You are on page 1of 7

Journal of Nuclear Materials 460 (2015) 37–43

Contents lists available at ScienceDirect

Journal of Nuclear Materials


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

Interaction of h1 0 0i dislocation loops with dislocations studied


by dislocation dynamics in a-iron
X.J. Shi a,⇑, L. Dupuy a, B. Devincre b, D. Terentyev c, L. Vincent a
a
CEA, DEN, SRMA, F-91191 Gif-sur-Yvette, France
b
Laboratoire d’Etude des Microstructures, CNRS-ONERA, 29 av. de la Division Leclerc, 92322 Châtillon Cedex, France
c
SCK–CEN, Nuclear Materials Science Institute, Boeretang 200, B-2400 Mol, Belgium

h i g h l i g h t s

 Interactions between edge dislocations and radiation-induced loops were studied by dislocation dynamics.
 Dislocation dynamics results are directly compared to molecular dynamics results.
 The complex elementary reactions are successfully reproduced.
 The critical shear stress to overcome individual loops if reproduced quantitatively.

a r t i c l e i n f o a b s t r a c t

Article history: Interstitial dislocation loops with Burgers vector of h1 0 0i type are formed in a-iron under neutron or
Received 17 September 2014 heavy ion irradiation. As the density and size of these loops increase with radiation dose and
Accepted 23 January 2015 temperature, these defects are thought to play a key role in hardening and subsequent embrittlement
Available online 3 February 2015
of iron-based steels. The aim of the present work is to study the pinning strength of the loops on mobile
dislocations. Prior to run massive Dislocation Dynamics (DD) simulations involving experimentally
representative array of radiation defects and dislocations, the DD code and its parameterization are val-
idated by comparing the individual loop–dislocation reactions with those obtained from direct atomistic
Molecular Dynamics (MD) simulations. Several loop–dislocation reaction mechanisms are successfully
reproduced as well as the values of the unpinning stress to detach mobile dislocations from the defects.
Ó 2015 Elsevier B.V. All rights reserved.

1. Introduction the remainder have Burgers vector ½h1 1 1i, in good agreement
with the elastic self-energies of loops [6]. We have therefore
In nuclear Pressurized Water Reactors (PWR), special attention focused our study to the interaction between a dislocation and a
is paid to the hardening and embrittlement processes existing in h1 0 0i interstitial loops.
the low-alloyed a-iron steel that constitutes the pressure vessels. Therefore, SIA loops are thought to be elements contributing to
The latter structural component, which cannot be replaced, is the hardening process observed in irradiated material, (i) either as
indeed an essential barrier between the fissile fuel and the a source of internal stresses opposed to dislocation mobility or (ii)
environment. Hence, a better understanding of the mechanical as a source of additional contact reactions pinning mobile disloca-
properties of irradiated a-iron is critical before considering possi- tions and increasing the forest strengthening [7–12].
ble lifetime extension of the nuclear reactors. Even though MD simulations provide basic and necessary infor-
During exposure to fast-neutron irradiation, point defects are mation about the interactions between SIA loops and mobile dislo-
created in displacement cascades and a high proportion of them cation, they have important time scale limitations preventing a
regroup in clusters [1]. In bcc iron, the self-interstitial atoms (SIAs) direct extrapolation of their results at the macroscopic scale. Then,
form platelets that are equivalent to nanometer-scale interstitial a multi-scale modeling strategy must be elaborated to develop
dislocations loops with Burgers vector b parallel to the loop axis constitutive plastic laws with longer-term prediction capability
[2–4]. Transmission electron microscopy study [5] showed that [13]. In the past years, Dislocation Dynamics (DD) simulations
above 300 °C, these loops have mainly Burgers vector h1 0 0i, while proved to be an effective method to proceed with this up-scaling
task. DD simulations are based on the elastic dislocation theory
and can capture the most important features observed at the
⇑ Corresponding author. atomic scale [12,14,15,22]. Furthermore, they can simulate plastic

http://dx.doi.org/10.1016/j.jnucmat.2015.01.061
0022-3115/Ó 2015 Elsevier B.V. All rights reserved.
38 X.J. Shi et al. / Journal of Nuclear Materials 460 (2015) 37–43

