Theory of Strengthening in FCC High Entropy Alloys Pre-Print 1

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176..

The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

Theory of Strengthening in fcc High Entropy Alloys

Céline Varvennea,1,∗, Aitor Luquea , William A. Curtina


a Laboratory for Multiscale Materials Modelling, Institute of Mechanical Engineering, École Polytechnique Fédérale de
Lausanne, Lausanne CH-1015, Switzerland

Abstract
High Entropy Alloys (HEAs) are a new class of random alloys having impressive strength and toughness.
Here, a mechanistic, parameter-free, and predictive theory for the temperature-, composition-, and strain-
rate-dependence of the plastic yield strength of fcc HEAs is presented, validated, and applied to understand
recent experiments. To first order, each elemental component in the HEA is considered as a solute embedded
in the effective matrix of the surrounding alloy. Strengthening is then mainly achieved due to dislocation
interactions with the random local concentration fluctuations around the average composition. The theory
is validated against molecular simulations on model Fe-Ni-Cr alloys. Hall-Petch-corrected yield strengths
in Ni-Co-Cr-Fe-Mn fcc HEAs are then predicted using only available experimental information, and good
quantitative agreement is achieved. The theory demonstrates the origins of the high strength and detailed
trends with composition, materials parameters, temperature, thus identifying the key measurable/calculable
material properties needed for design and optimization of fcc HEAs, and is a general model for fcc random
alloys.
Keywords: High Entropy Alloys, Mechanical properties, Solution Strengthening theory, Yield stress,
Molecular Simulations

1. Introduction the strengthening of HEA alloys remain unknown


[2, 3, 4, 5]. More broadly, although the behavior
High Entropy Alloys (HEAs) are random solid- of random alloys has been a major topic in metal
solution alloys with many components. For an N - physics over the last fifty years, there is no predic-
component alloy, the nominal composition is 1/N tive theory for alloy strength with high elemental
for each component, with the disorder presumed concentrations.
to facilitate fabrication of single phase materials Here, we develop and validate a mechanistic the-
[1, 2, 3]. Recently fabricated fcc HEAs have very ory for the yield stress of general fcc random al-
high tensile strength at low temperatures and high loys at arbitrary composition, thus including fcc
retained strength at elevated temperatures [4, 5]. HEAs, and we demonstrate its predictive capabil-
Furthermore, these fcc HEAs can have high ten- ity for the well-studied fcc Ni-Co-Fe-Cr-Mn alloys
sile elongation/ductility and exceptional fracture [4, 5, 8]. The theory rationalizes experimental re-
toughness [1, 2, 6]. The spectrum of impressive sults, quantitatively and qualitatively, and identi-
mechanical properties makes this new broad class fies the underlying material properties that control
[7] of structural materials attractive for many ap- strength, thus also serving as a basis for design and
plications. However, the physical mechanism(s) of optimization of new fcc HEAs.
The remainder of this paper is organized as fol-
∗ Corresponding author lows. Section 2 introduces some basic concepts for
Email address: varvenne@univ-mrs.fr (Céline our model of random alloys. Section 3 presents the
Varvenne) solute strengthening theory. Section 4 discusses val-
1 Present address: Centre Interdisciplinaire des
Nanosciences de Marseille, Aix-Marseille University-
idation of the concepts and model using molecular
CNRS, Campus de Luminy, case 913, Marseille F-13288, simulations. Section 5 presents a reduced model
France based on elasticity. In Section 6, we apply the
Preprint submitted to Acta Materialia March 22, 2016
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

reduced model to predict strengthening in the Ni- to local variations in solute chemical environment
Co-Fe-Cr-Mn alloys and compare with experiments. and structure around the average matrix environ-
Further discussions on implications and application ment.
of the theory are given in section 7. Several ap- Strengthening in the true random alloy, i.e. in-
pendices contain valuable derivations or additional creased stress required to move a dislocation, arises
detail that expand on results shown in the main from the totality of the interaction energies between
text. the solutes and an individual dislocation. The ef-
fective medium approximation for the matrix allows
2. Key Model Concepts us to consider a straight dislocation in that matrix,
surrounded by solutes. The interaction energy for a
We consider an N -component random fcc solute of type n located at position (xi , yj , zk ) rel-
th
HEA
PN with concentration cn of the n element ative to the center of a straight dislocation lying
( n=1 cn = 1). Typical HEAs have many el- along z is denoted as U n (xi , yj , zk ) (Fig.1a,b), and
emental components at high concentrations and may depend on the local environment of the type
with local structural/compositional disorder, and n solute. Contributions to this interaction energy
so analytical theories of solute strengthening for arise from the elastic interaction of the dislocation
dilute alloys with non-interacting solutes [9, 10, stress field with the misfit strain tensor of the so-
11, 12, 13, 14] would not seem to apply. How- lute and from chemical-specific interactions when
ever, to first order, each elemental alloy compo- the solute lies in the dislocation partial cores or
nent can be seen as a “solute” embedded in the along the stacking fault [13, 24, 25, 26, 27]. The
average effective medium “matrix” of the surround- random distribution of solutes in the lattice gives
ing material; such an effective medium approxima- rise to local fluctuations in solute concentrations
tion is well-established in different contexts such and distortions, and the dislocation is attracted
as in electronic structure theory, with the Vir- to energetically-favorable fluctuations and repelled
tual Crystal and Coherent Potential Approxima- by energetically-unfavorable fluctuations. When a
tions [15, 16, 17, 18, 19, 20], and in Embedded Atom straight dislocation segment of length ζ at initial
Method potentials [21, 22]. position x = 0 glides over a distance w, the change
We thus define here an average reference material in the position of the dislocation relative to the so-
for the HEA that has all the average properties of lutes leads to a potential energy change of
the true alloy: lattice constant a, elastic constants XX h
{Cij } including shear modulus µ and Poisson’s ra- ∆Utot (ζ, w) = snijk U n (xi − w, yj , zk )
tio ν, and stable/unstable stacking fault energies i,j,k n
i
γSF and γUSF , all of which depend on the average − U n (xi , yj , zk ) , (1)
alloy composition. The gliding {111}(110) dislo-
cations in the effective fcc matrix are like disloca- where snijk = 1 if a type-n solute is at position
tions in√elemental fcc metals, with a Burgers vector (xi , yj , zk ) and 0 otherwise. The magnitude of the
b = a/ 2, dissociation into two Shockley partials typical energy decrease when the dislocation seg-
separated by an intrinsic stacking fault, and gliding ment ζ moves into a region of favorable solute fluc-
at very low Peierls stresses [22, 23]. Each individ- tuations is the standard deviation of the potential
ual solute then interacts with the dislocation in the energy change σ∆Utot (ζ, w), given by
average matrix, shown schematically in Fig. 1a,b,
thus accounting for the solute interactions with the
D E D E2  21
2
average chemical neighborhood. The chemically- σ∆Utot (ζ, w) = ∆Utot (ζ, w) − ∆Utot (ζ, w) .
controlled (non-Hall-Petch) strengthening in HEAs (2)
is then primarily the strengthening of a random
solid solution with 100% solute concentration in the σ∆Utot (ζ, w) can be computed for a random dis-
average matrix. This fundamental approximation tribution of solutes, as shown in Appendix A. In
of solutes in an effective matrix is the basis for our particular, since the straight dislocation segment of
model of solute strengthening at arbitrary solute length ζ is invariant along the line direction z, an
concentrations. Moreover, while the model starts average over local variations in the interaction ener-
from the effective medium approximation, the de- gies due to both local distortions and local chemical
velopment below includes the additional effects due environments among the sites zk can be rigorously
2
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

performed (using a mean field approximation). The


final outcome of the averaging process can be ex-
pressed in the form
A

B   12
ζ
C σ∆Utot (ζ, w) = √ ∆Ẽp (w), (3)
3b

with
" 
X n n
2
∆Ẽp (w) = cn U (xi − w, yj ) − U (xi , yj )
i,j
n
# 21
2
+ σ∆U n
ij
. (4)

(a) real HEA n


Here, U (xi , yj ) is the average value of
U n (xi , yj , zk ) over the local environments along
the line zk at in-plane position (xi , yj ), and
σ∆Uijn is the associated standard deviation, which
embodies the contributions due to the local envi-
ronment/structural fluctuations along zk . ∆Ẽp (w)
is the key quantity for strengthening in the theory
presented in the next section.

3. Solution Strengthening Model


(b) effective medium
The theory for solution strengthening follows the
lines of a predictive model for dilute alloys [12, 27,
28]. Here, we present the important steps in the
derivation, highlighting key features that emerge in
the context of HEAs.

