You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/320297028

Computing the Absolute Gibbs Free Energy in Atomistic Simulations:


Applications to Defects in Solids

Article · October 2017


DOI: 10.1103/PhysRevB.97.054102

CITATIONS READS
20 232

2 authors:

Bingqing Cheng Michele Ceriotti


University of Cambridge École Polytechnique Fédérale de Lausanne
38 PUBLICATIONS   473 CITATIONS    225 PUBLICATIONS   5,903 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

MARVEL NCCR - Advanced Sampling and Machine Learning View project

Static and Dynamical Quantum Effects in Atomistic Simulations of Materials View project

All content following this page was uploaded by Bingqing Cheng on 12 October 2017.

The user has requested enhancement of the downloaded file.


Computing the Absolute Gibbs Free Energy in Atomistic Simulations:
Applications to Defects in Solids
Bingqing Cheng∗ and Michele Ceriotti
Laboratory of Computational Science and Modeling, Institute of Materials,
École Polytechnique Fédérale de Lausanne, 1015 Lausanne, Switzerland
(Dated: October 10, 2017)
The Gibbs free energy is the fundamental thermodynamic potential underlying the relative sta-
bility of different states of matter under constant-pressure conditions. However, computing this
quantity from atomic-scale simulations is far from trivial. As a consequence, all too often the poten-
tial energy of the system is used as a proxy, overlooking entropic and anharmonic effects. Here we
discuss a combination of different thermodynamic integration routes to obtain the absolute Gibbs
arXiv:1710.02815v1 [cond-mat.mtrl-sci] 8 Oct 2017

free energy of a solid system starting from a harmonic reference state. This approach enables the
direct comparison between the free energy of different structures, circumventing the need to sample
the transition paths between them. We showcase this thermodynamic integration scheme by com-
puting the Gibbs free energy associated with a vacancy in BCC iron, and the intrinsic stacking fault
free energy of nickel. These examples highlight the pitfalls of estimating the free energy of crystal-
lographic defects only using the minimum potential energy, which overestimates the vacancy free
energy by 60% and the stacking-fault energy by almost 300% at temperatures close to the melting
point.

I. INTRODUCTION known free energy. For example, the Helmoltz free en-
ergy of a crystal can be obtained by TI from an Einstein
Knowledge of the Gibbs free energy is crucial in pre- crystal, whose free energy can be expressed analytically,
dicting the relative stability of different states of mate- to the fully interacting system [6–10]. The Gibbs free
rials and molecules, and underlies a plethora of physi- energy of liquid water can be computed by following a
cal and chemical phenomena including phase diagrams, thermodynamic path from a Lennard-Jones model to the
solubility, equilibrium concentration of defects, and so real potential [11]. However, each TI route is associ-
on. However, free energy calculations in atomistic sim- ated with different complications which have been exten-
ulations are often technically challenging and/or compu- sively discussed in the literature, such as the singularity
tationally demanding. For this reason, in many cases around T = 0 when integrating with respect to the sys-
– particularly those involving solid phases – the poten- tem temperature [12], and the pathological divergence of
tial energy of the local minimum configuration associated dH(λ)/dλ that is often observed at the end points of the
with a given state is used as a simple proxy, and at times integral when switching between the real and the refer-
a harmonic correction is also included as the entropic ence systems [5, 7, 13, 14]. As such, in general it is
term. not trivial to find an optimal route for TI for a specific
One class of standard free energy techniques, such as system.
metadynamics, umbrella sampling, and transition path The free energies associated with crystallographic de-
sampling [1–3], relies on the concept that the phase space fects (e.g. vacancies, dislocations, grain boundaries, sur-
of a system can be divided into a number of states using faces, etc.) are extremely important in predicting the
a choice of reaction coordinates (or collective variables), micro-structures and the properties of crystalline materi-
and the free energy difference between two states can als. For example, the free energy of stacking faults is cru-
then be computed by sampling both states, as well as cial in predicting the dislocation nucleation rate [15], the
the transition paths that connect them. free energies associated with different surface reconstruc-
Another class of free energy methods concentrates on tions determine the surface phase diagram [16], grain
computing the “absolute” free energy of a system by per- boundary free energy affects the rate of boundary mi-
forming a thermodynamic integration (TI) [4–6]. The gration [17]. However, computing the free energies asso-
TI can be performed along a physical path for exam- ciated with the defects is a particularly challenging prob-
ple along temperature or pressure, or via an unphysi- lem, which reveals many of the shortcomings of standard
cal path between the physical system and a reference free energy methods. Determining a physical or a vir-
system over a switching parameter λ. For the latter tual transition path to introduce or destroy a defect in-
route, the parameterized Hamiltonian can be taken to side a crystal is often complicated, ruling out techniques
be H(λ) = (1 − λ)Href + λH, where H is the actual such as metadynamics or umbrella sampling. Thermody-
Hamiltonian and Href is for the reference system with a namic integration over λ using a harmonic reference often
leads to divergences at high temperatures, when diffusive
and anharmonic behaviors become dominant [13]. Due
to these difficulties in free energy estimations, the de-
∗ bingqing.cheng@epfl.ch fect free energies are usually approximated by just the
2