deformation in large micrometric volume elements to investigate


the collective properties of microstructures made of many disloca-
tions [16–18]. In practice, DD simulations provide an enhancement
of more than a thousand times in terms of size and timescale when
compared to MD simulations.
Yet, DD simulations require the definition of ‘‘local rules’’ to
account for the dislocation core properties which cannot be repli-
cated with the elastic theory in the same way as it occurs in atom-
istic simulations. For instance, in the present study one needs to
identify and justify a set of DD simulation parameters defining
the reactions that take place when a mobile dislocation contacts
an SIA loop. Hence, prior to run massive DD simulations dedicated
to realistic microstructures with a large density of SIA loops, the
parameters controlling the loop–dislocation reactions in the DD
simulation must be validated. The identification of these rules for
h1 0 0i dislocation loops is the aim of the present paper.
Several MD studies addressing the interaction of mobile dislo-
Fig. 1. Schematic illustration of the simulation condition. The dark plane here
cations with ½h1 1 1i loops were published in the past years
presents the main glide plane for the dislocation which lies along the direction of Y
[7,9,19,20], but not so many for the more complex interaction axis, and arrows touched with dislocation lines present their Burgers vectors. Two
involving h1 0 0i loops [7]. However, at the PWR relevant tempera- blue arrows indicate the direction of applied shear strain. (For interpretation of the
ture range mainly h1 0 0i loops are formed under neutron-irradia- references to colour in this figure legend, the reader is referred to the web version of
tion [21]. Consequently, in the present paper we restricted our this article.)

investigations to the most critical MD results reported in [7], i.e.


the four interactions discussed with some details in terms of mech-
the nomenclature for the interaction geometry defined in [7] was
anisms and strength. The goal is to achieve the parameterization
used in the following. Those configurations are listed in Table 1.
allowing one to reproduce both the interaction mechanisms and
Periodic boundary conditions were applied along the X and Y
critical resolved shear stress as obtained from MD simulations.
directions as in [7]. In agreement with a previous analysis made
The paper is organized as follows. Firstly, the DD simulation
in [27], the dimension of the simulated volume in the glide direc-
model and computational scheme is briefly presented. The results
tion (X) was considered large enough to neglect the long-range
of the DD simulation results are then reported and systematically
interactions with the image dislocations associated to the use of
compared with the relevant MD results. Finally, the critical DD
periodic boundary conditions. Tests have been made to verify the
parameters controlling the loop–dislocation interaction are identi-
validity of such simplification.
fied and discussed.
Acknowledging that little is known about the dislocation mobil-
ity in the numerous slip systems observed in a-iron at high tem-
peratures, we used simple linear over-damped viscous laws with
2. The DD simulation model the form:

hseff  s0s ibs


The DD simulation code NUMODIS (e.g. [22]) was used in the vs ¼ ð1Þ
Bs
present study. This code is similar in many aspects to the PARADIS
code [23,24]. It uses a nodal representation to discretize the dislo- where hi are the McCauley brackets (hui = u, if u > 0 and hui = 0
cation lines into a series of inter-connected linear segments. The otherwise), bs is the dislocation burgers vector, seff is the effective
computation of the internal elastic stress field and corresponding resolved shear stress calculated at the integration points along the
nodal forces, is based on the non-singular isotropic continuum dislocation lines [28], s0 is a constant friction stress which means
elastic theory of dislocations proposed by Cai et al. [25]. It should that the dislocation will move only when se ff > s0S . Bs is a viscous
be mentioned that particular attention was paid to the proper han- coefficient whose value is function of the dislocation systems ‘s’.
dling of the numerous topological changes that occur when dislo- As little exact and consistent information can be found in the liter-
cations interact with complex objects such as dislocation loops. ature to identify the Bs and s0s values [27,29–37], a direct identifica-
In order to compare the results of MD and DD simulations, the tion of these parameters was made to reproduce the results of MD
simulation conditions defined in Ref. [7] (including the initial, the simulations. The results of such identification are reported in Table 2
loading and the boundary conditions) were systematically repro- and will be discussed in Section 4. For the moment, it can be noted
duced with the DD simulations. As illustrated in Fig. 1, the simu- that the dislocation velocity form adopted in the DD simulations
lated crystal orientation is such that its three axes X, Y and Z are does not account for the non-Schmidt effects existing in BCC metals.
1
parallel to the [1 1 1], ½1  2 and ½1 1
 0 directions, respectively. Also, These effects are indeed not expected to influence irradiation-hard-
the dimensions of the DD simulated volume correspond to the vol- ening processes at elevated temperature [20].
ume used in [7], i.e. LX = 30 nm, LY = 41 nm and LZ = 20 nm. The iso- In the following, all computations have been made with a sim-
tropic elastic coefficients used in our simulations were calculated ulation time step equal to 5 1015 s and the imposed strain rate e_ xz
from the elastic constants associated to the interatomic potential was set to 107 s1. Given the applied simulation geometry the
used in [7] and with a procedure first proposed by Scattergood equilibrium steady-state dislocation velocity is 48 m/s.
and Bacon [26]. Hence, l = 62.9 GPa and m = 0.43.
An edge dislocation with burgers vector ½[1 1 1] and glide plane 3. Results
 0Þ is initially set at the center of the simulation cell. A square
ð1 1
SIA loop (with a side of 2.6 nm length) with Burgers vector h1 0 0i The simulation results reported in this section illustrate the
is then positioned in front of the line at a distance of 5 nm. In capacity of the DD simulation model NUMODIS at reproducing
accordance with the calculations reported in [7], different configu- the complex dislocation–loop reactions first studied in [7]. For each
rations of SIA loop have been tested. For the sake of consistency, SIA loop configuration, details on the interaction processes
X.J. Shi et al. / Journal of Nuclear Materials 460 (2015) 37–43 39

