Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Nuclear Materials 573 (2023) 154112

Contents lists available at ScienceDirect

Journal of Nuclear Materials


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

Influence of loading orientation on deformation localization of


irradiated tungsten
Zhijie Li, Yinan Cui∗
Applied Mechanics Lab., School of Aerospace Engineering, Tsinghua University, Beijing 100084, PR China

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

Article history: The plasticity of tungsten in low or medium temperature regimes is dominated by the screw dislocation
Received 24 August 2022 mobility, which is well known to exhibit a strong non-Schmid effect and loading orientation dependence.
Revised 20 October 2022
As a typical plasma-facing material, many questions regarding the role of these plasticity features on the
Accepted 31 October 2022
irradiation degradation performance of tungsten remain unanswered. Considering that deformation local-
Available online 2 November 2022
ization is a continuing critical problem in the nuclear application, in the current work, a systematic set of
Keywords: crystal plasticity calculations is designed to quantify the degree of deformation localization in irradiated
Tungsten tungsten, which is found to be sensitive to the loading orientation. The underlying physical mechanisms
Irradiation are carefully analyzed, highlighting the important role of the number of active slip systems, the non-
Loading orientation Schmid effect controlled by the evolution of local stress state, and the high resistance stress induced by
Deformation localization irradiation defects. Based on the micro-scale physical mechanisms, two physical indicators are proposed
to provide guidance for predicting the orientation-dependent deformation localization in irradiated tung-
sten, which agrees well with the simulation results.
© 2022 Elsevier B.V. All rights reserved.

1. Introduction Previous work demonstrates that dislocation channels form


along favorably oriented slip planes [16] and are controlled by
Irradiation with energetic particles leads to the formation of ir- the slip system softening [17]. Considering that loading orienta-
radiation defect clusters, which harden the materials by inhibiting tion plays a crucial role in influencing the plastic slip on a micro-
dislocation motion [1,2,3], but often promote the appearance of de- scopic scale, deformation localization is expected to develop in an
formation localization [4,5]. This is a significant concern in nuclear orientation-dependent manner [18]. Recent discrete dislocation dy-
applications because it is one of the main origins of irradiation em- namics (DDD) work discloses that dislocation channel formation is
brittlement [6] and stress corrosion cracking [7,8]. inhibited for a loading orientation that activates multiple slip sys-
Deformation localization of irradiated metals continues to tems, which requires a very high irradiation dose for the transi-
receive considerable experimental and theoretical attention tion from multi-slip mode to single-slip mode [13]. Barton et al.
[9,10,11,12]. The previous scenario has gained a consensus that observed that deformation localization develops very early in irra-
deformation localization in irradiated metals is induced through diated Fe single crystal at off-[100] orientation [9] due to the sup-
the formation of dislocation channels due to the clearing of pression of even distribution of slip. For neutron-irradiated Zr alloy,
radiation-induced obstacles by moving dislocations. Localization transmission electron microscope investigations show that dislo-
is prone to occur at high irradiation doses [13,10]. Widening cation channels are observed along basal planes during transverse
dislocation channels prefer to be formed in high stacking fault tensile tests at 350 °C, while both prismatic and pyramidal dislo-
energy materials where the thermally activated processes of cross cation channels are observed during axial tensile tests [19,20].
slip [14] is easy, depending on dose and stress state [15]. How- Almost no studies, to date, have systematically explored the ori-
ever, a thorough understanding of deformation localization and entation effect on deformation localization of irradiated tungsten,
dislocation channel formation is still lacking due to its complex even though W is considered as the most potential plasma-facing
nature and multiple contributing factors. One example is that the material in the divertor of nuclear fusion devices [21] and the for-
loading orientation (LO) effect on the susceptibility and features of mation of dislocation channels in W is observed [22,10]. It is in-
localization remains poorly understood. triguing to investigate this problem because many new physical in-
sights may appear. As a typical body center cubic (BCC) metal, the
plasticity of unirradiated tungsten is well-known to exhibit strong

Corresponding author. orientation dependence, and the well-known Schmid law breaks
E-mail address: cyn@mail.tsinghua.edu.cn (Y. Cui).

https://doi.org/10.1016/j.jnucmat.2022.154112
0022-3115/© 2022 Elsevier B.V. All rights reserved.
Z. Li and Y. Cui Journal of Nuclear Materials 573 (2023) 154112

down. The non-planar core structure of screw dislocations leads where a superposed dot means a time derivative. As for the ma-
to the non-Schmid effect [23,24], which means that the non-glide terials with the isotropic thermal expansion, the thermal deforma-
components of an applied stress tensor also affect the disloca- tion gradient F T can be expressed with the thermal stretch λ(T ),
tion behavior by modifying the dislocation core spreading pattern and the thermal velocity gradient LT and plastic velocity gradient
and affecting the critical shear stress for the onset of the disloca- L p can be simplified as [32,33],
tion slip. The deviations from the Schmid law can be quantified
F T = λ(T )I (6)
from the deviations of the screw dislocation trajectory away from
a straight path between equilibrium configurations, and a modi-
1 ∂λ
fied parameter-free Schmid law has been proposed by comparing LT = F e F˙ T F T,−1 F e,−1 = T˙ I = α T˙ I (7)
well with experimental variations and first-principles calculations λ (T ) ∂ T
of the dislocation Peierls stress as a function of crystal orienta-
tion [25]. Previous work suggests that the non-Schmid effect pro- L p = F e F T F˙ p F p,−1 F T,−1 F e,−1 = F e F˙ p F p,−1 F e,−1 = F e Lˆ p F e,−1 (8)
motes deformation localization [11,26]. However, how it affects the
orientation-dependent deformation localization under irradiation is Lˆ p = F˙ p F p,−1 (9)
far from well understood.
It is of great importance to answer these questions to im- where I is the second-order identity tensor, α is the linear expan-
prove the anti-localized deformation capacity of irradiated tung- sion coefficient.
sten structures. Firstly, a systematic set of crystal plasticity cal- In the current dislocation-based crystal plasticity model, it is
culations is designed to study the LO effect on the deformation assumed that all the plastic deformation is induced by dislocation
localization. The adopted crystal plasticity model is carefully val- slip. The plastic velocity gradient Lˆ p can be calculated as [32,34],
idated to be able to capture the loading orientation dependence  β
LˆP = PS γ˙ β (10)
of the mechanical properties of unirradiated tungsten. Afterwards,
β
the underlying mechanisms of deformation localization at different
β
loading orientations are studied. Finally, based on these insights, PS = mβ  nβ (11)
the deformation localization indicators are proposed.
β
where γ˙ β is the slip rate on the β th slip system, and PS is a geo-
2. Crystal plasticity model: description and validation metric projection tensor, representing the Schmid (geometric) pro-
jection of the strain rate contribution from the β th slip system de-
The investigations on the dislocation channel formation in irra- fined by the plane normal nβ and slip direction mβ (both unit vec-
diated tungsten requires a theoretical model, where several criti- tors). Based on the Orowan equation, the slip rate γ˙ β is calculated
cal and fundamental physical processes should be formulated rea- from the mobile dislocation density ρ β and dislocation velocity vβ
sonably, including the temperature and LO dependent plasticity of in β th slip system [32],
tungsten associated with the non-planar core structure of screw γ˙ β = bρ β vβ (12)
dislocations, the evolution of dislocation density, and the evolution
of irradiation defects and their collective interactions with moving As for tungsten, plastic deformation is dominated by screw dis-
dislocations. To achieve this goal, we have recently incorporated locations in the intermediate and low temperature regime [35,36].
all these physical processes into a crystal plasticity model [11] Edge dislocations mainly influence the initial micro-yielding defor-
based on the insights gained from molecular dynamics (MD) stud- mation stage at low temperature [29,11,37]. In the current work,
ies about the thermal-activated dislocation mobility law [27], DDD the contribution of both screw dislocation and edge dislocation is
studies about the evolution of dislocation and irradiation defects considered [11]. The slip rate γ˙ β is calculated as the sum of their
and their interactions [28], and the experimental studies about the contributions,
determinations of some model parameters [29,30]. We begin with
 
