Category Archives: Comprehensive nuclear materials

Recoil-energy effect

The recoil energy enters the PBM through the cas­cade parameters er and eg (see eqn [138]). Direct experimental evaluation of the recoil energy effect on void swelling was made by Singh et a/.,133 who compared the microstructure of annealed copper irradiated with 2.5 MeV electrons, 3 MeV protons, and fission neutrons at ^520 K. For all irradiations, the damage rate was ^10—8dpas—1. The average recoil energies in those irradiations were estimated133
to be about 0.05, 1, and 60keV for electron, proton, and neutron irradiations, respectively, thus, produc­ing the primary damage in the form of FPs (elec­trons), small cascades (protons), and well-developed cascades (neutrons). The cascade efficiency, 1 — er, hence, the real damage rate, was highest for electron irradiation (no cascades, the efficiency is equal to unity) and minimal for neutron irradiation (~0.1, see Section 1.13.3). If dislocation bias is the mecha­nism responsible for swelling, the swelling rate is proportional to the damage rate and therefore must be highest after electron and lowest after neutron irradiation. However, just the opposite was found; the swelling level after neutron irradiation was ^50 times higher than after electron irradiation, with the value for proton irradiation falling in between (see Figure 5). These results represent direct experimen­tal confirmation that damage accumulation under cascade damage conditions is governed by mechan­isms that are entirely different from those under FP production.

The results obtained in this study can be under­stood as follows. Under electron irradiation, only the first term on the right-hand side of eqn [138] oper­ates, as eg = 0. The swelling rate is low in this case because of the low dislocation density, as discussed in Section 1.13.6.2.1. Under cascade damage condi­tions, the damage rate is smaller because of the low cascade efficiency. In this case eg = 0 and the second term on the right-hand side of eqn [138] plays the main role, which is evident from the theoretical treatment of the experiment carried out in the fol­lowing section.24

Calculation techniques

A microcrack in the configuration shown on the right side of Figure 13 is loaded monotonically. To mimic the triaxial stress field surrounding the microcrack, we modify the expressions for the stress fields given in Wang and Lee59 by subtracting the applied normal stress (oAj) and adding the uniaxial yield stress (ay) for cases when sAy > sy. The value of the slip plane angle is chosen such that predominantly brittle crack configurations are simulated. The applied load is incremented at a rate dK/dt = 0.01MPax/ms~1; a variable time step is used to make sure dislocations move as an array, and numerical instabilities are avoided. Dislocations are emitted from source posi­tions (here chosen as 4b, b = 2.54 A) along the slip planes. Since source positions are equidistant from the respective crack tips, simultaneous emission occurs. The emitted dislocations move away from the crack tip along the slip planes with a velocity calculated using eqn [24]. The stress fields of these dislocations then shield the crack tip from external load. Once moved to their equilibrium position, the amount of shielding from each dislocation is calcu­lated using expressions from Wang and Lee,59 and the total shielding at the crack tip is obtained by summa­tion. We ignore antishielding dislocations in our simulations, since any dislocations nucleated in the present configuration around the crack tip will be absorbed by the crack. Thus, the number of disloca­tion absorbed is small, and any blunting effects due to them are neglected here. Because in our case disloca­tions are emitted from crack-tip sources, dislocations emitted from either end contribute to shield both crack tips.62 The applied load is increased monotoni — cally and the crack-tip stress intensity is calculated at each time step. When the crack-tip stress inten­sity reaches a preassigned critical value, the crack is assumed to propagate and the corresponding

K ^2r

Microcrack

Macrocrack

image1000

^ Notional crack tip

X

p

Figure 14 Schematic illustration of the crack-tip blunting. is the radius of the blunted crack tip, 2p the microcrack size, and Xp the distance of the microcrack from macrocrack tip. The distance of the microcrack from the notional sharp crack tip increases from Xp to Xp + p.

applied load is the microscopic fracture stress (sf). The following two predominantly brittle microcrack systems are considered here: (1) cleavage plane (001), crack front [110], slip system 1[T11](112), hence a’ = 35° 16′, and (2) cleavage plane (110), crack front [110],^Ш^Ш), hence a’ = 54°44′.

A schematic illustration of the macrocrack is shown on the left-hand side of Figure 14. The macrocrack is assumed to be semi-infinite, with dis­location sources close to the crack tip. Dislocations are emitted simultaneously along the two slip planes, symmetrically oriented to the crack plane. The plas­tic zone formed by emitted dislocations produces a field equivalent to an elastic-plastic crack with small scale yielding.61 The slip plane angle in this case is chosen to be 70.5° (the direction of maximum shear stress of the elastic crack-tip field), compared to the present slip plane angles, which match best with continuum estimates.61 The initial source position (x0) along the slip plane is chosen to be 4b, where b is the magnitude of Burgers vector. (For each positive dislocation emitted, a negative one is assumed to move into the crack.) The dislocation sources on either side of the crack plane are at equivalent posi­tions x0 and operate simultaneously.

