Multi-component solid solution and cluster hardening –Mn–Si alloys of Al Qinglong Zhao

advertisement
Materials Science & Engineering A 625 (2015) 153–157
Contents lists available at ScienceDirect
Materials Science & Engineering A
journal homepage: www.elsevier.com/locate/msea
Multi-component solid solution and cluster hardening
of Al–Mn–Si alloys
Qinglong Zhao a, Bjørn Holmedal a,n, Yanjun Li a, Espen Sagvolden b, Ole Martin Løvvik b
a
b
Department of Materials Science and Engineering, Norwegian University of Science and Technology, Trondheim N-7491, Norway
SINTEF Materials and Chemistry, Oslo N-0314, Norway
art ic l e i nf o
a b s t r a c t
Article history:
Received 19 September 2014
Received in revised form
28 November 2014
Accepted 2 December 2014
Available online 10 December 2014
Tensile tests on Al–Mn–Si ternary alloys show that a small amount of Si increases significantly the
strength compared to Al–Mn binary alloys with the same concentration of Mn. This cannot be explained
by classical theories for multi-element substitutional solid solution hardening under the assumption of
no interaction between different alloying elements. A new simplified cluster strengthening model which
addresses both the chemical and size misfit effects of atom dimers is proposed this work. The binding
energies and misfit of dimers were estimated by first principles atomistic simulations. The prediction
results of the model are reasonably consistent with the experimental results. It shows that the main
strengthening contribution is due to the misfit of dimers.
& 2014 Elsevier B.V. All rights reserved.
Keywords:
Aluminium alloys
Cluster
Strengthening mechanism
Modelling
1. Introduction
The interaction between solute atoms and dislocations is an
important strengthening mechanism for alloys. Solid solution
hardening in binary alloys has been extensively studied and its
principle has been well understood [1,2]. However, multi-element
solution hardening, which is essential for commercial alloys, has
been paid much less attention. The solute strengthening contribution from solute atoms of type i in binary alloys is denoted Δτi.
Under the assumption that there is no interaction between solute
atoms [3], the superposition of multi-element solution hardening
is
Δτp ¼ ∑ Δτpi
ð1Þ
i
Here 1 r pr2. Eq. (1) is a general superposition law at finite
temperatures and in the athermal limit [4]. The value p ¼1 is
justified when a high density of weak obstacles (e.g. solute atoms)
are mixed with strong ones such as forest dislocations and nonshearable precipitates [5]. The choice p ¼ 2 is suitable for obstacles
of similar strengths [5]. In literatures p ¼1.5 is often used for solute
strengthening [6–8], which can be obtained from Labusch theory [9].
Simulations suggest that the value of p depends on the strength and
density of obstacles in the athermal limit [10]. Eq. (1) describes well
the multi-element solution hardening in several ternary alloys, for
instance Cu–Si–Ge alloys [6], Pb–In–Tl alloys [7] and Mg–Gd–Y [11].
n
Corresponding author. Tel.: þ 47 735 94904; fax: þ47 735 50203.
E-mail address: bjorn.holmedal@ntnu.no (B. Holmedal).
http://dx.doi.org/10.1016/j.msea.2014.12.006
0921-5093/& 2014 Elsevier B.V. All rights reserved.
However, the predictions by Eq. (1) were found not to match
experimental results of Al–Fe–Si alloys [12]. This is probably because
the interactions between solute atoms in term of clustering cannot
be neglected.
Solute clustering is found to occur in many aluminium alloys
[13–17]. Clusters will influence the strength, for instance, the rapid
hardening in Al–Cu–Mg alloys during ageing is attributed to the
formation of co-clusters (containing various elements) [14,16,18].
Measurements of 3D atom probe tomography in [16,17] revealed
that the majority of atom clusters in a naturally aged AlMgSi alloy
and in as quenched or slightly aged AlMgCu alloys consisted of 2–4
atoms. Hence, it is common in theoretical analysis only to consider
the dimers. In ternary alloys the dimers can contain two atoms of
the same alloying element or one of each solute element.
