Sie sind auf Seite 1von 6

Scripta Materialia 129 (2017) 88–93

Contents lists available at ScienceDirect

Scripta Materialia
journal homepage: www.elsevier.com/locate/scriptamat

Regular Article

High throughput combinatorial method for fast and robust prediction of


lattice thermal conductivity
Pinku Nath a , Jose J. Plata a , Demet Usanmaz a , Cormac Toher a , Marco Fornari b , Marco Buongiorno Nardelli c ,
Stefano Curtarolo d, *
a
Dept. of Mechanical Engineering and Materials Science, Duke University, Durham, NC 27708, USA
b
Dept. of Physics and Science of Advanced Materials Program, Central Michigan University, Mount Pleasant, MI 48858, USA
c
Dept. of Physics and Dept. of Chemistry, University of North Texas, Denton, TX, USA
d
Materials Science, Electrical Engineering, Physics and Chemistry, Duke University, Durham, NC 27708, USA

A R T I C L E I N F O A B S T R A C T

Article history: The lack of computationally inexpensive and accurate ab-initio based methodologies to predict lattice
Received 18 July 2016 thermal conductivity, without computing the anharmonic force constants or time-consuming ab-initio
Received in revised form 7 September 2016 molecular dynamics, is one of the obstacles preventing the accelerated discovery of new high or low thermal
Accepted 26 September 2016 conductivity materials. The Slack equation is the best alternative to other more expensive methodologies
Available online 27 October 2016
but is highly dependent on two variables: the acoustic Debye temperature, ha , and the Grüneisen parame-
ter, c. Furthermore, different definitions can be used for these two quantities depending on the model or
Keywords:
approximation. In this article, we present a combinatorial approach to elucidate which definitions of both
High-throughput
variables produce the best predictions of the lattice thermal conductivity, j l . A set of 42 compounds was
Accelerated materials development
Quasi-harmonic approximation used to test the accuracy and robustness of all possible combinations. This approach is ideal for obtaining
Lattice thermal conductivity more accurate values than fast screening models based on the Debye model, while being significantly less
expensive than methodologies that solve the Boltzmann transport equation.
© 2016 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

1. Introduction 50% [13,14]. Furthermore, the extension of the Green-Kubo formal-


ism to multiscale models has been shown to be nontrivial [15].
Lattice thermal conductivity, j l , plays an important role in Characterization of the anharmonic forces constants and its use in
multiple applications and technologies [1]. Thermoelectric materi- the solution of the Boltzmann transport equation, BTE, is extremely
als [2], heat sink materials [3], rewritable density scanning-probe expensive which makes it unfeasible in a HT approach. Thus, for
phase-change memories [4] or thermal medical therapies [5] are an effective prediction of j l with HT methods, there is an advan-
some examples in which thermal transport is the technological tage in approximated methods based on ab-initio characterization.
enabling property. During the last three decades, several theoret- i. Ab-initio is fully self-consistent and does not need experimental
ical models and methodologies have been developed to calculate data or the use/generation of force fields; ii. There exists HT frame-
j l [6–10]. It is a trade-off: while the quickest approaches can works, such as AFLOW [16–18], that can monitor, manage, correct,
only predict trends, accurate methods are computationally expen- and post-process the information obtained from different quantum
sive and can not be implemented in high throughput (HT) frame- mechanical codes.
works, hindering the discovery of new materials with better per- There are three main families of approximated models based on
formance [1,11,12]. For instance, semi-empirical models predict first principles. They can be classified depending on performance,
thermal properties at reduced computational cost but require some accuracy and robustness.
experimental data [6–8]. Additionally, classical molecular dynamics
combined with Green-Kubo equations also produces reliable results.
Although this method includes high-order scattering processes, the • The “GIBBS” quasiharmonic Debye model is the least compu-
use of semi-empirical potentials leads to errors on the order of tationally expensive approach to identify trends and simple
descriptors for thermal properties, such as the Grüneisen param-
eters, c, and the Debye temperature, hD [19]. The Automatic-
* Corresponding author. GIBBS-Library (AGL) framework combines this model with the
E-mail address: stefano@duke.edu (S. Curtarolo). Slack equation [20] based on work with noble gas crystals by

http://dx.doi.org/10.1016/j.scriptamat.2016.09.034
1359-6462/© 2016 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
P. Nath et al. / Scripta Materialia 129 (2017) 88–93 89