potential energy of defects or by harmonic approxima-


tions [18, 19]. Recently, a number of studies have revealed
significant temperature dependency of the stacking fault
free energy due to entropic effects [15, 20, 21]. It has also
been shown that vacancy formation energy at high tem-
peratures is strongly affected by anharmonicity [22, 23].
Overall, the community is becoming more aware that the
accuracy of the prediction from the minimum potential
energies may deteriorate at high temperatures.
In this paper we combine a number of established
thermodynamic integration methods (Figure 1) in or-
der to devise a strategy for computing the absolute free
energy of a solid system with and without crystallo-
graphic defects in an efficient, easy to implement, and
generally-applicable way. We build a connection be-
tween the TI routes under the canonical (NVT) ensemble
and the isothermal-isobaric (NPT) ensemble, in order to
freely transform between the Helmoltz free energy and
the Gibbs free energy of a system. In addition, we in-
corporate several enhanced sampling methods and post-
processing techniques to improve the overall statistical
efficiency of the free energy estimation. Finally, we show-
case how to use thermodynamic integration to calculate FIG. 1. An illustration of the different thermodynamic in-
the free energy of a vacancy in BCC iron and the intrin- tegration routes employed in the present paper. Under the
canonical (NVT) ensemble, the yellow arrow indicates the
sic stacking fault free energy in FCC nickel. We demon-
switching between an harmonic reference system (λ = 0) and
strate that the anharmonic effects can significantly affect a real system (λ = 1), and the red arrow illustrates TI with
the free energies of defects in solids at high temperatures. respect to temperature. The dashed blue arrow shows the
transformation between the Helmholtz free energy and the
Gibbs free energy. The green arrow denotes TI over temper-
II. THEORY ature under the isothermal-isobaric (NPT) ensemble.

The statistical-mecanical expression for the free energy


of a system is closely related to the partition function, Gibbs free energy
which in turn depends on the thermodynamic boundary Z    
conditions defining the ensemble. Under the canonical PV A(N, V, T )
G(N, P, T ) = −kB T ln dV exp − exp − .
(NVT) ensemble, the partition function of a bulk system kB T kB T
that has N indistinguishable particles and is contained (3)
in a volume V is given by [4, 24] Computing directly the partition function for an arbi-
trary potential is impractical, and for this reason prac-
VN
Z  
U (q) tical routes for computing A or G typically use a refer-
Q(N, V, T ) = 3N dq exp − , (1)
Λ N! D(V ) kB T ence system for which the phase-space integral can be
computed analytically, followed by one or more thermo-
where the potential energy U is a function of the atomic dynamic integration steps. Since TI does not require a
coordinates q = {q1...N }, D(V ) denotes the spatial do- smooth transformation of the atomic coordinates, but
main defined by the containing volume [4], and Λ = just the evaluation of free-energy derivatives as a func-
tion of a change in the thermodynamic conditions, many
p
2π~2 /mkB T is the thermal de Broglie wavelength. The
expression for the Helmholtz free energy of the system is different paths can be used, and combined to obtain the
thus most efficient protocol.
In the case of a solid system, which is the main focus of
A(N, V, T ) = −kB T ln Q(N, V, T ) the present work, we found it effective to take a harmonic
crystal reference, and follow the TI routes illustrated in
VN
Z  
U (q) Figure 1. In a nutshell, for evaluating the Helmholtz free
= −kB T ln 3N − kB T ln dq exp − ,
Λ N! D(V ) kB T energy A of a solid system, we propose to first integrate
(2) along λ between the harmonic and the real crystal and
then do an integration with respect to the temperature
where the first term in the last line is the free energy of an T , which correspond to the yellow and the red arrows in
ideal gas. When the isothermal-isobaric (NPT) ensemble Figure 1, respectively. To calculate the Gibbs free energy
is used instead, the system can be characterized by the G of the system, we first obtain the Helmholtz free energy
3