Some models have been proposed to quantitatively explain
the hardening effect of clusters. A simplified model for modulus
hardening, which was originally proposed for precipitates [19], has
been adapted for co-clusters [18]. Starink et al. proposed an
analytic model, order strengthening model, considering the shortrange order of clusters in Al–Cu–Mg [18] and Al–Mg–Si [17]. They
demonstrated that the contribution of modulus hardening was
less than 10% of order strengthening. Zhao [20] recently proposed
a model that predicts that the elastic interactions between solute
atoms and dislocations are significantly increased if the atoms
are arranged in small few-atom clusters. The above-mentioned
models usually consider the interaction between clusters and
un-dissociated dislocations only. A more realistic picture should
include the localized influence of clusters on dislocation dissociation and stacking fault. Monte Carlo simulation suggested that
154
Q. Zhao et al. / Materials Science & Engineering A 625 (2015) 153–157
solute clustering in Al–Cu–Ag might affect extended dislocation
structure, increasing the separation between partial dislocations
[21]. The atomic-scale simulations of Proville et al. [22,23] suggested that the interaction between solute atoms/dimers and two
Shockley partials may not have the same magnitude, and that
solute dimers are significantly stronger obstacles than isolated
solute atoms for bypassing dislocations.
In the present work, the tensile stress of a set of Al–Mn binary
alloys and Al–Mn–Si ternary alloys are measured and to compared
with the prediction of Eq. (1). Atomistic calculations are performed
to estimate input parameters for simplified models for cluster
strengthening, which are applied to the Al–Mn–Si ternary alloys.
2. Experimental
Five alloys were cast, including the base material which was a
99.99% high purity Al (4NAl). First a pre-alloy was made where the
4NAl was mixed with 99.9% purity Mn in a ratio of 10:1. The prealloy was melted in a vacuum oven at a temperature of about
1673 K (1400 1C) and air cooled. At this temperature all the AlMn
phases were dissolved in the melt. A melting pot made of 99.7%
purity Al2O3 was applied to avoid pollutions. The pre-alloy and the
base material was then mixed in a larger melting pot made of
graphite clay but coated with Al2O3 and burned at 1473 K
(1200 1C) and further coated with boron nitride to reduce diffusion
of impurity elements into the melt. This was heated to about
1023 K (750 1C) to make sure that the AlMn6 phase in the pre-alloy
dissolved. A mould of 7 cm 7.5 cm 20 cm cooling from one end
was used for the final casting, where a uni-directional solidification structure with coarse grains with rotated [100] orientations
around the longitudinal direction of the ingot was achieved. The
chemical compositions were measured by a mass spectrograph by
Hydro R&D at Sunndalsøra and are listed in Table 1. The ingots
were cut into 20 mm thick slabs that were homogenized by fast
heating at 200 K/h to 908 K (635 1C), kept there for 16 h and finally
water quenched to ensure that all the Mn (and Si) were dissolved
in solid solution. Next the slabs were rolled at room temperature
by 16 passes of 1 mm thickness reduction each, in a laboratory
two-hi rolling mill with roll diameter of 205 mm to a thickness
of 4 mm. The thickness reduction provided stored deformation
energy for recrystallization, which was obtained by keeping
machined samples of the rolled material in a salt bath in 15 min
before water quenching. These specimens were used for tensile
testing at 78 K. The adequate temperatures of the salt baths were
determined by a series of hardness tests. The temperatures are
shown in Table 1, along with the measured grain sizes subsequent
to recrystallization. The temperatures were chosen only slightly
larger than the recrystallization temperature to avoid grain growth
and precipitation of Mn containing dispersoids. Electrical conductivity measurements for the AlMn1.0 alloy before and after the
recrystallization showed that the conductivity did not change.
The conductivity was measured using a Foerster Sigmatest and
remained 17.05 7 0.1 MS/m within the scatter of the
measurements. The electrical conductivity is sensitive to Mn,
indicating that the Mn level in solid solution was the same and
no
precipitation occurred. The AlMn0.25Si0.1 alloy was observed by
FEG-SEM, and no dispersoids were observed. The alloys for tensile
testing at room temperature were further rolled at room temperature from 4 mm to 1 mm thickness in three passes in the same
laboratory mill and then annealed to recrystallize. The details of
microstructure characterization have been described in [24].
Tensile specimens for the room temperature testing were
machined with a square cross section of 6 1 mm2 and gauge
length of 25 mm. The tensile tests performed in liquid nitrogen
had a cross section of 3 3 mm2 and a parallel region of 40 mm.