During the simulation, the applied stress intensity is increased in small increments and the positions of dislocations are determined. It is found that the dis­locations reach near-equilibrium positions. As the load is increased, more and more dislocations are emitted and the crack gets blunted. The radius of the blunted crack p is taken to be equal to Nb sin a, where N is the number of dislocations and a is the slip plane angle. The blunting of the macrocrack is illustrated in Figure 14. As blunting increases,
(1) the crack-tip fields are modified to be that of a blunted crack, (2) source position is chosen to be equal to the crack-tip radius, that is, x0 = p, and (3) the notional crack tip from which the image stress of dislocations is calculated is moved back to the center of curvature of the blunted crack. At each time step, the stress ahead of the crack at a distance Xp along the crack plane (the microcrack is assumed to be at this position) is calculated using the expres­sion from Trinkaus et at23 The fracture criterion is the tensile stress (o?) at Xp reaching oF calculated in the previous stage for the microcracks for the same temperature and the corresponding yield stress. Thus, when the microcrack tip stress intensity is K = Kjc, cleavage fracture of the matrix is assumed to occur. The applied load at the macrocrack then gives the fracture toughness (KF) at that given tem — perature/yield stress. Throughout the simulation, a microcrack of size 1 pm is used; this value is chosen as a typical (average) value in the range of sizes of microcracks found in experiments with which we compare our results.43 The detailed simulations reported here are obtained using Xp equal to 10 pm. However, the KF calculated using 20 pm is also shown in Figure 14. These values were chosen as typical of the order of grain sizes in this type of steel. It should be noted that changing these parameters will give different values for fracture toughness; however, the behavior of their variation with temperature will remain the same, as will be discussed next.

Dilute alloys

The point defect jump frequencies to be considered are those that are far from the solute, those leaving the solute, those arriving at a nn site of the solute, and those jumping from nn to nn sites of the solute. Diffusivities are approached by a series in which the successive terms include longer and longer looping paths of the point defect from the solute. Using the pair-association method, the whole series has been obtained for the vacancy mechanism in body — centered cubic (bcc) and face-centered cubic (fcc) binary alloys with nn interactions (cf. Allnatt and Lidiard6 for a review). However, there is still no accurate model for the effect of a solute atom on the self-diffusion coefficient.84 The L-coefficients as well as the tracer diffusivities depend on three jump frequency ratios in the fcc structure and two ratios in the bcc structure. For the irradiation studies, the same pair-association method has been applied to estimate the L-coefficients of the dumbbell diffusion mechanism in fcc85,86 and bcc83,87 alloys. Note, that the pair-association calculation in bcc alloys87 has recently been improved by using the self-consistent mean-field (SCMF) theory.83 The calculation includes the effect of the binding energy between a dumbbell of solvent atoms and a solute.

In the case of the vacancy diffusion mechanism, before the development of first-principles calculation methods, reliable jump frequency ratios could be estimated from a few experimental diffusion coeffi­cients (three for fcc and two for bcc). Data were calculated from the solvent and solute tracer diffu — sivities, the linear enhancement of self-diffusion with solute concentration, or the electro-migration en­hancement factors of tracer atoms in an electric field.8 Some examples of jump frequency ratios fitted to experimental data are displayed in tables.89 Currently, the most widely used approach is using first-principles calculations to calculate not only the vacancy, but also the interstitial jump frequencies.

1.18.3.4.2.1 Concentrated alloys

The first diffusion models devoted to concentrated alloys were based on a very basic description of the diffusion mechanism. The alloy is assimilated into a random lattice gas model where atoms do not interact and where vacancies jump at a frequency that depends only on the species they exchange with (two frequencies in a binary alloy). Using complex arguments, Manning8 could express the correlation factors as a function of the few jump frequencies. The approach was extended to the interstitial diffusion mechanism.9 At the time, no procedure was sug­gested to calculate these mean jump frequencies from an atomic jump frequency model. Such diffu­sion models, which consider a limited number of jump frequencies, already make spectacular correla­tion effects appear possible, such as a percolation limit when the host atoms are immobile.8,91,92 But they do not account for the effect of short-range order on the L-coefficients although one knows that RIS behavior is often explained by means of a com­petition between binding energies of point defects with atomic species, especially in dilute alloys. Some attempts were made to incorporate short — range order in a Manning-type formulation of the phenomenological coefficients, but coherency with thermodynamics was not guaranteed.93,94

