You are on page 1of 9

PHYSICAL REVIEW B 75, 014204 共2007兲

Solid-liquid phase diagrams for binary metallic alloys: Adjustable interatomic potentials

H.-S. Nam,1,* M. I. Mendelev,2 and D. J. Srolovitz1,3


1Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, New Jersey 08544, USA
2Materials and Engineering Physics, Ames Laboratory, Ames, Iowa 50011, USA
3
Department of Physics, Yeshiva University, New York, New York 10033, USA
共Received 11 August 2006; published 8 January 2007兲
We develop an approach to determining Lennard-Jones embedded-atom method potentials for alloys and use
these to determine the solid-liquid phase diagrams for binary metallic alloys using Kofke’s Gibbs-Duhem
integration technique combined with semigrand canonical Monte Carlo simulations. We demonstrate that it is
possible to produce a wide range of experimentally observed binary phase diagrams 共with no intermetallic
phases兲 by reference to the atomic sizes and cohesive energies of the two elemental materials. In some cases,
it is useful to employ a single adjustable parameter to adjust the phase diagram 共we provided a good choice for
this free parameter兲. Next, we perform a systematic investigation of the effect of relative atomic sizes and
cohesive energies of the elements on the binary phase diagrams. We then show that this approach leads to good
agreement with several experimental binary phase diagrams. The main benefit of this approach is not the
accurate reproduction of experimental phase diagrams, but rather to provide a method by which material
properties can be continuously changed in simulation studies. This is one of the keys to the use of atomistic
simulations to understand mechanisms and properties in a manner not available to experiment.

DOI: 10.1103/PhysRevB.75.014204 PACS number共s兲: 81.30.Bx, 82.60.Lf, 02.70.Uu

I. INTRODUCTION Because of their simplicity and applicability to systems with


Atomic-scale simulations have become an indispensable a wide range of properties, these potentials have been widely
tool for modern materials research. For example, molecular used, both for elemental and multicomponent systems.5,6
dynamics 共MD兲 and Monte Carlo 共MC兲 simulations are rou- However, such potentials are only realistic for very simple
tinely used to investigate phenomena that are not easily ac- materials, such as noble gases. In other materials, bonding is
cessible via experiment or to interpret experimental results.1,2 more complex. For example, for metals and alloys, it is well
The fundamental input to such simulations is a description of known that many-body effects play an important role.7,8 For
the interactions between atoms. While first-principles meth- such materials, potentials of the embedded-atom method9
ods accurately describe atomic bonding through quantum- 共EAM兲 type are widely used. Holian et al.10 proposed an
extension of the Lennard-Jones potential that allows for
mechanical treatments, they are usually limited to a rela-
many body interactions, like in EAM-type potentials. This
tively small number of atoms. Semiempirical or empirical
idea was further pursued by Baskes11–13 to treat a broad
interatomic descriptions are often motivated by quantum- range of metallic systems including alloys resulting in a po-
mechanical ideas but represent different materials through tential known as the Lennard-Jones embedded-atom method
parametrization schemes in which the constants are fit to 共LJ-EAM兲 potential. This potential represents a readily ad-
experimental 共and/or first-principles兲 data. While this ap- justable description of atomic interactions in metals and me-
proach has significant problems when the resultant potentials tallic alloys. Because of the relatively simple analytical form
are employed under conditions for which they were not fit- of the adjustable LJ-EAM potential, it can be used as a de-
ted, they can provide accurate results when applied carefully. scription of atomic interactions in systems with different,
Such potentials have the advantage that they can be easily controllable thermodynamics properties. As such, it provides
used for simulations involving a very large number of atoms a ready means to test the influence of thermodynamic prop-
共currently up to 109兲. Since few interesting materials are erties on materials behavior without the complexity of first-
pure, we focus on potentials for metallic alloys in this paper. principles approaches or all of the oversimplifications inher-
In order to determine the utility of potentials for alloys, we ent in pairwise potentials.
should insure that the potentials lead to the correct phases at In this paper, we focus on the systematic determination of
temperatures and compositions of interest. In this paper, we binary phase diagrams for LJ-EAM potentials. The ability to
describe the determination of binary phase diagrams for a predict phase diagrams for such adjustable interatomic po-
particularly flexible choice of potentials for metallic alloys. tentials is an important first step in identifying a set of atomic
While there has been a long tradition of modeling mate- interactions required to reproduce the requisite phases
rials using potentials that can be adjusted to represent differ- needed to describe specific phenomena and systems at non-
ent types of materials, such readily adjustable potentials are zero temperature. This is also important for determining
commonly pairwise. As such, they do not provide a reason- which thermodynamic properties are important in a particu-
able description of metals 共e.g., surface relaxation兲. Ideally, lar type of phenomena observed experimentally. We were
we will be able to easily adjust a potential to give the desired motivated to pursue this study through our own attempts to
metallic phase diagram. perform molecular dynamics simulations of liquid-metal em-
Examples of easily adjustable pairwise interatomic poten- brittlement in alloys.14 There have been several contradictory
tials include the Lennard-Jones3 共LJ兲 and Morse4 potentials. suggestions as to what type of thermodynamic behavior is