at a low temperature, switch from the NVT to the NPT B. The Helmoltz free energy of an anharmonic
ensemble, and finally do TI along the temperature T (the crystal
yellow, the blue, and the green arrows in Figure 1). In
the followings, we discuss in detail how each step can be Starting from a reference crystal (a) with a known free
computed conveniently and efficiently. energy, one can obtain the Helmholtz free energy of the
Before we start the detailed discussion, note that in real crystal (b) using thermodynamic integration in the
TI it is advantageous and often necessary to constrain NVT ensemble, as indicated by the yellow arrow in Fig-
the center of mass (CM) of the system. The Helmholtz ure 1. using a parameter λ to perform the switch between
free energy difference between the unconstrained and the the harmonic Hamiltonian Hh and the actual Hamilto-
constrained crystalline system under periodic boundary nian H. In practice, one should run multiple simulations
conditions can be expressed as [25] with the hamiltonian H(λ) = (1 − λ)Hh + λH at different
 V values of λ, so as to switch between the harmonic Hamil-
3 1 
∆Acm (N, V, T ) = −kB T ln + ln N + ln 3 , (4) tonian Hh and the actual Hamiltonian H [7]. The free
N 2 Λ energy of the real system with a fixed CM can then be
which can be considered as a finite size effect. Therefore, evaluated using
when we perform TI we focus solely on systems with fixed Z 1
CM, and at the end of the calculation the term ∆ACM can A(V, T0 ) − Ah (V, T0 ) = dλ hU − Uh iV,T0 ,λ , (6)
be added to retrieve the free energy of the unconstrained 0
system, although at times the influence may be negligi-
ble. We will also discuss in more detail other finite-size where h. . .iV,T0 ,λ denotes the ensemble average over NVT
effects in Section II F. In the other sections – since we simulations using the Hamiltonian H(λ).
will always work under the constant-number-of-particles In practice, to avoid severe statistical inefficiencies
framework in TI – we omit N when denoting thermody- and singularities in the integral, one should perform this
namic states. step at a low temperature T0 when the system is quasi-
harmonic and when diffusive or rotational degrees of free-
dom are completely frozen. If T0 is sufficiently low and
A. An absolute reference: the Helmoltz free the real and the reference systems are very similar, one
energy of the Debye crystal also has the option to evaluate A(V, T0 ) − Ah (V, T0 ) us-
ing the free energy perturbation method, eliminating the
Strictly speaking, only the relative free energy of a sys- integration error altogether. One only needs to run sim-
tem with respect to a reference can be defined without ulations for the reference harmonic crystal, and obtain
any ambiguity. The “absolute” free energy in this pa- the free energy of the real system using
per refers to the fact that the free energies of the chosen   
reference systems are analytic, and can be meaningfully U − Uh
A(V, T0 )−Ah (V, T0 ) = −kB T0 ln exp − ,
compared between distinct reference systems including kB T0 V,T0 ,λ=0
the ones that contain different numbers of particles. (7)
A harmonically-coupled crystal of N atoms with a con- where h. . .iV,T0 ,λ=0 denotes the ensemble average for the
strained center of mass constitutes a convenient refer- harmonic crystal at T0 and V , and U and Uh denote the
ence system for a solid (point a in Figure 1). Taking the real and harmonic potentials, respectively. Note that in
phonon frequency for the crystal to be {ωi=1...3N −3 }, [26] order to ensure the statistical efficiency of this perturba-
one can obtain an expression for the classical free energy tive approach, the standard deviation of U − Uh has to
of such a Debye crystal at the temperature T0 : be of the order of kB T0 [29].
3N −3
X ~ωi
Ah (V, T0 ) = kB T0 ln . (5) C. The Helmholtz free energy as a function of
i=1
kB T0
temperature
Note that from the standpoint of performing thermody-
namic integration, the reference can be any harmonic Thermodynamic integration from a Debye crystal to
crystal that has the same number of particles as the real the fully-anharmonic potential tends to become very in-
system. For instance, one could even take a reference in efficient as the temperature increases. For this reason, it
which all particles are independently coupled to the lat- is often useful to perform a thermodynamic integration
tice sites with a constant spring term (i.e. an Einstein with respect to temperature to from a low to a high tem-
crystal) [6, 7, 27]. However, for better statistical effi- perature under the desired thermodynamic conditions.
ciency, it is better to choose a reference harmonic crystal Let us start by discussing how to perform this step un-
that has the same frequency modes and equilibrium con- der the NVT ensemble, which is the process indicated
figuration as the real crystal, both of which can be deter- by the red arrow (b to c) in Figure 1. The Helmholtz
mined for example via local energy minimization followed free energy of a system that has N atoms and a fixed
by a diagonalization of the Hessian matrix [28]. CM can be expressed by the well-known thermodynamic
4

integration expression the efficiency of the TI procedure by choosing wisely the


discretization points or, equivalently, by performing a
Z T1 hU iV,T + hKiV,T
A(V, T1 ) A(V, T0 ) change of variables that yields a smoother integrand [30].
= − dT, (8)
kB T1 kB T0 T0 kB T 2 In this case, it is convenient to perform a change of vari-
ables that ensures that the statistical error in the inte-
where hU iV,T and hKiV,T are the ensemble averages of grand is roughly constant at all temperatures. Assuming
the potential energy and the kinetic energy, respectively. the temperature dependence of the fluctuations in the
One way to improve the convergence of Eqn. (8) is to anharmonic potential energy is similar to its harmonic

2
consider that the ensemble average of the classical kinetic counterpart, i.e. δU 2 V,T − hδU iV,T ∼ T , the required
energy of a system that has N atoms and a fixed CM is change of variable is y = ln(T /T0 ), which transforms the
analytic: hKiV,T = (3N − 3)kB T /2. Furthermore, one integral into the form
can also consider that if the potential was harmonic, also
hU i would take the same value. Thus, one can take Z T1
hδU iV,T
Z ln(T1 /T0 )
hδU iV,T0 ey
2
dT = dy. (13)
kB T T0 T 0 T0 ey
hδU iV,T = hU iV,T − A(V, 0) − (3N − 3) (9)
2 In other words, one should select temperatures that
that measures the temperature-dependent anharmonic are equally spaced in ln(T ) in simulations. Coinciden-
part of the potential energy. Note that A(V, 0) = tally, this selection is also optimal for performing replica
hU iV,0 . After performing analytically some of the in- exchanges between the systems at different tempera-
tegrals, Eqn. (8) becomes tures [31] - which should be done whenever possible as it
will greatly benefit statistical convergence.
A(V, T1 ) A(V, 0) A(V, T0 ) − A(V, 0) Another advantage of performing parallel tempering is
= + that it requires sufficient overlap between adjacent repli-
kB T1 kB T1 kB T0 cas at temperatures Ti and Ti+1 . Under these circum-
Z T1
T1 hδU iV,T stances, hU iV,T for Ti < T < Ti+1 can be evaluated via
− (3N − 3) ln − dT. (10)
T0 T0 kB T 2 re-weighting, such as
 