The current diffusion models, including short — range order, are based on either the path probability method (PPM)95-97 or the SCMF theory.84,92,98-100 Both mean-field methods start from an atomic diffu­sion model and a microscopic master equation. While the PPM considers transition variables, which are deduced from a minimization of a pseudo free — energy functional associated to the kinetic path, the SCMF theory introduces an effective Hamiltonian to represent the nonequilibrium correction to the
distribution function probability and calculates the effective interactions by imposing a self-consistent constraint on the kinetic equations of the distribution function moments. Diffusion models built from the PPM were developed in bcc solid solution and ordered alloys, using nn pair transition variables, which is equivalent to considering nn effective inter­actions, and a statistical pair approximation for the equilibrium averages within the SCMF theory.95,96 The SCMF theory extended the approach to fcc alloys8,91,92 and provided a model of the composition effect on solute drag by vacancies.98 The interstitial diffusion mechanism was also tackled.10,83,101 For the first time, an interstitial diffusion model including short-range order was proposed in bcc concentrated alloys.10 It was shown that the usual value of 1 eV for the binding energy of an interstitial with a neighbor­ing solute atom leads to very strong effects on the average interstitial jump frequency and, therefore, on the L-coefficients.

1.18.3 Continuous Models of RIS
partial diffusion coefficients, as new constants asso­ciated with a temperature and a nominal alloy com­position. The first authors to express the partial diffusion coefficients in terms of the L-coefficients were Howard and Lidiard103 for the vacancy in dilute alloys, Barbu86 for the interstitial in dilute alloys, and Wolfer11 in concentrated alloys. In the following, we use the formulation of Wolfer because the approxi­mations made to calculate the L-coefficients and driving forces are clearly stated. In a binary alloy (AB), fluxes are separated into two contributions: the first one induced by the point defect concentra­tion gradients, and the second one appearing after the formation of chemical concentration gradients near the point defect sink11:

JA = — (dAV CV + dAI CI)FV CA + CA(dAV V CV — dAIr Ci)

Jv = —(Ca^av + Cb4bv)VCv + Cv Ф (dA V V Ca + dBV V Cb)

Ji = — ( Ca dAi + Cb dBi )VCi — Ci®(4AiVCa + dBI V Cb ) [24]

with partial diffusion coefficients defined in terms of the L-coefficients and the equilibrium point defect concentration

image1052

image1053

RIS is a phenomenon that couples the fluxes of defects created by irradiation and the alloy compo­nents. In RIS models, point defect diffusion mechan­isms alone are considered, although it is accepted that displacement cascades produce mobile point defect clusters, which may contribute to the RIS. In the first section, we present the expressions used to simulate the RIS, the main results, and limits. Some examples of application to real alloys are discussed in the second part. The last section suggests some possible improvements in the RIS models.

Choices to make

Whatever the system considered and the code used, one needs to provide more inputs than just the atomic positions and types. Most codes suggest some values for these inputs. However, their tuning may still be necessary as default values may very well be suited for some supposedly standard situations and irrele­vant for others. Blind use of ab initio codes may thus lead to disappointing errors. Indeed, not all these choices are trivial, so mistakes can be hard to notice for the beginner. Choices are usually made out of experience, after considering some test cases needing small calculation time.

One can distinguish between choices that should be made only once at the beginning of a study and calculation parameters that can be tuned calculation by calculation. The main unchangeable choices are the exchange and correlation functional and the pseudopotentials or PAW atomic data for the various atomic types in the calculation.

First, one has to choose the flavor of the exchange and correlation functional that will be used to describe the electronic interactions. Most of the time one chooses either an LDA or a GGA func­tional. Trends are known about the behavior of these functionals: LDA calculations tend to overestimate the bonding and underestimate the bond length in bulk materials, the opposite for GGA. However, things can become tricky when one deals with defects as energy differences (between defect-containing and defect-free cells) are involved. For insulating materi­als or materials with correlated electrons, the choice of the exchange and correlation functional is even more difficult (see Section 1.08.5).

The second and more definitive choice is the one of the pseudopotential. We do not mean here the choice of the pseudoization scheme but the choice of the pseudopotential itself. Indeed, calculated ener­gies vary greatly with the chosen pseudopotential, so energy differences that are thermodynamically or kinetically relevant are meaningless if the various calculations are performed with different pseudopo­tentials. The determination ofthe shape ofthe atomic basis set in the case of localized bases is also of importance, and it is close in spirit to the choice of the pseudopotential except that much less basis sets than pseudopotentials are available.

More technical inputs include

• the k-point sampling. The larger the number of k points to sample the Brillouin zone, the more accurate the results but the heavier the calcula­tions will be. This is especially true for metallic systems that need fine sampling of the Brillouin zone, but convergence with respect to the number of k points can be accelerated by the introduction of a smearing of the occupations of electronic levels close to the Fermi energy. The shape and width of this smearing function is then an addi­tional parameter.19

• the number ofplane waves (obviously for plane wave codes but also for some other codes that also use FFT). Once again the larger the number of plane waves, the more accurate and heavier the calculation.