1098-0121/2007/75共1兲/014204共9兲 014204-1 ©2007 The American Physical Society


NAM, MENDELEV, AND SROLOVITZ PHYSICAL REVIEW B 75, 014204 共2007兲

necessary for this phenomenon to occur. Therefore, we spe- such a way that the total energy of the reference structure as
cifically consider the solid-liquid regions of the phase dia- a function of dilation is described by a LJ potential. If we
grams for a wide range of metallic binary alloys. We focus include interactions up to second-nearest neighbors, the like-
on two main parameters in describing the alloys—relative atom pair potential ␾AA共r兲 for species A is given by

冉 冊
atomic size and the strength of the chemical bonds. Of
Z2
course, this is not the first attempt to systematically describe ␾AA共r兲 + ␾AA共ar兲 = ␺A共r兲 共4a兲
binary phase diagrams from an atomistic view. Earlier at- Z1
tempts have examined the phase diagrams of hard sphere15 or
and Lennard-Jones materials.16,17 The methods employed to
determine the phase behavior range from density-functional
theory18 to Gibbs-Duhem integration methods.19,20 We com-
bine LJ-EAM potentials, molecular dynamics, and Gibbs-
N
␾AA共r兲 = ␺A共r兲 + 兺 共− 1兲n
n=1
冋 冉冊 Z2
Z1
n

␺A共anr兲 , 共4b兲

Duhem integration methods in a Monte Carlo framework to where

冉 冊
develop a systematic understanding of the relationship be-
tween potential properties and the solid-liquid phase dia- 2
␺A共r兲 = ␾AA
LJ
共r兲 − FA„¯␳A0 共r兲… 共4c兲
gram. Z1
and
II. POTENTIALS

In this section, we outline the form of the LJ-EAM model


¯␳A0 共r兲 = ␳A共r兲 + 冉 冊
Z2
Z1
␳A共ar兲. 共4d兲
used in this work. For a binary material described by classi-
Here, Z2 is the number of second-nearest-neighbor atoms and
cal LJ pair potentials, the interatomic potential between at-
a is the ratio of the second-to first-nearest-neighbor distance.
oms i and j takes the form
The summation of N terms is carried out until ␾AA共r兲 con-

␾sLJis j共r兲 = 4 ⑀ sis j 冋冉 冊 冉 冊 册


␴ sis j
r
12

␴ sis j
r
6
, 共1兲
verges.
Because this potential was fitted to a LJ form, it has only
four adjustable parameters, ÂA, ␤ˆ A, ⑀A 共⬅⑀AA兲, and ␴A
where ⑀sis j and ␴sis j are the attractive well depth and the 共⬅␴AA兲 for the single component material, A. This potential
diameter for the LJ potential describing the interactions be- form can be easily extended to multicomponent systems. For
tween species si and s j 共s = A or B兲. The total energy of a EAM binary alloys, we must fix seven functions ␳A共r兲, ␳B共r兲,
binary LJ-EAM system is given by the usual EAM form FA共r兲, FB共r兲, ␾AA共r兲, ␾BB共r兲, and ␾AB共r兲. The first six of

i

E = 兺 Fsi共¯␳i兲 +
1
兺 ␾s s 共rij兲 ,
2 j⫽i i j 册 共2兲
these are transferable from the two monatomic systems. Fol-
lowing Baskes and Stan,13 we can obtain the remaining func-
tion, ␾AB共r兲, by fitting to a particular alloy structure. Like
them, we focus on the ordered L10 compound 共after correct-
where Fsi共¯␳i兲 is the embedding energy and ␾sis j共rij兲 is the
ing a small error in Ref. 13兲, as described in Appendix A,
pair interaction between atoms i and j separated by a dis-
tance rij. The embedded function is commonly chosen as11 1 1
␾AB共r兲 = ␾AB
LJ
共r兲 − 关FA„¯␳AL10共r兲… + FB„¯␳BL10共r兲…兴 − 关␾AA共r兲
1 8 4
Fsi共¯␳i兲 = ÂsiZ1⑀sisi¯␳i关ln共¯␳i兲 − 1兴, 共3a兲 3
2
+ ␾BB共r兲 − ␾AA
LJ
共r兲 − ␾BB
LJ
共r兲兴 − 关␾AA共ar兲
8
where the electron density at the site of the atom is
+ ␾BB共ar兲兴. 共5兲
1
¯␳i = 兺 ␳ j共rij兲 共3b兲
Z1 j⫽i If we set ÂA = 0 and ÂB = 0, ␾AA and ␾AB take exactly the
same form as the LJ potential.
and ␾AB共r兲 is fully determined by Eq. 共5兲, except for ␾AB LJ
共r兲.

␳ j共rij兲 = exp ␤ˆ s j 冋冉 rij


冑6 2␴s s
i j
−1 冊册 . 共3c兲
Therefore, we need two parameters, ⑀AB and ␴AB, to describe
the cross interaction between species A and B 共in addition to
the eight parameters needed to describe pure A and pure B兲.
The cross-species interaction parameters ⑀ and ␴ are, there-
Here, the dimensionless parameter Âs represents the strength fore, the knobs we use to control alloy properties. The
of the many-body term, the parameter ␤ˆ s quantifies the dis- Lorentz-Bertholet mixing rules21 are widely used for obtain-
tance over which the electron density decays away from an ing alloy parameters in LJ systems; i.e., ␴AB = 共␴A + ␴B兲 / 2
atom position, and Z1 is the coordination number of the ref- and ⑀AB = 冑⑀A⑀B. However, applying the Lorentz-Berthelot
erence state 共e.g., face-centered cubic兲. mixing rules to determine ⑀AB and ␴AB results in phase dia-
We can combine the EAM form of the total energy appro- grams that do not correspond to those observed for metallic
priate for metals with the convenience of the adjustable LJ alloys 关i.e., the solid-liquid two-phase region is quite square
pair potential by choosing the pairwise term in Eq. 共2兲 in with very limited solubility in the solid and the liquid兴. Even

014204-2
SOLID-LIQUID PHASE DIAGRAMS FOR BINARY… PHYSICAL REVIEW B 75, 014204 共2007兲

TABLE I. A comparison of several key properties for the LJ-EAM potential model 共for  = 0.7, ␤ˆ = 7兲 and
several fcc metals. All properties are normalized by using E0共=6⑀兲, r0共=冑6 2␴兲, and ⍀共=r30 / 冑2兲.