a streamlined presentation of the model for completeness. γ˙ β = b ρscβ vβsc + ρed
β β
ved (13)
For a given material point X in the reference configuration, it β β β β
can be mapped to x in the current configuration through a linear where ρsc , vsc , ρed and ved are the screw dislocation density, screw
transformation represented by the deformation gradient tensor F , dislocation velocity, edge dislocation density and edge dislocation
which can be multiplicatively decomposed into the elastic part F e , velocity, respectively.
the thermal part F T , and the plastic part F p [31], Dislocation mobility laws are established based on the mech-
anism of phonon drag mechanism for edge dislocations and the
∂x thermal-activated kink pair for screw dislocations [32,27],
F= = F eF T F p (1)
∂X ⎧  
⎨ τeβf f b Gkp (σ ,T )
where the superscripts e, T and p denote the elastic part, ther- B(σ ,T )
exp − if Gkp (σ , T ) > 0
vβ ( σ , T ) = 2kT
(14)
mal part and plastic part, respectively, in the following. The cor- ⎩ τeβf f b
if Gkp (σ , T ) ≤ 0
responding velocity gradient L can be obtained by taking time B(σ ,T )
derivative of Eq. (1), including the elastic part Le , the thermal part   Gkp (σ ,T )

a 2a exp − +L
LT and the plastic part L p , 2kT
Gkp (σ , T ) > 0
B (σ , T ) = 2hL
Bk if (15)
L = F˙ F −1 = Le + LT + L p (2) B0 + B1 T if Gkp (σ , T ) ≤ 0
p
q
T

Gkp (σ , T ) = β
H0 1− (σ ) − (16)
T0
Le = F˙ e F e,−1 (3)
 
T e ˙T T,−1 e,−1 τeβf f + a1 σ : mβ  nβ1
L =F F F F (4) β
(σ ) =   
β β
a0 τ p − a2 σ : nβ × mβ  nβ − a3 σ : n1 × mβ  n1

L p = F e F T F˙ p F p,−1 F T,−1 F e,−1 (5) (17)

2
Z. Li and Y. Cui Journal of Nuclear Materials 573 (2023) 154112

where Gkp (σ , T ) is the activation enthalpy of kink-pair as de- This model has been validated in our recent work [11] to be
fined in [27], which considers the non-Schmid effect. k is Boltz- able to accurately predict the mechanical properties of unirradi-
mann’s constant, a is the lattice constant, h is the kink height, L = ated and irradiated tungsten single crystals for different tempera-

( ρ β |nβ · mβ | )−1 is the length of dislocation segment. B0 and B1 tures and irradiation doses [29,48,49] by developing the user sub-
β routine in ABAQUS. In the current work, this model is further veri-
are two material constants, which adopt different values for screw fied to check its capability of capturing the LO dependence. Beard-
dislocations and edge dislocations [38]. 0 ≤ p ≤ 1 and 1 ≤ q ≤ 2, a0 , more and Hull [50] carried out a series of experiments to study
a1 , a2 and a3 are fitting parameters based on molecular dynamics the influence of LOs on the mechanical response of tungsten single
β
calculations [27]. n1 is a unit vector, which forms an angle of −60◦ crystals at 295 K. In their experiments, the tested strip specimens
with the reference slip plane defined by nβ . For screw dislocations, had the gage section of 0.034 inches width and 0.014 inches thick-
the velocity is highly depending on the temperature and increases ness. The LO was described through angles  and  , as indicated
with the temperature. The fitting constant T0 can be interpreted as in the inset in Fig. 1(a). Tensile tests were carried out at a strain
an athermal transition temperature, above which the energy bar- rate of 4 × 10−4 /s. Here, the simulation setup strictly follows the
rier for kink-pair nucleation is guaranteed to vanish independent experiments, and the corresponding element size is about 8 μm.
β
of stress [38]. τe f f is the effective shear stress, acting as the driv- The convergence of the mesh size is firstly verified by using a
ing force for the dislocation motion, smaller mesh size for a given LO, and we found that the deforma-
tion process and the stress-strain relationship is basically the same.
τeβf f = τapp
β β β
− τ f or − τirr (18) The engineering stress-stain curves for different LOs are calculated
based on this model. The predicted yield strength is plotted as
β β
τapp = σ : PS (19) a function of LO to clearly show the LO effect. Fig. 1(a) shows a
good agreement with the experimental data of the yield strength
 (defined as the proportionality limit) [50,51,52]. Fig. 1(b) further
β

NS
compares the predicted hardening modulus at different engineer-
τ f or = αd μb ρα (20)
ing strains with the experimental data [50], and a good agreement
α =1
β
 between simulation and experiments is also observed.
τirr = αi μb N di (21)
β β β 3. Model predictions for loading orientation effect on
where τapp is the applied shear stress, τ f or and τirr are the resis- deformation localization
tance stress due to the dislocation forest and irradiation defects,
respectively. σ is the stress tensor, and NS is the number of all After the careful validations, this model is adopted to explore
considered slip systems. αd and αi are the hardening coefficients how LO influences the deformation localization of irradiated tung-
induced by dislocation forest and irradiation defects, respectively. sten. To the best of our knowledge, this problem has not been sys-
N is the irradiation defect density, and di is the average diameter tematically investigated theoretically or experimentally in previous
of irradiation defects. studies. The simulated specimens are cylindrical dog-bone shaped
The evolution of screw dislocation density ρ˙ sc follows the law to minimize the boundary effect. The diameter and gage length
proposed by Mecking and Kocks [39] and further developed by are 30 μm and 60 μm, respectively. The mesh size is 500 nm,
Beyerlein and Tome [40,41]: which has been validated to meet the convergence requirements