• the convergence criteria. The two major conver­gence criteria are the one for the self-consistent loop of the calculation of the ground-state electronic wave functions and the one to signal the convergence of a relaxation calculation (with some threshold depending on the forces acting on the atoms).

Peierls stress at zero temperature

Dislocations in the dominant slip system in bcc metals have (111)/2 Burgers vectors and {110} slip planes. Here, we consider an edge dislocation with Burgers vector b = 1/2 [111] (along x-axis), slip plane normal [110] (along j-axis), and line direction [T 12] (along z-axis). To prepare the atomic configuration, we first create a perfect crystal with dimensions 30[111], 40[H0], 2[ 112] along the x-, y-, z-axes. We then remove one-fourth of the atomic layers normal to the y-axis to create two free surfaces, as shown in Figure 8(a).

We introduce an edge dislocation dipole into the simulation cell by displacing the positions of all atoms according to the linear elasticity solution of the displacement field of a dislocation dipole. To satisfy PBC, the displacement field is the sum of the contributions from not only the dislocation dipole

image588Подпись: DislocationПодпись: bПодпись: Cut planeПодпись:Подпись: (b)<№

image594

(a)

Figure 8 (a) Schematics showing the edge dislocation dipole in the simulation cell. b is the dislocation Burgers vector of the upper dislocation. Atoms in shaded regions are removed. (b) Core structure of edge dislocation (at the center) and surface atoms in FS Ta after relaxation visualized by Atomeye.41 Atoms are colored according to their central-symmetry deviation parameter. Adapted from Li, J. In Handbook of Materials Modeling;Yip, S., Ed.; Springer: Dordrecht, 2005; pp 1051-1068; Mistake-free version at http://alum. mit. edu/www/liju99/Papers/05/Li05-2.31.pdf; Kelchner, C. L.; Plimpton, S. J.; Hamilton, J. C. Phys. Rev. B 1998, 58, 11085.

inside the cell, but also its periodic images. Care must be taken to remove the spurious term caused by the conditional convergence of the sum.26,40-42 Because the Burgers vector b is perpendicular to the cut — plane connecting the two dislocations in the dipole, atoms separated from the cut-plane by <|b|/2 in the x-direction need to be removed. The resulting struc­ture contains 21414 atoms. The structure is subse­quently relaxed to a local energy minimum with zero average stress. Because one of the two dislocations in the dipole is intentionally introduced into the vacuum region, only one dislocation remains after the relaxa­tion, as shown in Figure 8(b).

The dislocation core is identified by central symmetry analysis,13 which characterizes the degree of inversion-symmetry breaking. In Figure 8(b), only atoms with a central symmetry deviation (CSD) parameter larger than 1.5 A2 are plotted. Atoms with CSD parameter between 0.6 and 6 A2 appear at the center of the cell and are identified with the disloca­tion core. Atoms with a CSD parameter between 10 and 20 A2 appear at the top and bottom of the cell and are identified with the free surfaces.

The edge dislocation thus created will move along the x-direction when the shear stress axy exceeds a critical value. To compute the Peierls stress, we apply shear stress axy by adding external forces on surface atoms. The total force on the top surface
atoms points in the x-direction and has magnitude of Fx = axyLxLz. The total force on the bottom sur­face atoms has the same magnitude but points in the opposite direction. These forces are equally distributed on the top (and bottom) surface atoms. Because we have removed some atoms when creating the edge dislocation, the bottom surface layer has fewer atoms than the top surface layer. As a result, the external force on each atom on the top surface is slightly lower than that on each atom on the bottom surface.

We apply shear stress axy in increments of 1 MPa and relax the structure using the conjugate gradient algorithm at each stress. The dislocation (as identified by the core atoms) does not move for axy < 27 MPa but moves in the x-direction during the relaxation at axy = 28 MPa. Therefore, this simulation predicts that the Peierls stress of edge dislocation in Ta (FS potential) is 28 ± 1 MPa. The Peierls stress computed in this way can depend on the simulation cell size. Therefore, we will need to repeat this calculation for several cell sizes to obtain a more reliable prediction of the Peierls stress. There are other boundary conditions that can be applied to simulate disloca­tions and compute the Peierls stress, such as PBCs in both x — and j-directions,42 and the Green’s function boundary condition.44 Different boundary conditions have different size dependence on the numerical error of the Peierls stress.

The simulation cell in this study contains two free surfaces and one dislocation. This is designed to minimize the effect of image forces from the bound­ary conditions on the computed Peierls stress. If the surfaces were not created, the simulation cell would have to contain at least two dislocations so that the total Burgers vector content was zero. On application of the stress, the two dislocations in the dipole would move in opposite directions, and the total energy would vary as a function of their relative position. This would create forces on the dislocations, in addi­tion to the Peach-Koehler force from the applied stress, and would lead to either overestimation or underestimation of the Peierls stress. On the contrary, the simulation cell described above has only one dislocation, and as it moves to an equivalent lattice site in the x-direction, the energy does not change due to the translational symmetry of the lattice. This means that, by symmetry, the image force on the dislocation from the boundary conditions is iden­tically zero, which leads to more accurate Peierls stress predictions. However, when the simulation cell is too small, the free surfaces in the y-direction
and the periodic images in the x-direction can still introduce (second-order) effects on the critical stress for dislocation motion, even though they do not produce any net force on the dislocation.