Property Normalized quantity LJ LJ-EAM Cu Ag Au fcc metals

Bulk modulus B⍀ / E0 8.5 8.5 3.0 3.92 5.05 2.0–6.0


Melting point k BT m / E 0 0.082 0.033 0.033 0.037 0.029 0.024–0.041
Cauchy discrepancy 共c12 − c44兲 / B 0 0.48 0.30 0.425 0.69 −0.4– 0.7
Vacancy-formation energy Evf / E0 1.0 0.31 0.37 0.39 0.23 0.2–0.5
共100兲 surface energy E共100兲r20 / E0 0.67 0.26 0.24 0.23 0.22 0.15–0.25

though the Lorentz-Berthelot mixing rules may be appropri- LJ-EAM material with those of several fcc noble metals. For
ate for LJ potentials, there is no reason to expect this to be the LJ-EAM parameters employed in Table I, the LJ-EAM
true for LJ-EAM potentials. This is because in the LJ-EAM potential yields reasonable properties for the fcc metals ex-
potentials the pairwise interaction represents just part of the cept for the bulk modulus. Unlike other properties, the bulk
bonding and it is the total LJ-EAM potential that must repro- modulus is solely determined by the second derivative of the
duce the LJ potential for dilation. Therefore, to construct the energy as a function of lattice dilation 共regardless of the
pairwise interaction term in LJ-EAM, we apply the Lorentz- many-body interactions兲. Since the LJ-EAM potential was
Berthelot mixing rule to ␾AB共r兲 in Eq. 共5兲, rather than to fitted to a LJ form 共by definition兲, both the LJ and LJ-EAM
␾AB
LJ
共r兲 itself. Doing this, we find the well depth ⑀AB should models yield similar bulk moduli. The bulk modulus of LJ
be systems is known to be too large compared with the fcc
metals.11 One could obtain better bulk modulus values by
1 冑
⑀AB = 共ÂA⑀A + ÂB⑀B兲 + 兩共1 − ÂA兲⑀A共1 − ÂB兲⑀B兩. 共6兲
fitting the pair interactions to another form, e.g., the univer-
sal binding-energy relation.27
2

The derivation of this rule is given in Appendix B.


Although LJ-EAM potentials are not as widely used as III. GIBBS-DUHEM INTEGRATION METHOD
the classic LJ potential, the LJ-EAM potential has already There are several computational methods that can be used
been widely applied in the literature.11–13,22–25 The equilib- to determine the phase diagram of a system described by any
rium structure of a LJ crystal at 0 K is face centered cubic choice for the interatomic potentials.19,28 Here, we directly
共fcc兲 and the melting point depends solely on the cohesive construct the phase diagram using the Gibbs-Duhem integra-
energy and the interaction range of the potential.26 However, tion technique proposed by Kofke.19,20 In this method, two or
both the ground-state structure and melting temperature of more coexisting phases are simulated 共semigrand canonical
the single-component LJ-EAM system are strongly depen- Monte Carlo simulations兲 independently at the same tem-
dent on the two many-body interaction parameters  and ␤ˆ .11 perature and pressure. Once a single point of the coexistence
As the strength of many-body interaction  increases, the curve between two phases is known, the rest of the curve can
melting points of the pure elements decrease although the be computed 共without any free-energy calculations兲 by inte-
cohesive energy remains unchanged.11 Experimentally, most grating the equivalent of the Clausius-Clapeyron equation for
fcc metals exhibit normalized melting temperatures kBTm / E0 coexistence during the course of the simulations. The Clap-
in the range from 0.025 to 0.04, where kB is the Boltzmann eyron equation for equilibrium between two binary phases
constant and E0 is the cohesive energy of the solid at zero 共e.g., liquid and solid兲 at constant pressure is given by
pressure. If we set  = 0.7 and ␤ˆ = 7 in the LJ-EAM potentials d␤ 共xBl − xBs兲
for both components, pure A and pure B will be fcc at T = 0 = , 共7兲
and the normalized melting points fall within this tempera- d␰B ␰B共1 − ␰B兲共Hl − Hs兲
ture range for fcc metals.12 where ␤ is the reciprocal temperature, 1 / kBT, and T is the
The melting point of an elemental fcc solid is strictly absolute temperature, ␰B is the fugacity fraction of species B,
proportional to the well depth ⑀ 共for any choice of  and ␤ˆ 兲. ␰B = f B / f A, f i is the fugacity of species i in solution, xB is the
For the reference element A, we set ⑀ = 0.6 eV and ␴ mole fraction of species B, and H is the molar enthalpy. The
= 2.5 Å with which the melting point Tm = 1405 K was ob- right-hand side of Eq. 共7兲 can be integrated numerically to
tained. However, this choice of ⑀ and ␴ is rather arbitrary, find an equation for ␤ as a function of ␰B if we have an initial
since only the ratios of well depths ⑀B / ⑀A and diameters condition describing the temperature, fugacity fraction, en-
␴B / ␴A are relevant in determining the binary phase diagrams thalpies, and compositions at one coexistence point.
for LJ-EAM potentials. The potential interactions were trun- The Gibbs-Duhem integration method is a more efficient
cated between the second- and third-nearest neighbors such approach to determining phase equilibrium in solid systems
that the relaxed cohesive energy is E0 = 6.0⑀ for the fcc ref- than the Gibbs-ensemble method,2,28 because the important
erence state 关see Eq. 共4a兲兴. Monte Carlo move is changing the elemental identity of a
Table I shows a comparison of the basic properties of the particle rather than inserting or removing a particle from the

