You are on page 1of 12

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

net/publication/277895576

Self-consistent phonon calculations of lattice dynamical properties in cubic


SrTiO$_{3}$ with first-principles anharmonic force constants

Article  in  Physical Review B · June 2015


DOI: 10.1103/PhysRevB.92.054301 · Source: arXiv

CITATIONS READS

248 1,243

2 authors, including:

Terumasa Tadano
National Institute for Materials Science
63 PUBLICATIONS   1,763 CITATIONS   

SEE PROFILE

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

First-principles simulation of phonon anharmonicity View project

All content following this page was uploaded by Terumasa Tadano on 20 August 2015.

The user has requested enhancement of the downloaded file.


Self-consistent phonon calculations of lattice dynamical properties in cubic SrTiO3
with first-principles anharmonic force constants
Terumasa Tadano1 and Shinji Tsuneyuki2, 3
1
Department of Applied Physics, The University of Tokyo, Tokyo 113-8656, Japan
2
Department of Physics, The University of Tokyo, Tokyo 113-0033, Japan
3
Institute for Solid State Physics, The University of Tokyo, Kashiwa 277-8581, Japan
We present an ab initio framework to calculate anharmonic phonon frequency and phonon lifetime
that is applicable to severely anharmonic systems. We employ self-consistent phonon (SCPH) theory
with microscopic anharmonic force constants, which are extracted from density-functional calcula-
tions using the least absolute shrinkage and selection operator technique. We apply the method
to the high-temperature phase of SrTiO3 and obtain well-defined phonon quasiparticles that are
free from imaginary frequencies. Here we show that the anharmonic phonon frequency of the an-
tiferrodistortive mode depends significantly on the system size near the critical temperature of the
arXiv:1506.01781v1 [cond-mat.mtrl-sci] 5 Jun 2015

cubic-to-tetragonal phase transition. By applying perturbation theory to the SCPH result, phonon
lifetimes are calculated for cubic SrTiO3 , which are then employed to predict lattice thermal con-
ductivity using the Boltzmann transport equation within the relaxation-time approximation. The
presented methodology is efficient and accurate, paving the way toward a reliable description of
thermodynamic, dynamic, and transport properties of systems with severe anharmonicity, including
thermoelectric, ferroelectric, and superconducting materials.

PACS numbers: 63.20.dk,63.20.kg,63.20.Ry,77.80.B-

I. INTRODUCTION tained efficiently and systematically using either den-


sity functional perturbation theory (DFPT) [14] or the
Lattice anharmonicity plays an important role in finite-displacement approach [15]. Using the cubic terms,
characterizing various physical properties of solids and phonon linewidth can be obtained by evaluating the bub-
molecules, including the temperature-dependence of vi- ble diagram [Fig. 1(b)]. This type of calculation has been
brational frequencies, thermal expansion and phase sta- performed to predict the lattice thermal conductivity of
bility of solids [1]. It is also responsible for the finite many solids [3, 4, 16, 17] and can also be applied to com-
phonon linewidth and the lattice thermal conductivity plex materials [18]. To estimate the phonon frequency
κL , which is a key quantity when optimizing the thermo- shift due to lattice anharmonicity, one also needs to com-
electric figure-of-merit ZT [2]. The magnitude of anhar- pute the loop diagram [Fig. 1(a)] using the quartic terms.
monicity varies significantly for different materials. For The calculation of the quartic terms can, in principle, be
example, covalent materials such as silicon, diamond and achieved using the finite-displacement approach. How-
graphene are very harmonic and show high thermal con- ever, since the number of quartic parameters increases
ductivities [3, 4]. Conversely, thermoelectric and ferro- rapidly as the number of atoms in the supercell increases,
electric (FE) materials often show severe anharmonicity, such calculations have only been reported for simple sys-
demonstrated by inelastic neutron scattering spectra and tems [19, 20].
ultralow κL values [5–7]. Anharmonic effects can also be The perturbative approach is valid only when the an-
significant in superconductors [8–10] and materials under harmonic self-energies are sufficiently small compared
extreme conditions [11, 12]. To develop a robust under- with the harmonic frequency. Therefore, one cannot ex-
standing of anharmonic properties of solids, a reliable and pect this technique to yield accurate results for severely
versatile computational method is required. Therefore, anharmonic systems. High-temperature phases of FE
the development of first-principles methods to calculate material are typical cases where the perturbation ap-
anharmonic properties of solids and molecules has been proach fails because of the imaginary frequencies of har-
the subject of intense research in recent years. monic phonons. To overcome this limitation, it is nec-
Many-body perturbation theory is one approach for essary to employ a non-perturbative approach to treat
treating lattice anharmonicity. This technique consid- anharmonic effects.
ers the anharmonic effects as self-energies [13]. The self- Methods based on ab initio molecular dynam-
energies can be calculated using a systematic approxima- ics (AIMD) can consider anharmonic effects non-
tion to the Feynman diagrams, where the lowest-order perturbatively. From the velocity-velocity autocorrela-
approximation is usually employed in the ab initio calcu- tion function calculated using the trajectory of an AIMD
lations based on density-functional theory (DFT). Per- simulation, one can obtain the vibrational density of
forming this calculation requires the cubic and quartic states with full anharmonicity. To obtain the anhar-
force constants, which are the third- and fourth-order monic frequency and linewidth of individual phonons, the
derivatives of the Born-Oppenheimer potential energy velocity should be projected onto the phonon eigenvec-
surface, respectively. The third-order terms can be ob- tor [12]. Inherent in this procedure is the assumption
2