Leibfried and Schlömann [21] and Julian [22] to predict j l [23]. It


reproduces correctly the ordinal ranking of the thermal conduc-
tivity for several different classes of semiconductor materials
using only energy-volume curves, but suffers in comparing
families with different structures.
• More accurate Grüneisen parameters can be obtained using
the quasiharmonic approximation, QHA, and then used in the
Slack equation [24]. However, harmonic force constants have
to be calculated to build the dynamical matrix that describe
the vibrational modes of the system. An alternative method
proposed by Madsen et al. [24] consists of the use of lat-
tice dynamics calculations to compute approximate relaxation
scattering times at hD . Both methods give j l in reasonable
quantitative agreement with experiments.
• Third order interatomic force constants (3rd IFCs) are required Fig. 1. Combinatorial diagram for lattice thermal conductivity.
to calculate the phonon scattering times included in the solu-
tion of the Boltzmann transport equation, BTE [9,10,25]. This
is the most computationally expensive method but it is the
most used for highly accurate results. Once scattering pro- average atomic mass, Mav . The thermal conductivity at any tempera-
cesses are computed using 3rd IFCs, different schemes have ture is estimated by [20,24,28]:
been proposed to solve the BTE. The relaxation time approxi-
mation, RTA, is the simplest and predicts values on average 10%
smaller than experimental quantities. The full solution requires ha
jl (T) = jl (ha ) . (2)
a self-consistent iteration, but produces values very close to T
experiment while adding only a small computational cost com-
pared to RTA solutions [26]. The bottle neck of both methods
Eqs. (1) and (2) show that only the acoustic Debye temperature
comes from on the computation of the 3rd order IFCs. Recently,
and the Grüneisen parameter are needed to calculate j l at any
some authors have proposed the computation of these forces
temperature. Various groups have used different models to pre-
using compressive sensing. However, the cost to obtain reliable
dict this quantity using the Slack equation. It seems that Slack
results is still high [27].
equation combined with the Debye model tends to underestimate
j l [23], while when combined with lattice dynamics calculations, j l
Despite the different approaches developed in the last few is overestimated [24]. We propose the combination of both mod-
decades, there is a lack of inexpensive, accurate, and robust methods els to offset the errors and obtain values closer to the experimental
that can be used routinely. In this article, we use different definitions results.
of ha and c based on the Debye model and the QHA to obtain values Our implementation is based on a combinatorial approach where
for j l using the Slack equation. Using a phenomenological approach, the lattice thermal conductivity is a function of two variables: acous-
we compare the results obtained with our combinatorial schema to tic Debye temperature, ha , and Grüneisen parameter, c: j l (ha , c). We
available experimental data to decide which description of these two use different formulations based on the Debye model and quasi-
variables best predict values for j l with qualitative and quantitative harmonic approximation to compute these two variables and then
accuracy in a high-throughput approach. combine them to obtain different values for j l (hxa , cy ) (see Fig. 1).
We use a set of 42 materials, all belonging to different space groups
and presenting a range of four orders of magnitude for j l , as a test
2. Methodology to determine the best combination of variables from a quantitative
point of view. This approach maximizes the flexibility of the method,
2.1. Lattice thermal conductivity optimizing the results without extra cost beyond the harmonic force
constants calculations.
The lattice thermal conductivity1 is computed using the Slack The first definition for c is extracted from Ref. [24]:
equation (Eq. 1) because of its simplicity:

√  y <k ha
0.849 × 3 4
3 
iq
B
jl (ha ) =  × (1)  2
ciq Ciq
 
20p3 1 − 0.514c−1 + 0.228c−2 (1)  q i
ca 2
= c̄a =  , (3)
  1 
yiq

<kB ha
kB ha 2 kB Mav V 3
× . Ciq
 c2 q i

This equation predicts j l at the acoustic Debye temperature (ha ) where ciq is defined as:
using the Grüneisen parameter, c, primitive cell volume, V, and the

0K
Veq ∂ Dq
ciq = − e∗iq e . (4)
2y2qj i
∂ V iq
1
List of symbols: ha acoustic Debye temp.; c Grüneisen parameter; j l lattice ther-
mal cond.; hD Debye temp.; V prim. cell volume; Mav average atomic mass; yiq
frequency mode i at q-point q; kB Boltzmann constant; Ciq specific heat; V0Keq , prim. cell
volume at 0 K; eiq eigenvector; Dq dynamical matrix; n number atoms in prim. cell;
Since the acoustic bands provide the majority of the contribution to
yD Debye freq.; s Poisson ratio; F(V) free energy; Feq eq. free energy; Veq eq. volume; j l , the sum is performed over the modes, i, and q-points, q, having
B bulk modulus; Bp derivative of B over P. an energy less than kB ha . We propose a second definition where the
90 P. Nath et al. / Scripta Materialia 129 (2017) 88–93

(a) (e) (i)


1000
a γ ) (W/(mK))

κl(θ(1) (1)
a γa ) κl(θ(1) (2)
a γa ) κl(θ(1) (3)
a γ )

100
(y)
κl(θ(x)

