Вы находитесь на странице: 1из 14

Journal of Nuclear Materials 411 (2011) 114

Contents lists available at ScienceDirect

Journal of Nuclear Materials


journal homepage: www.elsevier.com/locate/jnucmat

Ab-initio based modeling of diffusion in dilute bcc FeNi and FeCr


alloys and implications for radiation induced segregation
S. Choudhury a, L. Barnard a, J.D. Tucker c,1, T.R. Allen b, B.D. Wirth d,e, M. Asta f, D. Morgan a,
a

University of Wisconsin Madison, 1509 University Avenue, Madison, WI 53706, USA


University of Wisconsin Madison, 1500 Engineering Drive, Madison, WI 53706, USA
c
Knolls Atomic Power Laboratory, P.O. Box 1072, Schenectady, NY 12301, USA
d
University of California Berkeley, 4165 Etcheverry Hall, Berkeley, CA 94720, USA
e
Department of Nuclear Engineering, University of Tennessee, Knoxville, TN 37996-2300, USA
f
Department of Materials Science and Engineering, University of California, Berkeley, CA 94720, USA
b

a r t i c l e

i n f o

Article history:
Received 3 September 2009
Accepted 20 December 2010
Available online 12 January 2011

a b s t r a c t
A combination of ab-initio calculations and statistical mechanical models has been used to study diffusion
behavior in dilute ferritic FeCr and FeNi alloys. A full set of Onsager matrix coefcients (Lij) and tracer
diffusion coefcients (D) were calculated both for vacancy and interstitial mediated diffusion. The key
results are: (1) Cr is the fastest diffusing species by both vacancy and interstitial mediated transport followed by Ni for vacancy and Fe for interstitial mediated diffusion, respectively; (2) weak interactions
exist between Ni and Cr with vacancies as rst nearest neighbors; (3) the calculated D predict opposite
trends of radiation induced segregation (RIS) of Cr by vacancy (depletion) and interstitial (enrichment)
diffusion mechanisms, perhaps explaining the lack of clear trends in experimentally determined Cr RIS
proles; (4) unlike the widely used Darken and Manning approaches, the calculated Lij can reliably predict the direction of the vacancy defect ux, particularly at low temperatures; (5) the Lij calculated for
interstitial mediated transport indicate that the contribution of interstitial ux can be signicant in determining RIS in these alloys; and (6) solute drag is unlikely to occur for vacancy mediated diffusion. In addition, the LAB off-diagonal terms for interstitial transport, which are typically assumed to be small, are
shown to be as large as the diagonal elements LBB. These results provide a basis for understanding the
complex radiation induced segregation behavior of Cr and Ni in ferritic/martensitic alloys.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Ferritic/martensitic steels with Cr as the major alloying element
are likely to form the basis of many materials used in the next generation of nuclear reactors. These alloys are known to exhibit better swelling and thermal shock resistance than the austenitic
counterparts and hence are considered as excellent candidate
materials for nuclear fuel cladding as well as rst wall and breeder
blanket structures for future fusion reactors [1,2]. Improvement in
the lifetime and performance of these alloys as structural materials
for nuclear reactors requires a fundamental understanding of the
microstructural evolution in a wide range of temperatures and
radiation conditions.
Materials used in the radiation environments of nuclear reactors have been known to form large concentrations of point defects, e.g., vacancies and interstitials, as well as more extended
structural defects, e.g. dislocations and voids [1,3]. Formation of
Corresponding author.
1

E-mail address: ddmorgan@wisc.edu (D. Morgan).


The bulk of this work was done while at University of Wisconsin.

0022-3115/$ - see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.jnucmat.2010.12.231

extended defects may degrade structural properties of these materials. One radiation-induced phenomenon that potentially plays a
role in microstructural changes and materials degradation is radiation induced segregation (RIS). RIS is the process by which the local composition of an alloy is altered due to preferential
participation of certain species with the vacancy and/or interstitial
ux to sinks [1,3]. For example, it is well established that Cr depletes at the grain boundary in austenitic stainless steels, most
likely due to an inverse Kirkendall mechanism [4,5]. However,
experimental observation of Cr RIS at the grain boundaries in ferritic/martensitic alloy does not show any clear trend, and in fact
a much more complex segregation behavior as a function of irradiation dose. Both enrichment and depletion of Cr at the grain
boundary has been observed [2,69]. Further, Ohnuki et al. [8] observed that the presence of other alloying elements also modify the
RIS behavior of Cr. For example, Ohnuki et al. observed that Cr enriches at grain boundaries in an Fe13Cr1Si alloy, but depletes in
an Fe13Cr1Ti alloy when irradiated to 57 dpa with 200 keV C+
ions. However, for ferritic steel with a similar Cr composition
(Fe12Cr1Mo0.2C) Brimhall et al. [10] reported no measurable
segregation of Cr at the grain boundary when irradiated to about

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

1 dpa with 5 MeV Ni++. It is not clear if these variations in RIS are
due to intrinsic changes in the transport properties to the grain
boundary, or perhaps other microstructural differences between
different alloys (e.g., it has been argued that formation of Cr precipitates with other solutes at the grain boundary may alter the measured RIS proles [1].)
The inability to explain the RIS behavior in ferritic steels may be
in part due to a lack of complete understanding of the species
dependence of the interstitial ux. Interstitials are produced in
equal quantities as vacancies under irradiation and the interaction
between solutes and interstitial ux could play a signicant role in
determining the RIS proles of a solute. For example, it has been
argued [1] that in austenitic steels oversized (compared to Fe)
atoms like Cr tend to deplete at the grain boundary as the binding
energy of Cr with Fe self-interstitials is negative. Such a hypothesis
may help explain RIS behavior of Cr in austenitic steels, but does
not resolve the variation in experimentally observed RIS prole
of Cr in ferritic steels. To help explain results in the ferritic steels
it has been suggested that the presence of alloying elements can
change the average lattice parameter of the Fe, and that this variation can explain experimentally observed RIS prole based on the
fact that Cr could change from an oversized to an undersized atom
depending on the lattice parameter of the alloy [1]. However, these
qualitative arguments are highly uncertain and require a more
quantitative foundation.
Among the minor alloying elements in ferritic steels the Ni content plays an important role in determining the mechanical properties of these alloys [11]. Similar to Cr no clear RIS trend has
been observed in case of Ni. For example, in a ferritic steel alloys
Little and Stoter [12] observed that Ni enriches at the defect sinks,
while Hosseni and Jones [13] reported depletion of Ni at the grain
boundary in Fe3.6 wt.% Ni alloy. Similar to the case of Cr-rich precipitates, formation of Ni3Fe precipitates further complicates the
measured RIS prole [13]. While it is realistic to expect that formation of precipitates, the presence of various alloying elements, radiation dose and temperature may alter the evolution of
concentration of Cr and Ni in ferritic/martensitic steels, experimentally it is difcult to separate these complicated interaction effects.
In particular, it is difcult to determine what one might consider
the intrinsic RIS behavior in FeCr and FeNi alloys, which is due
to the fundamental species dependence of the bulk vacancy and
interstitial diffusion in the alloy. The goals of this work it to use
ab-initio methods to extract information about the bulk transport
and intrinsic RIS behavior of Cr and Ni in these alloys.
The intrinsic RIS tendencies in FeCr and FeNi alloys can be
established through the determination of diffusion properties of
solutes and the solvent. Such diffusion properties should take into
account the complex solutesolvent interactions as well as interactions between solutes with the defect uxes. For the sake of simplicity in this paper, we choose dilute alloys of Cr and Ni in
body-centered cubic Fe, as modeling the diffusion in more concentrated alloys requires a signicantly different approach [14,15].
This paper will focus on diffusion behavior in ferritic bcc alloys,
and results will be compared with previously reported modeling
of fcc alloys [16]. Results will also be compared to a recent study
[17] on bcc FeCr that used a similar approach to that discussed
here for vacancy diffusion and empirical potentials for studying
the interstitials (the current work will use ab-initio methods for
both the vacancies and the interstitials, and model their kinetics
using similar levels of approximation in their statistical treatment).
We adopt the framework of what are commonly referred to as
multi-frequency models to treat the statistical mechanics of diffusion of vacancies [18] and interstitials [19]. These models yield kinetic parameters from hopping frequencies, which in turn can be
calculated from ab-initio methods. The results will be compared
with known experimental data. The main results of this paper

are calculated values for tracer diffusion coefcients, Di , and transport or Onsager coefcients, Lij (sometimes called phenomenological coefcients). The Lij are less commonly used and to understand
the utility of these parameters we here briey review their role in
transport modeling.
Transport of species in alloys is commonly described using
Ficks rst law:

Ji 

N
X

Dij nj

j1

where Ji is the ux for species i, Dij is an element in the chemical diffusion coefcient matrix, and ni is the number of atoms of species i
per unit volume [20]. Eq. (1) is a useful form of the ux equation because ni are available experimentally. However, in this expression
the kinetic and thermodynamic information about the system are
inseparably combined in the diffusion coefcient matrix D. The
thermodynamic driving force for diffusion is in fact the gradient
in chemical potential, l, and not the gradient in concentration,
which leads to the following, alternative expression for ux:

Ji 

N
X

Lij lj

j1

where the kinetics are contained in the quantity Lij, the transport or
phenomenological coefcients which are more commonly known as
the Onsager matrix [20].
The Lij are the fundamental kinetic quantities and very useful
for both understanding and quantitative prediction of RIS and
other transport behavior. Among the variables described above,
of particular interest are the off-diagonal terms Lij(ij) containing
kinetic information about coupled uxes between species i and j.
These cross terms are essential for a quantitative understanding
of RIS as they include information about solutedefect coupling,
e.g., vacancy drag, which is not available from just the diagonal
terms. The off-diagonal terms will be used below to demonstrate
the absence of vacancy drag mechanisms. Lij are generally not directly measurable, making them difcult to obtain from
experiment.
One common path to obtaining Lij values is using either the Darken [21] or Manning [22] relations to determine Lij indirectly from
experimentally measured tracer diffusion coefcients (D). However, such an approach has serious drawbacks. First, tracer diffusion coefcients (D) are often available only over a limited range
of temperature and composition. Furthermore species dependent
experimental diffusion data is typically available only for vacancy
mediated diffusion [23,24], and hence the transport coefcients
involving interstitials cannot be determined [25]. In addition, while
Darkens approach provides a simple expression to calculate the
diagonal terms from measured D, it does not supply Lij(ij). Hence,
critical information about the defect-solute interactions as well as
the interaction between different atoms is absent. While it is possible to calculate the off-diagonal term using Mannings approach,
this only provides an approximate estimate of Lij, since it assumes
that the defect jumps are independent of the environment surrounding the defect. Moreover, it is not possible to predict negative
Lij within Mannings framework, which can occur at least in theory.
We will use our calculated Lij to assess the validity of the Darken
and Manning approximations for the Lijs in the FeCr and FeNi
systems.
Lijs are particularly interesting to determine for interstitial
transport. Under radiation, interstitial dumbbells are formed in
much higher concentrations than thermally formed interstitials
and transport of atoms through interstitial mechanisms could play
a signicant role in determining RIS. Due to the lack of experimental D for interstitial mediated diffusion it is almost impossible to