Stable Defect Formation

Initial work of Bacon and coworkers indicated that the number of stable displacements remaining at the end of a cascade simulation, ND, exhibited a power — law dependence on cascade energy.84 For example, their analysis of iron cascade simulations between 0.5 and 10 keV at 100 K showed that the total number of surviving point defects could be expressed as

Nd = 5-67E MD79 [3]

where EMD is given in kiloelectronvolts. This rela­tionship is not followed below about 0.5 keV because true cascade-like behavior does not occur at these
low energies. Subsequent work by Stoller64-67 indi­cated that Nd also begins to deviate from this energy dependence above 20 keV when extensive subcascade formation occurs. This is illustrated in Figure 9(a) where the values of ND obtained in cascade simula­tions at 100 K is plotted as a function of cascade energy. At each energy, the data point is an average of between 7 and 26 cascades, and the error bars indicate the standard error of the mean. It appears that three well-defined regions with different energy dependencies exist. A power-law fit to the points in each energy region is also shown in Figure 9(a). The best-fit exponent in the absence of true cas­cade conditions below 0.5 keV is 0.485. From 0.5 to

image665Подпись:Подпись: Figure 9 Cascade energy dependence of stable point defect formation in iron MD cascade simulations at 100 K: (a) total number of interstitials or vacancies and (b) ratio of MD defects to NRT displacements. Data points indicate mean values at each energy, and error bars are standard error of the mean.image668(b)

ratio at low energies was first measured in 1978 and

is well known.57,85 The error bars again reflect the standard error and the dashed line through the points is only a guide to the eye. The MD/NRT ratio is greater than 1.0 at the lowest values of EMD, indicat­ing that the NRT formulation underestimates defect production in this energy range. This is consistent with the low-energy (near threshold) simulations preferentially producing displacements in the ‘easy’ directions.[8] [9] [10] [11] [12] [13] [14] The actual displacement threshold var­ies with crystallographic direction and is as low as ~19 eV in the [100] direction.20,84 Thus, using the recommended average value of 40 eV £d in eqn [2] predicts fewer defects at low energies. The aver­age value is more appropriate for the higher energy events where true cascade-like behavior occurs. In the cascade-dominated regime, the defect density within the cascade increases with energy. Although many more defects are produced, their close proxim­ity leads to a higher probability of in-cascade recom­bination and a lower defect survival fraction.

The surviving defect fraction shows a slight increase as the cascade energy increases above 20, and the indicated standard errors make it arguable that the increase is statistically significant. If signifi­cant, the increase appears to be associated with sub­cascade formation, which becomes prominent above 10-20 keV. In the channeling regions between the high-angle collisions that produce the subcascades shown in Figures 7 and 8, the moving atom loses energy in many low-angle scattering events that pro­duce low-energy recoils. These are essentially like low-energy cascades, which have higher-than-average defect survival fractions (Figure 9). These events could contribute to the incremental increase in defect sur­vival at the highest energies. The average defect sur­vival fraction of ^0.3 NRT shown for cascade energies greater than about 10 keV is consistent with values of Frenkel pair formation obtained from resistivity change measurements following low-temperature

neutron irradiation and ion irradiation.26,27,57,85

The effect of irradiation temperature is shown in Figure 10, which compares the defect survival fractions obtained from simulations at 100, 600, and 900 K. Although it is difficult to discern a consistent effect of temperature between the 600 and 900 K data points, the defect survival fraction at 100 K is always somewhat greater than at either of the two higher temperatures. A similar result for iron was reported in Bacon and coworkers.84 In addition to an interest in radiation temperature itself, the effect of temperature is relevant to the
simulations presented here because no thermostat was applied to the simulation cell to control tem­perature. As mentioned above, the energy intro­duced by the PKA will lead to some heating if the simulation cell temperature is not controlled by a thermostat. For example, in a 1 keV cascade simula­tion with 54 000 atoms, the average temperature rise will be about 140 K when all the kinetic energy of the PKA is distributed in the system. This change in temperature should be more significant at 100 K than at higher temperatures. The fact that defect survival at 600 and 900 K is lower than at 100 K suggests that the 100 K results may be

image669

Figure 10 Temperature dependence of stable defect formation in MD simulations: ratio of MD defects to NRT displacements.

somewhat biased toward lower survival values by the PKA-induced heating. This is in agreement with the effect of temperature reported by Gao and coworkers77 in their study of 2 and 5 keV cascades with a hybrid MD model that extracted heat from the simu­lation cell. On the other hand, the difference between the 100 and 600 K results is not large, so the effect of ^200 K of cascade-induced heating may be modest.