10

(b) (f) (j)


1000
a γ ) (W/(mK))

κl(θ(2) (1)
a γa ) κl(θ(2) (2)
a γa ) κl(θ(2) (3)
a γ )
100
(y)
κl(θ(x)

10
M
κl
κAGL
l
1

(c) (g) (k)


1000
a γ ) (W/(mK))

(3) (1) (3) (2) (3) (3)


κl(θa γa ) κl(θa γa ) κl(θa γ )
100
(y)
κl(θ(x)

10

(d) (h) (l)


1000
a γ ) (W/(mK))

(4) (1) (4) (2) (4) (3)


κl(θa γa ) κl(θa γa ) κl(θa γ )
100
(y)
κl(θ(x)

10

1
1 10 100 1000 1 10 100 1000 1 10 100 1000
exp exp exp
κl (W/(mK)) κl (W/(mK)) κl (W/(mK))

Fig. 2. Lattice thermal conductivity, j l (ha , c), for the proposed models. Dashed black lines represent ±40% error.
P. Nath et al. / Scripta Materialia 129 (2017) 88–93 91

sum is done only for the acoustic modes instead of over the modes modulus, B, is obtained by fitting the free energy, F(V, T), obtained
having an energy less than kB ha : using the QHA to the Birch-Murnaghan (BM) function:

 



BVeq (Veq /V )Bp Veq B
 3 F(V) = Feq + +1 − , (13)
  c2 C Bp Bp − 1 Bp − 1
 q i=1 a,iq iq
ca
(2) 2 
= c̄a =  . (5)

3
Ciq where, equilibrium free energy, Feq , bulk modulus, B, equilibrium
q i=1
volume, Veq and the derivative of the bulk modulus with respect to
pressure, Bp are used as the fitting parameters.
Recently, Madsen et al. also used the averaged squared Grüneisen
parameters obtaining similar results with their model [29]: 2.2. Computational details

2.2.1. Geometry optimization



  c2 C All structures are fully relaxed using the automated framework,
 q i iq iq
c(3) = 
c̄2 =

. (6) AFLOW [16–18], and the DFT Vienna ab-initio simulation package,
Ciq
q i
VASP [32]. Optimizations are performed following the AFLOW stan-
dards [18]. We use the projector augmented wave (PAW) pseu-
dopotentials [33] and the exchange and correlation functionals
It is important to note that c(x) depends on ha since Ciq in Eqs. 3, 5, parametrized by the generalized gradient approximation proposed
and 6 is evaluated at T = ha . by Perdew-Burke-Ernzerhof (PBE) [34]. All calculations use a high
Phonon calculations can be also used to predict ha : energy-cutoff, which is 40% larger than the maximum recommended
cutoff among all component potentials, and a k-points mesh of 8000
∞ k-points per reciprocal atom. Primitive cells are fully relaxed (lattice
(1) 1/3 1/3 5 0 y2 g(y)dy parameters and ionic positions) until the energy difference between
ha =n hD = n . (7)
3kB 0∞ g(y)dy two consecutive ionic steps is smaller than 10 −4 eV and forces in
each atom are below 10 −3 eV/Å.
where n is the number of atoms per unit cell and g(y) is the phonon
(2)
density of states. Similar to ca , we get ha using only the acoustic 2.2.2. Phonon calculations
branches: Phonon calculations were carried out using the automatic phonon
library, APL, as implemented in the AFLOW package, using VASP
∞ to obtain the IFCs via the finite-displacement approach [35]. The
(2) 5 0 y2a g (ya ) dya magnitude of this displacement is 0.015 Å. Electronic self consis-
ha = . (8)
3kB 0∞ g (ya ) dya tent field (SCF) iterations for single point calculations were stopped
when the energy difference between last two step was less than
10 −8 eV. This threshold ensures a good convergence for the wave-
The classic definition for the Debye frequency, the maximum fre- function and accurate enough values for forces and harmonic force
quency in the vibrational spectra, can be also applied to obtain the constants. Non-analytical contributions to the dynamical matrix are
Debye temperature using the phonon density of states, pDOS: also included using the formulation developed by Wang et al. [36].
Frequencies and other related phonon properties are calculated on
 yD a 21 × 21 × 21 q-mesh in the Brillouin zone, which is sufficient to
g(y)dy = 3n, (9) converge the vibrational density of states, pDOS, and hence the val-
0
ues of thermodynamic properties calculated through it. The pDOS is
calculated using the linear interpolation tetrahedron method avail-
(3) yD able in AFLOW package. The derivative of dynamical matrix is
ha = n1/3 hD = n1/3 . (10)
kB obtained using the central difference method within a volume range
of ±0.03%.
The Debye model also predicts hD using the bulk modulus,
3. Results
B [19,30,31]:
A data set of 42 compounds has been used to validate our