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

assess their Lij from experiments. A few theoretical attempts


[17,19] have been made to understand diffusion of atoms through
interstitial dumbbell mechanism in ferritic steels and analytical
expressions for Lij in terms of atomic energies have been developed
[19]. However, no numerical values of interstitial Lij have been reported. One objective of this paper is therefore to determine a full
set of Lij Onsager coefcients for both vacancy and interstitial diffusion using the energies obtained from ab-initio calculations.
These values will then be used to determine the more usual tracer
diffusion coefcients, explore solute coupling and drag mechanism, and assess Darken and Manning approaches for calculating
L values. The implications for RIS will also be discussed.
2. Ab-initio methodology
All bulk crystal, vacancy and interstitial properties have been
calculated using the Vienna ab-initio Simulation Package (VASP)
[26,27]. In these calculations, we used the projector augmented
wave (PAW) [28] method with a plane wave cut off of 350 eV
and 300 eV for vacancy and interstitial calculations, respectively.
The calculations were spin-polarized and the PerdewBurkeErnzerhof [29] parameterization of the generalized gradient approximation (GGA) was used for the exchangecorrelation potential.
The PAW potentials were generated using the following electronic
congurations: 3p6 3d5 4s1 for Cr, 3p6 3d7 4s1 for Fe, 3p6 3d8 4s2
for Ni. We note that these PAW potentials include the p-electrons
in the valence. PAW potentials without the p-electrons produced
signicant changes in the interstitial energetics, often of over
100 meV/interstitial for interstitial binding energies. Defect calculations were performed with 54(1) atoms within a periodic
3  3  3 supercell of the bcc conventional cell. The cell shape
and volume are kept xed to that of pure Fe with a bcc structure
but internal ionic relaxations are allowed. Brillouin-zone sampling
was conducted using the Monkhorst and Pack scheme [30]. A
6  6  6 k-point mesh was used for vacancy calculations, while
3  3  3 k-point mesh was employed for interstitial calculations.
The ab-initio calculations for interstitials use a lower cut off and
k-points (compared to vacancies) as the calculations for interstitials were more computationally intensive and a much larger number were required than for the vacancies. For vacancy calculations
the errors associated with k-point mesh and energy cutoff are estimated to be less than 5 meV/defect for migration barriers, defect
formation energies and binding energies. To understand the effect
of cell size we calculated the vacancy formation energy and a few
representative vacancy migration barriers with a 4  4  4 supercell. The error for vacancy migration barriers and defect formation
energies associated with cell size is estimated to be less than
20 meV and 40 meV, respectively. We did not calculate the vacancy
binding energy as a function of cell size but using ab-inito approaches Vincent et al. [31] reported that the change in rst nearest neighbor binding energy is less than 10 meV when the cell size
is increased from 3  3  3 to 4  4  4 supercell. For interstitials
we require only migration energies as these are what inuence
the kinetic predictions. The migration energies are converged with
respect to k-points and cutoff to within 10 meV. The errors associated with cell size on interstitial migration are more complex to assess. However, comparing the values for FeCr interstitial
migration energies obtained in our 54 atom cell to those obtained
by Olsson et al. [32] in a 128 atom cell we nd changes of 60 meV.
Although this is not as small as might be desired, it would likely
correspond to a change of less than factor of three in diffusion coefcient at even a relatively low temperature for RIS of 400 C. A
thorough assessment of errors associated with ab-initio simulations for vacancy and interstitial energies are given in Refs.
[33,34]. Migration barriers for vacancy and interstitial hopping
were calculated using the nudged elastic band method (NEB) with

seven and three intermediate migration images, respectively. A cubic spline was tted to the migration energy prole and migration
barriers were calculated by taking the energy difference between
the saddle point with respect to the energy at the lattice point.
In this regard, a saddle point represents the conguration where
the energy is maximum in the migration energy prole.

3. Results
3.1. Ab-initio calculations of vacancy diffusion
The approach to calculating diffusion constants is to nd hopping rates from ab-initio barriers and then use those in multi-frequency statistical diffusion models. In order to understand the
effect of different vacancy exchange mechanisms on solute diffusion, in this work we adopted the multi-frequency framework
developed by LeClaire [18] as shown in Fig. 1. It is assumed that
the solute species is dilute enough that any solutesolute interactions may be neglected. The w0 is the rate of hopping in the pure Fe
(not shown). We calculated the activation energies of each of the
jump events, wi, for a vacancy in close proximity to a solute and
the results are presented in Table 1. The calculated migration barriers match well with migration barriers available in literature
[17,33,35]. The jump frequencies (wi) are calculated from the
migration barriers using the expression

wi mi exp 

!
DHi;mig
v
:
kT

We assumed a constant attempt frequency (mi) of 5  1012 s1


for all jumps and DHi;mig
is the migration barrier for the ith vacancy
v
jump. It is clear from Table 1 that the migration barriers are similar

w4
w2

w3

1 w
4

w4

w3

w3

Fe

Cr/Ni

Vacancy

Fig. 1. Illustration of multi-frequency model with rst nearest neighbor interaction


for vacancy jump in a dilute alloy with bcc structure. The numbers in the gures
indicates the nearest-neighbor positions to solute.

Table 1
Migration barriers of vacancies obtained from ab-initio calculations.
Jump type

w0
w2
w3
w4
w03
w04
w003
w004

Calculated migration energy (eV)


FeNi

FeCr

0.67
0.68
0.55
0.69
0.70
0.67
0.62
0.59

0.67
0.58
0.69
0.65
0.67
0.63
0.64
0.62

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

for FeCr and FeNi systems. In the ab-initio calculations we treated the associative and dissociative jumps to and from the 2nd, 3rd
and 4th nearest-neighbor positions explicitly (Table 1). However,
in order to t the ab-initio based rates into LeClaires framework
we replaced the rate of the dissociative and associative jumps from
or to the rst nearest neighbor site of the solute with an effective
eff
rate, denoted by weff
3 and w4 respectively [16,36]. The effective
rate is determined by an average weighted by the number of pathways to each nearest-neighbor distance as shown below.
0
00
7weff
3 3w3 3w3 w3

and
0
00
7weff
4 3w4 3w4 w4

The jump rates were used to calculate the vacancy Onsager


coefcients as a function of temperature in FeNi and FeCr alloy
using the relations presented in Ref. [37] (see Appendix A) and
the results are plotted in Fig. 2. In the gure A = Fe, B = Cr or Ni.
The dotted and the solid lines in the gure represent the off-diagonal and the diagonal Onsager coefcients, respectively. It should
be pointed out that based on the expressions of Lij in Appendix A
the value of LAA is different for FeCr and FeNi alloys, as Cr and
Ni have different solute enhancement factor (b). We observed that
the difference in value of LAA in FeCr and FeNi alloys is less than
5% within the range of temperature described in the gure. Hence
for the sake of clarity of the gure we only present LAA in FeCr alloy in Fig. 2. For the calculation of vacancy Onsager coefcients we
use a value for mole fraction of solute cB = 0.01, which corresponds
to nB = 8.5  1020 atoms/cm3. It can be observed that as the jump
rates are weakly species dependent the Onsager coefcients of Ni
and Cr are similar within the range of temperature described in
the gure. In the gure a change of slope in Lij occurs at
T = 1043 K because of the ferromagnetic phase transition, as explained later in this section.
The interaction between the solute and the vacancy may affect
the ability of the solute to diffuse through vacancy defect ux. For a
dilute alloy, if the vacancy solute interaction is negligible then the
solute diffuses though vacancysolute exchange (the inverse Kirkendall mechanism). However, if a strong attractive interaction exists between the solute and the vacancy then the solute and the
vacancy can migrate as a complex species, or in other words, the
solute is dragged in the direction of vacancy defect ux. In order
to understand how a vacancy binds with a solute atom (Cr, Ni) as
a function of separation between solute and vacancy we calculated
the solutevacancy binding energy as a function of the nearest
neighbor locations.

14

Log (Lij(1/meVs))

12
10
8
6
4
0.83

LFeFe in in
Fe-Cr
LFeFe
Fe-Cr
- abab-initio
initio
LFeCr
- abab-initio
initio
Fe-Cr
in Fe-Cr
LFeCr in
LCrCr
- ab ab-initio
initio
LCrCr in Fe-Cr
in Fe-Cr
LFeNi
- ab ab-initio
initio
in Fe-Ni
LFeNi in Fe-Ni
LNiNi
- ab initio
LNiNiinFe-Ni
in Fe-Ni
ab-initio
0.88

0.93

0.98

1.03

1.08

1/T ( x 10-3 K-1)


Fig. 2. Vacancy Onsager coefcients obtained from multi-frequency model.