ρ˙ sc (T ) = (k1 ρsc /b − k2 (ε˙ , T )ρsc )γ˙ (22) [11]. Testing temperature 373 K is taken as an example. A given
strain rate of 0.05/s is applied on the upper end to compress the
   sample. An irradiation dose of 0.6dpa is considered, and the corre-
k2 (ε˙ , T ) χ kT ε˙
= 1− ln (23) sponding irradiation defect density and average size are 7.2 × 1022
k1 g Db3 ε˙ 0 /m3 and 5.0 nm, respectively [53,54]. In the current work, the com-
pressive loading condition is chosen due to the following reasons.
D = D0 − φ ∗ T (24)
Firstly, a given tungsten structure is generally installed and fixed at
where k1 , k2 are the coefficients related to the generation and an- room temperature, and the higher working temperature is highly
nihilation process of dislocations, respectively. χ is the interaction possible to induce compressive loading state due to the thermal
parameter, g is the normalized activation energy, D is a proportion- expansion. Secondly, the compressive load will avoid the necking
ality constant, ε˙ 0 is a reference strain rate, ε˙ is the applied strain behavior of tungsten structures induced by tensile loading condi-
rate, D0 is the reference proportionality constant and φ is the ma- tion, and the deformation localization is mainly controlled by the
terial constant. In our recent work [11], the evolution of edge dis- irradiation condition. Lastly, recent experimental studies [10,22,55]
location density ρ˙ ed has been established and is expressed as, have indicated that the irradiated tungsten suffers from the local-
v √
 ized deformation under the nano-indentation (compression) con-
ρ˙ ed (T ) = sc
k1 ρed /b − k2 (ε˙ , T )ρed γ˙ (25) dition.
ved
Fig. 2 shows the plastic strain distribution of irradiated tung-
As for the evolution model of irradiation defects, the irradiation
sten single crystals under different LOs. Four kinds of deformation
defects can be either absorbed or swept away by the moving dis-
localization features are observed. (I) One clear shear band quickly
location within a prescribed capture distance y/2, which has been
appears across the whole body of samples, when the LO is [552],
verified by a wide range of numerical simulations, including DDD
[111], and [115]. The plastic strain in the deformation localization
[42,28,43] and MD [44,45,46]. Accordingly, the changing rate of ir-
zone is an order of magnitude larger than the applied engineering
radiation defect density N˙ is expressed as,
strain. (II) For [112], the deformation localization occurs until the
y
N˙ = N˙ p − λ N γ˙ (26) engineering strain achieves about 3%, much higher than type (I).
b (III) When the LO is [100], there are two local shear bands (marked
where N˙ p is the irradiation defect production rate and set as zero with "A" and "B" in Fig. 2(a-3)) near the transition zone of the
for post-irradiation deformation. λ describes the annihilation prob- sample. (IV) As for [210], no shear bands appear during the con-
ability. All the related parameters used in the model are listed in sidered engineering strain of 3%. The analysis above indicates that
Table 1 and have been discussed in our recent work [11]. deformation localization exhibits significant LO dependence.

3
Z. Li and Y. Cui Journal of Nuclear Materials 573 (2023) 154112

Table 1
Relevant parameters in the crystal plasticity model for the dislocation mobility law, resistance stress model and
the microstructure density evolution of tungsten [47,30,27,41].

Parameter[Units] a[Å] b[Å] h[Å] H0 [eV] p[-] q[-]



Value 3.16 2.72 a 2/3 1.63 0.86 1.69
Parameter[Units] α [/K] T0 [K] τ p [GPa] a0 [-] a1 [-] a2 [-]
Value 4.65 × 10−6 2956 2.03 1.50 1.15 2.32
Parameter[Units] a3 [-] Bk [Pa · s] 0 [Pa·s]
Bsc 1 [Pa · s/K]
Bsc 0 [Pa·s]
Bed 1 [Pa · s/K]
Bed
Value 4.29 8.3 × 10−5 9.8 × 10−4 0 4.26 × 10−4 8.7 × 10−5
Parameter[Units] αd [-] αi [-] k1 [-] χ [-] g[-] φ [MPa/K]
Value 0.6 0.2 0.2 0.9 0.01 16.0
Parameter[Units] D0 [MPa] ε˙ 0 [/s] λy/b[-]
Value 10,000 1.0 × 107 10

Fig. 1. Comparison of simulation results (marked with “Sim”) and experimental data (marked with “Exp”) on the loading-orientation dependence of yield strength and
hardening modulus at different engineering strains ε .

Fig. 2. Loading orientation effect on the deformation localization of irradiated tungsten single crystals at 373 K.

4
Z. Li and Y. Cui Journal of Nuclear Materials 573 (2023) 154112

Fig. 3. Deformation localization index (DLI) at three given engineering strains ε .  is the maximum slip rate ratio of the dominant slip system, and m is the strain rate
sensitivity coefficient. Different regimes describe different deformation localization mechanisms, as described in Section 4.1.

To further quantify the deformation localization extent for dif- generality, one can define a plastic contribution ratio ωβ to de-
ferent LOs, a deformation localization index (DLI) is adopted. DLI is scribe the contribution of β th slip system to the total plastic strain
defined as the volume percentage where the local plastic strain is along the loading direction at the yield point, which can be calcu-
smaller than the volume-averaged plastic strain [56]. It is easy to lated as,
prove that DLI ranges from 0 to 1, and higher DLI means a greater

NS
extent of localization. Systematic studies are carried out to under- ωβ = Mβ γ˙ β / Mα γ˙ α (27)
stand the LO effect on DLI, where LO is also described through α =1
angles  and  , as indicated in the inset in Fig. 3. For different
LOs, the evolution of DLI is calculated at three different engineer- where the plastic slip rate γ˙ β can be theoretically evaluated
ing strains (namely 1%, 2% and 3%) and shown in Fig. 3. DLI gen- through Eq. (12). At the yield point, ρ β can be taken as the ini-
erally increases with the deformation before the formation of dis- tial dislocation density, the dislocation velocity vβ is determined
location channels if the plastic deformation gradually concentrates by assuming that the total plastic strain rate is equal to the ap-
in the dislocation channel region. Once the dislocation channel ap- plied strain rate. According to Eq. (14), higher ωβ requires higher
β β
pears, DLI reaches a relatively stable value, because the subsequent effective shear stress τe f f and lower activation enthalpy Gkp . τe f f
deformation basically converts into the plastic deformation of the is positively correlated to |M|, while Gkp depends on the applied
limited channel zone. It can be seen that the DLI evolution is con- stress tensor and considers the non-Schmid effect.
sistent with the evolution of the plastic strain distribution in Fig. 2. Table 2 summarizes the calculated absolute value of Schmid
By comparing with Fig. 2, a clear shear band appears when DLI is factor |M| and the plastic contribution ratio of the active slip sys-
β β
higher than about 0.6. It is found that for all the considered LOs, tems for unirradiated tungsten (ωU ) and irradiated tungsten (ωI )
the evolution of the deformation localization is similar to the four under different LOs. The subscripts "U" and "I" are used to describe
typical cases described in the previous paragraph and Fig. 2. There- the unirradiated and irradiated tungsten, respectively. In order to
β
fore, it is reasonable to differentiate the four regimes, as indicated verify the theoretical calculation of ωI , the corresponding simu-
in Fig. 3. In the following, the underlying mechanism of these dif- β
lation results ωI,S are further listed in Table 2. A good agreement
ferent deformation localization features will be disclosed by taking β β
the typical cases shown in Fig. 2 as examples. between ωI and ωI,S is observed. Our results show several intrigu-
ing features.