In quasi-harmonic systems, one can further reduce the
h U 1 1 i
U exp − −
variance of the integrand. One can use again the an- kB T Ti
alytical expression for hKiV,T , together with the virial hU iV,T =   V,Ti . (14)
h U 1 1 i
theorem, to write exp − −
kB T Ti V,Ti
N
kB T 1X The integral in Eqn. (10) can thus be solved analytically
(3N − 3) = hKi = − hFi qi i . (11)
2 2 i=1 to give an exact (within statistical uncertainty) expres-
sion for the contribution to the integral from the [Ti , Ti+1 ]
Here qi and Fi are the position of atom i and the force window:
vector acting on it. Since the average force hFi i is zero,
one can also add an arbitrary reference position q̂i , and A(V, Ti+1 ) A(V, Ti )
− =
write [12, 14] kB Ti+1 kB Ti
  
* N
+ 3N − 3 Ti+1 U 1 1
1X − ln −ln exp − − ,
hδU iV,T = U + Fi (qi − q̂i ) − A(V, 0) (12) 2 Ti kB Ti+1 Ti V,Ti
2 i=1 (15)
V,T

If the potential energy surface is perfectly harmonic, and which effectively turns the thermodynamic integration
if one take q̂i equal to the equilibrium position of atom i, formalism into a sequence of free energy perturbations,
it is easy to verify that the virial term would cancel com- eliminating completely the integration error.
pletely the fluctuations in the potential energy. Even if
the potential is quasi-harmonic, as long as it is not dif-
fusive even at high temperatures, the use of the virial D. From the Helmholtz free energy to the Gibbs
reference and Eqn. (12) can substantially improve the free energy
statistical efficiency in the estimation of hδU iV,T . How-
ever, when the motion of atoms is strongly anharmonic More often than not, the isothermal–isobaric ensemble
or diffusive, an atom can start vibrating around a differ- (NPT) provides a more natural framework to describe
ent equilibrium position, and in that case the statistical the thermodynamic conditions of real systems than the
efficiency of the straightforward expression Eqn. (9) is NVT ensemble. However, the harmonic crystal (a in Fig-
better. ure 1) that was used as the absolute constant-volume ref-
Besides improving the convergence of each tempera- erence in previous sections does not extend naturally to
ture window, one can try to improve the accuracy and the NPT ensemble because its pressure is not well-defined
5

(e.g. the Einstein crystal is a system of independent par- E. The Gibbs free energy as a function of
ticles) [32]. As a result, it is not convenient to perform temperature
TI with respect to λ from the reference crystal to the real
crystal under the NPT ensemble. One way to avoid NPT Having converted a harmonic-reference Helmoltz free-
simulation involves performing multiple simulations at energy to a constant-pressure Gibbs free energy at a given
different constant volumes, and then compute the Gibbs temperature T0 , one can easily perform a thermodynamic
free energy and the equilibrium volume of the system by integration over temperature in the NPT ensemble (a
evaluating explicitly the integral (3). This is often done path indicated by the green arrow in Figure 1). For a
using a harmonic expression for the free energy at the dif- system with N atoms and a restricted CM, the expression
ferent volumes, leading to the so-called quasi-harmonic reads
approximation (QHA) [33, 34]. Alternatively, the equi- Z T1
librium volume can be computed by performing a single G(P, T1 ) G(P, T0 ) hHiP,T
= − dT, (19)
NPT simulation at the desired temperature, and then kB T1 kB T0 T0 kB T 2
A(hV iP,T , T ) is used as a proxy for G(P, T ) [32]. In this
section, we argue that the transformation between the kB T
where hHiP,T = hU iP,T + (3N − 3) + P hV iP,T is the
Helmholtz free energy and the Gibbs free energy, which 2
is the process marked by the dashed blue arrow in Fig- enthalpy. Starting from this expression, one can apply all
ure 1, can be conducted rigorously. This process effec- the techniques mentioned in Section II C, for example one
tively allows us to convert at will between the two ensem- can take
bles when performing thermodynamic integrations with kB T
respect to ensemble temperature. hδHiP,T = hHiP,T − G(P, 0) − (3N − 3) , (20)
2
The expression for the Gibbs free energy of a system
as an integral over the Helmholtz free energy is given by where G(P, 0) = hU iP,0 +P hV iP,0 . Doing the integration
Eqn. (3). This expression can be combined with that in Eqn. (19) explicitly leaves
for the distribution of volume fluctuations for the system
under the NPT ensemble G(P, T1 ) G(P, 0) G(P, T0 ) − G(P, 0)
    = +
PV A(V, T ) kB T1 kB T1 kB T0
exp − exp − Z T1
k T k T T1 hδHiP,T
ρ (V |P, T ) = B   B , , − (3N − 3) ln − dT. (21)
R PV A(V, T ) T0 T0 kB T 2
dV exp − exp −
kB T kB T
(16) In addition, one can also use the virial theorem
which is just the normalized probability of observing the (Eqn. (11)), the change of variable in the integration
system to have instantaneous volume V in a simulation (Eqn. (13)), and parallel tempering to further accelerate
under constant P and T. We can then write the convergence. When performing parallel tempering,
one can eliminate the thermodynamic integration error
G(P, T ) = A(V, T ) + P V + kB T ln ρ (V |P, T ) , (17) by using a free-energy perturbation to compute the in-
crement of G between two replicas at temperatures Ti+1
which is valid for arbitrary V .
and Ti :
In practice, one can run NPT simulations for a system
and compute ρ (V |P, T ) just by accumulating the his- G(P, Ti+1 ) G(P, Ti )
togram of the instantaneous volume of the system. After − =
that, one can select a volume V , preferably the one that kB Ti+1 kB Ti
   