Table 1
Definition of the SIA loop configurations considered in the simulations: The first column gives the name for the specific interaction geometry, the second one contains a schematic
description of the edge dislocation–SIA loop interaction, the third and fourth ones denote the Burgers vector of a square loop and orientation of its sides, respectively.

Configuration name Schematic SIA loop Burgers vector direction Side directions of the SIA loop
C2 b [1 0 0]  [0 1 1]
½0 1 1]
bL

C4U
b bL [1 0 0] [0 0 1] [0 1 0]

C5 b [0 1 0] [1 0 0] [0 0 1]

bL

Table 2 (see Figs. 2 and 3b). The upper side of the loop is then converted
List of the velocity coefficients used in the DD simulation for the different slip  1 1 by the following
into a segment with a Burgers vector b = ½½1
systems. For viscous coefficient Bs, the first slip system solution was directly fitted on
the results of [7], and the same order for the second and third systems [30]. The fourth favorable reaction:
system solution was the value from [36] when the last system solution was adjusted
by trial and error tests made with the DD simulations. For friction stress s0s , we used
1 1 
½1 1 1  ½1 0 0 ¼ ½1 1 1 ð2Þ
the values found in [20,31] for the first three systems, and set the last two values by 2 2
trial and error tests made with the DD simulations.  0 1
 plane
This new segment glides down across the loop in a ½1
~
b Burgers Planes Bs (105 Pa s) s0s (MPa) Reference and converts the remainder of the SIA loop to b = 1/2[1 1 1] by the
½h1 1 1i {1 1 0} 8 10 [7,20,31] reaction:
½h1 1 1i {1 1 2} 8 10 [20,30,31]
½h1 1 1i {1 2 3} 8 10 [20,30,31]
1  1
½1 0 0 þ ½1 1 1 ¼ ½1 1 1 ð3Þ
h1 0 0i {1 1 0} 80 300 [36] 2 2
h1 0 0i {1 0 0} 9000 300 Fitted
Finally, the SIA loop is completely incorporated on the mobile
edge dislocation in the form of a super-jog (see Figs. 2 and 3c). This
jogged dislocation is free to glide under a constant stress level
observed and calculation of the passing critical stress (i.e. the about 30 MPa (see Fig. 8). Such value is very close to the solution
applied stress needed for the edge dislocation to move away from found with the DD simulation for a jog-free edge dislocation glid-
the locking configuration made at contact with the SIA loop) are  0Þ plane.
ing in a ð1 1
reported.
3.2. C4U configuration
3.1. C2 configuration
The second tested configuration is the C4U. This configuration is
The C2 configuration is a square SIA loop with a burgers vector made of a square SIA loop with a Burgers vector [1 0 0] and sides
 1 directions. At the begin-
[1 0 0] and sides along the [0 1 1] and ½0 1 oriented along the [0 1 0] and [0 0 1] directions. In our calculation,
ning of the calculations, the center of the SIA loop lies in the edge complied with the conditions studied in [7], the lower side of the
dislocation glide plane (see Table 1). loop is initially located in the edge dislocation glide plane (see
In Figs. 2 and 3, three simulation snapshots captured at the Table 2).
same time in the DD and the MD simulations can be compared. From Fig. 5, we see that the C4U elementary reaction path pre-
We see that the reaction path obtained in the MD simulation is viously obtained with the MD simulation (see Fig. 4) is fairly well
precisely reproduced with the DD simulation. reproduced with the DD simulation model. In both MD and DD cal-
During this reaction, the mobile edge dislocation is first culations, the edge dislocation is first repelled by the SIA loop and
attracted by the C2 loop (see Figs. 2 and 3a) and hence bows bows backwards (see Figs. 4 and 5a). When the applied stress
toward the loop. Such elastic interaction justifies the rapid stress increases enough, the mobile dislocation line comes in contact
decrease observed in the stress–strain curve (see point r in with the nearest corner at the bottom side of the SIA loop. Then,
Fig. 8). This relaxation is a consequence of using constant strain the edge dislocation and the SIA loop lines start to react according
rate condition in the simulations. Then, the C2 SIA loop starts to to Eq. (2). This reaction converts the whole bottom side and a part
glide down until its upper side contacts the edge dislocation line of the left side of the loop to a segment with a Burgers vector