4. Discussions • For the LO of [552], [111] and [115] in regime I, there are sev-
eral active slip systems with a high absolute value of Schmid
4.1. Analysis of deformation localization mechanisms factor |M|, and the dominant slip system, contributing most to
β
the total plastic deformation, has a much higher ωI than that
Regarding the origin of LO dependence of deformation local- of other slip systems. The dominant slip system of unirradiated
ization, the natural idea is first to analyze the Schmid factor M tungsten is consistent with that of irradiated tungsten.
and the number of activated slip systems. Concerning tungsten, the • For the LO of [112] in regime II, despite there is also only one
non-Schmid effect makes the non-glide components of an applied dominant slip system, an unexpected phenomenon is observed.
stress tensor affect the dislocation motion. Therefore, the domi- The dominant slip system for unirradiated tungsten is differ-
nated slip system may not have the maximum M. Without loss of ent from that of irradiated tungsten, and there is a great dif-

5
Z. Li and Y. Cui Journal of Nuclear Materials 573 (2023) 154112

Table 2
β β
The absolute value of Schmid factor |M| and the plastic contribution ratio ω of active slip systems under different loading orientations, ωU and ωI correspond to unirradi-
β
ated and irradiated tungsten, respectively, and ωI,S is the simulation results for irradiated tungsten.

loading Orientation Slip System |M| ωUβ ωIβ β


ωI,S loading Orientation Slip System |M| ωUβ ωIβ β
ωI,S
[100] [1̄11](101 ) 0.41 0.25 0.25 0.25 [210] [1̄1̄1](011 ) 0.24 0.11 0.00 0.00
(Regime Ⅲ) [11̄1](1̄01 ) 0.41 0.25 0.25 0.25 (Regime Ⅳ) [1̄1̄1](101 ) 0.49 0.22 0.50 0.50
[1̄11̄](110 ) 0.41 0.25 0.25 0.25 [111](1̄01 ) 0.49 0.22 0.50 0.50
[111](1̄10 ) 0.41 0.25 0.25 0.25 [111](1̄10 ) 0.24 0.43 0.00 0.00
[552] [1̄1̄1](011 ) 0.42 0.80 0.81 0.91 [111] [1̄1̄1](011 ) 0.27 0.92 0.92 0.96
(Regime Ⅰ) [1̄1̄1](101 ) 0.42 0.19 0.19 0.09 (Regime Ⅰ) [1̄11](101 ) 0.27 0.04 0.04 0.02
[111](1̄01 ) 0.27 0.01 0.00 0.00 [1̄1̄1](101 ) 0.27 0.04 0.04 0.02
[112] [11̄1](011 ) 0.41 0.04 0.15 0.10 [115] [11̄1](011 ) 0.45 0.09 0.14 0.13
(Regime Ⅱ) [1̄11](01̄1 ) 0.14 0.67 0.00 0.00 (Regime Ⅰ) [1̄1̄1](011 ) 0.27 0.06 0.00 0.00
[1̄11](101 ) 0.41 0.17 0.85 0.90 [1̄11](01̄1 ) 0.30 0.32 0.00 0.00
[111](1̄01 ) 0.27 0.11 0.00 0.00 [111](1̄01 ) 0.42 0.50 0.80 0.85

β β
ference between ωU and ωI . The slip system with the maxi- that the orientation of dislocation channel plane agrees well with
β the {110} crystallographic plane [22,10]. Therefore, the dislocation
mum ωU has a lower absolute value of Schmid factor |M|, due
channels of irradiated tungsten are basically along the crystallo-
to the low activation enthalpy of kink-pair Gkp . However, the
β graphic plane of the dominant slip system. The formation of dis-
slip system with the maximum ωI is the same as that with location channels and the deformation localization have a strong
the maximum |M|, because the high resistance stress induced correlation with the parameter of maximum plastic contribution
by irradiation defects inhibits the dislocation motion in the slip ratio ω.
systems with a lower |M|. Therefore, only the slip system with
a higher |M| can develop into the dominant one in irradiated 4.1.2. Regime Ⅰ: Single-slip
tungsten, which further determines the orientation of the dis- The characteristic feature of Regime I is that there is only one
location channel. This phenomenon is consistent with the ex- dominant slip system, which has the high plastic contribution ratio
perimental observation [57,58] and simulation results [17,58,59] ωI and low activation enthalpy Gkp . Here, ωI and ωU means the
that the local region suffering from deformation localization al- β β
maximum value of ωI and ωU , respectively, as shown in bold in
ways has a higher |M|.
Table 2. When Gkp is low, relatively low effective shear stress τe f f
• For the LO of [100] in regime III, multiple dominant slip sys-
β can drive the gliding of dislocations to contribute to the required
tems are activated simultaneously, and the lower value of ωU
plastic strain rate. Taking LO of [552] as an example, Fig. 5(a-c)
β
is close to that of ωI . shows that τe f f is on the order of several MPa during the initial
β plastic deformation stage, and the plastic slip rate γ˙ of the domi-
• For the LO of [210] in regime IV, ωU is substantially different
β
from ωI , and there is more than one dominant slip system. nant slip system [1̄1̄1](011 ) is much higher than the others.
During the initial deformation stage, higher τe f f is observed in
In the following, we will first analyze the correlation between the near-surface region (such as point A in Fig. 5a), whose local
deformation localization and the crystallographic slip plane of plastic strain is higher, which further leads to the change of the
the dominant slip system, and then disclose how these different local stress state. When ε reaches about 1.25%, the non-axial stress
regimes contribute to different evolution features of deformation component is found to increase τe f f in the internal region (see
localization in irradiated tungsten. point B in Fig. 5a) to catch up with that in the near-surface re-
gion (see Fig. 5c). Therefore, γ˙ becomes comparable for points A
4.1.1. Crystallographic analysis on deformation localization and B. The continuous dislocation gliding along the dominated slip
For different LOs, the normal direction of dislocation channel plane dramatically absorbs or sweeps away the irradiation defects,
plane can be calculated from the simulation results of the irradia- boosting the formation of the dislocation channel across the whole
tion defect density distribution. The calculated normal vectors are sample (See Fig. 5a).
plotted in the pole figure of crystallographic orientations, as shown
in Fig. 4(a). The large gray circles represent the region within 10° 4.1.3. Regime Ⅱ: High kink-pair activation enthalpy
with respect to the normal vector of the dominated slip plane for Regime II is also manifested as having only one dominant slip
each case. For a given LO, Fig. 4(a) shows that the normal direc- system with the maximum plastic contribution ratio ωI . However,
tion of dislocation channel plane falls into the large gray circle, this slip system has a high activation enthalpy of kink-pair Gkp ,
proving that the dislocation channel plane is basically consistent because the slip system with low Gkp cannot be effectively acti-
with the corresponding dominated slip plane. The LO of [552] is vated since its low absolute value of Schmid factor |M| means low
taken as an example. It is known from Table 2 that the slip sys- applied shear stress, which cannot overcome the high resistance
β
tem of [1̄1̄1](011 ) has the maximum plastic contribution ratio ωI stress associated with irradiation defects. According to Eq. (14),
(0.81) for irradiated tungsten. Therefore, [1̄1̄1](011 ) is named as when Gkp is high, a remarkably high τe f f is required to drive the
the dominant slip system, and the corresponding normal vector plasticity. As expected, the simulation results in Fig. 5(f) for the LO
is (011 ). For this LO, the normal direction of dislocation channel of [112] show that τe f f is on the order of one hundred MPa, much
plane is (0.08 0.61 0.79) according to the simulation result, which higher than that for Regime I, while the plastic slip rate is compa-
is indicated by the black circle point in Fig. 4(a). The angle be- rable in Fig. 5(b) and (e) around the yield point.
tween these two normal vectors is about 8°, so they are essen- Similar to Regime I, the near-surface region has a higher τe f f
tially parallel. This agrees with the experimental observations that and plastic slip rate γ˙ initially. The difference between τe f f at
the localized deformation appears in the crystallographic slip plane points A and B gradually diverges, as shown in Fig. 5(f). This dif-
for irradiated tungsten [22,10] and other bcc metals [60,61]. As ference is much higher than Fig. 5(c) and cannot be eliminated
an example, Fig. 4(b) shows the nanoindentation experiment on by changing the local stress state during plastic deformation as
the helium-ion irradiated tungsten single crystals, which showed discussed for Regime I. The comparable plastic slip rate between