that the phonon eigenvectors are not altered by anhar- our approach. Finally, we conclude this work in Sec. V.
monic effects. Such an assumption, however, is valid
only for simple systems containing a few atoms in the
primitive cell. The temperature-dependent effective po- II. SELF-CONSISTENT PHONON THEORY
tential (TDEP) method [21] is another AIMD-based ap-
proach. The TDEP method optimizes the effective har- A. Potential energy expansion
monic force constants within an AIMD simulation at a
target temperature. This method should be useful in high The dynamics of interacting ions within the Born-
temperature because it allows both the phonon eigenvec- Oppenheimer approximation are described by the Hamil-
tors and the internal coordinate system to be changed by tonian H = T + U , where T is the kinetic energy and U
anharmonic effects. However, since the AIMD is based on is the potential energy of the system. When U is an an-
the Newton equation of motion, the MD-based methods alytic function of atomic displacements from equilibrium
cannot account for the zero-point vibration. Therefore, positions {u} , it can be expanded as a Taylor series with
these methods cannot be applied to superconductors and respect to u as
ferroelectric materials in the low-temperature range.
Self-consistent phonon (SCPH) theory [22] is another U = U0 + U2 + U3 + U4 + · · · , (1)
approach for including anharmonic effects beyond per- 1 X
turbation theory that considers the quantum effect of Un = Φµ1 ...µn (ℓ1 κ1 ; . . . ; ℓn κn )
n!
phonons. Other first-principles methods are able to {ℓ,κ,µ}
compute anharmonic phonon frequencies related to the × uµ1 (ℓ1 κ1 ) · · · uµn (ℓn κn ). (2)
SCPH theory: self-consistent ab initio lattice dynamics
(SCAILD) [23] and stochastic self-consistent harmonic Here, Un is the nth-order contribution to the po-
approximation (SSCHA) [24]. To avoid the cumbersome tential energy, uµ (ℓκ) is the atomic displacement of
calculation of quartic force constants, these methods em- the atom κ in the ℓth cell along the µ direction,
ploy real-space stochastic approaches and displace atoms and Φµ1 ...µn (ℓ1 κ1 ; . . . ; ℓn κn ) is the nth-order interatomic
in the supercell to model anharmonic effects. force constant (IFC). In Eq. (1) the linear term U1 is
In this study, we have developed an efficient first- omitted because atomic forces are zero in equilibrium.
principles method to treat lattice anharmonicity. The In the harmonic approximation, only the quadratic
method is based on the SCPH theory, and the anhar- term U2 is considered and cubic, quartic, and higher-
monic frequency is estimated from the pole of the Green’s order terms are neglected. This allows the Hamiltonian
function. The cubic and quartic force constants necessary H0 = T + U2 to be represented in terms of the harmonic
for the present SCPH calculations are extracted from the phonon frequency ω. To compute the phonon frequency
DFT calculations using the recently proposed compres- ω, one needs to construct the dynamical matrix
sive sensing approach [25]. By combining the perturba-
tion theory with the solution to the SCPH equation, we 1 X ′
Dµν (κκ′ ; q) = √ Φµν (ℓκ; ℓ′ κ′ )eiq·r(ℓ ) , (3)
can also estimate the phonon lifetime and lattice thermal M κ M κ′ ℓ′
conductivity of severely anharmonic materials.
To confirm the validity of our approach, the method is where Mκ is the mass of atom κ, Φµν (ℓκ; ℓ′ κ′ ) are the
applied to the high-temperature phase of SrTiO3 with harmonic IFCs, and r(ℓ) is a translation vector of the
cubic symmetry (c-STO). SrTiO3 is one of the most primitive lattice. By diagonalizing the dynamical matrix,
studied perovskite oxides and is known to undergo the one obtains harmonic phonon frequencies as
cubic-to-tetragonal phase transition at 105 K accompa- 2
nied by the freezing-out of the antiferrodistortive (AFD) D(q)eqj = ωqj eqj , (4)
soft mode [26–28]. The FE phase transition is not ob-
served, even at 0 K, because of the zero-point vibration. where the index j labels the phonon modes for each crys-
Our approach can describe the temperature dependence tal momentum vector q and eqj is the polarization vector
of the soft-mode frequencies and lattice thermal conduc- of the phonon mode qj.
tivity of the severely anharmonic c-STO.
This paper is organized as follows. First, we introduce
B. Dyson equation
the SCPH theory and details of our implementation in
Sec. II. We describe the details of the computational con-
ditions, including the compressive sensing of force con- To derive the SCPH equation, we employ the
stants in Sec. III. The method is applied to c-STO and the many-body Green’s function theory. The one-phonon
results are presented in Sec. IV. In Sec. IV B, we exam- imaginary-time Green’s function is given as
ine the size- and temperature-dependence of anharmonic D E
phonon frequencies and compare these results with ex- Gqj,qj ′ (τ ) = Tτ Aqj (τ )A†qj ′ (0)
H
perimental values. We also calculate the lattice thermal
conductivity of c-STO in Sec. IV C to show the validity of =Z −1
Tr{e −βH
Tτ [Aqj (τ )A†qj ′ (0)]}, (5)
3

where Tτ is the time-ordering operator, Aqj (τ ) = (a) (b)


eτ H Aqj e−τ H is the displacement operator in the Heisen-
berg picture, Z = Tre−βH is the partition function, and
β = 1/kT , where k is the Boltzmann constant and T is
the temperature. The displacement operator is defined
as Aqj = bqj + b†qj where bqj and b†qj are the annihilation
and creation operators of the phonon qj, respectively.
It is straightforward to show that the Green’s function FIG. 1. Diagrams of the self-energies considered in this study.
satisfies Gqjj ′ (τ ) = Gqjj ′ (τ + β) for −β < τ < 0 and (a) The first-order diagram associated with the quartic term.
Gqjj ′ (τ ) = Gqjj ′ (τ − β) for 0 < τ < β, where we simply (b) The second-order diagram associated with the cubic term.
denote Gqj,qj ′ as Gqjj ′ . Because of these properties, we
can also show the following result for the Fourier trans-
form of the Matsubara Green’s function: Eq. (11) and consider only the first-order contribution to
(a)
Z β the phonon self-energy Σq , which is independent of ω,
Gqjj ′ (iωm ) = dτ Gqjj ′ (τ )eiωm τ , (6) as will be described in Sec. II C. Nevertheless, the self-
0 consistency is retained in the SCPH approach because
where ωm = 2πm/β is the Matsubara frequency. To the self-energy is a function of phonon frequencies and
obtain the Green’s function for anharmonic systems, we polarization vectors, which themselves are updated by
need to solve the Dyson equation. When one obtains diagonalizing the matrix Vq .
Gqjj ′ (iωm ) within some approximations, it is possible to
obtain the retarded Green’s function Gqjj ′ (ω) by analytic
C. Anharmonic self-energy
continuation to the real axis as Gqjj ′ (ω) = Gqjj ′ (iωm →
ω + iǫ) with a positive infinitesimal ǫ. The function Gqjj ′
has a pole at the energy corresponding to the renormal- Solving the SCPH equation requires a diagrammatic
ized frequency Ωqj . In the case of the harmonic ap- approximation to the phonon self-energy Σq (ω). In this
proximation, one can readily obtain the expression for study, we consider anharmonicity up to the fourth order,
Gqjj ′ (ω) as i.e. H = H0 + U3 + U4 , where Un is the nth-order contri-
bution to the potential energy surface expressed in terms
2ωqj of the displacement operator A. This can be obtained by
G0qjj ′ (ω) = − 2 δjj .
′ (7)
ω 2 − ωqj substituting
Therefore, the free-phonon Green’s function is diagonal s
− 12
X ~
in the phonon polarization index j and can be obtained uµ (ℓκ) = (N Mκ ) Aq eµ (κ; q)eiq·r(ℓ) (12)
from the harmonic phonon frequencies. q
2ω q
To estimate the phonon Green’s function Gqjj ′ (ω), and
thereby obtain the anharmonic frequency Ωqj , we solve for Eq. (1), where q labels the phonon modes defined as
the Dyson equation q = (q, j) and −q = (−q, j), and N is the number of q
points. We then obtain the following result:
[Gq (ω)]−1 = [G0q (ω)]−1 − Σq (ω). (8)
 n
Here we denote the retarded Green’s functions in the ma- 1 ~ 2X Φ(q1 ; . . . ; qn )
Un = ∆(q1 + · · · + qn ) √
trix form and Σq (ω) is the phonon self-energy, which can n! 2 ωq1 · · · ωqn
{q}
be estimated within a systematic diagrammatic approxi-
mation. Since the left-hand side of Eq. (8) becomes zero × Aq1 · · · Aqn . (13)
at the frequencies of the renormalized phonons, finding
The function ∆(q) becomes 1 if q is an integral mul-
the solution {Ωqj } is equivalent to solving the following
tiple of the reciprocal vector G and is 0 otherwise.
equation
Φ(q1 ; . . . ; qn ) is the reciprocal representation of the nth-
det {[G0q (ω)]−1 − Σq (ω)} = 0. (9) order IFCs defined by
1
By multiplying det (Λq2 ) from the left and right of Eq. (9) Φ(q1 ; . . . ; qn )
with the diagonal matrix Λqjj ′ = 2ωqj δjj ′ , one obtains n 1
X
= N 1− 2 (Mκ1 · · · Mκn )− 2 eµ1 (κ1 ; q1 ) · · · eµn (κn ; qn )
the following SCPH equation: {κ,µ}
2
det {ω − Vq (ω)} = 0, (10)
X
× Φµ1 ...µn (0κ1 ; . . . ; ℓn κn )ei(q2 ·r(ℓ2 )+···+qn ·r(ℓn )) .
1 1
2 ℓ2 ,...,ℓn
Vqjj ′ (ω) = ωqj δjj ′ − (2ωqj ) (2ωqj ′ ) Σqjj ′ (ω).
2 2 (11)
(14)
This equation needs to be solved self-consistently because
the self-energy is a function of the solution ω. How- In solving the SCPH equation, we consider only the
ever, in the present study we neglect the ω-dependency in first-order contribution to the phonon self-energy due to
4