The tensile tests were performed in a servo hydraulic machine
under a constant ramp rate, giving a strain rate of 10 3 s 1 at
room temperature and in a screw machine at the same strain rate
at liquid nitrogen temperature (78 K). The tension tests at room
temperature had 3–6 parallel tests, whereas only one test of each
alloy was made at 78 K.
Flow stresses of Al–Mn binary alloys and Al–Mn–Si ternary
alloys at room temperature at various tensile strain levels are
plotted in Fig. 1a. The flow stress of the ternary alloy is significantly higher than the binary alloy containing the same amount of
Mn. The tensile tests performed at 78 K show the same phenomenon as illustrated in Fig. 1b.
3. Modelling
In the modelling part only the yield stress will be considered
in order to avoid further complexity in the mixing of strength
contributions from the dislocations created during work hardening. A part of the critical resolved shear stress is the Hall–Petch
contribution due to small grain sizes. In order to compare various
grain sizes, this part of the stress is corrected for in the analysis,
and the remaining part of the critical resolved shear stress, i.e.
Table 1
The compositions of alloys, temperatures for recrystallization (TREX), the grain
diameter (DG) and Taylor factor (M) of tensile specimens.
Alloy
Mn (wt%) Si (wt%) TREX (K) DG, 78 K (μm) DG, RT (μm) M
Al99.99
AlMn0.25
AlMn0.5
AlMn1.0
AlMn0.25Si0.1
AlMn0.5Si0.1
0.00
0.22
0.44
0.93
0.22
0.46
o 0.01
o 0.01
o 0.01
o 0.01
0.10
0.11
598
623
623
673
623
648
185
108
116
150
152
276
65
29
25
37
33
–
3.0
2.9
2.9
2.7
3.0
2.9
Fig.1. Flow stress at various strain levels as a function of Mn content in solid
solution (binary alloys in black; ternary alloys in red). Tensile tests were performed
(a) at room temperature and (b) at 78 K. (For interpretation of references to colour
in this figure legend, the reader is referred to the web version of this article.)
Q. Zhao et al. / Materials Science & Engineering A 625 (2015) 153–157
155
where D is the grain size from Table 1, and σ is the measured
tensile stress. At room temperature the Hall–Petch parameter k
equals 1.27 MPa mm1/2, as given by [25]. The precise k value at
78 K is not known, but the same value as for room temperature
was applied here. The Taylor factor M (Table 1) was calculated
from the texture measured by EBSD mapping (of larger than
1 mm2).
The increase of the critical resolved shear stress because of the
added amount of Mn and Si in solid solution results in a difference
between the critical resolved shear stresses of a coarse grained
alloy τMn þ Si and the critical resolved shear stress of the pure
aluminium τPA , given by
the Vienna ab initio simulation package (VASP) [28,29]. The PBE
generalized gradient approximation was employed [30], using a
plane wave basis set with an energy cut-off of 350 eV and the
projector augmented wave method to treat the core region [31].
Solutes were introduced into cubic super-cells consisting of 108
atoms (3 3 3 conventional cubic unit cells). Electronic selfconsistency required difference in energy between consecutive
iterations less than 10 6 eV. The density of integration points in
the Brillouin zone was less than 0.17 Å 1 in each direction
(3 3 3k points). The super cells were relaxed in size simultaneously as the atomic coordinates, and the relaxation criterion was
that the forces had to be smaller than 1 meV/Å.
Pairs of solutes based on Si and Mn were investigated. The
interaction energy between such pairs was defined as the difference between the electronic energy of a model with two solutes
(i.e. X and Y) and that of the sum of electronic energies of models
with single solutes, balanced by the number of Al atoms:
ΔτMn þ Si ¼ τMn þ Si τPA :
Eint ¼ EðAl106XY Þ þ EðAl108 Þ–EðAl107X Þ–EðAl107Y Þ
the critical resolved shear stress for coarse grained alloys, was
estimated from the flow stress by
1
k
τ¼
σ pffiffiffiffi ;
ð2Þ
M
D
ð3Þ
This strength contribution will first be modelled from classical
theory without solute interactions. Next atomistic modelling is
presented and models taking into account these modelling results
are included.
