concentration <strong>of</strong> vacanciesis typically 10 −6 , which is very small. The theory was thereforenot generally accepted until an elegant experiment by Smigelskas <strong>and</strong>Kirkendall (Fig. 7).Fig. 7: <strong>Diffusion</strong> couple with markersThe experiment applies <strong>to</strong> solids as well as cible liquids. Considera couple made from A <strong>and</strong> B. If the diffusion fluxes <strong>of</strong> the two elementsare different (|J A | > |J B |) then there will be a net flow <strong>of</strong> matter past theinert markers, causing the couple <strong>to</strong> shift bodily relative <strong>to</strong> the markers.This can only happen if diffusion is by a vacancy mechanism.An observer located at the markers will see not only a change inconcentration due <strong>to</strong> intrinsic diffusion, but also because <strong>of</strong> the Kirkendallflow <strong>of</strong> matter past the markers. The net effect is described bythe usual Fick’s laws, but with an interdiffusion coefficient D which isa weighted average <strong>of</strong> the two intrinsic diffusion coefficients:D = X B D A + X A D Bwhere X represents a mole fraction. It is the interdiffusion coefficientthat is measured in most experiments.Structure Sensitive <strong>Diffusion</strong>Crystals may contain nonequilibrium concentrations <strong>of</strong> defects suchas vacancies, dislocations <strong>and</strong> grain boundaries. These may provide easydiffusion paths through an otherwise perfect structure. Thus, the grainboundary diffusion coefficient D gb is expected <strong>to</strong> be much greater thanthe diffusion coefficient associated with the perfect structure, D P .


δrFig. 8: Idealised grainAssume a cylindrical grain. On a cross section, the area presentedby a boundary is 2πrδ where δ is the thickness <strong>of</strong> the boundary. Notethat the boundary is shared between two adjacent grains so the thicknessassociated with one grain is 1 2δ. The ratio <strong>of</strong> the areas <strong>of</strong> grain boundary<strong>to</strong> grain is thereforeratio <strong>of</strong> areas = 1 2 × 2πrδπr 2 = δ r = 2δdwhere d is the grain diameter (Fig. 8).For a unit area, the overall flux is the sum <strong>of</strong> that through thelattice <strong>and</strong> that through the boundary:so that2δJ ≃ J P + J gbd2δD measured = D P + D gbdNote that although diffusion through the boundary is much faster, thefraction <strong>of</strong> the sample which is the grain boundary phase is small. Consequently,grain boundary or defect diffusion in general is only <strong>of</strong> importanceat low temperatures where D P ≪ D gb (Fig. 9).Thermodynamics <strong>of</strong> diffusionFick’s first law is empirical in that it assumes a proportionalitybetween the diffusion flux <strong>and</strong> the concentration gradient. However,diffusion occurs so as <strong>to</strong> minimise the free energy. It should thereforebe driven by a gradient <strong>of</strong> free energy. But how do we represent thegradient in the free energy <strong>of</strong> a particular solute?


Dgbln{D}DPD measuredD gb2δd0.7T m1/TFig. 9: Structure sensitive diffusion. The dashed linewill in practice be curved.<strong>Diffusion</strong> in a Chemical Potential GradientFick’s laws are strictly empirical. <strong>Diffusion</strong> is driven by gradients<strong>of</strong> free energy rather than <strong>of</strong> chemical concentration:J A = −C A M A∂µ A∂xso thatD A = C A M A∂µ A∂C Awhere the proportionality constant M A is known as the mobility <strong>of</strong> A.In this equation, the diffusion coefficient is related <strong>to</strong> the mobility bycomparison with Fick’s first law. The chemical potential is here definedas the free energy per mole <strong>of</strong> A a<strong>to</strong>ms; it is necessary therefore<strong>to</strong> multiply by the concentration C A <strong>to</strong> obtain the actual free energygradient.The relationship is remarkable: if ∂µ A /∂C A > 0, then the diffusioncoefficient is positive <strong>and</strong> the chemical potential gradient is along thesame direction as the concentration gradient. However, if ∂µ A /∂C A < 0then the diffusion will occur against a concentration gradient!

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!