3.1. Minimum-energy dislocation configuration

A long straight dislocation of length L will re-


duce its total potential energy by adopting a wavy
(c) Burgers vector distribution configuration, where some segments of character-
istic length ζc reside in regions of favorable solute
Figure 1: Effective medium approach for dislocation/solute
fluctuations. These segments lie at the minima in a
interactions. (a), Fully-random 3-component HEA con-
taining a dissociated edge dislocation; (b), Effective ma- potential energy landscape having typical minima
trix material of the same alloy, with an embedded A and maxima spaced by some glide distance wc . The
“solute” at position (xi , yj , zk ) relative to the dislo- pinned segments ζc are connected through addi-
cation centered at the origin, with interaction energy
U A (xi , yj , zk ); (c), Normalized Burgers vector distribution
tional segments ζc (Fig. 2). Dislocation bowing has
−1

x−d/2
2
−1

x+d/2
2
an energy cost ∆ELT (ζc , wc ) = Γwc2 /2ζc (wc  ζc ,
db
dx
∼ e 2 σ
+e 2 σ
along the glide plane verified ex post facto), where Γ is the dislocation
of edge dislocation in the effective matrix, with d the dis-
location dissociation distance and σ describing the partial
line tension in the effective matrix.
spreading (see Appendix D). To determine the length scales ζc and wc , we first
consider the total energy for arbitrary ζ and w, and
then minimize that energy. The total energy is the
3
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

sum of the potential energy and elastic bowing en- favourable to an unfavorable potential energy fluc-
ergy, and for arbitrary ζ and w is given by tuation over a distance wc , minus the gain in the
  line energy ∆ELT ,
L
∆Etot (ζ, w) =∆ELT (ζ, w) − σ∆Utot (ζ, w)
2ζ ∆Eb =∆Eb0 − ∆ELT ,
"  12 # 
! 31
w2

ζ L wc2 Γ∆Ẽp2 (wc )
= Γ − √ ∆Ẽp (w) .
2ζ 3b 2ζ =1.22 . (8)
b
(5)

σ∆Utot (ζ, w) is the typical energy decrease by al- 3.2. Thermally-activated glide
lowing the dislocation segment ζ to move into a re- To glide, the dislocation must overcome the bar-
gion of favourable solute fluctuations, which is given rier ∆Eb by thermal activation. This is facilitated
by Eq. (3). ∆Etot is minimized with respect to ζ by the work −τ bζc x of an applied resolved stress
and w to determine the characteristic configuration τ done on the length ζc segment as it glides a dis-
(ζc , wc ). Minimization of Eq. (5) with respect to ζ tance x relative to the minimum energy position.
is analytic and yields For the sinusoidal energy well, the stress-dependent
!1 energy barrier is then accurately described [11] by
√ Γ2 w 4 b 3 the asymptotic result
ζc (w) = 4 3 . (6)
∆Ẽp2 (w)  3
τ 2
Subsequent minimization with respect to w is per- ∆E(τ ) = ∆Eb 1 − (9)
τy0
formed numerically, and reduces to the solution of
where τy0 is the zero-temperature flow stress, i.e.
the stress at which the energy barrier vanishes, and
∂∆Ẽp (w) ∆Ẽp (w)
= . (7) is given explicitly as
∂w 2w
! 13
Consequently, the minimized wc value is defined en- π ∆Eb ∆Ẽp4 (wc )
tirely by the potential energy function ∆Ẽp (w); this τy0 = = 1.01 . (10)
2 bζc (wc )wc Γb5 wc5
feature is important for validation using molecular
simulations, as discussed below. The dislocation At stresses τ < τy0 , and for quasi-static load-
thus adopts a waviness at a particular amplitude ing, the plastic strain-rate ε̇ is related to the energy
wc determined by the potential energy function and barrier through a thermally-activated Arrhenius
the long dislocation line then finds the value of ζc [11, 29] model: ε̇ = ε̇0 exp (−∆E(τ )/kT ). Using
that minimizes the total energy for a given line ten- the stress-dependent energy barrier of Eq. (9), we
sion Γ. invert this relation to obtain the finite-temperature,
In the minimum energy configuration, each seg- finite strain-rate flow stress τy (T, ε̇) as
ment ζc along the wavy dislocation line sits in a
local potential energy well. The depth of the well,  32 #
" 
kT ε̇0
relative to zero energy, is −σ∆Utot , and the width of τy (T, ε̇) = τy0 1 − ln . (11)
∆Eb ε̇
the energy well (min to max) is wc . This local po-
tential energy well is approximated as a sinusoidal
Here, ε̇0 is a reference strain-rate that is nomi-
function: with the minimum located at x = 0 along
nally connected to the dislocation density ρ, Burg-
the glide plane, the energy of the
h segment of ilength
∆E 0 πx
ers vector b, typical dislocation slip distance ds and
ζc at position x is E(x) = 2 b 1 − cos w , with attempt frequency ν0 via the Orowan relationship
0
√ c

∆Eb = 2σ∆Utot L/2ζc . The potential √ energy bar- ε̇0 = ρbds ν0 . The value of ε̇0 enters only through
rier ∆Eb0 is larger than σ∆Utot L/2ζc by 2, because a logarithm and so its precise value is of limited
the average barrier is the potential energy differ- importance. We set ε̇0 = 104 s−1 , consistent with
ence between the average minimum and the aver- previous work [12].
age maximum, not the average minimum and the The above framework dominates at low tem-
zero energy level. The total energy barrier ∆Eb peratures and high stresses. At higher temper-
corresponds to the energy cost of moving from a atures/lower stresses, the dislocation can explore
4
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

glide
direction

Figure 2: Schematic of the low-energy wavy configuration of the dislocation as it moves through the random field of solutes.
The configuration is characterized by segments of lateral length 2ζ of amplitude w along the length of the long dislocation.
The key quantity is the change in energy of a straight segment of length ζ as it glides a distance w through the random solute
field. The total dislocation energy is minimized with respect to both ζ and w to obtain the controlling characteristic lengths
ζc and wc .

additional longer-wavelength configurations super- energetically costly, even though such fluctuations
imposed on the underlying scale of ζc , wc , as first certainly exist.
proposed by Labusch [30] and recently formalized
and quantified by Leyson et al. [31]. However,
the thermally-activated finite-temperature strength The critical quantities, energy barrier ∆Eb and
still scales with the energy barrier ∆Eb , but zero-temperature yield stress τy0 , depend on the
strength given by [27]
key energy ∆Ẽp (wc ) which is computed here for
  random fcc alloys of any composition and number
1 kT ε̇0
τy (T, ε̇) = τy0 exp − ln . (12) of components (see Eq. (4)). The combination of
0.51 ∆Eb ε̇
Eqs. (4) and (8) - (11) then constitutes the general
theory for strengthening in HEAs, and is the first
The low-T/high-stress and high-T/low-stress re-
main result of this work.
sults above are essentially equal over the range
0.3 ≤ τy /τy0 ≤ 0.6.

3.3. Strengthening in HEAs The theory does not consider atomic fluctuations
at the scale of b < ζc , wc because the line tension
By starting from an effective medium matrix, the concept would not apply [24]. But, such fluctu-
theory first averages out the effects of all the solutes ations could occur - they are just not calculable
and then reintroduces the effects of solute fluctua- at the present time. In fact, Å-scale fluctuations
tions in attracting and repelling a dislocation. The in the dislocation position are indeed observed in
theory considers all possible scales of fluctuation simulations, and are likely due to additional terms
(ζ, w) in the random alloy, over scales where line in the potential energy that are not captured by
n
tension is suitable for evaluating the elastic energy either U (xi , yj ) or its fluctuations σ∆Uijn . How-
[24] of the non-straight dislocation configurations. ever, such small-scale fluctuations occur within the
The theory thus naturally identifies that there are larger mesoscale energy landscape, not in place of it.
mesoscale collective concentration/structural fluc- While not calculable, such fluctuations could gener-
tuations on the scale of (ζc , wc ) that create the dom- ate small additional energy barriers that would con-
inant energy barrier controlling the yield stress [32] tribute to strengthening at zero temperature, but
in the random alloy. The dislocation does not re- would be expected to be easily overcome at finite
spond to smaller-scale fluctuations because they are temperature.
5
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

4. Molecular Statics Validation random landscape varying on length scale wc is


1
1 √

SS π ζc 2
τy0 = 2 √ ∆Ẽp (wc ),
The model presented here is based on several ap- 2 bζc wc 3b
proximations, and so some independent validation 1.69
is valuable to assess the accuracy of the model. To = 3 1 ∆Ẽp (wc ). (13)
this end, we perform a comparison between key b 2 ζc2 wc
aspects of the model and direct molecular stat- For the assumed local sinusoidal potential energy
ics (MS) simulations of the flow stress. We study landscape, the theory also predicts the average dis-
Fe-Ni-Cr fcc alloys, described using the Embed- location position x versus stress τ as the stress
ded Atom Method (EAM) potentials developed pushes the dislocation up the underlying energy
in Ref. [33], as a convenient well-defined multi- SS
barrier, x = (wc /π) arcsin τ /τy0