3.1. Modelling without solute interactions
Eq. (1) applied to estimate the solid solution hardening of
ternary AlMnSi alloys, can be written.
ΔτMn þ Si p ¼ ΔτMn p þ ΔτSi p :
ð4Þ
Here ΔτMn and Δτ Si are the Mn and Si contributions. Based on the
tensile tests in Fig. 1 for binary AlMn alloys and by the use of
Eq.(2), the contribution to the strengthening from Mn in solid
solution was estimated to be ΔτMn ¼3.6 MPa for the case of
AlMn0.25Si0.1. The solid solution contribution from Si was extrapolated from a work on high-purity Al–Si binary alloys [26] as
ΔτSi 0.9 MPa with the Labusch theory applied. Alternatively ΔτSi
can be estimated to be in the range from 0.8 MPa to 1.2 MPa from
the tensile testing of water quenched or air cooled Al–Si alloys
reported in [27]. In this work ΔτSi ¼1.2 MPa is used as an upper
estimate. The pure solid solution strengthening under the assumption of no interactions is calculated from Eq. (4) as Δτ ¼ 4 MPa
using p ¼1.5. This value is smaller than the measured value
5.8 MPa, indicating the assumption of no interaction is questionable. Therefore, solute interaction/clustering is considered.
3.2. Atomistic modelling
ð5Þ
Eint is a function of the distance between the solutes within the
cluster, displayed in Fig. 2. Negative energies indicate attraction
between the atoms in a pair of solutes.
As can be seen from this figure, all pairs of solute atoms display
attraction at some level. The weakest attraction is between Si
atoms, where only 2nd nearest neighbours are attracted to each
other with |Eint| 41 kJ/mol. Si–Mn interactions are on the other
hand much stronger, with an attraction of more than 7 kJ/mol for
2nd nearest neighbours. Pairs of Mn are repulsive between 2nd
and 4th neighbours but attractive between nearest and 3rd
neighbours. There is thus a thermodynamic driving force towards
formation of clusters based on all the solute pairs that we studied.
The volume misfit ΔV was calculated from the definition of the
bulk modulus B:
ΔV ¼ ΔPV=B;
ð6Þ
where the pressure difference ΔP of a reference volume V (the
super cell used in the calculation) was calculated using VASP. The
experimental value of 79.4 GPa at 0 K was used for B [32]; this is
quite similar to the DFT calculated value of 78 GPa [33]. Three
different concentrations of solutes were investigated for each
impurity; 1, 2, and 4 solutes per super cell. The pressure difference
of a single solute was calculated from a linear fit to ΔP as a
function of solute concentration, as shown in Fig. 3. The calculated
misfit volume was 3.0 Å3 for Si, 14.2 Å3 for Mn, and 16.8 Å3 for the
Si–Mn pair. The latter value is rather close to the sum of misfit
volumes for the two constituent solutes.
Interaction energy (kJ/mol)
Interactions between solutes were quantified from first principles calculation using density functional theory as implemented in
8
6
4
2
0
-2
-4
-6
-8
-10
Si-Si
Si-Mn
Mn-Mn
1
2
3
Neighbour #
4
∞5
Fig. 2. The interaction energy Eint between the atoms in pairs of Si and Mn solutes.
The energy is in kJ/mol per solute pair, and is relative to solutes infinitely separated.
Fig. 3. Pressure versus solute atoms in a 108-atom super-cell for Si, Mn solutes and
Mn–Si dimers. The dashed line is linear fit for Mn–Si dimers, which slope is used in
Eq. (6).
156
Q. Zhao et al. / Materials Science & Engineering A 625 (2015) 153–157
3.3. Strengthening due to chemical effect
3.4. Strengthening due to size misfit of clusters
The clusters are formed due to the interaction between solute
atoms. A simplified model for chemical strengthening adapted
here is similar as described in [17,18]. A dislocation is located
between two {111} planes and the two atoms of the dimer will be
in one of these two planes or with one atom in each of them.
The clusters will in some cases be sheared apart by a dislocation
passing by, i.e. when the relative position of the two atoms in a
dimer is changed from first nearest neighbours to the second or
third nearest neighbours. If both the dimer atoms are located in
the {111} plane either below or above, then the dimer is simply
shifted one Burgers vector subsequent to the dislocation has
passed it. In some of the cases with one atom in each of the two
{111} planes the dimer can still be sheared into a similar configuration as before, say into another closest neighbour configuration.