6
Z. Li and Y. Cui Journal of Nuclear Materials 573 (2023) 154112

Fig. 4. Crystallographic orientations of dislocation channels. (a) Pole figure showing the normal vectors of the dislocation channel planes under different LOs. “A” and “B” for
the LO of [100] represent the two local shear bands shown in Fig. 2(a-3). (b) Experimental observation on the formation of dislocation channel, reprinted with permission
from [10].

points A and B is achieved, until the non-glide components of the level focuses on the dislocation motion behavior at a given slip sys-
applied stress tensor build up and decrease the kink-pair activa- tem, which is controlled by the irradiation resistance stress, the
tion enthalpy Gkp at point B, as shown in Fig. 5(f). The required applied shear stress and the non-Schmid effect. When the irradi-
engineering strain is about 2.2%, as marked by the dashed line in ation dose is high, the resistance stress induced by irradiation de-
Fig. 5(e). Afterward, the dislocation channel gradually appears, as fects is high. Therefore, a high applied shear stress is required to
shown in Fig. 5(d). obtain a positive effective shear stress τe f f . Because the magnitude
of the applied shear stress is positively correlated with the abso-
4.1.4. Regime Ⅲ: Multi-slip lute value of Schmid factor |M|, it is easier for the dislocations in
Regime III has the character of having multiple dominant slip the slip system with high |M| to active in irradiated tungsten. On
systems, which have a low kink-pair activation enthalpy Gkp and the other hand, the mobility of screw dislocation is controlled by
low maximum plastic contribution ratio ωI . The LO of [100] is the kink-pair mechanism. The kink-pair activation enthalpy Gkp
taken as an example. There are four equivalent slip systems, so represents the energy barrier to the dislocation motion. Therefore,
the plastic contribution ratio ωI of each dominant slip system is dislocation is prone to gliding in the slip system with the low
about 0.25, which is much lower than that with a single dominant Gkp . The non-Schmid effect manifests itself through the value of
slip system. This promotes the relatively homogeneous plastic de- Gkp . As consequence the slip system with the low |M| and low
formation, thus dramatically reducing the concentration of disloca- Gkp may serve as the dominant one for unirradiated tungsten. As
tion activities and irradiation defect destruction in each slip sys- for irradiated tungsten, the competitions between Schmid effect,
tem. During the initial deformation period, four equal slip systems non-Schmid effect, and irradiation barrier effect determine the dis-
contribute a similar amount to the total plastic strain. Afterward, location velocity for a given slip system. According to the analysis
one slip system gradually prevails over due to non-uniform defor- above, the absolute value of Schmid factor |M| and the kink-pair
mation in the transition zone of samples, just as shown in Fig. 5(h). activation enthalpy Gkp can serve as key physical quantities to
This process leads to the gradual appearance of a tiny dislocation control the dislocation mobility for a given slip system in high-
channel in Fig. 5(g). This process is consistent with the simulation dose irradiated tungsten.
results of the DDD observation [13] that the multi-slip mode can The second level focuses on the competition between differ-
gradually change into the single slip mode at a high irradiation ent slip systems. After the determination of dislocation mobility
dose. of each slip system, the dominant slip system can be determined.
The contribution of β th slip system to the plastic deformation can
4.1.5. Regime Ⅳ: Mixed-mechanisms be represented by the plastic contribution ratio ωβ , as defined in
Regime IV combines the mechanism of Regime II and Regime Eq. (27). Because the sum of ωβ for all slip systems is equal to one,
III. There are multiple dominant slip systems, which has a low the value of the highest ωβ (namely ω) will be low if there are
maximum plastic contribution ratio ωI but high kink-pair acti- multiple dominant slip systems, compared with the case having a
vation enthalpy Gkp . For the LO of [210], two equal slip sys- single dominant slip system. The simulation results have indicated
tems [1̄1̄1](101 ) and [111](1̄01 ) contribute most to the total plas- that the irradiated tungsten with multiple dominant slip systems
tic strain, inhibiting the formation of dislocation channel compared has better anti-deformation localization capability than that with a
with the single slip system dominated case. High Gkp calls for single dominant slip system. Therefore, the quantity of the highest
high effective stress for dislocation motion and makes it difficult plastic contribution ratio ω can also reflect the tendency of defor-
for the shear band to penetrate through the whole sample, com- mation localization and is adopted as another key physical quan-
pared with the case with a low Gkp . Therefore, no dislocation tity.
channel or shear band is observed in the considered strain range. Accordingly, the proposed four regimes can be distinguished ac-
cording to these three key physical quantities, as shown in Fig. 6(b-
4.1.6. Two levels of competition mechanisms for the proposed four e), respectively.
regimes
Based on the discussion above, the underlying mechanism of • Regime Ⅰ: The largest plastic contribution ratio ω is high, repre-
deformation localization regimes Ⅰ∼Ⅳ can be attributed to the two senting the plastic deformation is concentrated in less number
levels of competition mechanisms, as shown in Fig. 6(a). The first of dominant slip systems. The high absolute value of Schmid

7
Z. Li and Y. Cui Journal of Nuclear Materials 573 (2023) 154112

Fig. 5. Simulation results for typical loading orientation (a-c) [552] in Regime I, (d-f) [112] in Regime II, (g-i) [100] in Regime III. (b-c), (e-f), (h-i) give the evolution of plastic
slip rate, irradiation defect density, effective shear stress, and activation enthalpy of kink-pair for the points A and B, marked in (a), (d), (g), respectively. The vertical dot
line in (b-c), (e-f), (h-i) indicates the critical engineering strain for the initial formation of dislocation channels, which corresponds to the moment shown in (a), (d), (g),
respectively.

Fig. 6. Schematic for the two levels of competition mechanisms contributing to the four regimes of dislocation channels for irradiated tungsten. (a) Competition mechanisms
at the two levels and the corresponding key physical quantity. (b)∼(e) The four regimes of dislocation channels are schematically shown with the key physical quantity.