the quartic term D. Computational implementation

(a) 1 X ~Φ(qj; −qj ′ ; q1 ; −q1 ) In this section we describe the details of the compu-
Σqjj ′ (iωm ) = − √
2 q 4 ωqj ωqj ′ ωq1 tational implementation used to solve the SCPH equa-
1
tion efficiently. The most expensive part of the SCPH
× [1 + 2n(ωq1 )], (15) equation is the calculation of the quartic coefficients in
Eq. (15), which are changed in each cycle of the itera-
which corresponds to the loop diagram shown in tive algorithm through an update of the phonon eigen-
Fig. 1(a). Here, n(ω) = [eβ~ω − 1]−1 is the Bose-Einstein vectors. To avoid recalculating the quartic coefficient in
distribution function. Since we continue the iteration each cycle, we employ a unitary transformation of the
cycle of the self-consistent equation [Eq. (11)] until we eigenvectors, as will be described below. Our approach
obtain a convergence with respect to the anharmonic fre- is inspired by the method proposed by Hermes and Hi-
quencies, the SCPH equation automatically includes an rata for molecules [31], which we extended to periodic
infinite class of anharmonic self-energies that can be gen- systems at finite temperatures.
erated from the loop diagram. In this study, we consider First, we construct the dynamical matrix D(q) from
the off-diagonal components of the self-energy to allow the harmonic IFCs and calculate eigenvalues and eigen-
for polarization mixing (PM), which we found to be im- vectors {ωq2 , eµ (κ; q)} for the gamma-centered N1 × N2 ×
portant for c-STO, as will be discussed in Sec. IV B. If N3 q-point grid. We then calculate the matrix elements
(a) (a)
we neglect the off-diagonal elements, Σqjj ′ ≈ Σqjj ′ δjj ′ , Fqq1 ,ijkℓ = Φ(qi; −qj; q1 k; −q1 ℓ) by Eq. (14) using the
the SCPH equation can be reduced to the diagonal form harmonic eigenvectors and quartic IFCs. Here, the index
q is restricted to the irreducible points that are com-
Ω2q = ωq2 + 2Ωq Iq(a) , (16) mensurate with the supercell size, whereas the index q1
includes all of the N1 ×N2 ×N3 grid points. The next step
1 X ~Φ(q; −q; q1 ; −q1 )
Iq(a) = [1 + 2n(Ωq1 )]. (17) is to diagonalize the following SCPH equation, which can
2 q 4Ωq Ωq1 be obtained from Eqs. (9) and (15):
1

This equation is equivalent to the one derived by a varia- [1] 2 1X ~ [1 + 2n(ωq1 k )]


Vqij = ωqi δij + Fqq1 ,ijkk . (20)
tional approach where the anharmonic free-energy within 2 2ωq1 k
q1 ,k
the first-cumulant expansion is minimized with respect to
trial frequencies [29]. Here we added the superscript 1 to the matrix Vq to ex-
To calculate the phonon linewidth, one needs to con- plicitly show that it is the first iteration of the SCPH
sider the bubble self-energy shown in Fig. 1(b), which equation. Then, by diagonalizing the Hermitian matrix
has a contribution from cubic anharmonicity given as Vq as Vq = Cq Wq Cq† , we obtain the updated phonon
1
[1]
frequencies ωqi = Wqii 2
. The corresponding polariza-
(b) 1 X ~Φ(−qj, q1 , q2 )Φ(qj ′ , −q1 , −q2 ) tion vectors can be obtained from the unitary matrix
Σqjj ′ (iωm ) = √
2N q ,q 8 ωqj ωqj ′ ωq1 ωq2 [1]
Cq . Let Eq and Eq denote the s × s matrices de-
1 2
[1] [1] [1]
× ∆(−q + q1 + q2 )F (iωm , 1, 2). (18) fined as Eq = (eq1 , . . . , eqs ) and Eq = (eq1 , . . . , eqs ),
respectively, where s is the number of phonon modes.
[1]
Here we introduced the ω-dependent function F defined It can then be shown that Eq and Eq are unitary
as transformations of each other, which can be written as
[1] [1]
Eq = Eq Cq . Because the phonon polarization vectors
X  1 + n1 + n2 are modified in this manner, we need to modify Eq. (20)
F (iωm , 1, 2) =
σ=±1
iωm + σ(ω1 + ω2 ) for the next iteration of the SCPH equation. The equa-
 tion for the nth step of the iteration is given as
n2 − n1
+ , (19)
iωm + σ(ω1 − ω2 ) [n] 2 1 X [n−1]
Vqij = ωqi δij + Fqq1 ,ijkℓ Kq1 ,kℓ , (21)
2
q1 ,k,ℓ
where we symbolically denote n(ωqi ) and ωqi as ni and
ωi , respectively. We will consider the contribution from where K is defined as
this diagram in a perturbative manner whereby the equa-
[n] [n] [n−1]
tion (18) is evaluated using the phonon frequencies and Kq,ij = αKq,ij + (1 − α)Kq,ij , (22)
polarization vectors obtained as a solution to the SCPH [n]
X [n]∗ [n] [n]∗ [n]
equation. It should be noted that there is another second- Kq,ij = Cq,ki Cq,kj hQqk Qqk i
order diagram that contains two four-phonon vertexes. k
[n] 
Although we do not consider that contribution for com-