Based on results from first principle calculations in last section,
only dimers of two atoms that are either closest or second closest
neighbours will be considered. Guided by the atomistic binding
energies, only dimers consisting of one Si atom and one Mn atom
are considered here.
The breaking of a cluster will require extra energy ΔEs, which is
the energy barrier of the transition, i.e. the activation energy. The
value of ΔEs is difficult to determine experimentally, so atomistic
simulations are required. Ideally ΔEs corresponds to a saddle point
between the initial and the subsequent configuration. However,
the difference between the energies of each configuration is much
simpler to calculate from atomistic simulations and will here be
used as first estimates for ΔEs. In theory this energy is lowered by
the cases where a dimer is formed by that two atoms are brought
together by the dislocation. In a dilute alloy this contribution is
small and neglected. With increased strain many of the dimers
become sheared apart, which might affect the work hardening, but
this effect is not included in the modelling here.
The strengthening due to shearing clusters apart is clearly
different from classical order strengthening, which creates an antiphase boundary, and from chemical strengthening, which creates
two ledges of interface. The work per slip area done by a gliding
dislocation is W ord ¼ τord b. This work is assumed to equal the
change in binding energy due to shearing apart dimers across the
adjacent two slip planes. This energy equals the energy change per
broken dimer ΔEs, multiplied with the number of such dimers per
slip area. The dimers crossing the slip plane contains one atom
from each adjacent plane, so the number of dimers per slip area
equals the number of atoms per area that are part of such dimers
in any of the two adjacent atom planes. Let the concentration of
Mn atoms being part of any closest or second closest neighbour
dimer is ccl . A certain fraction P sh of the dimers are modified by the
gliding dislocation. The work energy balance gives
Zhao [20] proposed a model considering the elastic effect of
size misfit of small clusters, which is adapted from solute
strengthening theory. Basic assumptions are that the volume
misfit of a small cluster is assumed to be approximately the sum
of misfits of all the atoms in the cluster and the clusters can be
treated as point obstacles. Since the atomistic simulations confirmed the assumption that the misfits can be added for the
dimers considered in this investigation, the model of Zhao should
be applicable. The methodology by Leyson et al. [8] to predict
solute strengthening by combining the Labusch theory with firstprinciples calculations, are adapted here to estimate the strengthening due to the volume misfit of dimers in the matrix. It is
assumed that since the size of a dimer is small, its misfits can be
treated similar as single substitutional atoms.
According to Ref. [8], solute strengthening at 0 K is approximately given by
Δτord ¼
ΔE s
2c P
bA111 cl sh
Δτy0 A1 ΔV 4=3 c2=3
ð8Þ
Here ΔV is the volume misfit; c is atomic fraction of solutes; A1 is a
constant, 31.1 MPa Å 4. The yield stress will be reduced due to
the atomic thermal motion at elevated temperature. The temperature dependence is given by
"
#
kT
ε_ 0 2=3
Δτy ðTÞ ¼ Δτy0 1 exp
ln
ð9Þ
ΔEb ε_
or
Δτy ðTÞ ¼ Δτy0 exp kT
ε_ 0
;
ln
ε_
0:51ΔEb
ΔEb A2 ΔV 2=3 c1=3
ð10Þ
ð11Þ
T is temperature; k is Boltzmann constant; ΔEb is the energy
barrier; A2 and ε_ 0 are constants, A2 ¼ 1.31 eV Å 2 and
ε_ 0 ¼10 4 s 1. Eq. (9) is used for low temperature such as 78 K,
and Eq. (10) for more elevated temperature.
In this work, Eqs. (8)–(11) are adapted for misfit strengthening
of dimers. The dimer is treated similar to a single atom. Its volume
misfit is assumed to be approximately the sum of misfits of the
two atoms in the dimer,
ΔV cl ΔV Mn þ ΔV Si :
ð12Þ
The volume misfit of a cluster does not increase linearly with the
number of atoms per cluster, but linearity can be an approximation
for very small clusters. The atomistic simulation in Section 3.2
has verified Eq. (12). A combination of Eq. (12) with Eqs. (8)–(11)
provides an estimate of the misfit strengthening, with the concentration c in Eqs. (8) and (11) replaced by the cluster concentration ccl.