Fig. 2. Snapshot of the C2 reaction path in three stages as depicted with MD simulation. (Reprinted from [7] with permission of Elsevier.)
40 X.J. Shi et al. / Journal of Nuclear Materials 460 (2015) 37–43

Fig. 3. Snapshot of three different stages of the C2 elementary reaction found with the DD simulation. The time corresponding to image (a), (b) and (c) is 0.003, 0.004 and
0.04 ns, respectively.

Fig. 4. Snapshot of the C4U reaction path in four stages as depicted with MD simulation. (Reprinted from [7] with permission of Elsevier.)

Fig. 5. Snapshot of four different stages of the C4U elementary reaction found with the DD simulation. The time corresponding to image (a), (b), (c) and (d) is 0.25, 0.297, 0.35
and 0.36 ns, respectively.

b = ½½1 1 1 (see Fig. 5b and c). It should nevertheless be pointed out Again, the reaction path previously found with MD simulations
that the cross-slip of the dislocation observed in Fig. 4c in MD sim- (see Fig. 6) is successfully reproduced with the DD simulation (see
ulation could not be observed in DD simulation as this mechanism Fig. 7). More precisely, in both cases we see that the edge disloca-
is not currently available in our code. As the stress keeps on tion line is first repelled by the loop as the two defects are elasti-
increasing, the left side of the edge dislocation continues to bow cally repulsive. When increasing the applied stress, the
forward and finally breaks away after meeting the opposite SIA dislocation moves forward and is suddenly attracted and pushed
loop line section pinned at a loop corner. Eventually, two complex forward when touching the middle part of the SIA loop (see
dislocation loops with Burgers vector b = ½½1  1 1 and b = ½[1 1 1] Fig. 7a). Like in the C2 configuration, this non-contact interaction
intersected at 90° are left behind with the pre-existing common leads to a decrease of the applied stress on the stress–strain curve
[1 0 0] segments in the edge dislocation glide plane. The critical as reported in Fig. 8 (see point s). Then, the upper part of the
unpinning stress calculated with the DD simulation for this reac- intersected loop (Fig. 7a) reacts with the incoming dislocation to
tion is about 130 MPa and compares well with the MD result form a glissile dislocation with b = ½½1 1  1, which can therefore
(see Fig. 8). move downward along its prismatic cylinder (Fig. 7b):
1 1 
½1 1 1  ½0 1 0 ¼ ½1 1 1 ð4Þ
2 2
3.3. C5 Configuration This reaction ends when the remaining [0 1 0] segments of the
loop glide down to convert the newly formed ½½1 1  1 segments
The last tested configuration is C5. This configuration is a square to b = ½[1 1 1] following the reaction:
SIA loop with a Burgers vector [0 1 0] and its sides oriented along
the [1 0 0] and [0 0 1] directions (see Figs. 6 and 7). Like in the C2
1  1
½1 1 1 þ ½0 1 0 ¼ ½1 1 1 ð5Þ
configuration, the SIA loop center is initially located in the edge 2 2
dislocation glide plane (see Table 1).