8
Z. Li and Y. Cui Journal of Nuclear Materials 573 (2023) 154112

factor |M| and low kink-pair activation enthalpy Gkp promotes and strain rate hardening. The results in Fig. 1b shows that ori-
the dislocation motion and the formation of dislocation chan- entation dependent SHR of unirradiated tungsten does not show
nels. a good correlation with orientation dependent localization of irra-
• Regime Ⅱ: |M| and ω are high. However, the high kink-pair acti- diated tungsten, because the SHR of irradiated tungsten is domi-
vation enthalpy Gkp inhibits dislocation motion, reducing the nated by the interaction between irradiation defects and disloca-
tendency to localized deformation. tions, such as described by Eq. (41) [11]. However, it is found that
• Regime Ⅲ: High |M| and low Gkp promote the dislocation mo- orientation-dependent strain rate sensitivity of unirradiated tung-
tion, but the low ω contributes to more homogeneous deforma- sten 1/m correlates well with the orientation-dependent deforma-
tion. tion localization in irradiated tungsten, as shown in Fig. 3. This is
• Regime Ⅳ: Compared to Regime Ⅲ, the high kink-pair acti- understandable, because the strain rate sensitivity at a given low
vation enthalpy Gkp restricts the dislocation motion on the temperature (T<0.2Tm , Tm is the melting temperature) is controlled
multiple dominant slip systems, further improving the anti- by the thermally activated screw dislocation mobility [63], while
localized deformation capacity of irradiated tungsten. deformation localization is also sensitive to the screw dislocation
activities as discussed in Section 4.1. Therefore, 1/m can be used
Therefore, the LOs in Regime Ⅰ and Regime Ⅳ have the worst
to estimate the deformation localization resistance of irradiated
and best capacity to suppress the formation of dislocation chan-
tungsten.
nels, and those in Regime Ⅱ and Regime Ⅲ have the medium ca-
pacity. They are plotted using different transparency in Fig. 3.
5. Conclusions

4.2. Deformation localization indicator Understanding, predicting, and controlling deformation localiza-
tion in irradiated material promises significant benefits for nuclear
Irradiation experiments on tungsten are expensive and time- applications. However, it is still a challenging task due to the dif-
consuming, especially for neutron irradiation. Therefore, it will ficulty in unfolding the various contributions. In the current work,
benefit a lot if some indicators can conveniently estimate the re- the loading orientation effect on deformation localization is sep-
sistance of irradiated tungsten to deformation localization. arated out and carefully examined in irradiated tungsten through
Based on the insights gained in Section 4.1, for heavily irradi- systematic crystal plasticity studies.
ated tungsten at relatively low temperature, deformation localiza- The loading orientation dependence of deformation localiza-
tion prefers to occur when the dominant slip system has great de- tion in irradiated tungsten is disclosed. It is found that the dom-
formation capacity and contributes more to the total deformation. inated slip system in unirradiated tungsten is controlled by the
This means the dominant slip system on the one hand has a lower non-Schmid effect, but the dislocation channel in irradiated tung-
activation enthalpy of kink-pair Gkp , and on the other hand leads sten forms along favorably oriented slip planes with the maxi-
to the higher plastic slip rate γ˙ so as to destruct the irradiation mum Schmid factor, because the resistance stress induced by the
defects more rapidly according to Eq. (26) at a given applied strain high number of irradiation defects dominates the dislocation ac-
condition. Accordingly, a dimensionless indicator , representing tivities. This finding can also help to understand the behavior of
the maximum slip rate ratio of the dominant slip system, is pro- irradiated polycrystalline tungsten, such as the preference of dis-
posed and defined as follows, location channel formation in the grains with higher Schmid fac-
 
 γ˙   ωI  tor [59] and the orientation-dependent indentation response of

= =
ε˙ 0    M 
(28) helium-implanted tungsten [55].
Careful analysis of the underlying mechanism shows that the
where ε˙ 0 is the applied strain rate, the plastic slip rate γ˙ and single-slip mode has the highest susceptibility to localization.
the coefficient  are quantities corresponding to the dominant slip Whereas if the multi-slip mode is activated, and at the same
system.  is a dimensionless factor describing the effect of high time the active slip systems are controlled by the high irradia-
activation enthalpy Gkp , which is set to 2.0 when the dominated tion resistance stress and hence have high activation enthalpy of
slip system with maximum plastic contribution factor ωI in irradi- kink-pair, the best capacity to suppress localization is observed.
ated tungsten is different from that with maximum plastic contri- Based on these analyses, two deformation localization indicators,
bution factor ωU in unirradiated tungsten, and otherwise is set to namely the maximum slip rate ratio and strain-rate sensitivity co-
1.0. Obviously, higher  has a higher propensity to deformation lo- efficient, are proposed. The predicted orientation-dependent de-
calization. A similar idea of connecting the relative slip system ac- formation localization extents using these two indicators agree
tivities with deformation localization was previously proposed in well with the deformation distribution features calculated by the
[19], but they did not consider the contribution of Gkp to dis- crystal plasticity model and the available experimental observa-
location motion and the subsequent dislocation channel formation. tions. This work provides the link between orientation-dependent
The calculation results of  according to Table 2 are given in Fig. 3, deformation localization and the key physical quantities, so as
which show good agreement with the trend of DLI with respect to to guide the prediction of deformation localization in irradiated
the LOs. This indicates that  can serve as an indicator to estimate materials.
the tendency of deformation localization of irradiated tungsten un-
der different LOs, which can be theoretically calculated. Declaration of Competing Interest
On the other hand, according to the Considere-Hart criterion,
the propensity of metal for deformation localization α can be es- The authors declare that they have no known competing finan-
timated as α = (−SHR/σy − 1/m ) [62,63]. Here, SHR means strain cial interests or personal relationships that could have appeared to
hardening rate. m is the strain rate sensitivity coefficient, which is influence the work reported in this paper.
defined as ∂ (lnσy )/∂ (lnε˙ ) [64,62], where σy is the yield strength,
and ε˙ is the applied strain rate. Under the uniaxial loading con- CRediT authorship contribution statement
dition, m is obtained by calculating a series of yield strength σy
at different given applied strain rates through the crystal plasticity Zhijie Li: Investigation, Formal analysis, Writing – original draft.
model. Deformation localization prefers to occur when α is high, Yinan Cui: Conceptualization, Supervision, Formal analysis, Writing
due to the suppression of the stabilizing mechanisms of strain – original draft, Writing – review & editing.