014204-3
NAM, MENDELEV, AND SROLOVITZ PHYSICAL REVIEW B 75, 014204 共2007兲

system 共inserting a particle is a low acceptance probability


event兲. The Gibbs-Duhem integration approach has been
used to tackle a number of multicomponent, multiphase equi-
librium problems including calculating the phase diagram of
a binary Lennard-Jones fluid,17 and calculating phase dia-
grams for colloids in polymer solutions.15
In this paper, we determine the initial coexistence point by
performing a microcanonical ensemble molecular-dynamics
simulation of an elemental system containing both a solid
and liquid.29 Using this approach, we directly measure the
melting temperature of pure A or B 共the solid and liquid
fractions in the simulation cell evolve until the system
reaches the equilibrium melting point兲. Using this data as the
starting point, we employ Kofke’s Gibbs-Duhem integration
technique2,17 within the framework of semigrand canonical
Monte Carlo simulations.

IV. GENERIC BEHAVIOR OF BINARY PHASE DIAGRAMS


FOR LJ-EAM MATERIALS FIG. 1. Schematic phase diagrams for metallic binary alloys of
comparable melting points: 共a兲 azeotrope, 共b兲 simple eutectic, and
In this section, we investigate solid-liquid phase diagrams 共c兲 a combination of a eutectic, a monotectic, and a liquid-phase
for binary LJ-EAM systems. All of the temperature- miscibility gap. The symbols are as follows: L refers to a liquid
composition phase diagrams reported below are for binary solution of A and B, S to a solid solution of A and B, SA to a solid
alloys at atmospheric pressure. In particular, we focus on 共a兲 solution rich in A, and SB to a solid solution rich in B.
cases where the melting points of the two elemental systems
are identical and 共b兲 cases in which the elemental systems the temperature range examined兲. However, as the atomic-
have very different melting points. The former 共similar melt- size difference increases, the degree of phase separation in-
ing points兲 is a relatively common case while the latter is creases and the solid forms a miscibility gap that leads to a
more rare 共although common for systems exhibiting liquid- eutectic phase diagram starting at an atomic-size difference
metal embrittlement14兲. In each case, we examine how varia- between 8% and 10%. This crossover from solid solution to
tions of the ratio of Lennard-Jones diameters ␴B / ␴A and well eutectic phase diagrams occurs at a size difference that is
depth ⑀B / ⑀A affect the phase diagrams. We also compare ma- much smaller than reported for LJ 共Ref. 17兲 共14–15 %兲 or
jor trends observed in the simulated phase diagrams with hard sphere systems15 共12.5%兲. Moreover, the range of size
those measured experimentally 共in order to evaluate the ap- difference for this type of simple eutectic phase diagram is
propriateness of the LJ-EAM alloy model to mimic behavior
in real metallic systems兲.

A. Equal melting points


Figure 1 shows three types of binary phase diagrams com-
monly observed in metallic alloys; we refer to these as azeo-
tropes 关Fig. 1共a兲兴, simple eutectics 关Fig. 1共b兲兴, and liquid-
phase miscibility gap systems 关Fig. 1共c兲兴 of which there are
several types. We explicitly omit consideration of systems
exhibiting compounds 共for now兲. It is the atomic-size mis-
match and the cross-species pair interaction that determine
which phase diagram type pertains. To investigate the effect
of both atomic-size mismatch and cross-species pair interac-
tions, we considered three series of cases: 共1兲 like bonding
共⑀AB = ⑀A = ⑀B兲 with different atomic-size ratios ␴B / ␴A, 共2兲
equal atomic sizes 共␴B = ␴A = ␴AB兲 and equal well depths for
the elements 共⑀B = ⑀A兲 but vary the well depth describing the FIG. 2. Solid-liquid phase diagrams with various atomic-size
differences. Model alloys have the same melting point with zero
AB interactions ⑀AB, and 共3兲 vary both ␴B / ␴A and ⑀AB / ⑀A.
heat of mixing 共⑀B / ⑀A = 1.0, ⑀AB / ⑀A = 1.0兲. Solid-liquid phase dia-
The other parameters, Â and ␤ˆ , are fixed at reference values grams were calculated with a variation of atomic-size difference
0.7 and 7 as mentioned in Sec. II. 共␴B / ␴A = 1.05, 1.08, and 1.1兲. In this and subsequent phase dia-
Figure 2 shows temperature-composition phase diagrams grams, we estimate that each data point in the phase diagram has
for a series of binary systems of type 共1兲 共variation of atomic error bars of ⬇10° in temperature and ⬇3% 共although these errors
size兲. When the atomic-size difference is small 共less than have some variation depending on the relative stability of the dif-
8%兲, the solid region of the phase diagram is a solution 共for ferent phases at each point along the phase boundaries兲.