. The dislocation
component system. SS
escapes thus when x = wc /2 at τy0 .
The above pertains to the average local mini-
mum. Simulations will sample the statistical distri-
4.1. Flow-stress of a straight dislocation segment bution of local minimum, both weaker and stronger
binding, with variations in the local width of the po-
Direct molecular simulations cannot capture the tential energy landscape (e.g. a specific wc for each
evolution of wavy configurations of long disloca- case). Thus, for one given alloy composition, mul-
tions (L  ζc ) with energy barriers on the order of tiple simulations are needed to obtain the average
∼ 0.7 − 1.1 eV at experimental strain-rates. How- behavior for comparison with the theory. We now
ever, the stress required to move a single straight turn to the simulations on the model alloys.
dislocation segment of length ζ ≤ ζc out of its ini-
tial local energy minimum at T = 0K can be simu- 4.2. MS vs. Theory: Fe-Ni-Cr model alloys
lated and can be calculated using the theory. The- We focus on Fe(1−x)/2 Ni(1−x)/2 Crx fcc alloys,
oretically, at length ζ ≤ ζc the dislocation remains with different x values, because our studies reveal
straight and moves through the random potential that the Cr content controls many property varia-
energy landscape with no bowing. The potential tions. Within EAM description of alloys energet-
energy lanscape has maxima and minima spaced on ics, a rigourous analytical configurational averag-
the scale of wc . We can thus introduce a straight ing procedure [22] leads to an average-atom EAM
dislocation into the material, let it relax to its lo- potential that represents the effective medium ma-
cal energy minimum, and then measure the stress trix for the multicomponent alloy at any desired
required to push the dislocation out of the local composition. The resulting average-atom potential
minima, over the local maximum. captures many average properties of the random
The theoretical prediction for the stress required alloys, including defect properties, with all fluctu-
to move an initial straight dislocation of length ations averaged out analytically. It can be used to
ζ = ζc out of its local potential energy well is de- perform usual molecular simulations, and thus al-
rived by considering only the potential energy of lows to measure the Peierls stresses and all model
interaction with solutes, i.e. there is no bowing and inputs for the different Fe-Ni-Cr random alloys.
no line-tension energy cost associated with moving The Peierls stresses of an {111}(100) edge dislo-
of the straight dislocation. Therefore, an average cation in the effective matrix alloys are measured
dislocation straight segment starts in an average using standard methods, and are 10, 3.5, 1.5, 3 and
local energy minimum of depth σ∆Utot (absolute 3 MPa for x = 10, 20, 33, 40 and 50%, respectively.
value), relative to the zero energy level. The
√ dislo- These values are very small, as expected for single-
cation must escape over a barrier that is 2 larger element fcc materials. This result alone shows that
and located at a distance wc from the minimum. is it precisely the chemical and structural fluctua-
Again, an applied stress provides the energy (work) tions that cause the high strength in HEA materi-
necessary to escape over the barrier at zero tem- als. Figs. 3a,b show µ, γSF , and γUSF versus x for
perature. Thus, similar to Eqs. (9) and (10), the these alloys, demonstrating significant variations in
SS
predicted average zero-temperature flow stress τy0 key dislocation-related material parameters versus
for one segment of length ζ = ζc moving through a alloy composition.
6
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

The inputs to the theory are the average inter-


n (a)
action energies U (xi , yj ) between solutes n = Fe, 120 120
Ni, Cr and a straight edge dislocation in the effec-
110 110

µ (GPa)
tive matrix. These are computed directly as the

µ (GPa)
energy change between a simulation cell having a
substitutional single solute atom at the desired po- 100 100
sition near the dislocation into the effective medium
matrix, and a simulation cell with the same solute 90 90
but far from the dislocation, thus non-interacting (b)
with it [22]. The results are shown in Fig. 4 for 100 400

γUSF (mJ/m2)
γSF (mJ/m )
γSF (mJ/m22)
x = 10% and x = 50% and for solute positions in γUSF
and around one of the two dissociated partials (be- 75 320
havior around the other partial is the same). The
solute interactions also vary significantly with com- γSF
position: at x = 10%, interactions with the partial 50 240
core are nearly zero for Fe but very strong for Cr, (c)
whereas moderate interactions exist for all solutes 10 100
at x = 50%. The fluctuations in the interaction
8 80
w (Å)
energies due to local environments cannot be com-

ζc (Å)
wcc (Å)

puted in the average matrix, and are thus neglected, wc


2
i.e. σ∆U n = 0 here. The full theory then deter- 6 60
ij
mines (ζc , wc ), both of which also vary strongly with 4 ζc 40
composition x as shown in Fig. 3c. All these wide
variations in “matrix” properties, “solute” proper- 10 20 30 40 50
ties, and in characteristic mesoscale lengths across x
these model alloys precludes any a priori conclu-
sion on the flow strengths of these different EAM Figure 3: Effective matrix properties for different
alloys. The theory, however, automatically consid- Fe(1−x)/2 Ni(1−x)/2 Crx EAM alloys: (a), µVoigt , and (b),
γSF and γUSF . (c), Characteristic lengths (wc , ζc ) for the
ers all of these factors versus composition and has same alloys, as predicted by the full strengthening theory.
no adjustable parameters.

MS simulations of the flow stresses are performed


on twenty different random realizations, per alloy ning environment predicted by the theory. With
composition, of dislocation segments of length ∼ ζc this point in mind, Fig. 5a shows the applied shear
moving under the action of an increasing shear stress as a function of the measured average glide
stress (see Appendix C for details). It is important position of the dislocation x, relative to its initial
to note that, once it escapes from its local min- position in the energy minimum x = 0, averaged
imum, a straight dislocation segment will always over twenty samples at the two extreme composi-
be halted at a next-stronger-fluctuation arising sta- tions of Fe45 Ni45 Cr10 and Fe25 Ni25 Cr50 (see Ap-
tistically further along the glide plane. The stress pendix C for full data). Also shown are the the-
vs. position is thus always monotonically increas- oretical predictions for stress versus glide distance,
ing. The “strength” measured in the MS simula- with the dislocation predicted to unpin at x = wc /2
SS
tions therefore depends on the glide distance cho- at stress τy0 . The averaged simulations agree well
sen. The theory for escape from the average bar- with the theory for glide distances up to wc /2 and
rier predicts that the average strength is attained then show the slow steady increase for glide beyond
at a glide distance of wc /2, and so comparisons of wc /2 as expected due to features of the MS simula-
averages are made at this glide distance. The suc- tion discussed above. The theory thus captures the
cessive pinning in the MS of a single straight seg- relevant internal length scale wc controlling the dis-
ment does not arise in a real material because a location strength as a function of alloy composition,
very long dislocation line length L  ζc will bow- which is an important test of the theory. Moreover,
out around any rare stronger pinning regions and so Fig. 5b shows the predicted and simulated average
the strength will be controlled by the average pin- flow stress at x = wc /2. Excellent agreement is
7
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

Fe45Ni45Cr10 Fe25Ni25Cr50 of the parameter-free theory, in terms of not only


strength but also the distribution in strength and
Fe Fe
the length scales controlling the strength versus al-
loy composition, for a range of model materials hav-
ing considerable variations in many material prop-
erties, is the second main result of this paper.

Ni Ni 5. Reduced Model: Elastic Interactions

The full general model does not provide sig-


nificant insights into the role of the average ma-
trix or solute properties, nor dislocation struc-
ture, and full solute/dislocation interaction energies
n
Cr Cr U (xi , yj ) may not be easily computable or mea-
surable in real materials. We thus now reduce the
theory to a widely-usable form by considering only
the main elasticity contribution Ueln (xi , yj , zk ) =
−p(xi , yj )∆Vn (xi , yj , zk ) to the solute/dislocation
U (xi,yj) interaction energy, which is due to the interaction
between the pressure field p(xi , yj ) of the disloca-
tion at the solute site and the solute n misfit vol-
−200 −150 −100 −50 0 50 ume at this site. This contribution is common to
all fcc HEAs and, due to symmetry, local devi-
Figure 4: Interaction energies between one partial disloca- atoric misfit strains must average to zero and so
tion in the effective alloy matrix and each of the solutes Fe, would only appear in the secondary local fluctua-
Ni and Cr, in Fe(1−x)/2 Ni(1−x)/2 Crx EAM alloys and for
the limiting compositions x = 10 (right) and 50% (left). tion term. We express the dislocation pressure field
µ (1+ν)
as p(xi , yj ) = − 3π (1−ν) f (xi , yj ) where f (xi , yj ) is
the dimensionless anisotropic pressure field gener-
found between theory and simulations across the ated by the distribution of normalized Burgers vec-
five Fe-Ni-Cr compositions possessing a range of tor along the glide plane (see Fig. 1c) with µ and ν
underlying material properties. The low-Cr alloys the isotropic elastic constants introduced for scal-
are much stronger than the high-Cr alloys, and the ing purposes. Inserting these into Eq. (1), the key
wc much smaller, because the Cr/dislocation inter- energy in the theory becomes
action energies in the dislocation partial cores are   21
much larger in the low-Cr alloys (see Fig. 4). It is µ (1 + ν) X 2
important to keep in mind that in the absence of ∆Ẽp (w) = ∆fij (w)
3π (1 − ν) i,j
the theory there would be zero ability to predict ei-
ther this magnitude of strengths (recall the Peierls  
# 12
stresses are negligible) nor the detailed trends with
X 2 2
× cn ∆V + n σ∆V n
, (14)
alloy composition. n
If we wished to be more conservative in making
comparisons, we could measure the “strength” in where ∆fij (w) =  f (x i − w, y
j ) − f (xi , yj ). The
P 2 2
the simulations at x = 3wc /4, a point at which quantity n cn ∆V n + σ∆Vn is the key misfit
> 18 of the 20 samples have experienced some volume quantity that is closely related to the so-
very large jump forward in dislocation position, called misfit parameter δ, (see Appendix B), with
indicative of unpinning from a local energy well. ∆V n the average misfit volume of solute n and
The “strengths” measured in simulations using this σ∆Vn its standard deviation due to local fluctua-
larger distance are 15-25% larger than the theory, tions. Minimization of the total energy with re-
still in good agreement with, and following the spect to w to obtain wc is then determined only
trends of, the parameter-free theory. by ∆fij (w), which depends only on the dislocation
The overall qualitative and quantitative success core structure, independent of the solute properties.
8
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

Fe45Ni45Cr10 Sim. Sim.