Fig. 6. Snapshot of the C5 reaction path in three stages as depicted with MD simulation. (Reprinted from [7] with permission of Elsevier.)
X.J. Shi et al. / Journal of Nuclear Materials 460 (2015) 37–43 41

Fig. 7. Snapshot of the C5 reaction path in three stages as reproduced with the DD simulation, the time corresponding to image (a), (b), and (c) is 0.14, 0.32 and 0.49 ns,
respectively.

200 that the dislocation line tension computed in DD simulation based


on isotropic elasticity formula underestimates the one in the atom-
MD: C2 DD: C2
istic calculations. Nevertheless, as reported in Table 3, the pinning
150
MD: C4U DD: C4U strength values calculated for all SIA loops configurations with the
different simulations are always very close to the MD result. In
Stress (MPa)

100 MD: C5 DD: C5 average, the difference between the two simulation methods is
about 18%.
50

4. Remarks and observations


0 ĸ

The definition of the dislocation mobility in each slip systems is


ķ
-50 an essential input of the DD simulations. Unfortunately, the veloc-
0 0.2 0.4 0.6 0.8 1 ity laws and the corresponding law parameters required in the
Strain % present study can hardly be identified from existing experimental
Fig. 8. Comparison of the stress–strain calculated for the SIA loop configurations
or theoretical works. During the preparatory work of this study, it
C2, C4U and C5. Dotted lines are the results of MD simulations [7] and continuous was found that the results of our DD simulations were very sensi-
lines the results of DD simulations. tive to the amplitudes of the velocity considered for the slip system
h1 0 0i{0 0 1}. Indeed, the different solutions we tested for this slip
system could cause very different reactions for the C4U and the
Finally, the transformed SIA loop is completely absorbed on the C5 loop configurations. Hence, a special attention has to be paid
edge dislocation (see Fig. 7c), and as previously observed in the C2 to the identification of the mobility law for the slip system
configuration, a super-jog is formed along the mobile edge disloca- h1 0 0i{0 0 1}.
tion. At the end of the reaction the edge dislocation is relatively It is commonly admitted that the dislocation mobility in all slip
 1 2Þ plane
free to glide with, one side of the super-jog gliding in a ð3 systems except h1 0 0i{0 0 1} is relatively high in a-iron at high tem-