ð7Þ
pffiffiffi 2
Here A111 ¼ 3b =2 is the {111} area occupied per atom. The
statistical chance that an arbitrarily chosen atom at one of the
adjacent {111} planes is Mn or Si and part of a dimer equals 2ccl .
There are 12 different closest neighbour dimer configurations and
additional six second closest neighbour configurations to choose,
from which only three of each crosses the glide plane. When a
dislocation is gliding between these two planes, one of the closest
neighbour dimers would remain a closest neighbour, one would
become a second closest neighbour and the third would become
third closest neighbour. One second closest neighbour would
become closest neighbour while the two others would become
more distant. Hence three out of 18 dimers are modified from first
or second to more distant neighbours, i.e. P sh ¼ 1=6.
4. Application of the models
As shown in Section 3.2, Mn and Si attract each other with
interaction energies of 6–8 kJ/mol at the first and second nearest
neighbour sites. The attractions between Si–Si pairs and between
Mn–Mn pairs are much weaker, smaller than 2 kJ/mol for nearest
and 2nd nearest neighbours. The strong attraction between dimers
of Mn and Si suggests that these are favourable clusters, while
Si–Si and Mn–Mn pairs are less likely to occur. As a simplification, only Mn–Si pairs were thus considered as clusters in the
present work.
The interaction energy of Mn–Si pairs at the first nearest
neighbour site was found to be 6.4 kJ/mol from the first
principle calculations, i.e. the enthalpy of formation of Mn–Si pairs
Q. Zhao et al. / Materials Science & Engineering A 625 (2015) 153–157
157
5. Conclusion
It is found that the yield strength of an Al–Mn–Si ternary alloy
is larger than the multi-element hardening prediction under
the assumption of no interaction between alloying elements. This
implies that solute clustering may occur and contribute to higher
strengthening. Models for cluster strengthening were applied,
including chemical and elastic effects of clusters based on simplified assumptions and input from atomistic simulations. The predictions are reasonably consistent with experimental results.
The models should be applicable to other alloys containing
clusters. Atomistic simulations are required to further develop
the models for cluster strengthening. It is concluded that misfit
strengthening due to clusters makes the major contribution in the
investigated case.
Fig. 4. The predicted shear stress of Al–Mn (open symbols) and Al–Mn–0.1Si (filled
symbols) versus experimental values. Squares represent room temperature, triangles represent 78 K.
from a random solution. It is reasonable to assume that Mn–Si
clusters form and make a significant contribution to the strength.
The interaction energy of Mn–Si pairs at the second and third
nearest neighbour sites were found to be 7.5 kJ/mol and 0.6 kJ/
mol (repulsive), respectively, from the atomistic calculations. For
the sake of simplicity the value of ΔEs was set as 7 kJ/mol for both
first and second closest neighbour dimer configurations in
this work.
As the most extreme case it was assumed that all the Si atoms
formed clusters (pairs) with Mn atoms, i.e. spending all Mn and Si
atoms giving an atomic fraction ccl ¼ 0.11 at%. The experimental
strain rate was ε_ ¼ 10 3 s 1 . The volume misfits of Mn and Si for
misfit strengthening are taken from [8]. The total strengthening
becomes the sum of misfit strengthening and order strengthening,
given by
Δτclu ¼ Δτmis þ Δτord :
ð13Þ
The predicted values of ΔτMn þ Si of cluster strengthening in Al–
Mn–Si ternary alloys and pure solute strengthening in the Al–Mn
alloys are plotted in Fig. 4. Here only the measured yield strength
variations are considered. For the experiments performed at 77 K
the smallest strain of 0.01 are used as an estimate of the yield
strength. The results are qualitatively valid also for flow stress at
larger strains, but then also work hardening must be accounted
for. For cluster strengthening in Al–Mn–Si ternary alloys, the
contribution due to the binding energy of dimers is only
0.2 MPa. It can be seen that the misfit strengthening plays the
major role in the ternary alloy, while the cutting of the dimer
hardly contribute. The trend line of Al–Mn binary alloys illustrates
the prediction ability of the solute strengthening model [8].
As noted in Ref. [8], the prediction of Eq. (8) deviates up to 730%.