500 400
Th.
400
Th. 300

y0 (MPa)
τ (MPa)

y0 (MPa)
τ (MPa)

300
1 wc Fe25Ni25Cr50 Sim. 200

τMS
2

τMS
200
Th.
100
100 1 wc
2

0 0
0 2 4 6 8 10 10 20 30 40 50
dislocation position (Å) Cr content x
(a) (b)

Figure 5: Validation of the theory against Molecular Statics simulations for Fe(1−x)/2 Ni(1−x)/2 Crx EAM alloys. (a), Average
shear stress in MS simulations vs. average dislocation position, for x = 0.1 and x = 0.5 alloys. Red symbols indicate predictions
of the theory at the average depinning point wc /2. (b), Average zero-temperature flow stresses for the 5 alloys as measured
in simulations (grey squares) and as predicted using Eq. (13) (red diamonds). Error bars show the confidence intervals of the
mean strength obtained from simulations on 20 samples per alloy.

For a given core structure and thus given dimen- work. We first parameterize the continuous distri-
sionless pressure field f (xi , yj ), minimization of the bution of Burgers vectors along the glide plane of
total energy yields the (ζc , wc ), and then the theory the dissociated dislocation as two Gaussian peaks
predicts τy0 and ∆Eb as each of standard deviation σ = 1.5b separated by a
2
(2+ν)
  43 stacking fault of width d = γµbSF 24π(1−ν) [34] (see
− 31 1+ν Fig. 1c). This form is consistent with atomistic
τy0 =0.051α µ f1 (wc )
1−ν studies of fcc dislocations (see Appendix D). The
P  2
  23 dimensionless pressure field f (xi , yj ) can then be
2
c
n n ∆V n + σ∆Vn computed, and the minimized core coefficients in
× 6
 , (15)
b Eqs. (15) and (16) can be obtained as a function of
  23 partial spacing d, as shown in Fig. 6. Interestingly,
1
3 1+ν for moderate separations d/b > 10, the minimized
∆Eb =0.274α µb 3 f2 (wc )
1−ν “core” parameters are essentially independent of d,
P  2
  13 with the core coefficients for τy0 and ∆Eb being
2
c
n n ∆V n + σ∆Vn f1 (wc ) = 0.35 and f2 (wc ) = 5.70. A small discon-
× 6
 . (16) tinuity is seen at small d values, due to a change
b
in wc , which can only take on integer multiples of
where we introduced b/2. For typical HEA elastic moduli and with esti-
 the minimized core
5/2 P 2/3coeffi- mates of stacking fault energies in HEAs [35] being
b 2
cients f1 (wc ) = wc i,j ∆fij (wc ) and below 100mJ/m2 , and typically much smaller, we
h 2 P i1/3 conclude that the stacking fault energy value is not
f2 (wc ) = wbc 2
i,j ∆fij (wc ) , and where the important for the low-temperature strength in HEA
line tension is expressed here as Γ = αµb2 , with α materials. As an aside, we note that minimization
a dimensionless number. reveals the existence of a second minimum energy
To actually execute the minimization, we use configuration, with low τy0 and high ∆Eb ; this pro-
isotropic elasticity to predict both dislocation core vides a “plateau” stress relevant only at very high
structure and associated pressure field; extension temperatures above temperatures of interest here,
to anisotropic elasticity will be deferred to future and so this is not discussed further.

9
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

H-P contribution σH−P (dg ) can be subtracted from


0.4 8 the total experimental strength; this is standard
metallurgical practice.

∆Eb structure coefficient


τy0
τy0 structure coefficient

Polycrystalline elastic constants (µ, ν) versus


0.35 7 temperature (for NiCoFeCr and NiCoFeCrMn) and
at room temperature (for NiCo, NiCoFe, and NiC-
oCr) [4, 8] and the average Burgers vectors were ex-
0.3 6 perimentally measured (see Table 1). Average mis-
∆Eb fit volumes ∆V n of the Fe, Co and Cr solutes are
calculated from the measured average atomic vol-
0.25 5 umes of Ni1−cn Xcn binary fcc solid solutions with
X = Fe, Co and Cr [21, 38, 39]. The measurements
show a linear variation with cn in the binaries. Ap-
0.2 4 plying Vegard’s law to the average atomic volume of
10 d µb20 (2+ ν) 30 each solid solution, i.e. V = (1−cn )VNi +cn Vn , with
b = γSF 24π(1−ν) VNi = a3Ni /4 and V = a3Ni1−cn Xcn /4, we compute the
unknown individual atomic volume Vn for each el-
ement n, and get Vn = 10.94, 11.12, 12.09, 12.27Å3
Figure 6: Dependence of strength and energy barrier on dis- for Ni, Co, Fe and Cr, respectively. VMn = 12.60Å3
location dissociation distance d (related to γSF as indicated). is then obtained from the measured atomic vol-
Minimized dislocation structural coefficient for the energy
barrier ∆Eb vs. d appearing in Eq. (16) (red squares) and umes of equiatomic NiCoFeCrMn, NiCoFeMn, NiC-
for the zero-temperature strength τy0 , vs. d appearing in oCrMn, NiFeMn and NiCoMn alloys, using the
Eq. (15) (blue triangles). These coefficients are nearly in- previously-determined Vn for the other elements.
dependent of γSF . The discontinuity at small d is due to a
The misfit volume of an element n in a given P HEA
small change in the wc value emerging from the minimiza-
tion. Dashed lines indicate values used for predictions of is computed asP ∆V n = Vn − V , with V = n cn Vn
HEA strengths. (the sum rule n cn ∆V n = 0 is followed by con-
struction). The predicted average Burgers vector √
of each HEA is computed through b = (4V )1/3 / 2.
The reduced theory based on elasticity and mis- While magnetism in the binary alloys may affect the
fit volumes, leading to Eqs. (15) and (16), with the deduced atomic volumes, the values we obtain for
computed core coefficients independent of core par- this set of {Vn } give very good predictions for the
tial spacing, is not only simplified but is fully an- measured overall Burgers vector b of all the var-
alytic. Predictions require only elastic moduli, lat- ious HEAs, as shown in Table 1. We note that
tice constants, and accurate solute misfit volumes misfit volumes in HEAs are often computed us-
versus alloy composition; this is the third main re- ing textbook “atomic” radii for the elements [40],
sult of this work. independent of chemical, magnetic, or crystallo-
graphic structural details, and so the values here
6. Comparison to Experiments are a significant improvement over the standard es-
timates. Fluctuations in misfit volumes and local
We now apply the analytic theory to predict the deviatoric misfit strains are not available experi-
strengths of equiatomic fcc alloys in the Ni-Co-Fe- mentally and so are neglected in making predic-
Cr-Mn family. The uniaxial tensile yield strengths tions. The line tension parameter α = 0.123 is
versus temperature, strain-rate, and grain size dg obtained from atomistically-measured edge disloca-
in polycrystalline materials have been measured tion line tension in the EAM FeNiCr effective ma-
[4, 5, 8] and show two contributions: a grain-size- trix and is close to the coefficient for elemental Al
dependent Hall-Petch (H-P) contribution σH−P (dg ) [41]. The analytic theory of Eqs. (15) and (16) then
and a chemical/alloying effect σalloy . These two predicts the critical resolved shear strength versus
contributions to strength have distinctly different temperature and strain-rate for any alloy composi-
physical origins; the present theory accounts for the tion. The corresponding uniaxial yield stress σalloy
chemical/alloying contribution σalloy . We therefore for an equiaxed fcc polycrystal is obtained by mul-
compare our theoretical predictions of σalloy to ex- tiplying by the Taylor factor of 3.06.
periments on those alloys for which the measured Figs. 7a,b show the predicted and measured al-
10
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