X [n]∗ [n] ~ 1 + 2n(ωq1 k )
putational reasons, it is, in principle, possible to extend = Cq,ki Cq,kj [n]
. (23)
the theory to include higher-order diagrams [30]. k 2ωq1 k
5

Here we have used the fact that the mean square displace- the 12×12×12 Monkhorst-Pack k-point grid. We em-
ment of normal coordinate Qqj is given as hQ∗qj Qqj i = ployed the PBEsol exchange-correlation functional [37],
~ † ~ which was reported to work exceedingly well for predict-
2ωqj hAqj Aqj i = 2ωqj [1 + 2n(ωqj )]. In the classical limit
ing equilibrium volume and harmonic phonon frequency
(β → 0), the expectation value would be hQ∗qj Qqj i =
−2 of BaTiO3 and SrTiO3 [38]. The optimized lattice con-
kT ωqj . In addition, we introduced the mixing parame- stant is 3.896 Å, which agrees well with the experimental
ter α in Eq. (22) to improve convergence. value of 3.905 Å (Ref. 39, 293 K) and the previous DFT
[n] [n]
After we obtain Vq and Cq for all irreducible q result of 3.898 Å [38]. To consider the non-analytic part
points, we construct the new dynamical matrix as of the dynamical matrix, we calculated the Born effective
charges and the dielectric tensor of c-STO using DFPT.
Dq[n] = Eq[n] Wq[n] Eq[n]† Because the thermal expansion coefficient of c-STO is
= Eq Cq[n] Wq[n] Cq[n]† Eq† , (24) very small [39], we neglect thermal expansion effects in
this study.
[n] [n]
where Wqij = (ωqi )2 δij is the diagonal matrix. Using the
dynamical matrices, we construct dynamical matrices for
B. Estimation of force constants
the star of q using the unitary transformation:
[n]
DSq = Γq ({S|v(S)})Dq[n] Γ†q ({S|v(S)}). (25) To compute the harmonic phonon frequency, we ex-
tracted harmonic IFCs using the finite-displacement ap-
Here, Γq is the unitary matrix associated with the sym- proach [15]. The calculation was conducted with a
metry operation {S|v(S)} where S is the 3×3 rotation 2×2×2 cubic supercell containing 40 atoms as in Ref. 12.
matrix and v(S) is the translation vector. The detailed We displaced an atom from its equilibrium position by
expression for Γq can be found in Ref. 32. Finally, we 0.01 Å and calculated atomic forces for each displaced
construct the dynamical matrix in real space by taking configuration. We then extracted Φµν (ℓκ; ℓ′ κ′ ) by solv-
the inverse Fourier transformation ing the least-square problem
1 X [n] −iq·r(ℓ) Φ̃ = arg min kAΦ − F k22 , (27)
D [n] (r(ℓ)) = Dq e , (26) Φ
N q
as implemented in the alamode package [40]. Here, Φ =
[n] [n] [Φ1 , Φ2 , . . . , ΦM ]T is the parameter vector composed of
from which we obtain and ωqi for the dense N1 ×
Cq M linearly independent IFCs, F is the vector of atomic
N2 × N3 grid points, which are necessary for the next forces obtained by DFT calculations, and A is the matrix
iteration of the SCPH equation, by Fourier interpolation. composed of the atomic displacements.
For polar semiconductors, the non-analytic part of the To solve the SCPH equation and estimate the anhar-
dynamical matrix is accounted for using the mixed-space monic phonon frequencies of c-STO, one has to prepare
approach [33]. quartic IFCs. Cubic IFCs are also necessary to esti-
We iterate Eqs. (21)–(26) until convergence is achieved mate phonon linewidth and thermal conductivity, as will
for all phonon frequencies at the irreducible q points. be discussed in Sec. IV C. In principle, one can extend
We initialize the frequency and the unitary matrix as the finite-displacement approach to extract anharmonic
[0] [0]
ωqj = |ωqj | and Cq,ij = δij , respectively. Whenever terms, for which multiple atoms have to be displaced si-
we encounter an imaginary branch, we replace the fre- multaneously by an appropriately chosen displacement
quency with its absolute value. After the calculation has magnitude ∆u. However, finding an optimal value of ∆u
converged, the anharmonic frequencies and eigenvectors is not a trivial task, especially when imaginary modes
for a dense q grid, which are necessary for the subse- exist within the harmonic approximation, as in c-STO.
quent calculation of phonon lifetime and lattice thermal We found that the finite-displacement approach with
conductivity, can be obtained by Fourier interpolation. ∆u = 0.1 Å failed to yield reliable fourth-order IFCs that
could reproduce the double-well potential of the AFD
mode. To avoid this issue, one may alternatively em-
III. SIMULATION DETAILS ploy the AIMD simulation to sample the displacement-
force data set. This approach works particularly well for
A. DFT calculations simple systems such as Si and Mg2 Si [40]. However, it
should be noted that as long as one employs the typi-
Ab initio DFT calculations were performed using cal least-squares approach [Eq. (27)], an overfitting issue
the Vienna ab initio simulation package (vasp) [34], may arise unless the number of individual reference data
which employs the projector augmented wave (PAW) is fairly large compared with the number of parameters.
method [35, 36]. The adapted PAW potentials treat the Recently, Zhou et al. [25] proposed a more robust ap-
Sr 4s2 4p6 5s2 , Ti 3s2 3p6 3d2 4s2 , and O 2s2 2p4 shells as proach to estimate anharmonic IFCs. Noting that only
valence states. A cutoff energy of 550 eV was employed a small fraction of IFCs have a non-negligible contribu-
and the Brillouin zone integration was performed with tion to atomic forces, they employed the least absolute
6

shrinkage and selection operator (LASSO) technique. In 50


the LASSO technique, one solves the following equation: 40
n=3

30
Φ̃ = arg min kAΦ − F k22 + λkΦk1 , (28) 20
Φ
10
where the L1 penalty term is added to the least-squares 0
equation. Owing to the L1 penalty term, one can find 700
600 n=4
a sparse representation of the basis function, as demon- 500
strated by the cluster expansion method and the poten- 400
tial fitting [41, 42]. In this work, we followed the pro- 300
200

|Φ| (eV/Ån )
cedure of the previous study of Zhou et al. to solve the 100
LASSO equation. We initially conducted an AIMD sim- 0
1000
ulation at 500 K for 2000 steps with the time step of n=5
800
2 fs. From the trajectory of the AIMD simulation, we
600
then sampled 40 atomic configurations that were equally
400
spaced in time. For each configuration, we displaced all
200
of the atoms within the supercell by 0.1 Å in random
0
directions. The atomic forces for the configurations pre- 1400
pared in this manner were calculated using precise DFT 1200 n=6
1000
calculations, from which the matrix A and the vector F
800
in Eq. (28) were constructed. The LASSO equation was 600
solved using the split Bregman algorithm [41, 43], and 400
200
the optimal value of λ was selected from the four-fold
0
cross-validation score. To ensure that all of the terms 0 1 2 3 4 5 6
in the L1 term had the same dimension, we scaled the Distance (Å)
nth-order IFCs and atomic displacement by Φ → Φu0n−1
and u → u/u0 respectively, with u0 = 0.4 a0 (≈ 0.21 Å) FIG. 2. (color online). Absolute values of the third-, fourth-,
representing the order of the thermal nuclear motion. fifth-, and sixth-order anharmonic force constants estimated
by the LASSO technique plotted as a function of interatomic
distance. The onsite and two-body terms are indicated by
IV. RESULTS AND DISCUSSION circles and the three-body terms are indicated by triangles.