and the other side in a ð1 1 2Þ plane. In agreement with MD simu- perature [27,29–33]. Therefore, the dislocation velocity in all the
lations, the C5 configuration requires the highest unpinning stress, slip systems was calculated using Eq. (1), varying a damping coef-
being about 152 MPa. ficients B in the range 2.5 and 40  105 Pa s. Then, having in mind
that the ½h1 1 1i{1 1 0} slip system exhibits the highest dislocation
mobility, the corresponding Bs coefficient was adjusted to repro-
3.4. SIA loops pinning strength duce the MD measured critical shear stress (being about 30 MPa)
in the C2 configuration, by choosing the friction stress s0 to be a
It is worth to note that the three above discussed configurations constant value for BCC crystal (see Fig. 8). This fitting process
(C2, C4U and C5) were investigated with the identical set of DD allowed fixing a reference value of B = 8  105 Pa s which is very
simulation parameters. Hence, the demonstration has been made similar to the value obtained in [29]. Subsequently, the dislocation
that the NUMODIS simulation model can capture many details of mobility and the friction stress in the ½h1 1 1i{1 1 2} and
the edge dislocation–SIA loop reactions previously observed by ½h1 1 1i{1 2 3} slip systems were chosen to be the same, as reported
Terentyev et al. [7]. Now, the stress–strain curves computed with in Table 2. In the case of the ½h1 1 1i{1 1 2}, MD simulations have
both simulation methods and for the three different configurations recently shown that the friction stress for the edge dislocation is
are compared in Fig. 8. Analysis of the atomistic (MD) and elastic negligibly small at 250 K and above [30], just as in the well-studied
(DD) simulation results shows that the latter curves are systemat- ½h1 1 1i{1 1 0} slip system. Mobility and activation of the edge dis-
ically smoother than the MD simulation counterparts. This effect is location in the ½h1 1 1i{1 2 3} system was also studied in [34]
not surprising and comes from both the atomic vibration due to applying the same methodology as in [30]. It is also found that
finite temperature in MD simulations and the lower number of the friction stress for the edge dislocation reduces to several MPa
degrees of freedom used in DD simulation to integrate dynamics. as the ambient temperature reaches 200 K. Hence, the usage of
More important, the pinning strength hierarchy found for the similar B for the edge dislocation in the ½h1 1 1i{1 1 0},
three-tested configurations is independent of the simulation
method. Thus, the DD tool captures not only the details of the
interaction mechanisms but also reproduces the resulting Table 3
Critical stress to unpin an isolate edge dislocation in slip system ½h1 1 1i{1 1 0} from 3
strengthening. Nevertheless, small differences exist between the
different SIA loop configurations. Comparison is made between the results of MD
results of the simulations. Whatever the tested configuration, the simulations [7] and DD simulations.
stress–strain curves obtained with DD simulations always exhibits
Type MD (MPa) DD (MPa) Difference (%)
a larger strain at the end of the reaction. Such feature can very
likely be explained with dislocation line tension differences in both C2 35 30 14
C4U 110 130 18
types of simulations. Indeed, in DD simulations the passing dislo-
C5 128 152 19
cation curvature at the critical configuration is larger. This suggests
42 X.J. Shi et al. / Journal of Nuclear Materials 460 (2015) 37–43