The rst nearest neighbor binding energy between the solute


and the vacancy in a bcc iron matrix comprised of N atomic sites
is obtained from Eq. (6),

DHbind
v EN  2; 1Cr;1V; 1nn  EN  1; 1V  EN  1; 1Cr EN;
6
where EN  2; 1Cr;1V; 1nn is the energy of the supercell containing N  2 Fe atom and one Cr atom with a vacancy as a rst nearest neighbor [34]. Similarly, the second, third and the fourth nearest
neighbor binding energies were calculated and the results are presented in Fig. 3a. With the sign convention used here a negative
binding enthalpy means attractive interaction and a positive value
indicates repulsion. The gure shows that a weak attractive interaction exists between the vacancy and the solute in almost all cases.
The one exception is for the binding energy of Ni with the vacancy
when they are second nearest neighbors to each other, which shows
a quite strong attraction. The source of this attraction is not clear
but it is possibly due to a combined effect of size as well as chemical
interaction between the vacancy and the Ni atoms, as reported previously [31]. The calculated binding energies agree well (within
25 meV) with previous ab-initio calculation for the rst and second
nearest neighbors in the FeNi binary system [31]. These binding
energies suggest that the vacancy drag mechanism is unlikely,
although the effect of the second neighbor Ni interaction could lead
to vacancy drag.
To make a more quantitative assessment of the vacancy drag
the Lij can be used to determine the vacancy wind (G), which is often used to understand the effect of vacancies on the diffusion of
solute atoms [38]. The parameter G is related to the off-diagonal
Onsager coefcient (LAB) through the relation G = LAB/LBB [39]. It
should be noted that the Onsager coefcient LvB in a binary alloy
can be expressed as LvB = LBB (G + 1). It is clear that for G < 1,
LvB is positive, i.e. the vacancy and the solute diffuse in the same
direction as a complex species. Fig. 3b shows the vacancy wind
as a function of temperature for FeCr and FeNi system. It can
be seen that G > 1 at all temperatures of interest for both the alloys. However, it has been observed that G < 1 for T < 150 K (not
shown). We conclude that at all temperatures of interest the vacancy drag of substitutional solute is unlikely to occur for the
FeCr and FeNi systems, and the diffusion occurs through a vacancysolute exchange mechanism. Such a conclusion is consistent
with the generally weak binding between the solute and the vacancy (Fig. 3a). We note that If Ls are derived by applying Darkens
approach to D values then LAB = 0 and the vacancy wind (G) is
strictly equal to zero.
Fig. 3b demonstrates that the LFeCr/LCrCr ratio changes sign at
around 550 K for the FeCr system. Our calculation shows that LCrCr
remains positive below this temperature as it must by the second
law of thermodynamics [20]. Therefore, it is LFeCr which changes
sign as the temperature is decreased below 550 K. We note that
such change in sign of LAB is not predicted by applying Mannings
approach to obtain L from the D values, which gives LAB P 0 (a
negative value is also not possible to predict in Darkens approximation, which gives LAB = 0). Moreover, very little experimental
D data are available at low temperatures to calculate even the
diagonal terms reliably using either Darkens or Mannings approach. Thus, the multi-frequency approach provides a powerful
tool to determine Lij, including cross terms, over a wide temperature range.
Experimental validation of the energies calculated from ab-initio calculations can be done by comparing the ab-initio calculated
vacancy tracer diffusion coefcients (D) of Fe, Cr and Ni with
experimental measurements. The vacancy tracer diffusion coefcient of Fe and the solutes as a function of temperature can be calculated using the expressions given below [18].

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

0.1

Binding Energy (eV)

0.05

Cr

Repulsion
1st

2nd

3rd

Ni

(b)

4th

-0.05
-0.1

0.5

G, Vacancy Wind

(a)

Binding

-0.15

-0.5

-1

Fe-Cr
Fe-Ni

-0.2
-1.5
400

-0.25

700

Nearest Neighbor Position

1000

1300

T (K)

Fig. 3. (a) Solutevacancy binding energy vs. nearest neighbor position; (b) vacancy wind, G, for FeCr and FeNi.

!
DHfv
;
kB T

DFe cB DFe 01 bcB ;


DCr=Ni cB a2 f2 w2 exp

!
DHfv DHbind
v ;
kB T

where a is the lattice parameter, T is the temperature, DHfv and


DHbind
are the formation energy of the vacancy in pure Fe and the
v
rst nearest neighbor binding energy between a solute and a vacancy, respectively. DFe (0) is the tracer diffusion coefcient of Fe
in pure Fe, while DFe (cB) and DCr=Ni (cB) represent the tracer diffusion
coefcient of Fe and solutes in the dilute alloy, respectively. Note
that in the expression for D in Eqs. (7) and (9) the entropy terms
are set to zero, which is a simplifying approximation. We will discuss the effect of entropy terms on D later in this paper. Assuming
the alloys to be dilute in solute, the pure Fe lattice parameters are
used for all tracer diffusion calculations. The calculated vacancy
formation energy (2.23 eV) in pure Fe matches well with vacancy
formation energies available in literature [17,33,40]. For pure bcc
Fe, we used correlation factor f0 = 0.7272 [18]. The correlation factor
f2 is obtained as f2 = v/(2(w2/weff
3 ) + v) where, v = 7  (1 + 0.512(w0/
1
1
1
weff
 2(1 + 1.536(w0/weff
 (1 + 3.584(w0/weff
. [41] Be4 ))
4 ))
4 ))

fore the D values can be determined we must address the issue
of magnetic changes in the Fe.
It is well known that pure iron when cooled below 1043 K
undergoes a phase transition from a paramagnetic to a ferromagnetic state [42]. This phase transformation affects the activation

Tc=1043( K)

(b) 12

Fe -ab initio
Ni - ab initio
Cr - ab initio
Fe - 10 exps.
Ni - 3 exps.
Cr - 4 exps.

-16

-20

-22
0.85

Paramagnetic

0.9

Ferromagnetic

0.95

1/T ( x 10-3 K-1)

1.05

1.1

D*(Cr)/D*Fe - ab initio
D*(Ni)/D*(Fe) - ab inito
D*(Cr)/D*(Fe)- exps.

D*(Ni)/D*((Fe) - exps.
Tc=1043( K)

6
4

Paramagnetic

-18

10

Ferromagnetic

Log (D*(m2/s))

(a) -14

energy of vacancy diffusion in pure iron and its alloys. It has been
observed that in the paramagnetic state D for pure iron follows an
Arrhenius relationship, but below the Curie temperature the D
deviates downwards from the Arrhenius type relationship extrapolated from the paramagnetic state. This deviation has been attributed to temperature dependence of the activation energy for
vacancy diffusion arising due to the change of magnetization in
pure Fe and its alloys [43]. In other words, in the ferromagnetic
state the correct activation energy for vacancy diffusion is a function of the spontaneous magnetization. We have modied the activation energies calculated with ab-initio approach to account for
the magnetic changes with temperature using an empirical relationship. The details of this approach are given in Appendix B.
The calculated tracer diffusion coefcients for pure Fe (Eq. (7)),
and for Cr and Ni (Eq. (9)) solutes in Fe, are shown by the solid lines
in Fig. 4a. In the gure D(cB) (see Eq. (8)) for Fe has not been shown
separately as the solute enhancement factor (b) is weak at high
temperatures. In Fig. 4a scattered symbols show the average of
the experimentally measured diffusion coefcients ([4448], references within [23,24]). It should be pointed out that among the list
of references; we ignored references in which diffusion data are
presented in Arrhenius form through the ferromagnetic transition
temperatures. It is clear from the gure that the slope of the ab-initio based diffusion coefcients as a function of inverse temperature
(i.e. the activation barrier) matches well with experimental measurements. Further, the relative behavior of the diffusion coefcients between pure Fe, Cr and Ni predicted from the ab-initio
based approach agrees well with experimentally observed behavior, i.e. both experimentally and ab-initio calculated D show that
Cr is the fastest diffusing species by a vacancy mechanism while

D*(Cr, Ni) (cB) /D*(Fe) (cB )

DFe 0 a2 f0 w0 exp

2
0
400

700

1000

1300

T (K)

Fig. 4. (a) Vacancy tracer diffusion coefcients of pure Fe, Ni and Cr in alloy. In the gure the solid and the dotted lines shows the tracer diffusion coefcients obtained from
ab-initio calculations and average of experimental measurements, respectively; (b) ratio of Cr and Ni vacancy tracer diffusion coefcients relative to Fe tracer diffusion
coefcients.

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

the diffusivity of Fe and Ni are quite close to each other. This result
can be explained based on the migration energy (see Table 1) for
the jump involving direct exchange between the vacancy and the
species, w2. The migration energy for w2 jump is 0.58 eV and
0.68 eV for Cr and Ni, respectively, while the migration barrier
for Fe is 0.67 eV for the w0 jump. In Fig. 4b, we present the ratio
of calculated and measured tracer diffusion coefcients of solutes
to that of Fe as a function of temperature. We should point out that
in plotting calculated D for Fe we have taken into consideration
the solute enhancement factor b (Eq. (8)), as it was observed by
Wong et al. [17]. that b is signicant at low temperatures. The ratio
of D is particularly important in predicting RIS proles in an irradiated environment as the concentration prole of atoms depends
on the relative value of the tracer diffusion coefcients rather than
the absolute tracer diffusion coefcients. It can be observed that
within the temperature range in which experimental data are
available the calculated tracer diffusion coefcients match well
with experimental measurements except for the FeCr system below the ferromagnetic transition temperature. Although the absolute ratios are quite close, the experimental D(Cr)/D(Fe) does not
seem to be rising with decreasing temperature, as seen in the abinitio based model. Such error may arise from ab-initio calculated
energies. For example, an error of 20 meV in ab-initio calculated
migration barrier (see Section 2) may explain the discrepancy between calculated and measured ratio of D in FeCr system at
T = 900 K. Interestingly, for the FeNi system, within the same temperature ranges D(Fe) is slightly greater than D(Ni). However, at
low temperature D(Ni) is greater than D(Fe).
We should point out that in similar ab-initio based modeling for
FeCr in Ref. [17], the diffusion coefcients were calculated based
on nine frequencies. In the present calculations, although eight different jumps (see Table 1) were considered, only four effective frequencies were used in calculating tracer diffusion coefcients and
Onsager coefcients. To understand the error incorporated using
four effective jump frequencies on RIS prediction we calculated
the ratio of tracer diffusion coefcient of Cr with respect to Fe
using the jump barriers presented in Ref. [17], rst with all the nine
frequencies as presented in Ref. [17] and latter with four effective
frequencies calculated from the jump barriers in Ref. [17]. We observed that independent of temperature the D(Cr)/D(Fe) only varies within a factor of 3 between the two approaches. As the
analytical expressions to calculate the Onsager coefcients for vacancy transport provided in reference [37] only includes a limited
number of frequencies, we used four effective frequencies throughout this work to ensure the consistency of the calculations of D
and Lij throughout this work.
It can be observed in Fig. 4a that ab-inito sets of data under predict the experimental diffusion coefcients quite signicantly. For
example, at T = 1193 K for pure iron, the ratio (R) of the average D
measured experimentally to ab-initio calculated D is R = 684.2. As
an empirical correction to the ab-initio based values we simply
multiply the ab-initio calculated D for pure iron and solutes for
all temperatures with the ratio R. We observed that the calculated
new D now lies within the error bar of the measured D over the
range of temperatures presented in Fig. 4a for Fe, Cr and Ni (not
shown). The discrepancy between the measured and calculated
tracer diffusion coefcients could arise from several factors (or a
combination of factors) that have been neglected in calculating
the tracer diffusion coefcients and/or from errors in the calculated
ab-initio values. The neglected factors include the entropy of vacancy formation and migration, both of which can have complex
contributions from vibrations, magnetism and electronic sources
[16]. For example, using molecular dynamics simulations the entropy of vacancy formation in ferromagnetic iron has been estimated to be 1.78kB, [49] which would enhance the D values
about a factor of 6. Using frozen-phonon calculations, Huang