Table 1: Experimental data on the fcc Ni-Co-Fe-Cr-Mn family high entropy alloys. Burgers vector b and isotropic elastic
constants [4] (µ, ν) measured at room temperature, and elastic constants versus temperature for equiatomic fcc NiCoFeCr [4]
and NiCoFeCrMn [36, 37] alloys. Burgers vectors obtained with the procedure described in the text are given for comparison.
Estimates of the intrinsic stacking fault energies γSF are obtained by a combination of experiments and DFT calculations [35].
(2+ν)
The corresponding dissociation distances for edge dislocation db = γµb 24π(1−ν) are given between parenthesis.
SF

Alloy b (Å) Elastic constants


exp. pred. µ (GPa) ν γSF (mJ/m2 )
NiCo 2.499 2.499 84 0.29 −
FeNi 2.533 2.535 62 0.34 ∼ 100 (8)
NiCoFe 2.524 2.525 60 0.35 −
NiCoCr 2.517 2.529 87 0.30 −
NiCoFeCr 2.525 2.541 84 0.28 ∼ 30 (28)
NiCoMn 2.544 2.537 − − −
NiFeMn 2.557 2.561 − − −
NiCoFeMn 2.540 2.547 − − −
NiCoCrMn 2.538 2.550 − − −
NiCoFeCrMn 2.545 2.555 80 0.26 ∼ 20 (41)
Elastic constants T = 0K 77K 203K 293K 473K 673K
NiCoFeCrMn µ (GPa) 85 85 83 80 75 68
ν 0.259 0.259 0.258 0.259 0.264 0.271
NiCoFeCr µ (GPa) 93 91 87 84 79 72

loy contributions σalloy to the yield stress for the We use a linear correlation between the Hall-Petch
NiCoFeCr and NiCoFeCrMn alloys over the com- trend for HV and for the yield stress to obtain σalloy
plete temperature range studied at the experimen- for these materials. Fig. 7c shows the parameter-
tal strain-rate ε̇ = 10−3 s−1 . The predictions are free prediction of σalloy versus the measured val-
very good, with no fitting parameters. The exper- ues for all six alloys at T = 293K. The overall
imental strengths of NiCoFeCr and NiCoFeCrMn agreement is again very good, with accuracy levels
are nearly identical, as predicted. The NiCoFe- similar to those achieved in simpler dilute binary
CrMn alloy has a larger misfit contribution but a alloys [14, 27, 28]. The theory also preserves the
lower µ. In the theory, these two factors nearly observed ordering of strength versus composition,
cancel for the zero-temperature strength and give with the NiCoCr alloy being the strongest of all the
a slightly lower energy barrier, thus rationalizing alloys tested. The quantitative success of the ana-
the similarity in measured strengths. The predic- lytic model with no fitting parameters, over a wide
tions at the lowest temperature (T = 77K) are be- temperature range for two HEAs and at T = 293K
low the experiments. As mentioned earlier, low- for six HEAs, is strong evidence of the predictive
T behavior could be affected by the neglect of Å- capability of the model. This is the fourth main
scale fluctuations. Our predictions could also be result of this work.
refined using specific solute-core interactions, and
the interaction-fluctuation contribution σ∆Uijn , but
7. Implications and design guidelines for
these quantities are not available to date. Figs. 7a,b
HEAs
also show predictions using a line tension parame-
ter α = 0.06125, demonstrating that the strength The theory answers many open questions about
predictions are nearly independent of line tension strengthening in HEAs: (i) strength does not di-
except at T = 77K. Low-T predictions would thus rectly depend on the number of components N ,
require a more precise determination of the line ten- and is not maximized by the equi-atomic compo-
sion coefficient α. sition, (ii) the strongest and most temperature-
For four other alloys (NiCo, NiFe, NiCoFe, NiC- insensitive materials are achieved by maximizing
oCr), the Hall-Petch strengthening was only mea- the concentration-weighted mean-square misfit vol-
sured at T = 293K by Vickers Hardness (HV ) [5]. ume quantity and/or increasing the shear modulus,
11
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

NiCoFeCr NiCoFeCrMn T = 293K


300 300

experimental σalloy (MPa)


Exp. Exp.
Theory Theory 100

σalloy (MPa)
σalloy (MPa)

200 200
NiCo
FeNi
50 NiCoFe
100 100
NiCoCr
NiCoFeCr
NiCoFeCrMn
0 0 0
0 200 400 600 0 200 400 600 0 50 100
Temperature (K) Temperature (K) predicted σalloy (MPa)
(a) (b) (c)

Figure 7: Quantitative comparison between experiment and theory. (a) and (b), Yield stress vs. temperature for the NiCoFeCr
and NiCoFeCrMn equiatomic alloys [4, 8], as measured (black symbols) after subtraction of the Hall-Petch contribution to
strength and as predicted by the theory (red symbols). The dashed lines show predictions using a line tension of one-half the
original value (i.e. α is reduced by 1/2), showing the weak sensitivity of the model to the line tension except at very low T; (c),
Experimental [4, 5, 8] vs. predicted strengths for HEA alloys in the Ni-Co-Fe-Cr-Mn family at T = 293K, after subtraction of
the Hall-Petch contribution to strength.

(iii) the stacking fault energy has little influence on line tension) can, in principle, be computed. To
yield strength; and (iv) local chemical environments properly account for chemistry and magnetism, and
and structural disorder should generate an addi- to explore new potential alloys, ab initio approaches
tional contribution to the strength (see Eqs. (4), are probably required. Currently, Cij , b, ∆V n ,
2
(15) and (16)), even though such fluctuations may σ∆V n
, and γSF can be computed by ab initio meth-
be difficult to measure. ods. The dislocation core structure, direct so-
The present model accounts for the strength of lute/dislocation core interactions, and the disloca-
N -component random alloys at arbitrary compo- tion line tension, are far more challenging, and per-
sition, based only on solute/dislocation interac- haps prohibitively expensive, to compute. Thus, ab
tion energies. The use of average interactions pro- initio-based design is likely best pursued using the
vides the lowest-order approximation, and the in- elasticity-based model (Eqs. (14)-(16)) with dislo-
clusion of fluctuations in interaction energy repre- cation structures parameterized by b, γSF and γUSF ,
sents the first step towards inclusion of local chem- and a line tension scaled by µb2 . Within the sim-
ical and structural effects. Higher order effects as- plified elasticity model, higher strengths will gener-
sociated with multi-solute interactions can be in- ally be obtained for a larger solute misfit parameter
hP  2
i1/2
cluded within the same effective medium matrix ap- n cn ∆V n + σ 2
∆Vn and/or larger µ, with
proximation. Similarly, Short Range Order (SRO) γSF of less importance. New alloys designed using
could also be included (although SRO is only rarely the elasticity-based theory with computable mate-
observed in HEAs to date [42]). Both aspects would rial inputs will, while not exact, provide a clear
significantly complicate the analysis, due to statis- physical framework for identifying promising com-
tics and correlations, but remain within the scope positions to achieve higher-performance HEAs.
of the general approach here.
The present validated theory provides insights Appendix A. Standard deviation of the po-
and guidance for understanding and designing tential energy change
strengthening in fcc HEAs. The full theory does
not have any adjustable parameters; all inputs (so- In this appendix, we derive the standard devi-
lute/dislocation interaction energies, elastic con- ation of the potential energy change σ∆Utot (ζ, w)
stants, Burgers vector, dislocation core structure, when a dislocation segment of length ζ glides over
12
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

a distance w in a random field of solutes, in a N - The quantity of interest in the strengthening


component alloy at arbitrary composition. model is the standard deviation of the total inter-
The interaction for a solute of type n at posi- action energy change as defined in Eq. (2), where
tion (xi , yj ) relative to the center of the dislocation now both the concentrations and the interaction en-
and position zk along the dislocation is denoted as ergies of each solute n can fluctuate. We have, as-
U n (xi , yj , zk ). Positions (xi , yj , zk ) refer to the ef- suming that two solutes located in different atomic
fective medium atomic sites, and we denote Ns the rows are uncorrelated and ommitting the depen-
number of atomic sites along z direction for a dis- dancies in ζ and w for the sake of brevity
location segment of length ζ, at each atomic row "
(xi , yj ), with 1 ≤ k ≤ Ns . For fcc material √ and X X D E D E2 
2
{111}(110) dislocations we have Ns = ζ/ 3b. The σ∆Utot
= nnα
ij
2
− nnα
ij
nα 2
∆Uij
total potential energy of the dislocation segment is i,j n,α
X hD nβ
E D ED Ei
XX + nnα
ij nij − nnα
ij nnβ
ij

∆Uij nβ
∆Uij
Utot (ζ) = snijk U n (xi , yj , zk ), (A.1) n
α,β6=α
i,j,k n #
X hD mβ
E D ED