½h1 1 1i{1 1 2}, ½h1 1 1i{1 2 3} slip systems at room temperature is comparable to experimental conditions, could be considered
believed to be a justified approximation. using the defined here DD tool and its parameterization.
For the h1 0 0i{1 1 0} slip system, we used the value obtained 3. In a multi-scales modeling approach, the main benefit of the DD
directly by MD simulations [35,36] about 8  104 Pa s. Finally, for simulations is of course their capacity to investigate the collec-
the h1 0 0i{0 0 1} slip system, atomistic simulations have not shown tive response by considering the interaction of a dislocation
any direct evidence of dislocation glide [37]. Nevertheless, some array with a realistic loop microstructure (in terms of density,
dislocation reactions observed in the MD simulations suppose line size and spatial distribution), which corresponds to specific
displacement in the h1 0 0i{0 0 1} slip system [7,35]. To reproduce irradiation conditions. Such type of simulations have already
this observation, the dislocation mobility in h1 0 0i{0 0 1} slip sys- been published, but with DD simulation codes using constitu-
tem must be thermally activated and associated with a rather high tive rules simplifying a lot the elementary mechanisms tested
activation energy. The modeling of thermally activated dislocation in the present study [41–43] with the notable exception of
mobility in DD simulations is problematic when calculations [44]. Work is in progress to explore if this model can extend
involve co-existence of dislocations with significantly different or modify the already established DD predictions.
mobility [38–40], a mobility law governed by Eq. (1) was therefore
considered for the h1 0 0i{0 0 1} slip system by assigning a very high
drag coefficient. Trial and error simulations were finally made to Acknowledgements
identify the h1 0 0i{0 0 1} Bs value, which allows one reproducing
the MD result. Considering all the aspects above, Bs = 9  102 Pa s This work was mainly performed in the framework of the PER-
was finally taken to complete the parameterization of dislocation FORM60 project. The authors acknowledge the support of the
mobility. For both h1 0 0i{0 0 1} and h1 0 0i{0 1 1} slip systems, our French Agence Nationale de la Recherche (ANR) under reference
simulations are not sensitive to the friction stress s0. We set it to ANR-10-COSI-0011 (OPTIDIS) for the numerical development of
be about 300 MPa following the best agreement with MD results code NUMODIS. We are also grateful to the other developers of
for the simulation involving the interaction of the ½h1 1 1i{1 1 0} NUMODIS, M. Blétry, M. Fivel, E. Ferrié, A. Etcheverry and O. Cou-
edge dislocation with a ½h1 1 1i hexagonal loop inclined to the dis- laud, and also acknowledge E. Clouet for fruitful discussions.
location slip plane. The details of this study will be discussed in an
upcoming publication.
It must be noticed that once the parameters of the slip system Appendix A. Supplementary material
velocity laws are set (see Table 2), the DD simulation results are
essentially insensitive to the time step used to integrate dislocation Supplementary data associated with this article can be found, in
dynamics. Therefore, timeframe and crystal dimensions thousand the online version, at http://dx.doi.org/10.1016/j.jnucmat.2015.01.
times bigger than in the MD simulations can be easily explored 061.
in the DD simulations without affecting the accuracy of the
simulation results. References
Finally, we simulated the C6 configuration also investigated in
[7]. The DD simulation parameters were identical to those used [1] D.J. Bacon, F. Gao, Y.N. Osetsky, Nucl. Instrum. Methods Phys. Res., Sect. B 153
(1999) 87.
to simulate C2, C4U and C6 configurations. Unfortunately, the reac- [2] Y.N. Osetsky, D.J. Bacon, A. Serra, B.N. Singh, S.I. Golubov, Phil. Mag. 83 (2003)
tion path predicted by the MD could not be reproduced. Detailed 61.
analysis of the reaction pathway in the MD and DD simulations [3] Y.N. Osetsky, M. Victoria, A. Serra, S.I. Golubov, V. Priego, J. Nucl. Mater. 251
(1997) 34.
has shown that the reproduction of the reaction by the NUMODIS [4] Y.N. Osetsky, A. Serra, V. Priego, Mater. Res. Soc. Symp. Proc. 25 (1998) 59.
tool requires implementation of new types of dislocations reac- [5] A.C. Nicol, M.L. Jenkins, M.A. Kirk, Mater. Res. Symp. Proc. 650 (2001) R131.
tions. Work is in progress to develop and implement this feature [6] S.L. Dudarev, R. Bullough, P.M. Derlet, Phys. Rev. Lett. 100 (2008) 135503.
[7] D. Terentyev, P. Grammatikopoulos, D.J. Bacon, Y.N. Osetsky, Acta Mater. 56
in the NUMODIS code. (2008) 5034–5046.
[8] Z. Rong, Y.N. Osetsky, D.J. Bacon, Philos. Mag. 85 (2005) 1473–1493.
5. Conclusions and perspectives [9] D.J. Bacon, Y.N. Osetsky, Z. Rong, Philos. Mag. 86 (2006) 392.
[10] D. Terentyev, Y.N. Osetsky, D.J. Bacon, Scr. Mater. 62 (2010) 697.
[11] J. Marian, B.D. Wirth, R. Schäublin, G.R. Odette, J. Manuel Perlado, J. Nucl.
The present DD study was focused on the elementary reactions Mater. 323 (2003) 181–191.
and strengthening mechanisms involving an edge dislocation [12] D. Terentyev, G. Monnet, P. Grigorev, Scr. Mater. 69 (2013) 578–581.
interacting with the h1 0 0i dislocation loops in irradiated a-Fe, [13] G. Monnet, L. Vincent, B. Devincre, Acta Mater. 61 (2013) 6178–6190.
[14] C.S. Shin, M.C. Fivel, D. Rodney, R. Phillips, V.B. Shenoy, L. Dupuy, Le J. Phys. IV
most commonly observed upon neutron irradiation at PWR tem-
11 (2001).
perature and above. Comparison with the results of MD simula- [15] E. Martinez, J. Marian, A. Arsenlis, M. Victoria, J.M. Perlado, Phil. Mag. 88 (2008)
tions published in [7] proved the capability of the present DD 809–840.
[16] B. Devincre, L.P. Kubin, Mater. Sci. Eng. A 8 (1997) 234–236.
simulation model to quantitatively reproduce all the important
[17] B. Devincre, L.P. Kubin, C. Lemarchand, R. Madec, Mater. Sci. Eng. A 309–310
elementary features observed in more fundamental simulations (2001) 211–219.
at the atomic scale. Currently, we can draw the following conclu- [18] V.V. Bulatov, L.L. Hsiung, M. Tang, A. Arsenlis, M.C. Bartelt, V. Cai, J.N. Florando,
sions and perspectives: M. Hiratani, M. Rhee, G. Hommes, T.G. Pierce, T. Diaz de la Rubia, Nature 440
(2006) 1174–1178.
[19] A. Nomoto, N. Soneda, A. Takahashi, S. Ishino, Mater. Trans. 46 (2005) 463.
1. Quantitative difference for the unpinning stress measured by [20] D. Terentyev, L. Malerba, D.J. Bacon, Y.N. Osetsky, J. Phys.: Condens. Matter 19
the MD and DD simulations amounts to 18% in average. Consid- (2007) 456211.
[21] M.L. Jenkins, Z. Yao, M. Hernandez-Mayoral, M.A. Kirk, J. Nucl. Mater. 389
ering the computation accuracy and the physical approxima- (2009) 197.
tions behind both simulation methods, this gap is believed to [22] J. Drouet, L. Dupuy, F. Onimus, F. Mompiou, S. Perusin, A. Ambard, J. Nucl.
be satisfactory and no further investigation was made to Mater. 449 (2014) 252–262.
[23] V.V. Bulatov, W. Cai, Computer Simulations of Dislocations, Oxford University
improve the DD simulation models. Press, Oxford, 2006.
2. The present DD simulation results show the solid reliability of [24] A. Arsenlis, W. Cai, M. Tang, M. Rhee, T. Oppelstrup, G. Hommes, T.G. Pierce,
the elastic theory to reproduce dislocation properties. Hence, V.V. Bulatov, Mater. Sci. Eng. 15 (2007) 553–595.
[25] W. Cai, C. Weinberger, V.J. Bulatov, Mech. Phys. Solids 54 (2006) 561–587.
simulations that may be too difficult to realize at the atomic
[26] R.O. Scattergood, D.J. Bacon, Phil. Mag. 31 (1975) 179–198.
scale, for example to explore much lower deformation rates [27] Y.N. Osetsky, D.J. Bacon, Mater. Sci. Eng. 11 (2003) 427.
X.J. Shi et al. / Journal of Nuclear Materials 460 (2015) 37–43 43