Подпись: Figure 11 Effect of cascade heating on defect formation in 10 keV cascades at 100 K.

A simple assessment of this cascade-induced heat­ing was carried out using 10 keV cascades at 100 K. Two independent sets ofsimulations were carried out, seven simulations in a cell of 128 k atoms and eight simulations in a cell of 250 k atoms. A 10 keV cascade will raise the average temperature by 604 and 309 K, respectively, for these two cell sizes. The results of these simulations are summarized in Figure 11, where the parameters plotted are the surviving defect fraction (per NRT), the fraction of interstitials in clusters (per NRT), and the fraction of interstitials in clusters (per surviving MD defect). In each case, the range ofvalues for the two populations are shown, along with their respective mean values with the standard error indicated. The mean and standard error for the combined data sets is also shown. Although the heating differed by a factor of two, it is clear that the defect survival fraction is essentially identical for both populations. There is a slight trend in the interstitial clustering results, which indicates that a higher temperature (due to a smaller number of atoms) promotes interstitial clustering. This is consis­tent with the results that will be discussed below.

Nucleation of point defect clusters

Nucleation of small clusters in supersaturated solu­tions has been of significant interest to several genera­tions ofscientists. The kinetic model for cluster growth and the rate of formation of stable droplets in vapor and second phase precipitation in alloys during aging was studied extensively. The similarity to the con­densation process in supersaturated solutions allows the results obtained to be used in RDT to describe the formation of defect clusters under irradiation.

The initial motivation for work in this area was to derive the nucleation rate of liquid drops. Farkas63 was first to develop a quantitative theory for the so-called homogeneous cluster nucleation. Then, a great number of publications were devoted to the kinetic nucleation theory, of which the works by Becker and Doring,64 Zeldovich,65 and Frenkel66 are most important. Although these publications by no means improved the result of Farkas, their treat­ment is mathematically more elegant and provided a proper background for subsequent works in for­mulating ME and revealing properties of the clus­ter evolution. A quite comprehensive description of the nucleation phenomenon was published by Goodrich.67,68 Detailed discussions of cluster nucle — ation can also be found in several comprehensive reviews.69,70 Generalizations of homogeneous cluster nucleation for the case of irradiation were developed by Katz and Wiedersich71 and Russell.72 Here we only give a short introduction to the theory.

For small cluster sizes at high enough tempera­ture, when the thermal stability of clusters is rela­tively low, the diffusion of clusters in the size-space governs the cluster evolution, which is nucleation of stable clusters. In cases where only FPs are produced by irradiation, the first term on the right-hand side of eqn [18] is equal to zero and cluster nucleation, for example, voids, proceeds via interaction between mobile vacancies to form divacancies, then between vacancies and divacancies to form trivacancies, and so on. By summing eqn [18] from x = 2 to 1, one finds

^ = J (x)k=i = Jnucl [24]

1

where Nc = ^2 f (x) is the total number of clusters.

x=2

The nucleation rate in this case, Jcnucl, is equal to the rate of production of the smallest cluster (divacancies in the case considered); hence the flux J(x)|x=1 is the main concern.

When calculating Jcnucl, one can obtain two limit­ing SDFs that correspond to two different steady — state solutions of eqn [18]: (1) when the flux J(x, t) = 0, for which the corresponding SDF is »(x), and, (2) when it is a constant: J(x, t) = Jc, with the SDF denoted as g(x). Let us first find »(x). Using equation P(x)n(x) — Q(x + 1)»(x + 1, t) = 0 and the condition »(1) = C, one finds that

P(y)

Q (y + 1)

Using function »(x), the flux J(x, t) can be derived as follows

J(x, t) = P(x)n(x)

The SDF g(x) corresponding to the constant flux, J(x, t) = Jc, can be found from eqn [26]:

1

PG0«M

Using the boundary conditions g(1) = »(1) = C one finds that Jcnucl is fully defined by »(x):

nucl _ ______ 1_______

Jc 1

[P(x)n(x)]—1

x=1

Generally, »(x) has a pronounced minimum at some critical size, x = xcr, and the main contribution to the denominator of eqn [28] comes from the clusters with size around xcr. Expanding »(x) in the vicinity of xcr up to the second derivative and replacing the sum­mation by the integration, one finds an equation for Jcnucl, which is equivalent to that for nucleation of second phase precipitate particles.64,65 Note that eqn [28] describes the cluster nucleation rate quite accu­rately even in cases where the nucleation stage coexists with the growth which leads to a decrease of the concentration of mobile defects, C. This can be seen from Figure 2, in which the results of numerical integration of ME for void nucleation are compared with that given by eqn [28].73