et al. [50] recently calculated the various entropy terms for diffusion of impurities like Ta and Zr in pure Fe. It was found that the
inclusion of these entropy terms yields good agreement with
experimental data for their diffusion constants. Accurate estimation of all the factors neglected in calculating the D, or precise
assessment of error originating from ab-initio values is beyond
the scope of this paper, because the primary focus of this research
is to develop parameters for RIS modeling in FeCrNi alloys. As
mentioned above, RIS depends on the ratio of D of the solute with
respect to the solvent, rather than the absolute D.
In summary, our ab-initio calculations accurately predict the
relative relationship of tracer diffusion coefcients between Fe,
Cr and Ni, but the calculated tracer diffusion coefcients are significantly lower that the experimental values. Calculated Lij for vacancy mediated diffusion shows that vacancy drag is unlikely to
occur both in FeNi and FeCr bcc alloys at all temperatures of
interest.
3.2. Ab-initio calculations of interstitial diffusion
It is well known that in metals the formation energy of interstitial dumbbells (4 eV) is much higher than the vacancy formation
energy (12 eV), which generally results in very low concentration of interstitials at thermal equilibrium. However, under irradiation, interstitials and vacancies are produced in equal numbers, so
a signicant part of diffusion of atomic species under irradiation
can occur through interstitial mediated diffusion. As mentioned
in the introduction it is often difcult experimentally to obtain diffusion data for interstitials in multi-component alloys. Recently,
such data are increasingly obtained using ab-initio techniques
[16,32,51]. In this paper, we adopt the multi-frequency Barbe
and Nastar [19] model based on self-consistent mean eld theory
to calculate diffusion properties in FeCr and FeNi alloys through
interstitial mechanisms using ab-initio techniques.
Interstitials dumbbells in dilute bcc alloys can assume a number
of orientations, typically categorized by their crystallographic
alignment. Ab-initio calculations have shown that the Fe self-interstitial in pure bcc Fe is most stable in the h1 1 0i dumbbell orientation [33,52]. However, in case of self-interstitials in pure bcc Cr the
h1 1 1i is the most stable conguration [53]. It is to be noted that an
alternative h2 2 1i stable conguration has also been proposed recently for self-interstitials in pure bcc Cr [32]. Formation energies
calculated elsewhere[34,54,55] reveal that the h 110 i orientation
is more stable than the h1 1 1i orientation for all the interstitials
and interstitial-solute complexes that are included in our calculation of the interstitial phenomenological coefcients for dilute
FeCr and FeNi alloys. It is notable that the FeNi h1 1 0i mixed
interstitial is unstable relative to an FeFe h1 1 0i interstitial with
a Ni nearest neighbor, suggesting qualitatively that Ni migration
through interstitials could be unfavorable [54].
We used the model developed by Barbe and Nastar [19] for diffusion of h1 1 0i interstitial dumbbells in dilute bcc metals to calculate interstitial transport coefcients for dilute FeCr and FeNi
alloys. Within the framework of this model the hopping event of
an h1 1 0i AB interstitial dumbbell involves the displacement of
the AB atom pair towards a target atom C to form a new dumbbell
BC (if B is the atom that jumps), while atom A now occupies the
substitutional lattice site of the initial AB dumbbell. As AC can
be different types of atoms; such a hopping mode may change
the composition of the dumbbell. Following the work of Barbe
and Nastar we only considered the rst nearest neighbor jumps
and allowed displacements of atoms by simple translation of
atoms or a combination of translation and rotation by 60. In this
context, a target site is dened as the nearest neighbor site that
an atom in the dumbbell can hop to directly, while a non-target
site is a nearest neighbor site that an atom in the dumbbell can

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

only hop to if the interstitial rst undergoes an onsite rotation. It


should be noted that displacement through translationrotation
mechanism results in an orientation of the BC dumbbell different
than the initial AB dumbbell. Further, based on Ref. [19], eight
types of dumbbell congurations are considered. Each of the
dumbbell congurations can displace through a translational
migration (denoted sixi) and a translation combined with a rotation (denoted xi), for a total of 16 jump types. These jump types
are depicted schematically in Fig 5. Jump types 3, 5, and 7 are
not presented because these are symmetric counterparts to 2, 4,
and 6, respectively (i.e., jump type 3 is identical to jump type 2
in reverse). Using ab-initio methods we calculated the migration
barriers of all the 16 types of hopping events in FeCr and FeNi
alloys and the results are shown in Table 2. The migration barrier
of the jump type i is dened as the energy difference between
the initial conguration for jump type i and peak transition energy
for jump type i. The results are compared with calculated migration barriers obtained from literature. Unlike the vacancy migration barriers it is clear from Table 2 that the some of the
interstitial migration barriers are strongly species dependent. In
particular, jump barriers involving dumbbells in which the starting
conguration is mixed (x1, x2, s1x1 and s2x2 differ signicantly
between the species.
As an aside, we note that we observed two transition states,
with a minimum in the middle, in the migration energy prole
for s2x2 and s3x3 type of displacement in FeCr alloy (not shown).
These multiple maxima and minima were found even when the
number of images was increased to 7 in the NEB calculations.
The local minimum at the middle corresponds to the h1 1 1i orientation of the FeCr dumbbell. From the local energy minimum the
dumbbell can, if we ignore any possible new hopping paths that
might be available from the intermediate minimum, proceed further and complete the jump or can jump back to its original conguration. To take this possibility into account we assume that no
signicant time is spent in the intermediate local minimum and
use an effective jump frequency equal to one half of the original

jump frequency. However, this approximate treatment is inadequate as it is possible for the dumbbell to move from the new
intermediate h1 1 1i state to other dumbbell orientations, including the possibility of a h1 1 1i translation, without completing the
s2x2 and s3x3 type hops. To include these effects would require
signicant modication of the basic model, which is beyond the
scope of this paper. Therefore, to determine the impact of ignoring
these effects, we assess how changes in s2x2 and s3x3 type hops
affect the interstitial D values. The s2x2 and s3x3 are both translation hops (see Fig. 5), without any rotation. In both cases their
associated translationrotation hops, x2 and x3, involve much
lower barriers. This leads to a very weak dependence of D on
the exact values of s2x2 and s3x3 (for example, a change of a factor
of 10 in s2x2 and s3x3 changes D by only 2%). Given that the s2x2
and s3x3 type hops have relatively little impact on the D values
for the cases modeled here, the simple approximation of reducing
their hop rates by two is used in this work.
Similar to the vacancy case, interaction between the solute and
the interstitial affects the interstitial defect ux. It has been proposed that undersized solutes tend to bind more strongly with
an interstitials than vacancies. Moreover, a soluteinterstitial complex is often much easier to transport than a solutevacancy complex [56]. It has been reported both experimentally [57] and from
ab-initio calculations [55] that strong binding of Cr with self-interstitials in a concentrated FeCr alloys can reduce the mobility of
the interstitial dumbbell.
The migration barriers presented in Table 2 are used to calculate the jump frequencies using Eq. (3). In all the calculations we
used a xed attempt frequency (m) of 5  1012 s1, the same as
for the vacancy. The jump frequency is later used to calculate the
jump probability using the following relation:

W i c i xi

10

where ci is probability of nding an interstitial in the initial conguration corresponding to jump type i. The expressions for calculating ci are listed in Appendix C. The calculated jump probabilities are

(a)

(b)

Fig. 5. Schematic diagrams of the interstitial dumbbells hopping events. (a) By translation; (b) by translationrotation mechanisms. In the gures, the dark and grey circles
represent the solvent and the solute atoms respectively. The migration barriers of the individual jumps are presented in Table 2. It is to be noted that in the gure jump types
3, 5, and 7 are not presented separately as these jumps are symmetric counterparts to 2, 4, and 6, respectively (i.e., jump type 3 is identical to jump type 2 in reverse).

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

Table 2
Migration barriers of interstitial dumbbells obtained from ab-initio calculations.
Jump type

Description

Calculated migration energy


(eV), FeNi

s0x0
s1x1
s2x2
s3x3
s4x4
s5x5
s6x6
s7x7
x0
x1
x2
x3
x4
x5
x6
x7

AA ? AA
AB ? BA
BA ? AAkB
AAkB ? AB
AAkB ? AA
AA ? AAkB
AA\B ? AA
AA ? AA\B
AA ? AA
AB ? BA
BA ? AAkB
AAkB ? AB
AAkB ? AA
AA ? AAkB
AA\B ? AA
AA ? AA\B

0.84
0.69
0.47
0.71
0.80
0.71
0.40
0.41
0.36
0.41
0.09
0.33
0.33
0.28
0.36
0.34

Published value (eV),


FeNi
0.46 [54]

0.45 [54]
0.13 [54]
0.45 [54]

Calculated migration energy (eV),


FeCr

Published value (eV), FeCr

0.84
0.48
0.58
0.68
0.79
0.79
0.32
0.35
0.35
0.25
0.33
0.42
0.39
0.26
0.36
0.37

0.80
0.42

0.37
0.23
0.33

0.22

[54], 0.78 [52]


[32]

[54], 0.34 [54], 0.33 [76]


[32]
[32]

[32]

Symbols: AAkB means an AA dumbbell with a B on a target site. AA\B means an AA dumbbell with a B on a non-target site. AB means a mixed AB dumbbell. AA means an
unmixed AA dumbbell without a B on a nearest neighbor site.

utilized to calculate the interstitial Onsager coefcients (Lij) using


the expressions provided in Appendix C. The calculated Onsager
coefcients as a function of temperature are presented in Fig. 6.
Similar to the case of vacancy we used A = Fe, B = Cr or Ni. It is notable from Fig. 6 that the off-diagonal term in the L matrix, LAB, is
slightly higher than the main diagonal term LBB for both Cr and
Ni. In most instances the off-diagonal terms are small with respect
to the main diagonal terms, and consequently they are often neglected. In this case however neglecting the off-diagonal terms will
result in signicant error in estimating diffusion ux [19]. The physical consequence of this phenomenon is that a gradient in the chemical potential of the solute (either Cr or Ni) will result in a greater
ux of Fe than of the solute species. It is not clear to what extent
this surprisingly large contribution from the off-diagonal terms
would persist in the concentrated alloy as the Lij can be strongly
concentration dependent. For vacancy mediated diffusion the effect
of the ferromagnetic transition on the activation energy of diffusion
was estimated from the experimental measurement of D through
the ferromagnetic transition temperature (see Appendix B). However, similar measurement of D for interstitial mediated diffusion
are not available. Hence, for interstitial mediated diffusion we have
not taken into account the effect of ferromagnetic transition on our
calculated Onsager and tracer diffusion coefcients.
In case of vacancy mediated diffusion we observed that
Lv B LBB LAB =LBB 1, implying that the sign and magnitude of

LFeCr Crab-initio
LAB,
- ab initio
ab-initio
LFeNi Ni
LAB,
- ab initio
LCrCr Cr
ab-initio
LBB,
- ab initio
ab-initio
LNiNi Ni
LBB,
- ab initio
LFeFe Feab-initio
LAA,
- ab initio

Log (Lij (1/meVs))

-2

-4
0.83

0.88

0.93

0.98

1.03

1.08

1/T ( 10-3 K-1)


Fig. 6. Calculated values of interstitial Lijs based on ab-initio energetics.