014204-4
SOLID-LIQUID PHASE DIAGRAMS FOR BINARY… PHYSICAL REVIEW B 75, 014204 共2007兲

FIG. 5. Schematic phase diagrams of binary alloys where the


two components have very different melting points. Two such cases
are 共a兲 a eutectic with significant solubility in the solid and 共b兲 a
diagram with two eutectic points and a miscibility gap in the liquid
phase.

FIG. 3. Solid-liquid phase diagrams for different values of ⑀AB. gram would simply be a horizontal line at T = 1405 K. On the
The pure metals A and B have the same melting point and the same
other hand, when ⑀AB / ⑀A ⬍ 1, the heat of mixing is positive
atomic size 共⑀B / ⑀A = 1.0, ␴B / ␴A = 1.0兲. The solid-liquid phase dia-
共endothermic solid solution兲 and the A-B atoms “dislike each
grams were calculated with different values of the cross-species
other.” Under this circumstance, the liquidus curve is con-
interaction parameter 共⑀AB / ⑀A = 1.1, 1.05, and 0.95兲.
cave for 1 ⬍ ⑀AB / ⑀A ⬍ 0.95 共not shown兲. For ⑀AB / ⑀A 艋 0.95,
the two species are no longer miscible at all compositions
very narrow; further increase in this difference leads to a and a large miscibility gap appears in both the solid- and
miscibility gap in the liquid. When a miscibility gap exists in liquid-phase regions, as shown in the phase diagram in Fig.
the liquid, the phase diagram becomes more complex 关simi- 3.
lar to Fig. 1共c兲兴, with negligible solubility in solid phases. These results suggest that the heat of mixing alone does
Therefore, eutectic phase diagrams with a deep eutectic point not control the melting point 共liquidus兲. Rather, atomic-size
cannot be obtained by increasing the size difference in the difference also plays an important role, especially for form-
LJ-EAM model. This is not the case for LJ or hard sphere ing eutectic phase diagrams. However, both the size differ-
materials. ence and heat of mixing affect the tendency towards mixing
Figure 3 shows temperature-composition phase diagrams and can act to compensate each other. Therefore, controlling
for binary mixtures with the same atomic size 共␴B / ␴A = 1兲 atomic-size difference together with the heat of mixing can
and different values of cross-species interaction parameter. lead to a wide range of types of binary phase diagrams. For
When ⑀AB / ⑀A ⬎ 1, heat of mixing is negative 共exothermic example, increasing the liquidus temperature by increasing
solid solution兲 and A and B atoms “like each other.” In these ⑀AB / ⑀A can be compensated by increasing the atomic size
cases, the phase diagrams form continuous solid solutions difference. When large atomic-size differences are combined
over the whole composition range and the liquidus appears to with relatively large ⑀AB / ⑀A, eutectic phase diagrams with
be parabolic. The maximum temperature for which the solid deep eutectics can be formed, as shown in Fig. 4. This trend
and liquid coexist increases as ⑀AB / ⑀A increases 共the heat of is quite interesting because it seems to be related with the
mixing is more negative the larger the ⑀AB / ⑀A ratio兲. For a
binary mixture with a well-depth ratio of unity ⑀AB / ⑀A = 1, all
of the atoms are indistinguishable—hence, the phase dia-

FIG. 6. Solid-liquid phase diagrams for binary alloys with dif-


ferent melting points of component B, as controlled by different
choices of ⑀B / ⑀A 共=0.06, 0.5, and 0.4兲. These alloys have the same
FIG. 4. Solid-liquid phase diagrams for binary alloys with atomic sizes 关␴B / ␴A = 1.0, and the parameter ⑀AB is described by Eq.
atomic-size differences of 10%–13% and ⑀AB / ␴A = 1.03. 共6兲兴.

014204-5
NAM, MENDELEV, AND SROLOVITZ PHYSICAL REVIEW B 75, 014204 共2007兲

FIG. 7. Solid-liquid phase diagrams with different atomic-size FIG. 8. Solid-liquid phase diagrams with different atomic-size
ratios 共␴B / ␴A = 0.09, 1.0, and 1.1兲. The melting point of B was held ratios 共␴B / ␴A = 0.09, 1.0, and 1.1兲. The melting point of B was held
constant by fixing ⑀B / ⑀A = 0.5 and ÂB = 0.7兲. constant by fixing ÂB = 0.9 and ⑀B / ⑀A = 1.