9
Z. Li and Y. Cui Journal of Nuclear Materials 573 (2023) 154112

Data Availability [27] G. Po, Y.N. Cui, D. Rivera, D. Cereceda, T.D. Swinburne, J. Marian, N. Ghoniem,
A phenomenological dislocation mobility law for bcc metals, Acta Mater. 119
(2016) 123–135.
Data will be made available on request. [28] Y.N. Cui, G. Po, N.M. Ghoniem, A coupled dislocation dynamics-continuum bar-
rier field model with application to irradiated materials, Int. J. Plast. 104 (2018)
54–67.
Acknowledgement [29] D. Brunner, Temperature dependence of the plastic flow of high-purity tung-
sten single crystals, Int. J. Mater. Res. 101 (2010) 1003–1013.
[30] R. Lowrie, A.M. Gonas, Single-crystal elastic properties of tungsten from 24 to
This material is based upon work supported by the National 1800 C, J. Appl. Phys. 38 (1967) 4505–4509.
Natural Science Foundation of China under Grant No. 12172194, [31] E.H. Lee, Elastic-plastic deformation at finite strains, J. Appl. Mech. 36 (1969)
1–6.
12222205, 12102217, and National key laboratory of shock wave
[32] D. Cereceda, M. Diehl, F. Roters, D. Raabe, J.M. Perlado, J. Marian, Unravel-
and detonation physics JCKYS2021212004. ing the temperature dependence of the yield strength in single-crystal tung-
sten using atomistically-informed crystal plasticity calculations, Int. J. Plast. 78
(2016) 242–265.
References [33] C. McAuliffe, H. Waisman, A unified model for metal failure capturing shear
banding and fracture, Int. J. Plast. 65 (2015) 131–151.
[1] D.J. Bacon, Y.N. Osetsky, Modelling dislocation–obstacle interactions in metals [34] M.E. Gurtin, A finite-deformation, gradient theory of single-crystal plasticity
exposed to an irradiation environment, Mat. Sci. Eng. A 400 (2005) 353–361. with free energy dependent on densities of geometrically necessary disloca-
[2] X. Hu, Recent progress in experimental investigation of neutron irradiation re- tions, Int. J. Plast. 24 (2008) 702–725.
sponse of tungsten, J. Nucl. Mater. (2022) 153856. [35] D. Cereceda, A. Stukowski, M.R. Gilbert, S. Queyreau, L. Ventelon,
[3] M.A.O. Vrielink, V. Shah, J.A.W. Van Dommelen, M.G.D. Geers, Modelling the M.C. Marinica, J.M. Perlado, J. Marian, Assessment of interatomic potentials for
brittle-to-ductile transition of high-purity tungsten under neutron irradiation, atomistic analysis of static and dynamic properties of screw dislocations in W,
J. Nucl. Mater. 554 (2021) 153068. J. Phys. 25 (2013) 085702.
[4] Y. Dai, X. Jia, J.C. Chen, W.F. Sommer, M. Victoria, G.S. Bauer, Microstructure of [36] Y.N. Cui, N. Ghoniem, G. Po, Plasticity of irradiated materials at the nano &
both as-irradiated and deformed 304L stainless steel irradiated with 800 MeV micro-scales, J. Nucl. Mater. (2020) 152746.
protons, J. Nucl. Mater. 296 (2001) 174–182. [37] H.W. Schadler, Mobility of edge dislocations on {110} planes in tungsten single
[5] E.H. Lee, T.S. Byun, J.D. Hunn, M.H. Yoo, K. Farrell, L.K. Mansur, On the origin crystals, Acta Metall. Mater. 12 (1964) 861–870.
of deformation microstructures in austenitic stainless steel: Part I - Microstruc- [38] G. Po, Y. Cui, D. Rivera, D. Cereceda, T.D. Swinburne, J. Marian, N. Ghoniem,
tures, Acta Mater. 49 (2001) 3269–3276. A phenomenological dislocation mobility law for bcc metals, Acta Mater. 119
[6] L. Vincent, L. Gelebart, R. Dakhlaoui, B. Marini, Stress localization in BCC poly- (2016) 123–135.
crystals and its implications on the probability of brittle fracture, Mat. Sci. Eng. [39] H. Mecking, U.F. Kocks, Kinetics of flow and strain-hardening, Acta Metall.
A 528 (2011) 5861–5870. Mater. 29 (1981) 1865–1875.
[7] P.O. Barrioz, J. Hure, B. Tanguy, Effect of dislocation channeling on void growth [40] I.J. Beyerlein, C.N. Tome, A dislocation-based constitutive law for pure Zr in-
to coalescence in FCC crystals, Mat. Sci. Eng. A 749 (2019) 255–270. cluding temperature effects, Int. J. Plast. 24 (2008) 867–895.
[8] M.D. McMurtrey, B. Cui, I. Robertson, D. Farkas, G.S. Was, Mechanism of dislo- [41] D. Terentyev, X.Z. Xiao, A. Dubinko, A. Bakaeva, H.L. Duan, Dislocation-medi-
cation channel-induced irradiation assisted stress corrosion crack initiation in ated strain hardening in tungsten: thermo-mechanical plasticity theory and
austenitic stainless steel, Curr. Opin. Solid St. M. 19 (2015) 305–314. experimental validation, J. Mech. Phys. Solids 85 (2015) 1–15.
[9] N.R. Barton, A. Arsenlis, J. Marian, A polycrystal plasticity model of strain lo- [42] A. Arsenlis, M. Rhee, G. Hommes, R. Cook, J. Marian, A dislocation dynamics
calization in irradiated iron, J. Mech. Phys. Solids 61 (2013) 341–351. study of the transition from homogeneous to heterogeneous deformation in
[10] S. Das, H.B. Yu, E. Tarleton, F. Hofmann, Hardening and strain localisation in irradiated body-centered cubic iron, Acta Mater. 60 (2012) 3748–3757.
helium-ion-implanted tungsten, Sci. Rep.-Uk 9 (2019) 1–14. [43] T.D. De la Rubia, H.M. Zbib, T.A. Khraishi, B.D. Wirth, M. Victoria, M.J. Caturla,
[11] Z. Li, Z. Liu, Z. Zhuang, Y. Cui, Temperature dependent deformation localization Multiscale modelling of plastic flow localization in irradiated materials, Nature
in irradiated tungsten, Int. J. Plast. (2021) 103077. 406 (20 0 0) 871–874.
[12] M.D. McMurtrey, G.S. Was, L. Patrick, D. Farkas, Relationship between localized [44] A.V. Korchuganov, K.P. Zolnikov, D.S. Kryzhevich, V.M. Chernov, S.G. Psakhie,
strain and irradiation assisted stress corrosion cracking in an austenitic alloy, MD simulation of plastic deformation nucleation in stressed crystallites under
Mat. Sci. Eng. A 528 (2011) 3730–3740. irradiation, Phys. Atom. Nucl. 79 (2016) 1193–1198.
[13] Y. Cui, G. Po, N. Ghoniem, Suppression of localized plastic flow in irradiated [45] A.A. Leino, G.D. Samolyuk, R. Sachan, F. Granberg, W.J. Weber, H.B. Bei, J. Liu,
materials, Scr. Mater. 154 (2018) 34–39. P.F. Zhai, Y.W. Zhang, GeV ion irradiation of NiFe and NiCo: insights from MD
[14] R. Santos-Güemes, B. Bellón, G. Esteban-Manzanares, J. Segurado, L. Capolungo, simulations and experiments, Acta Mater. 151 (2018) 191–200.
J. LLorca, Multiscale modelling of precipitation hardening in Al–Cu alloys: dis- [46] D. Rodney, G. Martin, Y.J.M.S. Bréchet, E. A, Irradiation hardening by interstitial
location dynamics simulations and experimental validation, Acta Mater. 188 loops: atomistic study and micromechanical model, Mat. Sci. Eng. A 309 (2001)
(2020) 475–485. 198–202.
[15] T. Byun, N. Hashimoto, Strain localization in irradiated materials, Nucl. Eng. [47] Y. Cui, G. Po, N.M. Ghoniem, A coupled dislocation dynamics-continuum bar-
Technol. 38 (2006) 619–638. rier field model with application to irradiated materials, Int. J. Plast. 104 (2018)
[16] A. Patra, D.L. McDowell, Continuum modeling of localized deformation in irra- 54–67.
diated BCC materials, J. Nucl. Mater. 432 (2013) 414–427. [48] D. Brunner, V. Glebovsky, The plastic properties of high-purity W single crys-
[17] T.O. Erinosho, F.P.E. Dunne, Strain localization and failure in irradiated zircaloy tals, Mater. Lett. 42 (20 0 0) 290–296.
with crystal plasticity, Int. J. Plast. 71 (2015) 170–194. [49] L.M. Garrison, Y. Katoh, N.A.P.K. Kumar, Mechanical properties of single-crys-
[18] A. Reichardt, A. Lupinacci, D. Frazer, N. Bailey, H. Vo, C. Howard, Z. Jiao, tal tungsten irradiated in a mixed spectrum fission reactor, J. Nucl. Mater. 518
A.M. Minor, P. Chou, P. Hosemann, Nanoindentation and in situ microcompres- (2019) 208–225.
sion in different dose regimes of proton beam irradiated 304 SS, J. Nucl. Mater. [50] P. Beardmore, D. Hull, Deformation and fracture of tungsten single crystals, J.
486 (2017) 323–331. Less-Common Met. 9 (1965) 168–180.
[19] F. Onimus, J.L. Bechade, A polycrystalline modeling of the mechanical behavior [51] R.G. Garlick, H.B. Probst, Investigation of room-temperature slip in
of neutron irradiated zirconium alloys, J. Nucl. Mater. 384 (2009) 163–174. zone-melted tungsten single crystals, Trans. Metall. Soc. Aime 230 (1964)
[20] F. Onimus, I. Monnet, J.L. Bechade, C. Prioul, P. Pilvin, A statistical TEM inves- 1120–1128.
tigation of dislocation channeling mechanism in neutron irradiated zirconium [52] R.M. Rose, D.P. Ferriss, J. Wulff, Yielding and plastic flow in single crystals of
alloys, J. Nucl. Mater. 328 (2004) 165–179. tungsten, Trans. Metall. Soc. Aime 224 (1962) 981–989.
[21] D. Terentyev, C.-.C. Chang, C. Yin, A. Zinovev, X.-.F. He, in: Neutron Irradia- [53] X.X. Hu, T. Koyanagi, M. Fukuda, N.A.P.K. Kumar, L.L. Snead, B.D. Wirth, Y. Ka-
tion Effects On Mechanical Properties of ITER Specification Tungsten, Tungsten, toh, Irradiation hardening of pure tungsten exposed to neutron irradiation, J.
2021, pp. 1–19. Nucl. Mater. 480 (2016) 235–243.
[22] S. Das, H.b. Yu, K. Mizohata, E. Tarleton, F. Hofmann, Modified deforma- [54] T. Koyanagi, N.A.P.K. Kumar, T. Hwang, L.M. Garrison, X.X. Hu, L.L. Snead,
tion behaviour of self-ion irradiated tungsten: a combined nano-indentation, Y. Katoh, Microstructural evolution of pure tungsten neutron irradiated with
HR-EBSD and crystal plasticity study, Int. J. Plast. 135 (2020) 1–32. a mixed energy spectrum, J. Nucl. Mater. 490 (2017) 66–74.
[23] D. Cereceda, M. Diehl, F. Roters, P. Shanthraj, D. Raabe, J.M. Perlado, J. Mar- [55] S. Das, H. Yu, E. Tarleton, F. Hofmann, Orientation-dependent indentation re-
ian, Linking atomistic, kinetic Monte Carlo and crystal plasticity simulations of sponse of helium-implanted tungsten, Appl. Phys. Lett. 114 (2019) 221905
single-crystal tungsten strength, GAMM-Mitteilungen 38 (2015) 213–227. (221901)-(221905).
[24] Z. Li, T. Wang, D. Chu, Z. Liu, Y. Cui, A coupled crystal-plasticity and phase-field [56] Y.N. Cui, G. Po, N. Ghoniem, Size-tuned plastic flow localization in irradiated
model for understanding fracture behaviors of single crystal tungsten, Int. J. materials at the submicron scale, Phys. Rev. Lett. 120 (2018) 1–6.
Plast. 157 (2022) 103375. [57] M.N. Gussev, K.G. Field, J.T. Busby, Deformation localization and dislocation
[25] L. Dezerald, D. Rodney, E. Clouet, L. Ventelon, F. Willaime, Plastic anisotropy channel dynamics in neutron-irradiated austenitic stainless steels, J. Nucl.
and dislocation trajectory in BCC metals, Nat. Commun. 7 (2016) 11695. Mater. 460 (2015) 139–152.
[26] H. Lim, C.R. Weinberger, C.C. Battaile, T.E. Buchheit, Application of generalized [58] M.D. McMurtrey, G.S. Was, B. Cui, I. Robertson, L. Smith, D. Farkas, Strain local-
non-Schmid yield law to low-temperature plasticity in bcc transition metals, ization at dislocation channel–grain boundary intersections in irradiated stain-
Model. Simul. Mater. Sci. 21 (2013) 1–23. less steel, Int. J. Plast. 56 (2014) 219–231.