LAB plays a key role in assessing whether solute transport by vacancies occurs with the vacancy ux (vacancy drag) or against it (inverse Kirkendall). While a formally equivalent analysis can
performed for interstitials to yield LIB LBB LAB =LBB 1 it is not
clear what physical mechanism would allow solute to ow opposite the interstitial ux direction (LIB < 0). In fact, inspection of
the expressions for LAB and LBB (Eq. (C.2) and (C.3) in Appendix C)
reveals that both the coefcients are positive at all temperatures
and irrespective of the magnitude of the jump frequencies, as the
factor hw always less than unity. Thus the interstitial Onsager coefcient LIB is always greater than zero and interstitial mediated solute ux is in the same direction as the overall interstitial ux.
After the Lij transport coefcients are determined, tracer diffusion coefcients can be calculated as a function of temperature
using the following expression [20], thereby making a connection
between atomistic jump rates and macroscopic transport
characteristics:

DB LBB

kB T
nB

11

where DB is the tracer diffusion coefcients via the interstitial


mechanism for species B, respectively, and nB is the number of
atoms of species B per unit volume. The results are plotted in
Fig. 7a. In the case of Fe, the tracer diffusion coefcient is calculated
by evaluating the transport coefcient LBB by Eq. (C.3) with every
migration frequency set equal to the value for pure Fe. In other
words, every xi and sixi are replaced by x0 and s0x0, respectively.
It can be seen from Fig. 7a that although Cr and Fe have similar
interstitial tracer diffusion coefcients, D(Ni) is at least two orders
of magnitude lower than for both Cr and Fe. To better understand
the implications of the calculated D on RIS, we plotted in Fig. 7b
the ratio of interstitial tracer diffusion coefcients of the solutes
to that of Fe as a function of temperature. The gure shows that
Cr is the fastest diffuser followed by Fe and Ni. Moreover, relative
to Fe the diffusivity of Cr increases with decreasing temperature.
However, the interstitial-mediated diffusivity of Ni shows the
opposite trend as temperature decreases. Wong et al. [17] also calculated D(Cr)/D(Fe) for interstitial mediated diffusion using
molecular dynamics simulations. Using two different semi-empirical potentials they predicted that D(Cr)/D(Fe) is close to unity
between 400 and 1100 K (more specically, Wong et al. found a ratio of approximately 1.1 within this temperature range for the potential that best agreed with the ab-initio calculated binding
energies). Thus our predicted value of D(Cr)/D(Fe) differs

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

(a) -22
-24

16

D*(Cr)/D*(Fe) - ab initio
D*(Ni)/D*(Fe) - ab initio

0.35
0.30

-26
-28
-30

0.25

12
10

0.20

0.15

0.10

D*( Ni)/D*(Fe)

14

D*(Cr)/D*(Fe)

Log (D*(m2/s))

(b) 18

Fe - ab initio
Cr - ab initio
Ni - ab initio

4
-32

0.05

2
-34
0.83

0.88

0.93

1/T (

0.98

10-3

1.03

1.08

0
400

700

1000

0.00
1300

T (K)

K-1)

Fig. 7. (a) Interstitial tracer diffusion coefcients of Fe, Ni and Cr; (b) ratio of Cr and Ni interstitial tracer diffusion coefcients relative to Fe tracer diffusion coefcients.

signicantly with previously reported values, particularly at low


temperatures. In the next section we discuss the implications of
the interstitial and vacancy relative tracer diffusivities in predicting RIS as a function of temperature.
To summarize, we calculated the Lij and D for interstitial mediated diffusion in FeCr and FeNi bcc alloys. Calculated D shows
Cr is the fastest diffusing species followed by Fe and Ni. For the
interstitials it was observed that off-diagonal terms (LAB), which
are typically assumed to be small, are as large as the diagonal elements LBB.

faster than Fe by an interstitial mechanism, and thus it is expected


that Cr will enrich at the grain boundary if the diffusion occurs predominantly by an interstitial mechanism. Thus, it is clear that diffusion through vacancy and interstitial mechanisms will have
opposite effects on the Cr segregation behavior and whatever RIS
occurs at the grain boundary in FeCr bcc alloys will be a balance
between these two large and opposite RIS tendencies. The idea that
observed RIS is a balance of two large and opposite RIS tendencies
suggests that RIS proles might be strongly affected by changes in
either the interstitial or vacancy ux. There are a few ways these
changes might be realized.

4. Discussion of implications for radiation induced segregation


In this section we present the implications of the calculated tracer diffusion coefcients in predicting the spatial distribution of
atoms near sinks in an irradiated environment. Under a radiation
environment the local chemistry of the species near a sink depends
on the relative diffusivity of the solute compared to the solvent by
vacancy and interstitial diffusion mechanisms and the absolute
uxes of the interstitial and vacancy defect species. It should be
pointed out that although the full set of Onsager coefcients were
derived for vacancy and interstitial mediated diffusion, in this section we utilize the calculated D to discuss RIS behavior. This is
essentially equivalent to ignoring the cross terms in the Lij matrix.
Ignoring Lij for vacancy mediated diffusion is acceptable for qualitative analysis provided the cross terms do not cause any drag effects or dominate the problem, which we have previously shown to
be the case. For interstitial mediated diffusion ignoring the LAB may
cause an error in RIS prediction as LAB is of the same order of magnitude as LBB. In order to estimate the error caused by ignoring the
LAB term we calculated ux ratios (JB/JA) and JI for a delta function
peak of interstitials for the case with and without the LAB term
(see Appendix D). It was found that by ignoring the LAB term, JB/JA
changes approximately by a factor of 2 while the change in the
absolute interstitial defect ux is insignicant (<3%). Hence, in this
section we discuss the RIS implications based on only the calculated tracer diffusion coefcients.
For vacancy mediated diffusion, since the solute and vacancy
uxes are in opposite directions, a higher relative vacancy diffusivity of the solute compared to the solvent means that the solute is
expected to deplete at the grain boundary. Since the interstitial
and atom uxes are generally in the same direction, a higher solute
than solvent diffusivity by an interstitial transport mechanism will
generally cause the solute to enrich at the boundary. Fig. 4b shows
that at all temperatures, Cr has a higher vacancy tracer diffusion
coefcient than Fe. Hence, it is expected that Cr will deplete at
the grain boundary if the dominant defect diffusion occurs by a vacancy mechanism only. However, Fig. 7b shows that Cr diffuses

4.1. Solute concentration


Perhaps the simplest way the balance between vacancy and
interstitial RIS tendencies might be altered is by changes in composition altering the basic diffusion properties. For example, it has
been shown both experimentally [57] and using ab-initio [55] calculations that diffusion behavior of interstitials is dependent on
the Cr concentration in FeCr alloys. Such variation in diffusion
behavior originates from strong dependence of solutedefect binding energy [55] and interstitial migration barriers [58] with
increasing Cr concentration in the vicinity of the interstitial. While
these changes warrant further study they are beyond the scope of
this paper, as we use models restricted to relatively dilute limits.

4.2. Biased sinks


Another way the balance of RIS tendencies might be altered is
through interaction with other sinks. It is well known that there
are other sinks within the alloy that annihilate different point defects, often with a bias toward a certain defect type. In particular,
dislocations preferentially annihilate interstitials compared to
vacancies while voids tend to absorb a net ux of vacancies (which
are present in higher concentration than interstitials due to the
biased annihilation of interstitials at dislocations) [5961]. These
sinks can alter the RIS by a few mechanisms. The simplest effect
would simply be to reduce the overall ux to the grain boundaries,
which would reduce the RIS. Sinks might also alter the overall balance of interstitials and vacancies that make it to the grain boundary. However, it is not clear to what extent this is possible as it is
often assumed that the interstitial and vacancy uxes to a grain
boundary must be close to equal. Finally, sink bias and distribution
could alter the RIS through the presence of sinks near the grain
boundary altering the local concentrations. For example, large concentrations of voids are observed to form near the grain boundaries
in ferritic/martensitic steel [62]. As voids absorb vacancies they

10

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

may have signicant RIS at their surface, which in turn may inuence the measured RIS prole at the nearby grain boundary.
It is important to consider how the spatial distribution of sinks
might alter RIS in real alloys. Distributions of such sinks are not
uniform and can vary signicantly depending on processing and
operating conditions. For example, the dislocation density formed
during irradiation is not uniform throughout the alloy. Allen
et al. [63] observed that there is a signicant spatial variation in
dislocation density in the ferriticmartensitic alloy HCM12A, with
some regions free of dislocations while other areas contained
dense complex dislocation networks, and that this variation persists after irradiation. A similar observation was made by Gupta
et al. in the ferriticmartensitic alloy T91 [2]. Such spatial variation
of dislocation density may contribute to a variation in the grain
boundary RIS behavior measured in different regions of the sample.
Variations in sink properties with different processing conditions
could perhaps also lead to different RIS behavior in alloys with similar compositions.
4.3. Formation of interstitial clusters
The interstitial ux to the grain boundaries will also likely be affected by the presence of a signicant quantity of interstitial clusters, which have not been considered in this work (only diffusion of
single interstitial dumbbells has been included). Stoller and coworkers have shown that up to 50% of the interstitials that are initially produced in energetic displacement cascades are in the form
of interstitial clusters [6466]. Furthermore, numerous molecular
dynamics simulations have shown that interstitial clusters in bcc
Fe and dilute Fe-based alloys are highly mobile, with the largest
interstitial clusters having an a/2 h1 1 1i Burgers vector and
migrating in one dimension along their glide cylinder [6770]. Recent work of Terentyev et al. [71] shows that, depending on temperature, the self-interstitial clusters in Fe can get trapped and
become immobile for a time period of the order of few microseconds. However, they will eventually migrate through 1D fast glide
along h1 1 1i via unfaulting. Osetsky and co-workers have shown
that in Cu, for interstitial clusters up to about size 50, they are
highly mobile and migrate in one dimension, [69,70] Wong and
co-workers have recently investigated the effect of small interstitial clusters on Cr migration [72]. Nonetheless, a full understanding
of the behavior of such interstitial clusters in real engineering Fe
Cr based alloys, and in particular an understanding of the magnitude of the interstitial cluster uxes to grain boundaries and their
associated coupling to solute migration, remain outstanding
questions.
To summarize the discussion of effects of various microstructural features, we point out that variation in sink biases and their
distribution, and the formation of interstitial clusters could lead
to changes in the cancellation between two large intrinsic RIS tendencies of interstitial and vacancy ux and may help explain the
discrepancies in Cr RIS proles observed experimentally. The effects of these microstructural features could occur even in the absence of any precipitates or effects of other minor alloying
elements.
Based on the calculated D, the RIS prole for Ni is expected to
be temperature dependent. Except for high temperatures, the vacancy tracer diffusion coefcient of Ni is higher than that for Fe
(Fig. 4b). Hence, in the FeNi system Ni is likely to deplete at the
grain boundary by the vacancy diffusion mechanism. The interstitial tracer diffusion coefcient of Fe is about two orders of magnitude higher than Ni, and thus Ni is also expected to deplete at grain
boundaries by an interstitial transport mechanism. Our prediction
of depletion of Ni at grain boundaries in irradiated FeNi binary alloys agrees well with some experimental measurements [13] but
not others [12].