[28] L. Kubin, Dislocations, Mesoscale Simulations and Plastic Flow, vol. 5, Oxford [36] D. Terentyev, N. Anento, A. Serra, J. Nucl. Mater. 420 (2012) 9–15.
University Press, 2013. [37] N. Kumar, R. Tewari, P.V. Durgaprasad, B.K. Dutta, G.K. Dey, Comput. Mater. Sci.
[29] D. Terentyev, D.J. Bacon, Y.N. Osetsky, J. Phys.: Condens. Matter 20 (2008) 77 (2013) 260–263.
445007. [38] G. Monnet, B. Devincre, L. Kubin, Acta Mater. 52 (2004) 4317–4328.
[30] G. Monnet, D. Terentyev, Acta Mater. 57 (5) (2009) 1416–1426. [39] J. Chaussidon, F. Marc, R. David, Acta Mater. 54 (2006) 3407–3416.
[31] S. Queyreau, J. Marian, M.R. Gilbert, B.D. Wirth, Phys. Rev. B 84 (6) (2011) [40] S. Naamane, G. Monnet, B. Devincre, Int. J. Plast 26 (1) (2010) 84–92.
064106. [41] T. Nogaret, C. Robertson, D. Rodney, Philos. Mag. 87 (2007) 945–966.
[32] A.Y. Kuksin, A.V. Yanilkin, Phys. Solid State 55 (5) (2013) 1010–1019. [42] N.M. Ghoniem, S.H. Tong, J. Huang, B. N Singh, M. Wen, J. Nucl. Mater. 307
[33] N. Urabe, J. Weertman, Mater. Sci. Eng. 18 (1975) 41–49. (2002) 843–851.
[34] D. Terentyev, A. Bakaev, X.J. Shi, D. van Neck, Model. Sim. Mater. Sci. Eng., [43] B.N. Singh, N.M. Ghoniem, H. Trinkaus, J. Nucl. Mater. 307 (2002) 159–170.
submitted for publication. [44] A. Arsenlis, M. Rhee, G. Hommes, R. Cook, J. Marian, Acta Mater. 60 (2012)
[35] D.A. Terentyev, Y.N. Osetsky, D.J. Bacon, Acta Mater. 58 (2010) 2477. 3748–3757.

You might also like