Ei

with snijk = 1 if a type-n solute is at position + nnα
ij nij

− nij nij nα
∆Uij ∆Uij ,
(xi , yj , zk ), and 0 otherwise. Along the line di- n,m6=n
α,β
rection z of the dislocation, the dislocation struc-
ture does not change but there can be varying local (A.5)
chemical and structural environments that give rise
to fluctuations of the interaction energy of a solute which involves the variances and co-variances

nαof the
type n with the dislocation along the line. Thus, for random variables {nnαij }. By inspection, n ij =
each (xi , yj ) and solute n, we consider a statistical pnα cn Ns , where pnα is the probability of finding
distribution of interaction energies that exist along a type-n solute with local environment
Pαmax α, with
the line z. For ease of manipulation, we describe the normalization condition p
α=1 nα = 1. For
this distribution as a discrete distribution of αmax (n, α) 6= (m, β) we have, also by inspection,
different possible local environments for each solute D E

type n. The corresponding solute/dislocation in- nnα
ij nij = (pnα cn Ns ) (pmβ cm (Ns − 1)) . (A.6)
teraction energies are denoted as U nα (xi , yj ), with
1 ≤ α ≤ αmax . Different environments at different
The former term corresponds to the average number
sites are assumed uncorrelated. Eq. (A.1) is then
of type (n, α) solute in Ns sites, and the latter to
rewritten as
the average number of type (m, β) solute in Ns − 1
sites, given that at least one site is occupied by a
XX
Utot (ζ) = snα
ijk U

(xi , yj ), (A.2)
i,j,k n,α
type (n, α) solute. Finally, for (n, α) = (m, β) we
have
with snα
ijk = 1 if a type-n solute with local envi-
ronment α sits at position (xi , yj ) and 0 otherwise.
D E X Ns !
nnα
ij
2
= ckn q Ns −kn
Thus, the total potential energy change when a dis- kn !(Ns − kn )! n n
kn
location glides a distance w along the x direction is X kn !
× pknα q kn −knα knα
2
XX knα (kn − knα )! nα nα
∆Utot (ζ, w) = nnα nα
ij ∆Uij (w), (A.3) knα
i,j n,α X Ns !
= ckn q Ns −kn
kn !(Ns − kn )! n n
with kn

× p2nα kn2 + kn pnα (1 − pnα )


 

∆Uij (w) = [U nα (xi − w, yj ) − U nα (xi , yj )] ,
=p2nα c2n Ns2 + cn (1 − cn )Ns
 
(A.4)
+ pnα (1 − pnα )cn Ns
=Ns pnα cn + p2nα c2n (Ns − 1) ,
PNs  
and nnα
ij = nα
k=1 sijk the number of solutes n with (A.7)
local environment α along dislocation line of length
ζ at position (xi , yj ). with 0 ≤ kn ≤ Ns and 0 ≤ knα ≤ kn . After inser-
13
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

tion into Eq. (A.5), we obtain to the so-called misfit parameter [2, 3, 43, 44].
" ThisP connection arises through the “solute” quan-
2
X X
nα 2 tity n cn ∆V n 2 that emerges in the present the-
σ∆Utot =Ns pnα cn (1 − pnα cn )∆Uij
i,j n,α
ory. The average misfit volume ∆V n is related to
X nβ
the variation in volume with respect to alloy com-
− pnα pnβ c2n ∆Uij

∆Uij position as
n
α,β6=α  
X ∂V ∂V
∆V n = cm − , (B.1)
#
∂cn c̄ ∂cm c̄
X mβ

− pnα pmβ cn cm ∆Uij ∆Uij . m
n,m6=n
α,β with V the average atomic volume and where c̄ =
(A.8) (c1 , ..., cN ) refers to the composition of the N -
component alloy of interest. For an fcc alloy, we
Eq. (A.8) can be simplified since have V = a3 /4, with a the lattice parameter of the
PN Pαmax nα
c
n=1 n p
α=1 nα ∆Uij (w) = 0 (the so- alloy, and thus
lute/dislocation interaction energies are defined  
with respect to an effective medium at the overall 3V X ∂a ∂a
∆V n = cm − . (B.2)
composition) and using the normalization of the a m ∂cn c̄ ∂cm c̄
pnα . This leads to
If we apply Vegard’s law to theP lattice parameter
2 ζ XX X nα of the solid solution, i.e. a = n cn an with an the
σ∆U (ζ, w) = √ cn pnα ∆Uij (w)2 ,
tot
3b i,j n lattice parameter of element n, then
α
a 
(A.9) n
∆V n = 3V −1 . (B.3)
√ a
using Ns = ζ/ 3b. Finally, writing each specific The key solute quantity in the strengthening theory
class of interaction energy variation for a type n is thus
solute as a deviation from the average value, i.e.
nα n nα
P 2

∆Uij (w) = ∆U ij (w) + δUij (w), we obtain (with- 2 n cn rn 2
X
2
cn ∆V n = 9V − 1 = 9V δ 2 ,
out assuming that deviations are small) r 2
n
(B.4)
2 ζ XX  n

σ∆Utot
(ζ, w) = √ cn ∆U ij (w)2 + σ∆U
2
n ,
3b i,j n ij
where rn and r are the atomic radii of the ele-
(A.10) ment n and of an average atom, respectively. We
thus obtain a direct connection with the misfit pa-
2 nα 2
P
with σ∆U n =
α pnα δUij being the square of rameter δ used to qualitatively characterize HEAs
ij
the standard deviation of the distribution of the in- [2, 3]. However, δ is usually estimated using tables
teraction energy change {∆Uij nα
(w)}. Interestingly, of “accepted” atomic radii for the elements. For
even in the case of a zero average interaction energy, robust application of the strengthening theory, it is
a type-n solute can still contribute in the strength- much better to use the true solute misfit volumes in
ening due to the fluctuations of the interaction en- the actual alloy of interest, which are well-defined
ergy, which appear through the standard deviation quantities from both thermodynamics and mechan-
in Eq. (A.10). One specific case where this applies ics perspectives, and do not rely on the validity of
is for the interaction energy between local devia- Vegard’s law, nor require an assumption that the
toric strains of solutes n and the deviatoric stress different elements crystallize into the same struc-
field of the dislocation, where the average deviatoric ture.
strains must be zero by symmetry for substitutional
solutes in fcc materials. Appendix C. Atomistic details about MS
shear tests
Appendix B. Connection to the δ parameter
Atomistic measurements of the flow stresses for
Neglecting the standard deviation term in Fe(1−x)/2 Ni(1−x)/2 Crx EAM alloys are performed
Eq. (A.10), we can further connect σ∆Utot (ζ, w) using the open-source code Lammps [45]. For each
14
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

composition x, 20 different random realizations at


the overall composition are prepared. In each of 800 Fe45Ni45Cr10

(MPa)
ττ (MPa)
them, a straight dislocation is introduced as two 600
Shockley partials separated by the proper dissocia- 400
tion distance using the Volterra displacement field.
200
The system is relaxed to its local minimum energy
state, with traction free boundaries on the surfaces 0
800 Fe40Ni40Cr20
of the sample parallel to the glide plane, and peri-

(MPa)
ττ (MPa)
600
odic boundary conditions in both the glide and line
directions [46]. The position of the dislocation is 400
continuously monitored as the average of the posi- 200
tions of the two partial dislocations. The simulation 0
cell sizes are X × Y × Z = 100b × ζc × 48b, with X 800 Fe33Ni33Cr33

(MPa)
ττ (MPa)
the glide direction, Y the line direction and Z the 600
(111) direction, thus large enough to prevent any 400
finite-size effects. 200
We then apply an increasing resolved shear stress, 0
with steps of 5 MPa. We start from τ0 = 250 MPa (MPa)
ττ (MPa) 800 Fe30Ni30Cr40

for x = 0.1 , from 200 MPa for x = 0.2 and 0.33, 600
from 150 MPa for x = 0.4, and from 100 MPa for 400
x = 0.5. Some samples having initial dislocation
200
jumps xd > wc /2 were re-started with lower values
for the applied shear stress, so as to accurately de- 0
800 Fe25Ni25Cr50
scribe the behavior around the dislocation position
(MPa)
ττ (MPa)

600
xd = wc /2. Results are displayed in Fig. C.8 for
the five EAM alloys. 400
We then computed the average shear stress τ 200
vs. dislocation position by using a grid of point 0
xi = ib/2, with i = 1, 2, . . . . The window around 0 2 4 6 8 10
each grid point xi is ±b/4, consistent with the preci- dislocation position xd (Å)
sion in the theory: wc ∝ b/2. We set τ (xd = 0) = 0,
and the average shear stress for dislocation posi-
Figure C.8: Molecular Statics simulations of the flow stress
tions in [0, b/4[ is centered at x0 = b/8. The aver- for Fe(1−x)/2 Ni(1−x)/2 Crx EAM alloys. The 20 different
age curves are shown in black in Fig. C.8. In the samples for each targeted composition x are shown by the
x = 0.1, 0.4 and 0.5 Fe(1−x)/2 Ni(1−x)/2 Crx alloys, colored symbols and markers. The average applied stress τ
for the small values of xi , typically b/8 and b/2, vs. average dislocation position xd is shown as the black line.
The vertical dashed line indicates the value of wc /2 predicted
some samples already experience a big displace- by the theory as the point at which the dislocation becomes
ment jump at the initial stress τ0 , i.e. the shear unpinned, on average.
stress required to move the dislocation is smaller
than τ0 within that interval. In evaluating the av-
erage strength, we thus fixed for these samples the energy quantity ∆Ẽp (wc ) related to the standard
stress value to τ0 /2, so as to obtain a reasonable deviation of the interaction energy σ∆Utot , which
estimate of τ (taking only the other available data corresponds to the average for pinning in the lower
points would lead to an overestimation of τ ). The 33% of the probability distribution function. Vari-
less accurate sampling for the average stress is re- ations around this average strength are expected in
flected in the associated 70% confidence intervals simulations from the theory, and a large spreading
in Figs. 5a,b, which are then higher for the low xi in strength is indeed observed, reflecting the un-
values. For the x = 0.2 and 0.33 alloys, we exclude derlying statistical nature of the stengthening. The
the [0, b/4[ zone for the average strength curve, as simulations also reveal the underlying length scale
the sampling is not sufficient to define an average. wc that the theory predicts (Fig. C.8, dashed lines)
The theory assumes a Gaussian distribution for and its evolution with the overall alloy composition:
the fluctuations in potential energy, with the key most of the samples undergo a significant displace-
15
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

ment jump when x ' wc /2.