maximizes ρ (V |P, T ) for the sake of better statistical ef- 3N − 3 Ti+1 U + PV 1 1
− ln −ln exp − − .
ficiency in the determination of ρ (V |P, T ), and compute 2 Ti kB Ti+1 Ti P,Ti
A(V, T ) for the same system at that volume using the (22)
route a to b in Figure 1. Finally, the Gibbs free energy
can be obtained applying Eqn. (17).
For a solid system, in order to avoid residual strain and F. Finite size effects
elastic energy, one can vary the shape of the simulation
cell instead of using a fixed shape in NPT simulations Most of the time, one is interested in computing the
under a hydrostatic pressure [35]. To account for the free energy per atom of a bulk, infinite system, or the
degree of freedom associated with the variable cell, in excess free energy of a defect in the dilute limit. An
this case Eqn. (17) should be modified to read atomistic simulation, however, is inevitably restricted to
G(P, T ) = A(h, T ) + P det(h) + kB T ln ρ (h|P, T ) , (18) a finite system size, which can result in deviations from
the ideal case. Many of these finite-system-size effects
where h is a matrix that represents the dimensions of have been documented in the literature. First of all,
a simulation cell, and A(h, T ) is the free energy of the in the limit of small system size, the free energy of the
system evaluated at constant cell dimensions. system is not an extensive quantity. Taking the ideal
6

gas part of the Helmholtz free energy Aid (N, V, T ) = B. Simulation details and results
VN
−kB T ln 3N in Eqn. (2) for example, one can see
Λ N! We studied a BCC iron system using a widely used
that EAM potential [39, 40]. This potential was fitted with
  the BCC vacancy formation energy at 0K but lacks a
Aid (N, V, T ) V 1 ln N 1 thermally stable FCC phase [39, 41]. Iron exhibits phase
= 1 − ln + ln 3 − +O (23)
N kB T N Λ 2N N transitions between BCC α-iron, FCC γ -iron and a
BCC δ-phase when increasing the temperature at ambi-
using Stirling’s formula. The leading ln N/N term is a ent pressure, which are largely due to the magnetism of
well-documented finite size effect that reduces to zero in the material [41]. Since the stabilization of the austenitic
the thermodynamic limit [25, 36]. The Gibbs free energy phase is due to quantum mechanical effects and this EAM
per atom G(N, P, T )/N as well as kB T ln ρ (V |P, T ) /N in potential does not reproduce it, we neglect the FCC
Eqn. (16) also displays a similar dependence on ln N/N . phase in the present study, and performed the simula-
Constraining the center of mass of the system in simu- tions considering a perfect BCC crystal (250 atoms) and
lations also introduces a non-extensive correction to the a BCC crystal with a vacancy (249 atoms). In all the
free energy, simulations, the centers of mass of the systems were con-
strained.
∆Acm (N, V, T ) = A(N, V, T ) − Acm (N, V, T ), (24) At high temperatures, the computation of the free en-
ergy of the crystal with a vacancy is particularly prob-
lematic using the integration over λ in Eqn. (6) due to
where Acm denotes the Helmholtz free energy of the sys-
the onset of diffusion in simulations [13]. This difficulty is
tem with fixed center of mass [5, 25]. Fortunately, it
circumvented here as we performed the integration from
is easy to correct for this part, because the expression
the harmonic crystal to the real crystal at a low tem-
for ∆Acm in Eqn. (4) is analytic and trivial to compute.
perature T0 =100K and at the equilibrium cell size (the
More subtle sources of finite size effects come from the
yellow arrow in Figure 1). This harmonic crystal has
cutoff of potentials, and from the discretization of the
the same phonon modes and Hessian matrix as the real
vibrational phonon spectrum due to the size of the su-
system [28, 42]. Note that at this step the Helmholtz
percell in simulations [32]. To help with this issue, there
free energy of a crystal with a vacancy that sits at a
are interpolation techniques that help accelerate the con-
fixed lattice site is computed, as the vacancy does not
vergence of the computed phonon dispersion relation [37].
diffuse during the simulations at T0 =100K. After that,
It is worth stressing that for system sizes that can be we switched to the NPT ensemble (the blue arrow in
reached easily in simulations using empirical force fields, Figure 1), and ran simulations at different temperatures
finite-size effects may not be significant. However, one and zero hydrostatic pressure, using stochastic velocity
should always be aware of their presence and check for re-scaling for temperature control and the anisotropic
system-size convergence, particularly in ab initio calcula- Nose-Hoover barostat to vary the dimensions of the or-
tions where the number of atoms that can be simulated thorhombic periodic supercell [43, 44]. During this step,
is highly restricted. To minimize the impact of finite-size we obtained the temperature dependence of the free en-
effects one should always compare free energies between ergies using Eqn. (19) (the green arrow in Figure 1). At
systems of similar sizes, to benefit for a (partial) error high temperatures, the vacancy does diffuse but diffu-
cancellation. sion does not change the values of hδHiP,T compared
with the case when the vacancy is fixed at one site, due
to the translational symmetry of the lattice. As such, the
Gibbs free energy of a crystal with a fixed vacancy was
III. APPLICATION 1: VACANCY FREE
obtained after integration using Eqn. (19).
ENERGY
All the detailed procedures, the key data points, anno-
tated input files and python notebooks for data analysis
A. Introduction are included in the supplemental material [45]. The ab-
solute Gibbs free energies of the perfect BCC iron and
A vacancy is a type of point defects in a crystal, the crystal with a fixed vacancy are plotted in Figure 2.
in which an atom is removed from one of the lattice
sites. At any given temperature and pressure up to the
solid-liquid coexistence line, an equilibrium concentra- C. Gibbs free energy of a vacancy
tion exp [−Gv /kB T ] of vacancies exists, where Gv is the
Gibbs free energy of a vacancy. Often, particularly in The Gibbs free energy of a fixed vacancy in the crystal
materials produced by fast quenching, a non-equilibrium can be expressed as
concentration of vacancies can persist at low tempera-
ture, which can play an important role in technologically Nvacancy
Gv = Gvacancy − Gperfect . (25)
relevant solid-state transformations [38]. Nperfect
7