In the case of low temperature irradiation, when all vacancy clusters are thermally stable (C = Cv in the case) and only FPs are produced by irradiation,

Подпись:Подпись:Подпись: [29]Подпись: wCvDv Cvimage777Подпись: [30]Подпись: 1 1 - DiCi/DvCv image780Подпись: [31the void nucleation rate, eqn [21], can be calculated analytically. Indeed, in the case where the binding energy of a vacancy with voids of all sizes is infinite, Ev(x) = 1 (see eqn [75]), it follows from eqn [25] that the function n(x) is equal to

C1 DyCy x 1

x1/3 DiCi

Substituting eqn [29] in eqn [28], one can easily find that the nucleation rate, Jcnucl, takes the form

rnucl

where w = (48я2/02)1/3 is a geometrical factor of the order of 1020m-2 (see Section 1.13.5). The sum in the dominant eqn [30] is a simple geometrical progression and therefore it is equal to

DyCy

DvCv — Di Ci

Substituting eqn [31] to eqn [30], one can finally obtain the following equation

Jcnucl = wCv(DvCv — DiCi) [32]

Note that the function g(x) in this case takes a very simple form, g(x) = Cc/x1/3, and hence decreases with increasing cluster size. In contrast, in R-space,
g(R) (see eqn [16]) increases with increasing cluster size: g(R) = (36я/0)1/3 CcR (see also eqns [43] and [44] in Feder et a/.69).

The real time-dependent SDF builds up around the function g(x) with the steadily increasing size range (see, e. g., Figure 2 in Feder et a/.69). Also note that homogeneous nucleation is the only case where an analytic equation for the nucleation rate exists. In more realistic scenarios, the nucleation is affected by the presence ofimpurities and other crystal imper­fections, and numerical calculations are the only means of investigation. Such calculations are not trivial because for practical purposes it is necessary to consider clusters containing very large numbers of defects and, hence, a large number of equations. This can make the direct numerical solution of ME impractical. As a result several methods have been developed to obtain an approximate numerical solu­tion of ME (see Section 1.13.4.4 for details).

The equations formulated in this section govern the evolution of mobile and immobile defects in solids under irradiation or aging and provide a frame­work, which has been used for about 50 years. Appli­cation of this framework to the models developed to date is presented in Sections 1.13.5 and 1.13.6.

OKMC Example: Ag Fission Product Diffusion and Release in TRISO Nuclear Fuel

The second example demonstrates mesoscale KMC model simulations of the diffusion of silver (Ag) through the pyrolytic carbon and silicon carbide containment layers of a TRISO nuclear fuel particle. The model atomically resolves Ag, but provides a nonatomic, mesoscale medium of carbon and silicon carbide that includes a variety of defect features including grain boundaries, the carbon-silicon car­bide interfaces, cracks, precipitates, and nanocavities. These defect features can serve as either fast diffu — sional pathways or traps for the migrating silver. The model consists of a 2D slab geometry incorporating the pyrolytic carbon and silicon carbide layers, with incident silver atoms placed at the innermost pyro­lytic carbon layer, as described in more detail in Meric de Bellefon and Wirth.88

The key input parameters to the model (diffusion coefficients, trap binding energies, interface charac­teristics) are determined from available experimental data, or parametrically varied, until more precise values become available from lower length scale modeling. The predicted results, in terms of the time/temperature dependence of silver release dur­ing postirradiation annealing and the variability of silver release from particle to particle have been compared to available experimental data from the German High-Temperature Reactor (HTR) Fuel Program,89 as shown below in Figure 7, and studies performed by the Japan Atomic Energy Research Institute (JAERI).90

Figure 6 presents KMC simulation results, which shows the effect of different grain geometries in SiC on silver release during postirradiation annealing. In this model, the grains are considered to have a rect­angular geometry. The smaller dimension is parallel to the interfaces and has a fixed length of 1 pm. The longer dimension, parallel to the radial direction in a TRISO fuel particle, has a variable length that is uniformly distributed among grains over a range from 1 to 40 pm, as shown in the upper plot of Figure 6. Such a grain distribution mimics a highly columnar structure, as is often observed experimentally.91

The diffusion coefficient for silver transport within the grain boundaries has been assumed to be three orders of magnitude higher than in bulk SiC. As expected, within this model, the presence of grains provides fast diffusion paths for silver trans­port and accelerates the released fraction. Adding a

1

 

Decreasing grain width range

 

0.1

 

OPyC

 

image904
image905

0.001

 

image906image907

columnar structure in those grains further increases the release of silver; the released fraction increases from 50 to 80% when the grain distribution shifts from an isotropic structure (grains are 1 mm squares) to a highly columnar structure (length of grains uni­formly distributed from 1 to 40 mm) after 270 h of heating at 1700 °C.