(a)
2.5

σ/b
σ/b
Appendix D. Determination of the disloca- 2
tion partial spreading
1.5
We parameterize the Burgers vector distribution

USF
(b) 0.4

USF)
of the dissociated dislocation using two Gaussian

4π(1−ν)γ
2
peaks separated by a distance d and with stan-

σµb/(4πγ
µb
0.3
dard deviation of each Gaussian denoted by σ. The
Burgers vector distribution is divided into a discrete 0.2
set of fractional Burgers vector on the discrete lat- 10 20 30 40 50
tice points
x
 2  2
xi −d/2 xi +d/2
− 12 − 21
∆b(xi ) e σ
+e σ
=  2  2 , Figure D.9: (a), Partial core spreading σ in
b P+∞ − 12
xk −d/2
− 12
xk +d/2
Fe(1−x)/2 Ni(1−x)/2 Crx EAM alloys, measured by fitting the
k=−∞ e +e
σ σ
atomistically-determined Burgers vector distributions with
(D.1) the distribution of Eq. (D.1). (b), Dimensionless ratio of the
measured σ to the material quantity µb2 /(4π(1 − ν)γUSF )
where xi = nb/2 with n = 0, ±1, ±2... is the posi- for the various alloys, showing a constant scaling as expected
tion along the glide plane, b is the Burgers vector from a Peierls-Nabarro model [47].
of the edge perfect dislocation, and where d is the
partial separation distance. Isotropic elasticity pre-
2
(2+ν)
dicts d = γµbSF 24π(1−ν) for an fcc edge dislocation, Thus, the value of σ = 1.5b we used for HEAs is
with µ, ν the elastic constants and γSF the stable generally valid for these fcc systems.
stacking fault energy. Second, we consider the dilute solute-
The spreading within the partial cores is repre- strengthening predictions of Leyson et al. [27]
sented by the parameter σ. In a Peierls-Nabarro for six substitutional solutes (Mg, Cu, Si, Cr, Mn,
model for the core [47], the spreading of the partial Fe) in Al. In these systems, full DFT calculations
cores should scale with µb2 /(4π(1 − ν)γUSF ), where of the solute/dislocation core interactions and
γUSF is the unstable stacking fault energy. The pre- solute misfit volumes were performed. Using only
cise scaling depends on details of the entire gener- the misfit volume interaction we make predictions
alized stacking fault curve, and this is not available of τy0 as a function of core structure, using dis-
for the HEA materials. To determine a good value sociation distances d = 2.5b, 3b consistent with
for σ, we consider two different analysis. the DFT-computed core [49] and varying partial
First, we study the core structures in the random core spreading σ. All other Al parameters are
Fe(1−x)/2 Ni(1−x)/2 Crx EAM alloys with x = 10, 20, given in Ref. 27. The best agreement with the full
33, 40 and 50%, having different material parame- DFT-computed results is obtained for σ = 1.5b,
ters. Employing the effective medium approach for justifying again our choice of σ = 1.5b as a fixed
EAM potentials [22], we relax edge dislocations in parameter for the dislocation partial core spreading
the effective medium for each alloy and fit the re- in the HEA alloys.
sulting Burgers vector distributions to Eq. (D.1) to
obtain σ values for all alloys. Fig. D.9a,b show the
measured σ values and the same values normalized
by µb2 /(4π(1−ν)γUSF ), vs. Cr content x. The vari- Acknowledgments
ations in σ are small among the different materials,
ranging from 1.5b to 2.5b, and the near-constant
normalized value verifies the expected scaling. Us- Support for this work was provided through a Eu-
ing this constant value of ∼ 0.28 in Fig. D.9b for ropean Research Council Advanced Grant, “Predic-
pure Ni (µ = 79 GPa, γUSF = 113 mJ/m2 ) and for tive Computational Metallurgy”, ERC Grant agree-
pure Al (µ = 28.4 GPa, γUSF = 224mJ/m2 ), we ment No. 339081 - PreCoMet. Dr. F. Pavia is ac-
obtain σ = 1.74b and 1.26b, respectively (materials knowledged for performing the atomistic line ten-
parameters from Refs. 33 and 48 for Ni and Al). sion measurements.
16
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

References related physical systems, Rev. Mod. Phys. 46 (1974)