The main contribution from the clusters follows from the same
equation by replacing single atoms by dimers, and then also the
predicted value of ΔτMn þ Si fits well this trend line.
Acknowledgement
This work is funded by the Research Council of Norway, Hydro
and Sapa under the KMB project No. 193179/I40.
References
[1] P. Haasen, in: F.R.N. Nabarro (Ed.), Dislocations in Solids, North-Holland, New
York, 1979, pp. 155–189.
[2] M.Z. Butt, P. Feltham, J. Mater. Sci. 28 (1993) 2557–2576.
[3] L.A. Gypen, A. Deruyttere, J. Mater. Sci. 12 (1977) 1028–1033.
[4] R.C. Picu, R. Li, Acta Mater. 58 (2010) 5443–5446.
[5] U.F. Kocks, A.S. Argon, M.F. Ashby, Prog. Mater. Sci. 19 (1975) 1–281.
[6] J. Friedrichs, P. Haasen, Philos. Mag. 31 (1975) 863–869.
[7] L.A. Gypen, A. Deruyttere, J. Mater. Sci. 12 (1977) 1034–1038.
[8] G.P.M. Leyson, L.G. Hector Jr., W.A. Curtin, Acta Mater. 60 (2012) 3873–3884.
[9] R. Labusch, Phys. Status Solidi 41 (1970) 659–669.
[10] Y. Dong, T. Nogaret, W.A. Curtin, Metall. Mater. Trans. A 41 (2010) 1954–1960.
[11] L. Gao, R.S. Chen, E.H. Han, J. Alloy. Compd. 481 (2009) 379–384.
[12] Q. Zhao, M. Slagsvold, B. Holmedal, Scr. Mater. 67 (2012) 217–220.
[13] T. Kosugi, T. Kino, J. Phys. Soc. Jpn. 53 (1984) 3837–3842.
[14] S.P. Ringer, K. Hono, T. Sakurai, I.J. Polmear, Scr. Mater. 36 (1997) 517–521.
[15] L. Reich, S.P. Ringer, K. Hono, Philos. Mag. Lett. 79 (1999) 639–648.
[16] R.K.W. Marceau, G. Sha, R. Ferragut, A. Dupasquier, S.P. Ringer, Acta Mater. 58
(2010) 4923–4939.
[17] M.J. Starink, L.F. Cao, P.A. Rometsch, Acta Mater. 60 (2012) 4194–4207.
[18] M.J. Starink, S.C. Wang, Acta Mater. 57 (2009) 2376–2389.
[19] P. Gomiero, Y. Brechet, F. Louchet, A. Tourabi, B. Wack, Acta Metall. Mater. 40
(1992) 857–861.
[20] Q. Zhao, Scr. Mater. 84–85 (2014) 43–46.
[21] C. Lane Rohrer, J. Mater. Res. 10 (1995) 578–590.
[22] L. Proville, D. Rodney, Y. Brechet, G. Martin, Philos. Mag. 86 (2006) 3893–3920.
[23] L. Proville, S. Patinet, Phys. Rev. B 82 (2010) 1–19.
[24] Q. Zhao, B. Holmedal, in: H. Weiland, A.D. Rollett, W.A. Cassada (Eds.), ICAA13:
Proceedings of the 13th International Conference on Aluminum Alloys, John
Wiley & Sons Inc., Pittsburgh, 2012, pp. 1825–1829.
[25] N. Hansen, Acta Metall. 25 (1977) 863–869.
[26] G.J. Mahon, G.J. Marshall, J. Miner. Met. Mater. Soc. 48 (1996) 39–42.
[27] Q. Zhao, B. Holmedal, Mater. Sci. Eng., A 563 (2013) 147–151.
[28] G. Kresse, J. Furthmuller, Phys. Rev. B 54 (1996) 11169–11186.
[29] G. Kresse, J. Hafner, Phys. Rev. B 47 (1993) 558–561.
[30] G. Kresse, D. Joubert, Phys. Rev. B 59 (1999) 1758–1775.
[31] P.E. Blochl, Phys. Rev. B 50 (1994) 17953–17979.
[32] G.N. Kamm, G.A. Alers, J. Appl. Phys. 35 (1964) 327–330.
[33] P.H. Ninive, K.E. Moen, A. Strandlie, O.M. Løvvik, Unpublished, 2014.
Download