Figure 7 presents KMC simulation results of the transport of silver through a given PyC/SiC/PyC microstructure during postirradiation thermal an­nealing at 1600, 1700, and 1800 °C, as well as results from three German experimental release measure­ments performed at annealing temperatures of 1600, 1700, and 1800 °C. The simulated microstructures include reflective interfaces, trapping cavities, and a grain boundary structure in the SiC layer. The microstructures for the 1600 and 1700 °C simulations (particle 23 and 24) are identical and contain an isotropic grain geometry in SiC that consists of 1-mm long square grains with a grain boundary diffu — sivity 100x higher than in bulk. In the 1800 °C simu­lation (particle 25), the SiC grain characteristics are varied to match the measured release at 1800 °C. Faster transport through grain boundaries is required to match the experimental results, which is obtained by implementing a highly columnar structure in which the grains are 0.5 mm-wide and have a radial length between 10 and 40 mm into SiC, as well as a much higher grain boundary diffusivity that is 2000 times higher than in bulk.

Ionic Sublattice/Modified Quasichemical Model for Liquids

In contrast to using associates for liquid solutions is a sublattice approach in which cations and anions are mixed on respective lattice sites. With anions and cations assigned to specific sublattices, it is possible
to capture interactions and short-range order with species occupying the sites and additional energetic terms. The components can essentially be allowed to independently mix on each sublattice within the energetic constraints and the system free energy minimized.46 The approach has been successfully used by Gueneau et a/.47 to model the liquid in the U-O and Pu-O systems where ionic metal species reside on one lattice and oxygen anions, neutral UO2 or PuO2, charged vacancies, and O species on the other.

An improvement to the simple sublattice approach is the quasichemical method introduced by Fowler and Guggenheim48 and later further devel­oped by Pelton and coworkers.49-52 It approaches short-range order in liquids through the formation of nearest-neighbor pairs on a quasilattice. It thus differs significantly from the modified associate species approach such that in the quasichemical method short-range order is accommodated by com­ponents pairing and the energetically described extent of like and unlike components pairing. The technique thus avoids the paradox where a high

image1025
degree of short-range order in the associate species approach causes minimal end-member species con­tent and therefore fails in the limit to be a Raoult’s law solution.52

In the modified quasichemical approach for a sim­ple binary A—B system, the components are treated as distributed on a quasilattice and that an energetic term governs exchange among the pairs.

(A — A) + (B — B)=2(A — B) [19]

A parameter, Z, represents the nearest-neighbor coordination number such that each component forms Z pairs. For one mole of solution,

ZXA = 2nAA + 2nAB [20]

ZXb = 2»bb + 2»ab [21]

where the moles of each pair are nAA, nBB, and nAB. The relative proportions of each pair is Xj where

Xij = n, j / (»AA + «BB + »ab) [22]

The configurational entropy contribution is captured from the random distribution of the pairs over the quasilattice positions. The result is a heat of mixing of

Hmix = (Xab/2)Eqm [23]

where Eqm is the quasichemical model energetic parameter, and a configurational entropy

Sconfig = — R(Xa ln Xa + Xb ln Xb)

— R(Z/2)(XAAln(XAA/xA ) + Хвв1п(Хвв/Хв2)

+ Хав1п(Хав/2ХаХв)) [24]

Utilizing Gibbs free energy functions for the compo­nents and expanding EQm as a polynomial in a scheme for minimizing the system free energy pro­vide a system for optimization of the liquid using known thermochemical values and phase equilibria. Issues such as the displacement of the composition of maximum short-range order from 50% composition are dealt with by assuming different coordination numbers for each component. Greater accuracy is obtained by inclusion of the Bragg-Williams model, thus incorporating lattice interactions beyond nearest neighbors. The modification to the quasi­chemical model yields

Hmix = (Eqm/2)Xab + EbwXaXb [25]

The extension of the modified quasichemical model to ternary systems is directly possible using only binary model parameters.

An issue for the modified quasichemical model is that it fails at high deviations from random ordering, although that is generally not a problem because immiscibility will occur before the deviations grow too large. The model can also predict a large amount
of ordering that can result in a negative configu­rational entropy, a physical impossibility.53

Electron Irradiations

The unique feature ofelectron irradiations in compari­son to ions and neutrons is that they create defects in very low-energy recoil events. As a consequence, nearly all FPs are produced in isolation. This has been of foremost importance in developing our understanding of radiation damage, as it made studies of defect crea­tion mechanisms as well as the fundamental properties of FPs possible. Recall that the properties of vacancies and vacancy clusters, for example, formation and migra­tion energies, stacking fault energies, etc., could be determined from quenching studies. It is not possible, however, to quench in interstitials in metals. Very little was therefore known about this intrinsic defect prior to about 1955 when irradiation experiments became widely employed. In this section, we highlight some of the key findings derived from these past studies.