10
Z. Li and Y. Cui Journal of Nuclear Materials 573 (2023) 154112

[59] A. Patra, D.L. McDowell, Crystal plasticity investigation of the microstructural [63] Q. Wei, K.T. Ramesh, E. Ma, L.J. Kesckes, R.J. Dowding, V.U. Kazykhanov, R.Z. Va-
factors influencing dislocation channeling in a model irradiated bcc material, liev, Plastic flow localization in bulk-tungsten with ultrafine microstructure,
Acta Mater. 110 (2016) 364–376. Appl. Phys. Lett. 86 (2005) 101907.
[60] K. Farrell, T.S. Byun, N. Hashimoto, Deformation mode maps for tensile de- [64] R. Huang, Q.J. Li, Z.J. Wang, L. Huang, J. Li, E. Ma, Z.W. Shan, Flow stress in sub-
formation of neutron-irradiated structural alloys, J. Nucl. Mater. 335 (2004) micron BCC iron single crystals: sample-size-dependent strain-rate sensitivity
471–486. and rate-dependent size strengthening, Mater. Res. Lett. 3 (2015) 121–127.
[61] S.J. Zinkle, B.N. Singh, Microstructure of neutron-irradiated iron before and af-
ter tensile deformation, J. Nucl. Mater. 351 (2006) 269–284.
[62] P. Srivastava, K. Jiang, Y.A. Cui, E. Olivera, N. Ghoniem, V. Gupta, The influence
of nano/micro sample size on the strain-rate sensitivity of plastic flow in tung-
sten, Int. J. Plast. 136 (2021) 102854.

11

You might also like