A comparison of the diffusion behavior of the fcc Ni-based Ni


CrFe alloys [16] and the current bcc Fe-based FeCr and FeNi alloys shows that relatively weak vacancysolute binding exits in
both crystal structures. However, signicant differences have been
observed in binding between the Cr and interstitial dumbbell between these two alloys. For example, in the bcc alloy the binding
energy between Cr in an FeCr mixed dumbbell (relative to Cr on
the lattice and an FeFe dumbbell) is 0.059 eV [34,55] compared
to the much stronger binding energy of 0.92 eV for CrCr dumbbell and 0.49 eV for a NiCr mixed dumbbell in an fcc alloy [36].
Thus, compared to an FeCr dumbbell in the bcc Fe-based alloys, in
Ni-based fcc alloys it is much more difcult to completely dissociate a CrCr or NiCr dumbbell during diffusion. Stabilization of Cr
in the form of an interstitial dumbbell may increase the annihilation of these interstitials through vacancy recombination mechanism, potentially reducing the overall defect ux during
diffusion. Another point of interest is the vacancy drag. We nd
that vacancy drag is unlikely at all temperatures of interest in
the bcc FeCr and FeNi alloys. However, this may not be the case
in the fcc NiCr system where vacancy drag is expected at lower
temperatures (T < 460 K). However it should be pointed out that
the possibility of drag in the fcc case does not originate from vacancysolute binding energy but rather due to the fact that the
w1 rate is higher than the w3 rate, which makes the vacancy prefer
circling around the Cr atom over dissociation.
A comparison of the vacancy tracer diffusion coefcients shows
that in both bcc and fcc structures, Cr is the fastest diffuser. The
ordering of diffusivity of the species is primarily due to lower
migration energy of Cr than Fe and Ni in both the structures. However, unlike the fcc alloy, the vacancy migration barriers are rather
weakly species dependent in the bcc alloys. Hence, vacancy mediated D for all the three elements are more similar in bcc as compared to fcc alloys.
Similar to the vacancy tracer diffusion, we nd that interstitial
tracer diffusion coefcients for Cr are the highest in both bcc and
fcc structures. More importantly for RIS, the interstitial tracer diffusion coefcient of Cr with respect to the interstitial tracer diffusion coefcient of the solvent is much higher in a Ni-based fcc alloy
than in an Fe-based bcc alloy. For example, at T = 400 K, D(Cr)/
D(Fe)  16 in bcc alloy whereas D(Cr)/D(Ni) 40 in fcc alloy.
Thus the expected tendency for Cr to segregate though interstitial
mechanism is higher in a fcc alloy compared to a bcc alloy. It is suggested by the above D ratios that Cr RIS in both bcc and fcc structures involves a balance between transport of Cr though interstitial
and vacancy mechanisms. However, measured RIS proles exhibit
both enrichment and depletion in bcc alloys only, while the fcc alloys consistently show depletion of Cr at the grain boundaries. This
Cr depletion in fcc alloys is generally assumed to be due to an inverse Kirkendall mechanism, suggesting that the transport of Cr
through the interstitial mechanism may be suppressed. At this
point, the reason for suppression of transport of Cr though interstitial mechanism in fcc alloys is not clear and requires further
investigation.
The RIS trends predicted for interstitial Ni ux are quite different between bcc and fcc alloys. Ni diffuses much slower than Fe (at
least two orders of magnitude) in bcc alloy through the interstitial
mechanism. This can be compared to the fcc alloy where Ni and Fe
have a similar magnitude interstitial tracer diffusion coefcient.
Thus Ni tends to segregate through the interstitial mechanism
more in a bcc alloy compared to its fcc counterpart.
In summary, we have identied the contributions of vacancy
and interstitial ux in determining the RIS prole in dilute bcc
FeCr and FeNi alloys. In FeCr alloys these two contributions
show opposite tendencies, while in FeNi alloys the contributions
of vacancy and interstitial ux show similar trends in determining
RIS prole. The present models do not include effects of non-dilute

11

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

solute, other sinks, or defect clusters. The predicted balance between vacancy and interstitial tendencies in Cr RIS suggests that
the exact RIS prole may be inuenced by the spatial distribution
of biased absorbing sinks and traps, which can alter the balance between vacancy and interstitial ux. Such effects may help explain
the variations seen in Cr RIS proles in bcc FeCr based alloys. Further work is needed to understand the coupling of grain boundary
RIS to other sinks and to extend the analysis presented here to include non-dilute composition effects.

Acknowledgements
We gratefully acknowledge nancial support from the DOE Nuclear Engineering Research Initiative Consortium Program (NERIC), Award Number DE-FG07-07ID14894, DOE Nuclear Engineering
Research Initiative Grant Award DE-FG07-07ID14828 at UC Berkeley, and computing support from the National Science Foundation
(NSF) National Center for Supercomputing Applications (NCSA),
Award Number DMR060007. MA acknowledges support by the
US Department of Energy, Ofce of Nuclear Energy, contract number DR-FG07-071D14893.

5. Conclusions
In summary, we have calculated tracer diffusion coefcients and
the full set of phenomenological coefcients (Onsager matrix) for
diffusion of Fe, Cr and Ni atoms within an Fe matrix by both vacancy
and interstitial mediated transport. The vacancy tracer diffusion
coefcients were compared with experimental measurements.
We found that
1. Calculated D for vacancy mediated diffusion shows that ab-initio accurately predicts the relative relationship in diffusivity of
Cr and Ni with respect to Fe atoms as well as the activation
energy for diffusion. However the absolute magnitudes of the
calculated tracer diffusion coefcients are signicantly underpredicted as compared to the experimental tracer diffusion
coefcients. The discrepancy between experimental and ab-initio data may arise from a combination of several factors, as discussed in Section 3.1.
2. Based on the calculated tracer diffusion coefcients, Cr is the
fastest diffusing species at all temperatures both for vacancy
and interstitial diffusion, primarily because of low migration
barriers of Cr atoms. Fe diffuses much faster than Ni at all temperatures through the interstitial mechanism, while Ni diffuses
faster than Fe through vacancy mediated diffusion, except for
temperatures greater than 1000 K.
3. The calculated D for vacancy and interstitial mediated diffusion
suggest that the observed RIS behavior of Cr is a balance
between two opposing tendencies, whose exact balance depend
on the presence and spatial distribution of biased absorbing
sinks and traps, as well as the concentration of the solute. This
result suggests a possible explanation for the variation in experimentally observed RIS proles in FeCr alloys. In the case of
FeNi alloys, it is predicted that Ni will deplete at the grain
boundaries, which agrees well with some, but not all, experimental observations.
4. RIS behavior differs signicantly between Fe-based bcc alloys
and Ni-based fcc alloys. The species dependence of the diffusion
constants is observed to be signicantly greater for both vacancy
and interstitial mechanisms in fcc than bcc alloys, consistent
with the greater RIS tendency observed in fcc materials. Furthermore, a vacancy drag mechanism is unlikely to occur in either
FeNi or FeCr bcc alloys at all temperatures of interest, in contrast to the low-temperature vacancy drag predicted for fcc.
5. Using a multi-frequency approach, for the rst time a complete
set of Onsager coefcients has been calculated both for vacancy
and interstitial mediated diffusion in the bcc FeCr and FeNi
systems. A negative value of LAB is predicted in FeCr alloy for
vacancy mediated diffusion, a result which is not possible to
obtain either from the Darken (LAB = 0) or Manning (LAB P 0)
approaches to estimate the values of the L coefcients from
the D diffusivities. It has been shown that the coupling
between Fe and solute is substantial for interstitial mediated
diffusion; hence considering only the diagonal elements of Lij
may lead to signicant error in diffusion ux prediction in
dilute bcc FeNi and FeCr alloys.

Appendix A. Calculating vacancy Onsager coefcients from


atomic jump rates.
Below we give the expressions for the Onsager coefcients for
vacancy mediated diffusion in terms of quantities that are calculated with ab-initio methods in the text. These equations are based
on LeClaires model and from Ref. [37]

LAA

LAB

LBB

a2 w0 N expDHfv
1 bcB
kT

A:1

2a2 w2 expDHfv
kT


8u  7B6  32u2  1B5 nB
 expDHmig

;
v
4uu 1B7
w

A:2

a2 w2 expDHfv
kT


8u 7B6  32uu 1B5 nB
 expDHmig

v
4uu 1B7
w

A:3

where B5 8u2 77u 155, B6 32u3 324u2 810u 475,

2v 8u 7B6  32uu 1B5


;
4uu 1
!
8weff
w 1 50 eff4 expDHfv
w3

B7

weff
4
;
w0

w2
weff
3


1
6u 2v u  1  7
45 12u
b 58
w
u


8u  7B6
B6
 16u  1

2u 1B7
B7
In the above equations, we take A = Fe, B = Cr or Ni. b is the solute enhancement factor. Lij and w represent the Onsager coefcients and jump frequency, respectively. The expressions for
jump frequencies are given in Eqs. (3) and (4). The lattice parameter a used for pure iron is 0.286 nm. The Onsager coefcients were
determined assuming a mole fraction cB = 0.01. In Eq. (A.2) and Eq.
(A.3) nB = NcB, where N = 8.5  1022 cm3 is the number of atoms/
volume. In our calculations, we assume w = 1. We should mention
that the effect of ferromagnetic transition on the jump barriers has
been ignored in calculating u and v. Further, the entropy terms
were set to zero in calculating all the Onsager coefcients for vacancy mediated diffusion, as discussed in the text.
Appendix B. Effect of ferromagnetic transition on activation
energy for vacancy diffusion.
In this section, we present the expressions to consider the
change in the vacancy activation energy in the Fe ferromagnetic
state. The modied activation energy is later used in calculating
the vacancy tracer diffusion coefcients and Onsager coefcients.