A. Anharmonic force constants in cubic SrTiO3 0.06 4


(a) (b)
3
0.05
DFT energy (eV/atom)

To find a sparse representation of the basis function 2


DFT force (eV/Å)

0.04
for c-STO, we first prepared a large parameter vector 1

Φ that included anharmonic terms up to the sixth or- 0.03 0

der. For harmonic and cubic terms, we included all pos- 0.02
-1

sible IFCs present in the 2×2×2 supercell. The quartic -2


0.01
terms were considered up to third-nearest neighbor shells, -3

whereas fifth- and sixth-order IFCs were considered for 0.00


0.00 0.01 0.02 0.03 0.04 0.05 0.06
-4
-4 -3 -2 -1 0 1 2 3 4
nearest-neighbor pairs. We determined a set of linearly LASSO energy (eV/atom) LASSO force (eV/Å)
independent parameters by considering the space group
symmetry and the constraints for the translational in- FIG. 3. (color online). Comparison of (a) potential energy
variance [15, 40]. We fixed the harmonic terms to the and (b) atomic forces sampled by an individual AIMD sim-
values determined by the finite-displacement approach ulation at 300 K. The dashed lines indicate cases where the
[Eq. (27)] and employed the LASSO technique for esti- results are identical.
mating the remaining anharmonic terms. The number
of linearly independent anharmonic parameters M was
1053, from which a sparse representation was found by IFCs decays rapidly with increasing interatomic distance,
Eq. (28). which indicates the locality of anharmonic interactions.
Figure 2 shows the magnitude of the anharmonic IFCs The terms with the largest magnitude occur at a distance
estimated by solving the LASSO equation. Here, the of 1.95 Å and represent force constants between a Ti atom
distance for the IFCs related to more than two atoms is and one of the surrounding O atoms. Among the onsite
defined as the distance of the most distant atomic pairs. quartic terms, Φµµµµ νννν
Ti,Ti,Ti,Ti (µ = x, y, z) and ΦO,O,O,O ,
The absence of onsite force constants for the third- and where ν is the direction parallel to the Ti-O bond, are
fifth-order IFCs is due to the inversion symmetry of c- most significant. Compared with these terms, the other
STO. As shown in Fig. 2, the magnitude of anharmonic onsite IFCs, including those of the Sr atom, are one order
7

800 of q1 points. The results for the lowest-energy soft modes


at Γ, R, and M points are summarized in table I. Our re-
sults indicate that at least 8×8×8 q1 points are needed to
600 obtain convergence and a less dense 2×2×2 q1 -point grid
Frequency (cm−1 )

significantly overestimates the Ωq values. This occurs


because the anharmonic phonon-phonon interaction is
400
limited only between the zone-center and zone-boundary
phonons by the 2×2×2 q1 grid. Thus, our numerical re-
200 sults indicate the importance of including mode coupling
between longer-wavelength phonons to obtain a reliable
description of the phonon softening in c-STO. The same
0
size-dependence should also be inherent in the real-space
approaches because the available phonon modes are lim-
Γ X M Γ R M ited by the size of the employed supercell.
We also considered the role of the off-diagonal elements
FIG. 4. (color online). Anharmonic phonon dispersion of c- of the phonon self-energy that cause PM. In Fig. 5, we
STO at 300 K calculated using the SCPH theory with 8×8×8 compare the anharmonic phonon frequencies of two zone-
q1 points (solid lines). The dotted lines show the harmonic center optical modes, labeled TO1 and TO2, obtained
phonon dispersion and the open symbols are experimental using the SCPH equation with and without PM. We
values at room temperature adapted from Refs. 44 and 45. have shown the SCPH results with 2×2×2 q1 points, as
these results will subsequently be compared with those
obtained using an MD-based approach. In the SCPH
of magnitude smaller. equation without PM, we neglect the off-diagonal ele-
The accuracy of the IFCs estimated by the LASSO ments of the phonon self-energy [Eq. (15)], which is ob-
equation was assessed by preparing independent test data tained by substituting the unitary matrix Cq in Eqs. (23)
using an AIMD simulation at 300 K for 2000 steps. and (24) with the identity matrix. Therefore, the polar-
We then calculated the potential energy [Eq. (1)] and ization vectors are fixed to the initial harmonic values.
atomic forces using the atomic displacements {u} and Fig. 5 demonstrates that PM is vital to describe the anti-
the IFCs {Φ}. In Fig. 3 we compare the potential en- crossing of the TO1 and TO2 phonon modes, both of
ergy U − U0 and the atomic forces obtained from DFT which belong to the same irreducible representation Γ15 .
and with those calculated from the IFCs estimated by In the case when we neglect PM, an artificial crossing oc-
LASSO. The model potential well reproduced the DFT curs around 500 K and the frequencies significantly devi-
results for various atomic configurations. The relative er- ate from those with PM. Therefore, we conclude that the
rors for the test data were 1.4 and 6.1% for the potential harmonic polarization vectors should not be employed to
energy and the atomic force, respectively, which are as predict anharmonic phonon properties of cubic SrTiO3
small as those reported in Ref. [25]. and other perovskite oxides having the same symmetry.
In Fig. 5 we compare results obtained with the SCPH
method with those obtained with the temperature-
B. SCPH solution dependent effective potential (TDEP) method [21]. In
the TDEP method, atomic displacements and forces are
Using the harmonic and quartic force constants ob- sampled by AIMD simulations at a target temperature
tained from the finite-displacement and the LASSO tech- and are then used to extract effective harmonic force
niques, respectively, the SCPH equation [Eqs. (21)–(26)]
was solved numerically. Since we employed the 2×2×2
supercell in this study, the q point in Eq. (21) was lim- TABLE I. Anharmonic phonon frequency (cm−1 ) of the soft
ited to the irreducible points on the 2×2×2 grid. We modes at 300 K calculated using the SCPH equation with
changed the q1 grid to investigate the convergence of various q1 -grid densities. The harmonic phonon frequency is
the anharmonic phonon frequencies. The mixing param- also shown for comparison.
eter of α = 0.1 was employed for all temperatures ex- q1 points Γ15 (FE) R25 (AFD) M3
cept those near the critical temperature of the structural
2×2×2 144 69 103
phase transition, where a much smaller α was required.
Figure 4 shows the anharmonic phonon dispersion of 4×4×4 138 46 89
c-STO at 300 K obtained as the solution for the SCPH 6×6×6 136 39 86
equation. The phonon frequencies are increased by the 8×8×8 136 37 85
quartic anharmonicity, evident in the low-energy soft 10×10×10 135 36 85
modes at the Γ (0, 0, 0), R ( 12 , 21 , 21 ), and M ( 12 , 12 , 0) 12×12×12 135 35 85
points. We investigated the convergence of the anhar- Frozen phonon 58i 76i 21
monic phonon frequency Ωq with respect to the number
8