rule of thumb for making metallic glasses. In metallic glass Phase diagrams were calculated for several different bi-
systems, it has been established, empirically, that the ability nary systems, where we varied the melting point of species
to form glasses is greatest in multicomponent systems in B. Figure 6 shows the temperature-composition phase dia-
which the atomic-size difference is large and the heat of grams for systems with the atom-size ratio fixed as ␴B / ␴A
mixing is strongly negative 共large ⑀AB / ␴A兲.30 The present = 1 and well-depth ratios of ⑀B / ⑀A = 0.6, 0.5, and 0.4. When
phase diagram results suggest that this is also the description the melting point of B is comparable to A, the system forms
of the condition for the formation of deep eutectics. Deep a solid solution with a spindle-shaped solid-liquid two-phase
eutectics are also known as systems for which glass forma- region. As ⑀B / ⑀A decreases, A-B become weaker and the
tion is particularly easy. phase diagram evolves to a eutectic diagram. It is difficult to
see the solid-liquid two-phase region at the B-rich side of the
phase diagram in Fig. 6 because the eutectic point is close to
B. Phase diagrams of large melting-point difference
pure B. Nonetheless, we assure the reader that this is indeed
Solid-liquid coexistence is a key to many materials pro- a eutectic, just like in Fig. 5共a兲. As ⑀B / ⑀A decreases further,
cessing strategies involving solidification and in-service con- the solubility of B in the solid phase and the solubility of A
ditions where a solid metal is in contact with another, liquid in the liquid phase decrease. This trend agrees with the ob-
metal. In many of the latter cases, the liquid phases consist of servation that eutectic phase diagrams determined from ex-
low melting-point species such as Hg, Ga, Bi, Pb, and Sn. periment tend to show smaller solubilities as the ratio of the
Metallic binary systems in which the melting points of the melting points of the two species deviates further from unity.
two components differ greatly typically show one of two Phase diagrams were also determined for binary systems
types of simple solid-liquid phase diagram 共provided no in- with different atomic-size ratios. Figure 7 shows the
termetallic compounds form兲: these are eutectics phase dia- temperature-composition phase diagrams for binary mixtures
grams with or without a liquid-phase miscibility gap, as with a fixed well-depth ratio, ⑀B / ⑀A = 0.5, and several diam-
shown in Fig. 5. When there is a liquid-phase miscibility eter ratios, ␴B / ␴A = 0.9, 1.0, and 1.1. When there is no
gap, the solubility of the minority species in the solid phase atomic-size mismatch, the solubilities in the solid and liquid
is usually very small 关although it often appears exaggerated phases are quite large, despite the large melting-point differ-
in schematic phase diagrams such as Fig. 5共b兲兴. However, in ence. But, as the atomic-size mismatch decreases, these solu-
other binary systems, such as Al-Ga and Zn-Ga, the solubil- bilities decrease. Interestingly, when the atomic size of B is
ity in the solid phase can be significant over the entire tem- larger than that of A, the size effect is dominant 共i.e., the
perature range, in spite of the large melting-point difference solubility is negligible over the entire temperature range and
关see Fig. 5共a兲兴. a miscibility gap appears in liquid phase兲.

TABLE II. Cohesive energy E0, lattice parameter a0, and melting point Tm for column IB fcc metals
共Refs. 31 and 32兲 and the corresponding potential parameters ⑀, ␴, and Â.

Element E0 共eV兲 a0 共Å兲 Tm 共K兲 ⑀ ␴ Â

Cu 3.54 3.62 1357 0.59 2.277 0.7


Ag 2.85 4.09 1235 0.475 2.573 0.66
Au 3.93 4.08 1392 0.655 2.570 0.8

014204-6
SOLID-LIQUID PHASE DIAGRAMS FOR BINARY… PHYSICAL REVIEW B 75, 014204 共2007兲

The melting point of B can be set by the choice of the formed a systematic investigation of the effect of relative
many-body interaction parameter ÂB 共in addition to choosing atomic sizes and cohesive energies of the elements on the
the well-depth ratio ⑀B / ⑀A兲. The melting point of B decreases binary-phase diagrams. Finally, we demonstrated that this
approach leads to good agreement with several experimental
with increasing ÂB even though the cohesive energy remains
binary-phase diagrams. The main benefit of this approach is
unchanged. Figure 8 shows the solid-liquid phase diagram
not, in our opinion, to accurately reproduce the phase dia-
with ÂB = 0.9 for several different atomic-size ratios. The ef- grams of real materials. Rather, it is to provide a method by
fect of atomic-size mismatch is still valid for this kind of which material properties can be continuously changed in
solid-liquid pairs. The trends in the phase diagrams with
atomic-size mismatch are similar in this case to those shown
in Fig. 7.

C. Comparison with real binary-alloy systems


Since the phase diagrams shown above were determined
within the framework of generic interatomic potentials LJ-
EAM, it is interesting to inquire to what degree choosing
parameters in the potential can lead to phase diagrams that
are consistent with those found experimentally in real metal-
lic systems. We can compare the simulation data to experi-
mental results to verify the ability of the LJ-EAM model to
mimic behavior in real metallic systems. The elements found
in column IB of the periodic table, copper, silver, and gold,
are common fcc metals that are well described with the
embedded-atom-method framework. We determine the LJ-
EAM parameters, ⑀, ␴ and Â, to reproduce the cohesive en-
ergy, lattice parameter, and melting temperatures of these
column IB elements 共see Ref. 12兲, as shown in Table II.
We employ these parameters within the LJ-EAM frame-
work to calculate the corresponding Ag-Cu, Cu-Au, and
Ag-Au binary-phase diagrams 共Fig. 9兲. By only adjusting the
ratio ⑀AB / ⑀A 关note, the final values were very close to those
predicted by Eq. 共6兲兴, we are able to reasonably reproduce
the experimental phase diagrams 共each with a unique topol-
ogy兲. Although the agreement certainly is not perfect, the
phase-diagram type, the temperature ranges of the features of
the diagrams, and solubilities are in good agreement with
experiment. This is remarkable given that only one param-
eter was varied 共and the atomic sizes and cohesive energies
are available from experiment兲. This type of agreement is
possible for many binary systems, provided that they do not
exhibit intermetallic compounds.