465–543. doi:10.1103/RevModPhys.46.465.
[1] M.-H. Tsai, J.-W. Yeh, High-entropy alloys: A critical [16] J. Faulkner, The modern theory of alloys, Prog. Mater.
review, Mater. Res. Lett. 2 (3) (2014) 107–123. doi: Sci.s 27 (1–2) (1982) 1 – 187. doi:http://dx.doi.org/
10.1080/21663831.2014.912690. 10.1016/0079-6425(82)90005-6.
[2] Y. Zhang, T. T. Zuo, Z. Tang, M. C. Gao, K. A. Dah- [17] D. D. Johnson, D. M. Nicholson, F. J. Pinski, B. L. Gy-
men, P. K. Liaw, Z. P. Lu, Microstructures and proper- orffy, G. M. Stocks, Density-functional theory for ran-
ties of high-entropy alloys, Prog. Mater. Sci. 61 (2014) dom alloys: Total energy within the coherent-potential
1–93. doi:10.1016/j.pmatsci.2013.10.001. approximation, Phys. Rev. Lett. 56 (1986) 2088–2091.
[3] D. B. Miracle, J. D. Miller, O. N. Senkov, C. Woodward, doi:10.1103/PhysRevLett.56.2088.
M. D. Uchic, J. Tiley, Exploration and development of [18] F. Ducastelle, Order and Phase Stability in Alloys,
high entropy alloys for structural applications, Entropy North-Holland, Amsterdam, 1991.
16 (1) (2014) 494–525. doi:10.3390/e16010494. [19] D. Johnson, M. Asta, Energetics of homogeneously-
[4] Z. Wu, H. Bei, G. Pharr, E. George, Temperature de- random fcc al-ag alloys: A detailed comparison of com-
pendence of the mechanical properties of equiatomic putational methods, Comput. Mater. Sci. 8 (1–2) (1997)
solid solution alloys with face-centered cubic crystal 54 – 63, proceedings of the joint NSF/CNRS Work-
structures, Acta Mater. 81 (0) (2014) 428 – 441. doi: shop on Alloy Theory. doi:http://dx.doi.org/10.
http://dx.doi.org/10.1016/j.actamat.2014.08.026. 1016/S0927-0256(97)00016-5.
[5] Z. Wu, H. Bei, F. Otto, G. Pharr, E. George, Recovery, [20] B. Ujfalussy, J. S. Faulkner, N. Y. Moghadam, G. M.
recrystallization, grain growth and phase stability of Stocks, Y. Wang, Calculating properties with the
a family of fcc-structured multi-component equiatomic polymorphous coherent-potential approximation, Phys.
solid solution alloys, Intermetallics 46 (0) (2014) 131 Rev. B 61 (2000) 12005–12016. doi:10.1103/PhysRevB.
– 140. doi:http://dx.doi.org/10.1016/j.intermet. 61.12005.
2013.10.024. [21] R. W. Smith, G. S. Was, Application of molecular
[6] B. Gludovatz, A. Hohenwarter, D. Catoor, E. H. dynamics to the study of hydrogen embrittlement in
Chang, E. P. George, R. O. Ritchie, A fracture-resistant ni-cr-fe alloys, Phys. Rev. B 40 (1989) 10322–10336.
high-entropy alloy for cryogenic applications, Science doi:10.1103/PhysRevB.40.10322.
345 (6201) (2014) 1153–1158. doi:10.1126/science. [22] C. Varvenne, A. Luque, W. G. Nöhring, W. A. Curtin,
1254581. Average-atom interatomic potentials for random alloys,
[7] M. Poletti, L. Battezzati, Electronic and thermody- accepted to Phys. Rev. B.
namic criteria for the occurrence of high entropy al- [23] S. Rao, C. Woodward, T. A. Parthasarathy,
loys in metallic systems, Acta Mater. 75 (2014) 297 D. Dimiduk, D. B. Miracle, W. A. Curtin, Molecular
– 306. doi:http://dx.doi.org/10.1016/j.actamat. statics and molecular dynamics simulations of disloca-
2014.04.033. tion behavior in model fcc and bcc multicomponent
[8] F. Otto, A. Dlouhý, C. Somsen, H. Bei, G. Eggeler, concentrated solid solution alloys, in preparation.
E. George, The influences of temperature and mi- [24] D. J. Bacon, D. M. Barnett, R. O. Scatter-
crostructure on the tensile properties of a cocrfemnni good, Anisotropic continuum theory of lattice defects,
Prog. Mater. Sci. 23 (1980) 51–262. doi:10.1016/
high-entropy alloy, Acta Mater. 61 (15) (2013) 5743
– 5755. doi:http://dx.doi.org/10.1016/j.actamat. 0079-6425(80)90007-9.
[25] E. Clouet, The vacancy - edge dislocation interaction
2013.06.018.
[9] R. Labusch, A statistical theory of solid solution hard- in fcc metals, in: P. Gumbsch (Ed.), Proceedings of
the Third International Conference Multiscale Materi-
ening, Phys. Status Solidi A 41 (2) (1970) 659. doi:
10.1002/pssb.19700410221. als Modeling, Freiburg, Germany, 2006, pp. 708–711.
[26] E. Hayward, C. Deo, B. P. Uberuaga, C. N. Tomé,
[10] M. Zaiser, Dislocation motion in a random solid solu-
tion, Philos. Mag. A 82 (15) (2002) 2869–2883. doi: The interaction of a screw dislocation with point de-
fects in bcc iron, Phil. Mag. 92 (2012) 2759. doi:
doi:10.1080/01418610208240071.
[11] A. Argon, Strengthening Mechanisms in Crystal Plas- 10.1080/14786435.2012.674646.
ticity, Oxford Univ. Press, 2007. [27] G. P. M. Leyson, L. Hector Jr., W. A. Curtin, So-
[12] G. P. M. Leyson, W. A. Curtin, L. G. Hector, Jr., C. F. lute strengthening from first principles and application
Woodward, Quantitative prediction of solute strength- to aluminum alloys, Acta Mater. 60 (9) (2012) 3873
ening in aluminium alloys, Nat. Mater. 9 (9) (2010) – 3884. doi:http://dx.doi.org/10.1016/j.actamat.
750–755. doi:10.1038/NMAT2813. 2012.03.037.
[13] J. A. Yasi, L. G. Hector Jr, D. R. Trinkle, First- [28] G. P. M. Leyson, L. Hector Jr., W. A. Curtin, First-
principles data for solid-solution strengthening of mag- principles prediction of yield stress for basal slip in
nesium: From geometry and chemistry to properties, mg–al alloys, Acta Mater. 60 (13–14) (2012) 5197
Acta Mater. 58 (17) (2010) 5704 – 5713. doi:http: – 5203. doi:http://dx.doi.org/10.1016/j.actamat.
//dx.doi.org/10.1016/j.actamat.2010.06.045. 2012.06.020.
[14] D. Ma, M. Friák, J. von Pezold, D. Raabe, J. Neuge- [29] U. Kocks, A. S. Argon, M. F. Ashby, Prog. Mater. Sci.
bauer, Computationally efficient and quantitatively ac- 19 (1975) 1 – 281. doi:http://dx.doi.org/10.1016/
curate multiscale simulation of solid-solution strength- 0079-6425(75)90005-5.
ening by ab initio calculation, Acta Mater. 85 (2015) [30] R. Labusch, G. Grange, J. Ahearn, P. Haasen, Rate pro-
53 – 66. doi:http://dx.doi.org/10.1016/j.actamat. cesses in plastic deformation of materials, in: J. C. M.
2014.10.044. Li, A. K. Mukherjee (Eds.), Proceedings from the John
[15] R. J. Elliott, J. A. Krumhansl, P. L. Leath, The the- E. Dorn symposium, ASM, Metal Park, Ohio, 1975,
ory and properties of randomly disordered crystals and p. 26.

17
This is a pre-print of the following article: Varvenne, C.; Luque, A.; Curtin, W. A. Acta Mater. 2016, 118, 164–176.. The formal
publication is available at http://dx.doi.org/10.1016/j.actamat.2016.07.040

[31] G. P. M. Leyson, W. A. Curtin, Solute strengthening tion properties of aluminum, Phys. Rev. B 62 (2000)
at high temperatures, submitted. 3099–3108. doi:10.1103/PhysRevB.62.3099.
[32] R. Labusch, Cooperative effects in alloy hardening, [49] S. M. Keralavarma, A. F. Bower, W. A. Curtin,
Czech. J. Phys. 38 (5) (1988) 474–481. doi:10.1007/ Quantum-to-continuum prediction of ductility loss in
BF01597457. aluminium-magnesium alloys due to dynamic strain ag-
[33] G. Bonny, D. Terentyev, R. C. Pasianot, S. Ponce, ing, Nat. Commun. 5. doi:10.1038/ncomms5604.
A. Bakaev, Interatomic potential to study plasticity in
stainless steels: the fenicr model alloy, Modelling Simul.
Mater. Sci. Eng. 19 (8). doi:10.1088/0965-0393/19/8/
085008.
[34] J. P. Hirth, J. Lothe, Theory of Dislocations, 2nd Edi-
tion, Wiley, New York, 1982.
[35] A. J. Zaddach, C. Niu, C. C. Koch, D. L. Irving, Me-
chanical properties and stacking fault energies of nife-
crcomn high-entropy alloy, JOM 65 (12) (2013) 1780–
1789. doi:10.1007/s11837-013-0771-4.
[36] A. Haglund, M. Koehler, D. Catoor, E. George, V. Kep-
pens, Polycrystalline elastic moduli of a high-entropy
alloy at cryogenic temperatures, Intermetallics 58 (0)
(2015) 62 – 64. doi:http://dx.doi.org/10.1016/j.
intermet.2014.11.005.
[37] G. Laplanche, P. Gadaud, O. Horst, F. Otto,
G. Eggeler, E. George, Temperature dependencies of
the elastic moduli and thermal expansion coefficient of
an equiatomic, single-phase cocrfemnni high-entropy al-
loy, J. Alloys Compd. 623 (0) (2015) 348 – 353. doi:
http://dx.doi.org/10.1016/j.jallcom.2014.11.061.
[38] J. Bandyopadhyay, K. P. Gupta, Low temperature lat-
tice parameter of nickel and some nickel-cobalt alloys
and grüneisen parameter of nickel, Cryogenics (1977)
345–347.
[39] F. Abe, T. Tanabe, Change in lattice spacing of nickel
by dissolved chromium and tungsten, Z. Metall. 76 (6)
(1985) 420–425.
[40] N. J. D. Tilley, Understanding Solids: The Science of
Materials, Wiley, 2004.
[41] B. Szajewski, F. Pavia, Y. Dong, W. A. Curtin, Robust
line tension via atomistics, Modelling Simul. Mater. Sci.
Eng. 23 (8).
[42] C. Niu, A. J. Zaddach, A. A. Oni, X. Sang, J. W.
Hurt, J. M. LeBeau, C. C. Koch, D. L. Irving, Spin-
driven ordering of cr in the equiatomic high entropy
alloy nifecrco, Appl. Phys. Lett. 106 (16). doi:http:
//dx.doi.org/10.1063/1.4918996.
[43] I. Toda-Caraballo, P. E. J. Rivera-Dı́az-del Castillo,
Modelling solid solution hardening in high entropy al-
loys, Acta Mater. 85 (0) (2015) 14 – 23. doi:http:
//dx.doi.org/10.1016/j.actamat.2014.11.014.
[44] X. Li, H. Zhang, S. Lu, W. Li, J. Zhao, B. Johans-
son, L. Vitos, Elastic properties of vanadium-based al-
loys from first-principles theory, Phys. Rev. B 86 (2012)
014105. doi:10.1103/PhysRevB.86.014105.
[45] S. Plimpton, Fast parallel algorithms for short-range
molecular dynamics, J. Comput. Phys. 117 (1) (1995)
1 – 19. doi:http://dx.doi.org/10.1006/jcph.1995.
1039.
[46] Y. N. Osetsky, D. J. Bacon, An atomic-level model for
studying the dynamics of edge dislocations in metals,
Modelling Simul. Mater. Sci. Eng. 11 (2003) 427–446.
doi:10.1088/0965-0393/11/4/302.
[47] V. V. Bulatov, W. Cai, Computer Simulations of Dis-
locations, Oxford series on materials modelling, Oxford
University Press, 2006.
[48] G. Lu, N. Kioussis, V. V. Bulatov, E. Kaxiras,
Generalized-stacking-fault energy surface and disloca-

18

You might also like