12

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

We partitioned the tracer diffusion expressions in Eqs. (7) and


(9) into pre-exponential factors and activation energies to match
the Arrhenius form

DBA

DB0;A

DQ BA
exp 
kT

!
B:1

Fe
mig
2
Thus for pure Fe, DFe
DHfv and for
0;Fe a f0 m0 and DQ Fe DH v

solute tracer

Ni;Cr
f
2
DHmig
diffusion DNi;Cr
v DHv
0;Fe a f2 0 and DQ Fe
Ni;Cr
DFe
0;Fe and D0;Fe are the pre-exponential factors of tra-

DHbind
where
v
cer diffusion coefcients of pure iron and the solutes respectively
while DQ and DH are the corresponding activation energy and enthalpy. Enthalpy of vacancy formation, migration and solutevacancy binding are denoted using superscripts f, mig, and bind,
respectively. The enthalpy of vacancy formation is given in the text
while the enthalpy of vacancy migration and biding enthalpy are
given in Table 1 and Fig. 3, respectively. The activation energy for
vacancy diffusion though the transition temperature both for pure
Fe as well as in alloys is expressed as [73].

DQ F T DQ P 1 asT2

B:2

where QP is the activation energy in the paramagnetic state. s(T) is


the ratio of the spontaneous magnetization of pure iron at a given
temperature to the spontaneous magnetization at 0 K [42]. The constant a takes into account the change in formation and binding
energies due to magnetic transformation. The value of a is obtained
empirically in Refs. [43,45] for pure Fe and FeCr system to be 0.156
and 0.133 respectively. For the FeNi system a has been estimated
to be 0.12 obtained by tting to the D(Ni) measured by Borg and
Lai [44]. a can also be estimated using the relation given below [45]:

a 0:053DM12 0:163

B:3

where DM12 is the change in local magnetization of iron atoms


summed over the rst and second shells. The value of DM12 was
determined from ab-initio magnetic calculations in Ref. [74]. Using
this approach a is estimated to be 0.219. However a = 0.219 predicts that for some T < 1043 K, D(Ni) < D(Fe) which contradicts
experimental observation of Hirano et al. [75] that D(Ni) > D(Fe),
at all temperatures.

Appendix C. Calculating interstitial Onsager coefcients based


on jump rates.
By a process similar to that described in Appendix A, it is possible to determine Onsager coefcients for interstitial diffusion from
atomistic migration barriers. Utilizing the model developed by V.
Barbe and M. Nastar [19] for interstitial diffusion in dilute bcc metals, we have calculated transport coefcients for a dilute binary alloy of Cr and Ni in Fe for the interstitial mechanism. The primary
expressions for calculating transport coefcients are:

4kb T
2 s3 W 3  2 s4 W 4 2
LAA W A  6
2
na
2 s3 W 3 32 s4 W 4


8w2 2 s4 W 4 2 s3 W 3  2 s4 W 4
3 2s1 W 1 2 s2  hwW 2

C:1

C:4
w

2 s3 W 3
32 s4 W 4 2 s3 W 3

C:5

1 s2 2 2
1 s1

C:6

The central quantities in these expressions are the Wi, which are
dened as the jump probability of the atomistic jump type i. This
model considers eight interstitial congurations, each with both
a translational migration (denoted sixi) and a translation combined with a rotation (xi), for a total of 16 jump types. The jump
probability Wi is dened as follows:

W i c i xi

C:7

where xi is the jump frequency of jump type i, and ci is probability


of nding an interstitial in the initial conguration corresponding to
jump type i.
Inspection of Fig. 5 reveals that there are in fact only four distinct initial congurations: an unmixed dumbbell without a solute
on a nearest neighbor site (denoted cAA), a mixed dumbbell (cAB), an
unmixed dumbbell with a solute atom on a target-type nearest
neighbor site (cAA||B), and an unmixed dumbbell with a solute atom
on a non-target-type nearest neighbor site (cAA\B). These coefcients are calculated to the rst order in c0B as follows:

cAA 1  c0B P 
cAB c0B PAB 

cI
6

cI
6

cAAjjB c0B PAAjjB 

C:2

4kb T
2 s1 W 1 1 2s1 W 1 32 s2  hwW 2
LBB
na2
3 s1 W 1 2 s2  hwW 2

C:3

C:8
C:9

cI
6

C:10

cI
6

C:11

cAA?B c0B P AA?B 

where c0B is the concentration of solute B farther than 1 nn distance


from an interstitial dumbbell and cI is the concentration of interstitials. In a real alloy, where cB  cI, the simplication that c0B = cB can
be made. The quantities PAB, PAA||B, and PAA\B are Boltzmann-type
probability weights calculated as follows:



DH3  DH2 DH5  DH4
PAB exp 
kb T

C:12



DH 5  DH 4
PAAjjB exp 
kb T

C:13



DH 7  DH 6
PAA?B exp 
kb T

C:14

The quantities DHi are the migration barriers for translation


rotation jump type i. The quantity P is a global probability weight
introduced for convenience, dened as

P 2PAB 4P AAkB 4PAA?B :

4kb T
42 s2 2 W 1 4wW 4
4kb T
LAB

LBA
na2
3 2s1 W 1 2 s2  hwW 2
na2

where the quantities WA, w, and h are dened as:

W A 32 s0 W 0 32 s2 W 2 182 s5 c0B W 5 242 s7 c0B W 7

C:15

The jump frequencies xi are calculated with the expression in


Eq. (3). Finally, the coefcients si are dened simply as the ratio
of the translation frequency of jump type i to the translationrotation frequency:

si

s i xi
xi

C:16

13

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114

Appendix D. Qualitative relation of RIS and kinetic parameters.


Simple intuitive arguments based on tracer or chemical diffusion coefcients are often used to qualitatively estimate RIS trends.
For example, if species B diffuses faster than species A by interstitial ux then interstitial ux is likely to cause enrichment of B at a
sink. Here we put these intuitive arguments on more quantitative
footing in order to assess the role of off-diagonal Onsager coefcients in the arguments. This more quantitative analysis is necessary as the off-diagonal Onsager coefcients for the interstitial
ux were found to be as large as the diagonal elements in this
study. For concreteness and simplicity, we focus only on interstitial
ux, although a similar approach could incorporate vacancy diffusion as well.
For an interstitial mediated diffusion the eventual steady-state
concentration prole at a sink of species under irradiation depends
on (a) the absolute interstitial defect ux; (b) the ux ratio of solute compared to the solvent by interstitial diffusion mechanisms.
To predict the prole rigorously requires a full simulation of the
coupled ux equations. To provide an estimate of the trends in
RIS prole, in this appendix we present the change in the factors
(a) and (b) immediately after a radiation damage cascade formation on a plane far away from the grain boundary. We consider
only a one-dimensional problem. For the sake of simplicity we assume the vacancies are frozen and thus the diffusion is dominantly
through interstitial mechanism. We particularly focus on the effect
of including or ignoring cross-terms LAB on the contribution of
interstitial ux.
The ux expression (Eq. (2)) can be rewritten as

Ji 

N
X
j1

Lij

N
@ lj X
Lij Hjk ni

@x
j1

D:1

where l is the chemical potential, ni is the number of atoms of spe@l


cies i per unit volume, Hjk @n j , is the thermodynamic factor. For an
k
ideal solution, the chemical potential is expressed as
0
li li T; P kB T lnci and thermodynamic factor is expressed as

Hij kcBiT N knBiT for i j


0 for i j

D:2

Combining Eqs. (D.1) and (D.2), the explicit ux for interstitial


mediated diffusion can be written as



nA
nB
nI
J A NkB T LAA
LAB
LAI
nA
nB
nI

D:3

p
p1
mnB mnB

nA Dx
Dx
the cascade formation,

nA
nA

mnB . Thus for the plane p just after

nB

D:7

nB

Using the relations between the Onsager coefcients as well as


Eqs. (D.3), (D.4), and (D.7) the ratio of JB/JA is expressed as



nA
nB
nI
nA
nB
LAA
J B =J A LAB
LBB
LAB LBB
LAB
nA
nB
nI
nA
nB

nI
LAA LAB
nI





nI nB
nB nI
LAA LAB

LAB LBB
nI
nB
nB
nI
LAB LBB =LAA LAB
For the case when LAB  LAA, LBB the ux ratio JB/JA can be rewritten as

J B =J A

LBB nB DB
f
LAA nA DA

D:8

To introduce the tracer diffusion coefcients we have used the


fact that the diagonal Onsager coefcients are related to the tracer
diffusion coefcients through LBB knBBT DB and LAA fknAT DA , where f
B
is the correlation coefcient [20]. It is clear from Eq. (D.8) that
the ratio of ux JB/JA is proportional to the concentration ratio
times the ratio of tracer diffusion coefcients. RIS will occur when
the ux ratio deviates from the nB/nA ratio. If we assume f  1 then
the extent of RIS will depend on the ratio of DB /DA . Thus for the
case when solute and solvent uxes are dominated by the diagonal
Onsager coefcients, DB /DA can be used for qualitative RIS prediction. However, it can be seen from Fig. 6 that both for dilute Ni
and Cr the off-diagonal Onsager coefcient (LAB) is of the same order of magnitude as diagonal coefcient (LBB). Thus, for the Osager
coefcient values determined here, it is clear from Eq. (D.8) that
ignoring LAB by predicting RIS purely based on the ratio of tracer
diffusion coefcients will cause error in RIS prediction in dilute
FeCr and FeNi alloys. In order to estimate the error introduced
by ignoring LAB in our RIS prediction we calculate ux ratios (JB/
JA) and JI for the plane p far away from the grain boundary, for
the case with and without the LAB term. We assume LAB LBB and
LAA kLBB where k  100, which are similar values to what we nd
with A = Fe and B = Cr. The effect of LAB is estimated from



k 1
1 1
k

2
2k
2 2k

fJ B =J A WithoutLAB g=fJ B =J A WithLAB g

k1



nA
nB
nI
LBB
LBI
J B NkB T LAB
nA
nB
nI

D:4



nA
nB
nI
LBI
LII
J I NkB T LAI
nA
nB
nI

D:5

Using the relation JA + JB = JI it is possible to show

For k  1, it is clear, that neglecting the LAB term changes the


ux ratio of solute and solvent by a factor of 2.
Similarly, from Eqs. (D.5) and (D.7) it can be shown that for
large k there is an insignicant change in the total interstitial defect ux by ignoring the LAB term as given below.