V. CONCLUSION

We developed an approach to determining LJ-EAM po-


tentials for alloys and used these to determine the solid-
liquid phase diagrams for binary-metallic alloys using
Kofke’s Gibbs-Duhem integration technique combined with
semigrand canonical Monte Carlo simulations. Inclusion of
many-body interactions led to phase diagrams, which can be
quite different from those determined using LJ or hard
sphere materials. We demonstrated that it is possible to pro-
duce a wide -range of experimentally observed binary-phase
diagrams 共with no intermetallic phases兲 by reference to the
atomic sizes and cohesive energies of the two elemental ma- FIG. 9. Solid-liquid phase diagrams for 共a兲 Ag-Cu, 共b兲 Cu-Au,
terials and by judicious choice of a single parameter that and 共c兲 Ag-Au. Cohesive energy and LJ diameter were fitted to real
controls the pairwise interactions of these two elements. In properties of materials and ⑀AB / ⑀0AB was adjusted in the LJ-EAM
addition, we provided a formula that leads to good choices binary-alloy model. ⑀AB / ⑀0AB = 1.03 for Ag-Cu, ⑀AB / ⑀0AB = 1.03 for
for this one free parameter. Within this framework, we per- Cu-Au, ⑀AB / ⑀0AB = 1.0 for Ag-Au, where ⑀0AB is given in Eq. 共6兲.

014204-7
NAM, MENDELEV, AND SROLOVITZ PHYSICAL REVIEW B 75, 014204 共2007兲

simulation studies in order to develop understanding of 1


mechanisms and properties in a manner not available to ex- EL10共r兲 = 关FA„¯␳A0 共r兲… + FB„¯␳B0 共r兲…兴
2
periment. To this end, the relationship between the adjustable
potentials and the phase diagrams they imply is central. + 关␾AA共r兲 + 4␾AB共r兲 + ␾BB共r兲兴 = ␾AA
LJ
共r兲 + 4␾AB
LJ
共r兲
+ ␾BB
LJ
共r兲. 共B1兲
ACKNOWLEDGMENTS
In equilibrium at zero pressure, r = req共=冑2␴A = 冑2␴B兲,
6 6
The authors gratefully acknowledge the support of Korea ␳A0 共req兲 = ␳B0 共req兲 = 1, ␾AA
LJ
共req兲 = ⑀A, ␾BB
LJ
共req兲 = ⑀B, and
Science and Engineering Foundation 共H.-S.N兲 and the sup-
␾AB共req兲 = ⑀AB. Substituting these expressions into Eqs. 共3a兲
LJ
port of the U.S. Department of Energy, Office of Fusion
and 共4a兲 yields
Energy Sciences, Grant No. DE-FG02-011ER54628.
1
APPENDIX A: DERIVATION OF ␾AB„r… FA„¯␳A0 共req兲… = ÂAZ1⑀A = 6ÂA⑀A ,
2
To derive a formulation for ␾共r兲, L10 structure with c / a
= 1 was considered as a reference state. In LJ-EAM formal- 1
ism, the energy per atom of this structure as a function of FB„¯␳B0 共req兲… = ÂBZ1⑀B = 6ÂB⑀B , 共B2兲
2
dilation is given by
1 and
共r兲 = 关FA„¯␳AL10共r兲… + FB„¯␳BL10共r兲…兴 + 关␾AA共r兲
冉 冊
LJ-EAM
EL1
0 2 2
␾AA共req兲 = ␾AA
LJ
共req兲 − FA„¯␳A0 共req兲… = ⑀A − ÂA⑀A
3 Z1
+ 4␾AB共r兲 + ␾BB共r兲兴 + 关␾AA共ar兲 + ␾BB共ar兲兴,
2 = 共1 − ÂA兲⑀A ,
共A1兲
where ␾BB共req兲 = 共1 − ÂB兲⑀B . 共B3兲
1 1 The energy per atom at r = req then becomes
¯␳AL10共r兲 = 关2␳B共r兲 + ␳A共r兲兴 + ␳A共ar兲 共A2兲
3 2
1
and EL10共req兲 = 关6ÂA⑀A + 6ÂB⑀B兴 + 关共1 − ÂA兲⑀A
2
1 1
¯␳BL10共r兲 = 关2␳A共r兲 + ␳B共r兲兴 + ␳B共ar兲. 共A3兲 + 4␾AB共r兲 + 共1 − ÂB兲⑀B兴 = ⑀A + 4⑀AB + ⑀B .
3 2
共B4兲
If we rewrite the energy of this system in terms of LJ pair
potentials, Solving this for ⑀AB yields
LJ
EL1 共r兲 = ␾AA
LJ
共r兲 + 4␾AB
LJ
共r兲 + ␾BB
LJ
共r兲. 共A4兲 1
0
⑀AB = 共ÂA⑀A + ÂB⑀B兲 + ␾AB共req兲. 共B5兲
By setting LJ−EAM
EL1 共r兲 = EL1
LJ
共r兲, we can derive Eq. 共5兲. 2
0 0
Now, determining ⑀AB is a matter of determining ␾AB共req兲. As
APPENDIX B: MIXING RULE FOR ⑀AB described in the text, we obtain ␾AB共req兲 by applying the
In this appendix, we show the derivation of the expression Lorentz-Bertholet mixing rule to ␾共req兲 rather than to ⑀
we employ for describing the interaction between unlike spe- 共these are equivalent for the pairwise potentials兲.
cies. Given Eq. 共5兲, this reduces to the determination of ␾AB
LJ

The length-scale parameter ␴AB is simply the arithmetic av-


LJ
.