250 5

200 4

6×6×6
Frequency (cm−1)

150 3 2×2×2 4×4×4

Ω2q (THz2)
100
2

50
SCPH w/ PM (TO1) TDEP (TO1) 1
SCPH w/ PM (TO2) TDEP (TO2)
SCPH w/o PM (TO1) Servoin et al. (IR)
12 × 12 × 12
SCPH w/o PM (TO2) Yamada and Shirane (INS)
0 8×8×8
0 200 400 600 800 1000 0
0 100 200 300 400 500
Temperature (K)
Temperature (K)

FIG. 5. (color online). Temperature-dependence of the an-


FIG. 6. (color online). Temperature-dependence of the
harmonic phonon frequencies of two Γ15 modes calculated us-
squared phonon frequency of the R25 mode obtained from
ing the SCPH equation with and without PM, and with the
the SCPH theory with various q1 -point densities. The open
TDEP approach (see the text for details). The TO1 mode
circles are experimental values adapted from Ref. 45.
corresponds to the FE mode. The open symbols are experi-
mental values for the TO1 mode reported by Servoin et al. [46]
and Yamada and Shirane [27].
TO1 mode in Fig. 5).
Figure 6 shows a comparison of the temperature-
dependence of the squared frequency of the AFD mode
constants by numerical fitting. In our TDEP simula- and experimental measurements [45]. As can be seen in
tions, we performed MD simulations using the Taylor the figure, the frequency of the AFD mode is severely
expansion potential [Eq. (1)] instead of AIMD to re- size-dependent. For the 2×2×2 q1 grid, we do not ob-
duce computational costs. Anharmonic terms up to the serve a freezing-out of the AFD mode even at absolute
sixth order were considered and the force constants esti- zero. When we increase the q1 -grid density and allow
mated by the LASSO technique were employed. We con- interactions with longer-wavelength phonons, we observe
ducted the constant-temperature MD simulations with the precursor of the freezing-out of the AFD mode at
the 2×2×2 supercell and the temperature was controlled temperatures near 200 K. Although the soft-mode fre-
by the Berendsen thermostat [47]. We employed a time quency does not reach zero in the current simulation with
step of 1 fs and conducted the MD simulations for 50000 finite q1 points, we expect that this would occur in the
steps at each temperature. The last 40000 steps were thermodynamic limit (N → ∞) [48]. As we decrease
employed to extract effective harmonic IFCs by least- the temperature to the transition temperature, the num-
squares fitting [Eq. (27)]. Although the anharmonic fre- ber of q1 points necessary for the convergence increases,
quencies obtained using the TDEP approach are slightly which indicates an increase in the correlation length of
smaller than the SCPH results, they agree qualitatively the fluctuations near the critical point. Above approxi-
with the SCPH results, as shown in Fig. 5. We con- mately 300 K, the temperature-dependence can be reli-
sider this discrepancy to be reasonable for the following ably fitted by the equation Ω2q (T ) = a(T −Tc )2 . Applying
two reasons. First, the SCPH results include only anhar- this equation to the result obtained using the 12×12×12
monic self-energies that can be generated from Fig. 1(a), q1 grid, we obtain the Tc of the cubic-to-tetragonal phase
whereas the TDEP includes higher-order anharmonic ef- transition as 220 K.
fects. Among these higher-order terms, the first-order For comparison, we have plotted experimental results
contribution due to the cubic anharmonicity, as depicted in Figs. 4, 5, and 6 using open symbols. The SCPH equa-
in Fig. 1(b), should have the largest contribution. We tion reproduces the temperature-dependence of the soft
found that the effect of the diagram in Fig. 1(b) is to modes qualitatively, but not quantitatively, i.e., the fre-
reduce the anharmonic frequency for the FE mode. Sec- quencies of the FE and AFD modes are overestimated
ond, since the quantum effect of nuclear motion is not and underestimated, respectively. Because the ADF fre-
considered in the MD simulation,
the thermal average quency is underestimated, the transition temperature
of the squared normal coordinate Q∗q Qq is underesti-

predicted is twice as large as the experimental value
mated for temperatures below the Debye temperature in of 105 K. We consider this deviation to be acceptable
the TDEP approach. Therefore, the renormalization of because phonon-related properties of ferroelectric mate-
anharmonic effects is underestimated in the TDEP ap- rials are known to be sensitive to the lattice constant
proach, which explains why the deviation from the SCPH and exchange-correlation functional employed [38, 49].
result becomes larger with decreasing temperature (see In this study, we employed the PBEsol functional to
9