1.8
0K
-1050
1.6

G Vacancy [eV]
-1100
0K HAR QH
G [eV]

-1150 1.4
HAR

-1200 ANH 1.2


ANH
-1250
500 1000 1500 1
Temperature [K] 500 1000 1500
Temperature [K]
FIG. 2. The solid and the dashed lines indicate the Gibbs
free energy of the perfect BCC crystal and the crystal with FIG. 3. The Gibbs free energy associated with a fixed va-
a fixed vacancy, respectively. The black lines are the estima- cancy in BCC iron estimated using potential energy differ-
tions using just the potential energy at 0K, the blues lines ence (PE) at 0K, the harmonic approximation (HAR), and
show the results from harmonic approximations, and the red the thermodynamic integration method that considers anhar-
lines indicate the predictions that takes into account of all the monicity (ANH). QH indicates the quasi-harmonic approxi-
anharmonic effects using the TI approach. mation using the equilibrium configuration at 1500K. Statis-
tical uncertainties are indicated by the error bars.

We used three different methods to estimate this quantity


– namely a minimum-potential energy calculation, a har- the free energy increase that is associated with the pres-
monic free-energy estimate and the fully anharmonic TI ence of the intrinsic stacking fault plane. The stacking
– and plot the results in Figure 3 as a function of temper- fault free energy has a significant effect on the plastic
ature. Notice that the defect free energy is tiny compared deformation behavior of crystalline materials. For exam-
with the magnitude of the absolute free energy of both ple, metals with a low γSF form more stacking faults and
systems. The difference between the predictions from the twins, and more extended partial dislocations which have
harmonic approximation and the TI is largely negligible reduced mobility [46–48]. Stacking faults also provide a
at low temperatures, but becomes significant when the strong barrier to dislocation gliding [46–48].
temperature approaches the melting point 1772K for this
EAM potential system [39]. To investigate whether this
difference stems from a shift in the phonon spectra due B. Simulation details and results
to lattice expansion, or from anharmocity, we analyzed
0
the vibrational modes {ωi=1...3N −3 } using the equilib-
rium configuration of each system at 1500K, computed A widely used EAM potential that describes the Ni-Ni
the vacancy free energy under this quasi-harmonic ap- interactions was employed [40, 49]. Note that different
proximation, and plotted the result as the orange dot in EAM potentials usually yield quite divergent estimates
Figure 3. It can be seen that the harmonic contribution for the stacking fault energy, and a comprehensive com-
at 1500K cannot explain the difference, and therefore the parison between them together with a few DFT results
difference is mainly due to anharmonic effects. can be found in Ref. [19]. As such, the purpose of this
application is to highlight the temperature dependence
of γSF for a generic metallic system rather than a quan-
titative analysis for pure Ni, which would require a more
IV. APPLICATION 2: STACKING FAULT FREE reliable interatomic potential.
ENERGY
We performed the simulations of a perfect FCC crystal
and of a FCC crystal with a stacking fault layer sepa-
A. Introduction rately. The perfect FCC crystal used in simulations has
12 layers of {111} planes, and a total of Nperfect = 1440
A stacking fault (SF) is a defect in the planar stack- atoms. The defective crystal has 11 layers with the stack-
ing sequence of atoms in a crystal. While the perfect ing ABC|ABC|AB|ABC|, and a total of Nsf = 1320
stacking for FCC crystals metals on the {111} plane is atoms. We calculated separately the Gibbs free energy
ABC|ABC|ABC|, an intrinsic stacking fault changes the for both systems, and the results are reported in Figure 4.
arrangement to ABC|AB|ABC| as if one plane had been The centers of the mass of the systems were constrained
removed. The stacking fault free energy (γSF ) measures in all the simulations. All the detailed procedures and
8

60
-6000
0K

HAR

SFE [mJ /m 2 ]
-6500 40
G [eV]

-7000 0K QH
HAR 20
-7500 ANH ANH
500 1000 1500 0
Temperature [K] 500 1000 1500
Temperature [K]
FIG. 4. The solid and the dashed lines indicate the absolute
Gibbs free energy of the perfect FCC crystal and the crystal FIG. 5. The area-specific stacking fault free energy esti-
with a stacking fault, respectively. The black lines are the mated from potential energy difference at 0K, the harmonic
estimations using just the potential energy at 0K, the blues approximation (HAR), and the thermodynamic integration
lines shows the results from harmonic approximations, and method that considers anharmonicity (ANH). QH indicates
the red lines indicate the predictions that takes into account of the quasi-harmonic approximation using the equilibrium con-
all the anharmonic effects using thermodynamic integration. figuration at 1408K. Statistical uncertainties are indicated by
The statistical errors are minute at the scale of the plot. the error bars.