␾AB共req兲 = 冑兩␾AA共req兲␾BB共req兲兩 = 兩共1 − ÂA兲⑀A共1 − ÂB兲⑀B兩.
erage of those for the elements. Therefore, we only require a 共B6兲
mixing rule for the well-depth parameter ⑀AB. In this appen-
dix, we show the origin of the choice for this parameter that Substituting this into Eq. 共B5兲 gives the modified mixing rule
was quoted in Eq. 共6兲. 关Eq. 共6兲兴
In order to keep this analysis simple, we rewrite Eq. 共A1兲
for the special case of nearest-neighbor interactions, ␴A = ␴B 1 冑
⑀AB = 共ÂA⑀A + ÂB⑀B兲 + 兩共1 − ÂA兲⑀A共1 − ÂB兲⑀B兩. 共B7兲
and ␤ˆ A = ␤ˆ B, 2

014204-8
SOLID-LIQUID PHASE DIAGRAMS FOR BINARY… PHYSICAL REVIEW B 75, 014204 共2007兲

16 M.
*Electronic address: hnam@princeton.edu J. Vlot, J. C. van Miltenburg, H. A. J. Oonk, and J. P. van der
1
K. Ohno, K. Esfarjani, and Y. Kawazoe, Computational Materials Eerden, J. Chem. Phys. 107, 10102 共1997兲.
Science 共Springer, Berlin, 1999兲. 17
M. R. Hitchcock and C. K. Hall, J. Chem. Phys. 110, 11433
2 D. Frenkel and B. Smit, Understanding Molecular Simulation,
共1999兲.
2nd ed. 共Academic Press, San Diego, 2002兲. 18 W. G. T. Octoby, Mol. Phys. 72, 679 共1991兲.
3 J. E. Lennard-Jones, Proc. R. Soc. London, Ser. A 106, 463
19
D. A. Kofke, Mol. Phys. 78, 1331 共1993兲.
共1924兲. 20
D. A. Kofke, J. Chem. Phys. 98, 4149 共1993兲.
4
P. M. Morse, Phys. Rev. 34, 57 共1929兲. 21 J. S. Rowlinson, Liquids and Liquid Mixtures 共Butterworth Sci-
5
E. J. Jensen, W. D. Kristensen, and R. M. J. Cotterill, Philos.
Mag. 27, 623 共1973兲. entific, London, 1982兲.
22 E. Yurtsever and F. Calvo, Phys. Rev. B 62, 9977 共2000兲.
6 J. Q. Broughton and G. H. Gilmer, J. Chem. Phys. 79, 5095

共1983兲.
23
T. Wang, F. X. Zhou, and Y. W. Liu, Chin. Phys. 11, 139 共2002兲.
7 Atomistic Simulation of Materials Beyond Pair Potential, edited 24 M. I. Baskes, JOM 56, 45 共2004兲.
25
by V. Vitek and D. J. Srolovitz, 共Plenum, New York, 1989兲. C.-W. Pao and D. J. Srolovitz, Phys. Rev. Lett. 96, 186103
8
A. E. Carlsson, Beyond Pair Potentials in Elemental Transition 共2006兲.
Metals and Semiconductors 共Academic Press, Boston, 1990兲. 26
J. A. Pryde, The Liquid State 共Hutchinson University Library,
9 M. S. Daw and M. I. Baskes, Phys. Rev. Lett. 50, 1285 共1983兲. London, 1969兲.
10 27
B. L. Holian, A. F. Voter, N. J. Wagner, R. J. Ravelo, S. P. Chen, J. H. Rose, J. R. Smith, F. Guinea, and J. Ferrante, Phys. Rev. B
W. G. Hoover, C. G. Hoover, J. E. Hammerberg, and T. D. 29, 2963 共1984兲.
Dontje, Phys. Rev. A 43, 2655 共1991兲. 28 A. Z. Panggiotopoulos, N. Quirke, M. Stapleton, and D. J. Tildes-
11
M. I. Baskes, Phys. Rev. Lett. 83, 2592 共1999兲. ley, Mol. Phys. 63, 527 共1988兲.
12 S. G. Srinivasan and M. I. Baskes, Proc. R. Soc. London, Ser. A 29 J. R. Morris, C. Z. Wang, K. M. Ho, and C. T. Chan, Phys. Rev.

460, 1649 共2004兲. B 49, 3109 共1994兲.


13 M. I. Baskes and M. Stan, Metall. Mater. Trans. A 34, 435 30 A. Inoue, Acta Mater. 48, 279 共2000兲.

共2003兲. 31
C. Kittel, Introduction to Solid State Physics, 7th ed. 共Wiley, New
14
B. Joseph, M. Picat, and F. Barbier, Eur. Phys. J.: Appl. Phys. 5, York, 1995兲.
19 共1999兲. 32 Metal Reference Book, 5th ed., edited by C. J. Smith 共Butter-
15 W. G. T. Kranendonk and D. Frenkel, Mol. Phys. 72, 679 共1991兲.
worths, London, 1976兲.

014204-9

You might also like