avoid problems inherent to the local-density approxi- 102

mation (LDA) and the generalized-gradient approxima- 14

tion with the Perdew-Burke-Ernzerhof parameterization 101


(PBE) [50]; LDA tends to underestimate the equilibrium

τq (ps)
12
volume, whereas PBE tends to overestimate it. How- 100

κL (W/mK)
ever, our numerical results suggest that PBEsol cannot
10
give a quantitative description of c-STO. This issue is ex-
10−1
pected to be resolved, at least partially, by employing a 0 100 200 300 400 500 600 700 800
ωq (cm−1 )
hybrid functional. Wahl et al. [38] investigated the func- 8
tional dependence of the harmonic frequency in the FE
mode of c-STO and reported the results of 29i and 74i for 6
the PBEsol semilocal and the Heyd-Scuseria-Ernzerhof This work (12×12×12)
Expt. (Muta et al.)
(HSE) hybrid functionals [51], respectively. Since the Expt. (Popuri et al.)
2 2
harmonic frequency changes as ωHSE < ωPBEsol < 0, we 4
200 300 400 500 600 700 800 900 1000
expect that the Fock exchange can increase the depth Temperature (K)
of the double-well potential, thereby decreasing the an-
harmonic frequency of the FE mode. Wahl et al. also FIG. 7. (color online). Temperature-dependence of the lat-
reported that the energy gain for the AFD phase was tice thermal conductivity of c-STO. The computational result
smaller in HSE than in the PBEsol functional. This in- is compared with experimental values reported by Muta et
dicates that the depth of the double-well potential for al. [52] and Popuri et al. [53]. Lines are shown to guide the
the AFD mode can be decreased, and the anharmonic fre- eye. Inset: Calculated phonon lifetime of c-STO at 300 K.
quency can be increased by using HSE instead of PBEsol.
Therefore, we believe that the quantitative accuracy of
the present SCPH results could be improved by employ- linewidth Γq (ω) can be obtained from the imaginary part
ing a hybrid functional, which will be the topic of future of the phonon self-energy that results from the cubic an-
work. harmonicity [Eq. (18)], which is given explicitly as
In the present SCPH calculations we have not con-
π X ~|Φ(−q, q ′ , q ′′ )|2
sidered effects related to the cubic anharmonicity, such Γq (ω) =
as thermal expansion, relaxation of internal coordinates 2N ′ ′′ 8Ωq Ωq′ Ωq′′
q ,q
and intrinsic frequency shifts due to the bubble diagram. × [(nq′ + nq′′ + 1)δ(ω − Ωq′ − Ωq′′ )
However, these effects can, in general, become impor-
tant in severely anharmonic systems [48], and should −2(nq′ − nq′′ )δ(ω − Ωq′ + Ωq′′ )] . (30)
be considered, especially when one intends to quanti-
Here, the matrix element Φ(q, q ′ , q ′′ ) is calculated from
tatively compare theoretical results with experimental
cubic IFCs using Eq. (14) with eigenvectors {eµ (κ; q)} re-
data. Therefore, extending the present ab initio method
placed by the solution to the SCPH equation {ǫµ (κ; q)}.
to include these effects, either perturbatively or self-
The equations (29) and (30) are identical to those that
consistently, could be another important direction for
have commonly been employed in the thermal conductiv-
further research and development.
ity calculations except that harmonic phonon frequencies
and eigenvectors are substituted by anharmonic frequen-
cies and eigenvectors, respectively, obtained using the
C. Lattice thermal conductivity
SCPH equation.
Figure 7 compares the calculated thermal conductivity
The lattice thermal conductivity plays a major role in of c-STO with experimental results [52, 53]. The calcu-
optimizing the thermoelectric figure-of-merit ZT , and it lation was conducted using the 8×8×8 q1 grid for the
has been the subject of intense theoretical study in recent SCPH equation and the 12×12×12 q grid for the BTE-
years. To show the validity of our theoretical approach RTA equation [Eq. (29)]. Although we observed devi-
based on the SCPH equation, we estimated the lattice ations in soft-mode frequencies, the calculated thermal
thermal conductivity of c-STO. For this work, we em- conductivity agrees well with the experimental results,
ploy the Boltzmann transport equation (BTE) within the as can be seen in Fig. 7. We expect that the agree-
relaxation time approximation (RTA), where the lattice ment could be improved further by employing a finer q
thermal conductivity is given as grid and using a hybrid functional, although such cal-
culations were not performed because of computational
1 X
κµν
L (T ) = Cq (T )vqµ (T )vqν (T )τq (T ). (29) limitations. In the Fig. 7 inset, we also show the phonon
VN q lifetime τq at 300 K calculated by Eq. (30). The phonon
lifetimes of c-STO obtained from the perturbation the-
Here, V is the unit-cell volume, Cq is the lattice spe- ory [Eq. (30)] are found to be even smaller than those
cific heat, vq = dΩq /dq is the group velocity, and of PbTe [16], but the κL value of c-STO is higher be-
τq = [2Γq (Ωq )]−1 is the lifetime of phonon q. The phonon cause of the larger group velocities. The lifetime shows
10

a characteristic feature in the low-frequency region (< to-tetragonal transition temperature as Tc = 220 K that
100 cm−1 ): the phonon modes split into two separate was twice as large as the experimental value of 105 K.
regions in τq > 3 ps and τq ∼ 0.6 ps, where the former Although further theoretical investigations are required
corresponds to the acoustic modes that follow the fre- to understand the origin of this discrepancy, we expect
quency dependence of τ ∼ ω 2 , which has been observed that the quantitative agreement can be improved by em-
in other materials [16, 40], and the latter corresponds to ploying a hybrid functional. We also calculated the lat-
the phonon modes around the R point, which indicates tice thermal conductivity κL of cubic SrTiO3 using the
the severe anharmonicity of the AFD mode. Boltzmann transport equation within the relaxation-time
approximation. The calculated κL values reproduced ex-
perimental results especially in the high temperature re-
V. CONCLUSIONS gion. The underestimation of κL in the low temperature
region may be attributed to the overestimation (under-
We developed an ab initio method to compute anhar- estimation) of the ferroelectric (antiferrodistortive) soft
monic phonon frequencies and lifetimes that can be ap- mode, which will be addressed in a future work.
plied to severely anharmonic systems. The method em- The present method, which combines the SCPH theory
ploys anharmonic force constants up to the fourth order, with perturbation approach based on anharmonic force
which are extracted from DFT calculations using a com- constants, enables us to obtain the anharmonic phonon
pressive sensing approach. The frequency renormaliza- frequencies and lifetimes at various temperatures effi-
tion associated with the quartic anharmonicity is treated ciently just by changing the occupation number. The
non-perturbatively using the SCPH theory. By perform- system size dependency can be investigated using the re-
ing the perturbation calculation after the SCPH solution, ciprocal space formalism. Therefore, we believe that the
we also calculated phonon lifetimes that result from the present method paves the way for understanding lattice
three-phonon scattering processes. anharmonicity and related dynamical and thermodynam-
We applied the method to the high-temperature phase ical properties of thermoelectric, ferroelectric, and super-
of perovskite SrTiO3 . Unlike the harmonic phonon dis- conducting materials.
persion, the SCPH solution was free from the imaginary
branches in the entire Brillouin zone. We found that
including polarization mixing is important to correctly
account for the temperature dependence of the phonon VI. ACKNOWLEDGEMENTS
frequency of the ferroelectric soft mode of perovskite ox-
ides. In addition, we examined the size-dependence of the We wish to thank Takashi Miyake, Mitsuaki Kawa-
anharmonic frequencies of the soft modes and found that mura, and Takuma Shiga for fruitful discussions, and
long-wavelength phonons significantly reduced the anhar- Masato Okada for helpful suggestions regarding the com-
monic frequencies, especially for the antiferrodistortive pressive sensing. This study is partially supported by
mode near the transition temperature. The temperature- Tokodai Institute for Element Strategy (TIES) and also
dependence of the soft mode frequencies calculated using by Thermal Management Materials and Technology Re-
the SCPH theory agreed qualitatively well with the ex- search Association (TherMAT). The computation in this
perimental results. However, the quantitative accuracy work has been done using the facilities of the Supercom-
of the present calculations based on the PBEsol func- puter Center, Institute for Solid State Physics, The Uni-
tional was unsatisfactory, where we obtained the cubic- versity of Tokyo.