J I WithoutLAB =J I WithLAB

LAI LAA LAB and LBI LAB LBB and LII LAI LBI
LAA LAB LBA LBB

p
p1
nA nA

k 1
 1
k 3

D:6

We consider a plane (p) far away from the grain boundary at


time immediately after a radiation damage cascade creates a local
change in interstitial concentration. We assume that a total of Dn
interstitials are formed, comprised of A and B atoms in the same ratio as the local concentration around plane (p). We write the composition relationship in plane p as npA mnpB (and we assume
nearby planes have the same composition). For example, for
cB = 0.01 we get m = 99. The gradient of A between the planes p
and p-1 can be expressed as

References
[1] Z. Lu, R.G. Faulkner, G. Was, B.D. Wirth, Scr. Mater. 58 (2008) 878.
[2] G. Gupta, Z. Jiao, A.N. Ham, J.T. Busby, G.S. Was, J. Nucl. Mater. 351 (2006) 162.
[3] G.S. Was, Fundamentals of Radiation Materials Science, Springer, New York,
2007.
[4] T.R. Allen, J.T. Busby, G.S. Was, E.A. Kenik, J. Nucl. Mater. 255 (1998) 44.
[5] T.R. Allen, G.S. Was, Acta Mater. 46 (1998) 3679.
[6] R.E. Clausing, L. Heatherly, R.G. Faulkner, A.F. Rowcliffe, K. Farrell, J. Nucl.
Mater. 141 (1986) 978.
[7] Z. Lu, R.G. Faulkner, N. Sakaguchi, H. Kinoshita, H. Takahashi, P.E.J. Flewitt, J.
Nucl. Mater. 351 (2006) 155.

14
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]

[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]

S. Choudhury et al. / Journal of Nuclear Materials 411 (2011) 114


S. Ohnuki, H. Takahashi, T. Takeyama, J. Nucl. Mater. 103 (1981) 1121.
H. Takahashi, S. Ohnuki, T. Takeyama, H. Kayano, J. Nucl. Mater. 96 (1981) 233.
J.L. Brimhall, D.R. Baer, R.H. Jones, Nucl. Mater. 122 (1984) 196.
R.L. Klueh, D.S. Gelles, S. Jitsukawa, A. Kimura, G.R. Odette, Nucl. Mater. 307
(2002) 455.
E.A. Little, L.P. Stoter, Special Tech. Publ. 782 (1982) 207.
A.A. Hosseini, I.P. Jones, Phys. Status Solidi A 113 (1989) 57 (and references
therein).
A. Van der Ven, G. Ceder, Phys. Rev. Lett. 94 (2005) 045901.
A. Van der Ven, J.C. Thomas, Q.C. Xu, B. Swoboda, D. Morgan, Phys. Rev. B 78
(2008) 104306.
J.D. Tucker, R. Najafabadi, T.R. Allen, D. Morgan, J. Nucl. Mater. 405 (2009) 216.
K.L. Wong, H.J. Lee, J.Y. Shim, B. Sadigh, B.D. Wirth, J. Nucl. Mater. 386388
(2009) 227.
A.D. Leclaire, J. Nucl. Mater. 6967 (1978) 70.
V. Barbe, M. Nastar, Philos. Mag. 87 (2007) 1649.
A.R. Allnatt, A.B. Lidiard, Atomic Transport in Solids, Cambridge University
Press, Cambridge, 2003.
L.S. Darken, Trans. Am. Inst. Min. (Metall.) Eng. 175 (1948) 184.
J.R. Manning, Metal. Trans. 1 (1970) 499.
W.F. Gale, T.C. Totemeier, Smithells Metals Reference Book, Elsevier, San
Francisco, 2004.
J. Askill, Tracer Diffusion Data for Metals, Alloys and Simple Oxides, IFI/
Plenum, New York, 1970.
M. Nastar, V. Barbe, A Self-Consistent Mean Field Theory for Diffusion in Alloys,
Meeting on Atomic Transport and Defect Phenomena in Solids, Guildford,
England, 2006.
G. Kresse, J. Furthmuller, Phys. Rev. B 54 (1996) 11169.
G. Kresse, J. Hafner, Phys. Rev. B 47 (1993) 558.
P.E. Blochl, Phys. Rev. B 50 (1994) 17953.
J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865.
H.J. Monkhorst, J.D. Pack, Phys. Rev. B 13 (1976) 5188.
E. Vincent, C.S. Becquart, C. Domain, Instr. Meth. Phys. Res. B 228 (2005) 137.
P. Olsson, J. Nucl. Mater. 386 (2009) 86.
C. Domain, C.S. Becquart, Phys. Rev. B 65 (2002) 024103.
P. Olsson, C. Domain, J. Wallenius, Phys. Rev. B 75 (2007) 014110.
C.C. Fu, J. Dalla Torre, F. Willaime, J.L. Bocquet, A. Barbu, Nat. Mater. 4 (2005)
68.
J.D. Tucker, Thesis, University of Wisconsin, Madison, 2008.
Y. Serruys, G. Brebec, Philos. Mag. A 46 (1982) 661.
J.R. Manning, Phys. Rev. 139 (1965) A126.
J. Bocquet, G. Brebec, Y. Limoge, Diffusion in Metals and Alloys, Elsevier Science
BV, Amsterdam, The Netherlands, 1996.
J. Wallenius, P. Olsson, L. Malerba, D. Terentyev, Inst. Meth. Phys. Res. Sec. B
255 (2007) 68.
M.J. Jones, A.D. Leclaire, Philos. Mag. 26 (1972) 1191.

[42] J. Crangle, G.M. Goodman, Roy. Soc. London Ser. a-Math. Phys. Sci. 321 (1971)
477.
[43] H. Nitta, Y. Iijima, Philos. Mag. Lett. 85 (2005) 543.
[44] R.J. Borg, D.Y.F. Lai, Acta Metall. 11 (1963) 861.
[45] S. Takemoto, H. Nitta, Y. Iijima, Y. Yamazaki, Philos. Mag. 87 (2007) 1619.
[46] G. Hettich, H. Mehrer, K. Maier, Scr. Metall. 11 (1977) 795.
[47] Y. Iijima, K. Kimura, K. Hirano, Acta Metall. 36 (1988) 2811.
[48] C.G. Lee, Y. Iijima, T. Hiratani, K. Hirano, Mater. Trans. JIM 31 (1990) 255.
[49] J. Wallenius, P. Olsson, C. Lagerstedt, N. Sandberg, R. Chakarova, V. Pontikis,
Phys. Rev. B 69 (2004) 094103.
[50] S. Huang, D.L. Worthington, M. Asta, V. Ozolins, G. Ghosh, P.K. Liaw, Acta
Mater. 58 (2009) 1982.
[51] C. Domain, J. Nucl. Mater. 351 (2006) 1.
[52] C.C. Fu, F. Willaime, P. Ordejon, Phys. Rev. Lett. 92 (2004) 175503.
[53] D. Nguyen-Manh, A.P. Horseld, S.L. Dudarev, Phys. Rev. B 73 (2006) 020101.
[54] E. Vincent, C.S. Becquart, C. Domain, J. Nucl. Mater. 359 (2006) 227.
[55] T.P.C. Klaver, P. Olsson, M.W. Finnis, Phys. Rev. B 76 (2007) 214110.
[56] R.G. Faulkner, S.H. Song, P.E.J. Flewitt, M. Victoria, P. Marmy, J. Nucl. Mater. 255
(1998) 189.
[57] A.L. Nikolaev, V.L. Arbuzov, A.E. Davletshin, J. Phys.: Condens. Matter 9 (1997)
4385.
[58] D. Terentyev, P. Olsson, L. Malerba, J. Nucl. Mater. 386 (2009) 140.
[59] G.R. Odette, M.J. Alinger, B.D. Wirth, Ann. Rev. Mater. Res. 38 (2008) 471.
[60] W.G. Wolfer, L.K. Mansur, J. Nucl. Mater. 91 (1980) 265.
[61] T. Okita, W.G. Wolfer, J. Nucl. Mater. 327 (2004) 130.
[62] C.H. Zhang, J. Jang, H.D. Cho, Y.T. Yang, J. Nucl. Mater. 386 (2009) 457.
[63] T.R. Allen, L. Tan, J. Gan, G. Gupta, G.S. Was, E.A. Kenik, S. Shutthanandan, S.
Thevuthasan, J. Nucl. Mater. 351 (2006) 174.
[64] W.J. Phythian, R.E. Stoller, A.J.E. Foreman, A.F. Calder, D.J. Bacon, J. Nucl. Mater.
223 (1995) 245.
[65] R.E. Stoller, G.R. Odette, B.D. Wirth, J. Nucl. Mater. 251 (1997) 49.
[66] R.E. Stoller, J. Nucl. Mater. 276 (2000) 22.
[67] B.D. Wirth, G.R. Odette, D. Maroudas, G.E. Lucas, J. Nucl. Mater. 276 (2000)
33.
[68] N. Soneda, T.D. De la Rubia, Philos. Mag. A 81 (2001) 331.
[69] Y.N. Osetsky, D.J. Bacon, A. Serra, B.N. Singh, S.I. Golubov, Philos. Mag. 83
(2003) 61.
[70] Y.N. Osetsky, D.J. Bacon, B.N. Singh, B. Wirth, J. Nucl. Mater. 307 (2002) 852.
[71] D.A. Terentyev, T.P.C. Klaver, P. Olsson, M.C. Marinica, F. Willaime, C. Domain,
L. Malerba, Phys. Rev. Lett. 100 (2008) 14.
[72] K.L. Wong, J.H. Shim, B.D. Wirth, J. Nucl. Mater. 367 (2007) 276.
[73] L. Ruch, D.R. Sain, H.L. Yeh, L.A. Girifalco, J. Phys. Chem. Sol. 37 (1976) 649.
[74] B. Drittler, N. Stefanou, S. Blugel, R. Zeller, P.H. Dederichs, Phys. Rev. B 40
(1989) 8203.
[75] K. Hirano, M. Cohen, B.L. Averbach, Acta Metall. 9 (1961) 440.
[76] R.A. Johnson, Phys. Rev. Lett. 134 (1964) A1329.

Вам также может понравиться