the key data points used in the calculations are included We used three different methods to estimate γSF (T ),
in the supplemental material [45], along with annotated and juxtaposed them in Figure 5. We found the estima-
input files and python notebooks for data analysis. In tions from just the potential energy at 0K and the har-
brief, we first computed the Helmoltz free energy of a monic approximation are only accurate at very low tem-
real system at T0 =90K by thermodynamic integration peratures. The prediction of the quasi-harmonic approx-
with respect to λ starting from a reference system using imation using the equilibrium configuration at 1408K is
Eqn. (6). The reference system is a harmonic crystal that more accurate than the harmonic approximation, but still
has the same phonon modes and the Hessian matrix as falls short. The inclusion of all the entropic and anhar-
the real system at the minimum potential energy [28, 42], monic effects dramatically decreases the stacking fault
and the simulation cell is kept constant at the equilib- free energy at high temperatures, making the magnitude
rium size of the real system at 90K. Afterwards, at tem- of the stacking fault free energy at T >1200K less than
peratures that are equally spaced in ln(T ), independent one third of the value predicted from the 0K potential
molecular dynamics simulations were performed under energies.
the NPT ensemble with zero external stress [43, 44], in
order to compute the anharmonic contributions hδHiP,T
and employ thermodynamic integration with respect to
V. CONCLUSIONS
T using Eqn. (21). We also found the use of the virial
theorem to significantly reduce the fluctuations in the
estimation of hδHiP,T when we select q̂ to be the equi- In the present study we have discussed a number of
librium position of atoms in Eqn. (12). thermodynamic integration routes, and devised a strat-
egy to combine them in a way that makes the best use of
the similarity between a crystalline system and a har-
C. Stacking fault free energy estimations monic Debye crystal at low temperatures, and at the
same time fully take into account the anharmonic ef-
fects. Furthermore, we discuss how to convert between
The free energy excess (γSF × Area) that is associated
free energy values estimated under constant-volume and
with the stacking fault plane with a surface area equal
constant-pressure thermodynamic ensembles, and incor-
to the cross section of the simulation supercell is just the
porate several techniques to enhance the efficiency and
difference between the free energies of a crystal with a
accuracy of free energy estimation.
stacking fault and a perfect bulk FCC crystal that have
the same number of atoms: The thermodynamic integration method does not rely
on any approximation for anharmonic effects, thus it can
Nsf also serve as a benchmark for other approximation meth-
γSF × Area = Gsf − Gperfect . (26)
Nperfect ods such as quasi-harmonic approximations, self consis-
9

tent phonons [50–52], the Greens function approach [53]. to encourage and facilitate future efforts that aim to use
In addition, it is also possible to include the oft-neglected thermodynamic integration methods to accurately pre-
quantum nuclear effects by adding an extra TI route from dict the defect concentrations, phase diagrams, or any
classical to quantum mechanical system by employing of the many materials properties that are determined by
path-integral molecular dynamics [9, 54]. the Gibbs free energy.
We computed the free energy of a vacancy in BCC
iron and the stacking fault free energy of FCC nickel. In
both cases, the anharmonic effects significantly lower the ACKNOWLEDGMENTS
free energy of defects close to the melting point. These
results reinforce the notion that it is not sufficient to We thank William Curtin for insightful comments on
use potential energy minima, or the associated harmonic an early version of the manuscript. We also thank David
free energy, to estimate the stability of crystallographic Wilkins and Benjamin Helfrecht for critically reading
defects at high temperatures. In the supplementary ma- the manuscript. BC acknowledges funding from the
terials [45], We have included all the input files, work- Swiss National Science Foundation (project ID 200021-
flows, and data analysis routines that were employed to 159896). MC acknowledges financial support from the
produce the results discussed in the present paper, hoping CCMX Project “AM3 ”.