[1] D. C. Wallace, Thermodynamics of Crystals (Dover Pub- [8] Z. Hiroi, J. ichi Yamaura, and K. Hattori, J. Phys. Soc.
lications, INC., 1972). Jpn. 81, 011012 (2012).
[2] G. J. Snyder and E. S. Toberer, Nature Mater. 7, 105 [9] R. Mankowsky, A. Subedi, M. Först, S. O. Mariager,
(2008). M. Chollet, H. T. Lemke, J. S. Robinson, J. M. Glow-
[3] A. Ward, D. A. Broido, D. A. Stewart, and G. Deinzer, nia, M. P. Minitti, A. Frano, M. Fechner, N. A. Spaldin,
Phys. Rev. B 80, 125203 (2009). T. Loew, B. Keimer, A. Georges, and A. Cavalleri, Na-
[4] L. Lindsay, D. A. Broido, and N. Mingo, Phys. Rev. B ture 516, 71 (2014).
82, 115427 (2010). [10] I. Errea, M. Calandra, C. J. Pickard, J. Nelson, R. J.
[5] O. Delaire, J. Ma, K. Marty, A. F. May, M. A. McGuire, Needs, Y. Li, H. Liu, Y. Zhang, Y. Ma, and F. Mauri,
M. H. Du, D. J. Singh, A. Podlesnyak, G. Ehlers, M. D. Phys. Rev. Lett. 114, 157004 (2015).
Lumsden, and B. C. Sales, Nature Mater. 10, 614 (2011). [11] W. Luo, B. Johansson, O. Eriksson, S. Arapan, P. Sou-
[6] L.-D. Zhao, S.-H. Lo, Y. Zhang, H. Sun, G. Tan, C. Uher, vatzis, M. I. Katsnelson, and R. Ahuja, Proc. Natl. Acad.
C. Wolverton, V. P. Dravid, and M. G. Kanatzidis, Na- Sci. 107, 9962 (2010).
ture 508, 373 (2014). [12] T. Sun, D.-B. Zhang, and R. M. Wentzcovitch, Phys.
[7] T. Takabatake, K. Suekuni, T. Nakayama, and Rev. B 89, 094109 (2014).
E. Kaneshita, Rev. Mod. Phys. 86, 669 (2014).
11

[13] A. A. Maradudin and A. E. Fein, Phys. Rev. 128, 2589 Matter 22, 202201 (2010).
(1962). [34] G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169
[14] S. Baroni, S. de Gironcoli, A. Dal Corso, and P. Gian- (1996).
nozzi, Rev. Mod. Phys. 73, 515 (2001). [35] P. E. Blöchl, Phys. Rev. B 50, 17953 (1994).
[15] K. Esfarjani and H. T. Stokes, Phys. Rev. B 77, 144112 [36] G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).
(2008). [37] J. P. Perdew, A. Ruzsinszky, G. I. Csonka, O. A. Vydrov,
[16] Z. Tian, J. Garg, K. Esfarjani, T. Shiga, J. Shiomi, and G. E. Scuseria, L. A. Constantin, X. Zhou, and K. Burke,
G. Chen, Phys. Rev. B 85, 184303 (2012). Phys. Rev. Lett. 100, 136406 (2008).
[17] A. Togo, L. Chaput, and I. Tanaka, Phys. Rev. B 91, [38] R. Wahl, D. Vogtenhuber, and G. Kresse, Phys. Rev. B
094306 (2015). 78, 104116 (2008).
[18] T. Tadano, Y. Gohda, and S. Tsuneyuki, Phys. Rev. [39] A. Okazaki and M. Kawaminami, Mater. Res. Bull. 8,
Lett. 114, 095501 (2015). 545 (1973).
[19] G. Lang, K. Karch, M. Schmitt, P. Pavone, A. P. Mayer, [40] T. Tadano, Y. Gohda, and S. Tsuneyuki, J. Phys: Con-
R. K. Wehner, and D. Strauch, Phys. Rev. B 59, 6182 dens. Matter 26, 225402 (2014).
(1999). [41] L. J. Nelson, G. L. W. Hart, F. Zhou, and V. Ozoliņš,
[20] N. Bonini, M. Lazzeri, N. Marzari, and F. Mauri, Phys. Phys. Rev. B 87, 035125 (2013).
Rev. Lett. 99, 176802 (2007). [42] A. Seko, A. Takahashi, and I. Tanaka, Phys. Rev. B 90,
[21] O. Hellman, P. Steneteg, I. A. Abrikosov, and S. I. 024101 (2014).
Simak, Phys. Rev. B 87, 104111 (2013). [43] T. Goldstein and S. Osher, SIAM J. Imaging Sci. 2, 323
[22] N. R. Werthamer, Phys. Rev. B 1, 572 (1970). (2009).
[23] P. Souvatzis, O. Eriksson, M. I. Katsnelson, and S. P. [44] W. G. Stirling, J. Phys. C 5, 2711 (1972).
Rudin, Phys. Rev. Lett. 100, 095901 (2008). [45] R. A. Cowley, W. J. L. Buyers, and G. Dolling, Solid
[24] I. Errea, M. Calandra, and F. Mauri, Phys. Rev. B 89, State Commun. 7, 181 (1969).
064302 (2014). [46] J. L. Servoin, Y. Luspin, and F. Gervais, Phys. Rev. B
[25] F. Zhou, W. Nielson, Y. Xia, and V. Ozoliņš, Phys. Rev. 22, 5501 (1980).
Lett. 113, 185501 (2014). [47] H. J. C. Berendsen, J. P. M. Postma, W. F. van Gun-
[26] G. Shirane and Y. Yamada, Phys. Rev. 177, 858 (1969). steren, A. DiNola, and J. R. Haak, J. Chem. Phys. 81,
[27] Y. Yamada and G. Shirane, J. Phys. Soc. Jpn. 26, 396 3684 (1984).
(1969). [48] E. R. Cowley, Physica A 232, 585 (1996).
[28] R. A. Cowley, Phys. Rev. 134, A981 (1964). [49] W. Zhong, D. Vanderbilt, and K. M. Rabe, Phys. Rev.
[29] I. Errea, B. Rousseau, and A. Bergara, Phys. Rev. Lett. B 52, 6301 (1995).
106, 165501 (2011). [50] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev.
[30] R. S. Tripathi and K. N. Pathak, Il Nuovo Cimento B Lett. 77, 3865 (1996).
21, 289 (1974). [51] J. Heyd, G. E. Scuseria, and M. Ernzerhof, J. Chem.
[31] M. R. Hermes and S. Hirata, The Journal of Physical Phys. 118, 8207 (2003).
Chemistry A 117, 7179 (2013). [52] H. Muta, K. Kurosaki, and S. Yamanaka, J. Alloys
[32] A. A. Maradudin and S. H. Vosko, Rev. Mod. Phys. 40, Compd. 392, 306 (2005).
1 (1968). [53] S. R. Popuri, A. J. M. Scott, R. A. Downie, M. A. Hall,
[33] Y. Wang, J. J. Wang, W. Y. Wang, Z. G. Mei, S. L. E. Suard, R. Decourt, M. Pollet, and J.-W. G. Bos, RSC
Shang, L. Q. Chen, and Z. K. Liu, J. Phys: Condens. Adv. 4, 33720 (2014).

View publication stats

You might also like