[1] A. Laio and M. Parrinello, Proceedings of the National 112, 105501 (2014).
Academy of Sciences 99, 12562 (2002). [22] B. Grabowski, L. Ismer, T. Hickel, and J. Neugebauer,
[2] G. M. Torrie and J. P. Valleau, Journal of Computational Physical Review B 79, 134106 (2009).
Physics 23, 187 (1977). [23] A. Glensk, B. Grabowski, T. Hickel, and J. Neugebauer,
[3] P. G. Bolhuis, D. Chandler, C. Dellago, and P. L. Physical Review X 4, 011018 (2014).
Geissler, Annual review of physical chemistry 53, 291 [24] M. P. Allen and D. J. Tildesley, Computer simulation in
(2002). chemical physics, Vol. 397 (Springer Science & Business
[4] M. Tuckerman, Statistical mechanics: theory and molec- Media, 2012).
ular simulation (Oxford University Press, 2010). [25] J. M. Polson, E. Trizac, S. Pronk, and D. Frenkel, The
[5] C. Vega, E. Sanz, J. Abascal, and E. Noya, Journal of Journal of Chemical Physics 112, 5339 (2000).
Physics: Condensed Matter 20, 153101 (2008). [26] The zero-frequency translational modes are excluded due
[6] L. M. Ghiringhelli, J. H. Los, E. J. Meijer, A. Fasolino, to the constraint on CM.
and D. Frenkel, Physical review letters 94, 145701 (2005). [27] E. G. Noya, M. Conde, and C. Vega, The Journal of
[7] J. M. Polson and D. Frenkel, The Journal of chemical chemical physics 129, 104704 (2008).
physics 109, 318 (1998). [28] P. Y. Ayala and H. B. Schlegel, The Journal of chemical
[8] J. Anwar, D. Frenkel, and M. G. Noro, The Journal of physics 108, 2314 (1998).
chemical physics 118, 728 (2003). [29] M. Ceriotti, G. A. Brain, O. Riordan, and D. E.
[9] M. Rossi, P. Gasparotto, and M. Ceriotti, Phys. Rev. Manolopoulos, in Proc. R. Soc. A (The Royal Society,
Lett. 117, 115702 (2016). 2011) p. rspa20110413.
[10] L. Li, T. Totton, and D. Frenkel, The Journal of Chem- [30] M. Ceriotti and T. E. Markland, J. Chem. Phys. 138,
ical Physics 146, 214110 (2017). 014112 (2013).
[11] E. Sanz, C. Vega, J. Abascal, and L. MacDowell, Phys- [31] D. J. Earl and M. W. Deem, Physical Chemistry Chem-
ical review letters 92, 255701 (2004). ical Physics 7, 3910 (2005).
[12] S. G. Moustafa, A. J. Schultz, and D. A. Kofke, Physical [32] R. Freitas, M. Asta, and M. de Koning, Computational
Review E 92, 043303 (2015). Materials Science 112, 333 (2016).
[13] S. M. Foiles, Physical Review B 49, 14930 (1994). [33] J. Xie, S. de Gironcoli, S. Baroni, and M. Scheffler, Phys-
[14] A. J. Schultz, S. G. Moustafa, W. Lin, S. J. Weinstein, ical Review B 59, 965 (1999).
and D. A. Kofke, Journal of chemical theory and compu- [34] R. Ramı́rez, N. Neuerburg, M.-V. Fernández-Serra, and
tation 12, 1491 (2016). C. Herrero, The Journal of chemical physics 137, 044502
[15] D. H. Warner and W. Curtin, Acta Materialia 57, 4267 (2012).
(2009). [35] M. Parrinello and A. Rahman, Journal of Applied physics
[16] M. Valtiner, M. Todorova, G. Grundmeier, and J. Neuge- 52, 7182 (1981).
bauer, Physical review letters 103, 065502 (2009). [36] W. G. Hoover, The Journal of Chemical Physics 49, 1981
[17] D. Turnbull, Trans. Aime 191, 661 (1951). (1968).
[18] R. LeSar, R. Najafabadi, and D. Srolovitz, Physical Re- [37] P. Giannozzi, S. De Gironcoli, P. Pavone, and S. Baroni,
view Letters 63, 624 (1989). Physical Review B 43, 7231 (1991).
[19] J. A. Zimmerman, H. Gao, and F. F. Abraham, Mod- [38] S. Pogatscher, H. Antrekowitsch, M. Werinos,
elling and Simulation in Materials Science and Engineer- F. Moszner, S. S. A. Gerstl, M. F. Francis, W. A.
ing 8, 103 (2000). Curtin, J. F. Löffler, and P. J. Uggowitzer, Phys. Rev.
[20] S. Ryu, K. Kang, and W. Cai, Proceedings of the Na- Lett. 112, 225701 (2014).
tional Academy of Sciences 108, 5174 (2011). [39] M. Mendelev, S. Han, D. Srolovitz, G. Ackland, D. Sun,
[21] W. K. Kim and E. B. Tadmor, Physical review letters and M. Asta, Philosophical magazine 83, 3977 (2003).
10

[40] G. Bonny, R. C. Pasianot, N. Castin, and L. Malerba, [48] V. Yamakov, D. Wolf, S. Phillpot, A. Mukherjee, and
Philosophical magazine 89, 3531 (2009). H. Gleiter, Nature materials 3, 43 (2004).
[41] C. Engin, L. Sandoval, and H. M. Urbassek, Modelling [49] A. F. Voter and S. P. Chen, MRS Online Proceedings
and Simulation in Materials Science and Engineering 16, Library Archive 82 (1986).
035005 (2008). [50] D. Hooton, The London, Edinburgh, and Dublin Philo-
[42] M. Ceriotti, J. More, and D. E. Manolopoulos, Computer sophical Magazine and Journal of Science 46, 422 (1955).
Physics Communications 185, 1019 (2014). [51] S. E. Brown, I. Georgescu, and V. a. Mandelshtam, J.
[43] G. Bussi, D. Donadio, and M. Parrinello, The Journal Chem. Phys. 138, 044317 (2013).
of chemical physics 126, 014101 (2007). [52] I. Errea, M. Calandra, and F. Mauri, Physical Review
[44] S. Plimpton, J. Comp. Phys. 117, 1 (1995). B 89, 064302 (2014).
[45] “Supplemental material,” . [53] C. Campaná and M. H. Müser, Physical Review B 74,
[46] S. Allain, J.-P. Chateau, O. Bouaziz, S. Migot, and 075420 (2006).
N. Guelton, Materials Science and Engineering: A 387, [54] M. Rossi, W. Fang, and A. Michaelides, J. Phys. Chem.
158 (2004). Letters 6, 4233 (2015).
[47] H. Van Swygenhoven, P. Derlet, and A. Frøseth, Nature
materials 3, 399 (2004).

View publication stats

You might also like