(4)   2 1/2 1/3 B approach. The list of materials includes semiconductors and insula-
ha = n1/3 hD = n1/3 6p V n f (s) . (11) tors that belong to different structural prototypes such as diamond,
kB M
zinc blende, rock salt and fluorite. To maximize the heterogeneity of
the data set, materials have been selected containing as many differ-
Here, M is the mass of the unit cell, and f(s) is given by ent elements as possible from the s-, p-, and d-blocks of the periodic
table. Our results compare preferably to experimental measure-
⎧  ⎫1 ments of defect-free single-crystals. Scattering due to boundaries
⎨     −1 ⎬ 3
2 1 + s 3/2 1 1 + s 3/2 and defects, which are common in the samples of more complex
f (s) = 3 2 • + • , (12)
⎩ 3 1 − 2s 3 1−s ⎭ structures, cannot be included using this methodology. These effects
can dominate the lattice thermal conductivity experimental val-
ues so they are not the best candidates for the validation of our
within the assumption that the Poisson ratio, s, remains constant. combinatorial approach. The model proposed by Slack and Gal-
For the calculations described in this article, this value is set at 0.25, ginaitis [37] is more suitable for a direct comparison with these
which is the theoretical value for a Cauchy solid [19,31]. The Poisson complex structures as has been demonstrated by Madsen et al. [29]
ratio s for crystalline materials is typically in the range 0.2–0.3. Bulk and Yang et al. [38].
92 P. Nath et al. / Scripta Materialia 129 (2017) 88–93

Table 1
60
Pearson and (Spearman) correlations for the material data set. Highest Pearson
correlations in bold. κl(θ(2) (3)
a γ )
(1)
ha
(2)
ha
(3)
ha
(4)
ha
50 exp.
(1)
ca 0.87 0.97 0.82 0.93
40

κl (W/(m K))
(0.88) (0.89) (0.90) (0.88)
(2)
ca 0.80 0.95 0.64 0.83
(0.89) (0.89) (0.89) (0.89) 30
c(3) 0.97 0.99 0.91 0.97
(0.90) (0.90) (0.86) (0.90)
20

10
The comparison between our combinatorial approach and exper-
imentally reported values for j l is depicted in Fig. 2. The twelve
models predict qualitatively the main trend for j l . We used different 0
statistical quantities to measure the level of qualitative and quan- 500 1000 1500 2000 2500 3000
titative agreement between the models and the experimental data. T (K)
The Pearson and Spearman correlations measure the linear corre-
lation and the monotonicity of the relationship between the two Fig. 3. Predicted lattice thermal conductivity of MgO compared to the available
experimental data [39].
variables respectively (see Table 1). We have found high Pearson and
Spearman correlation values for some of the predicted models. For
     
(1) (2) (4)
instance, jl ha , c(3) , jl ha , c(3) and jl ha , c(3) present Pear-
son and Spearman correlations with experiments higher than 0.90. screening method, this method underestimates j l . This underestima-
To test the quantitative accuracy of the different models, we also tion can be as large as an order of magnitude which is particularly
compute the root mean square relative deviation, RMSrD, for each detrimental to compounds with high j l values.
(see Table 2). The results are far from the accuracy of the methods Finally, we tested the accuracy of our methodology for a wide
based on the exact solution of the BTE, however some of the com- range of temperatures to analyze the high-order anharmonic effects,
binations present which are not included
 in this
 model. We computed the temperature
 interesting
 results
 considering
 their lower
 compu- (2)
(2) (2) (2) (4)
tational cost. jl ha , ca , jl ha , c(3) and jl ha , c(3) present a dependence of jl ha , c(3) for MgO and we compare it with experi-
RMSrD lower than 100%. Moreover, in all cases, the major contribu- mental single-crystal measurements (see Fig. 3) [39]. Our framework
tion to this deviation corresponds to the materials with a lower j l , in performs well both above and below the ha (557 K) not only captur-
which a small absolute error could be transformed into a big relative ing the right trend but also being reasonably accurate.
error.
Combining both statistical  descriptors,
 we can conclude that the
(2) 4. Conclusions
best results are obtained for jl ha , c(3) . This is because the original
formulation of the Slack model depends onthe averageGrüneisen A high throughput combinatorial method for fast and robust pre-
(1)
parameter and not on the acoustic ones ca and c(2) . In addi- diction of lattice thermal conductivity is presented. Using the QHA-
(2)
tion, ha is computed explicitly using only the acoustic branches. APL implementation [35], we can compute different definitions for
(1) (2) (4)
However, ha , ha , and ha use the approximation: ha = n1/3 hD . the Debye temperature and the Grüneisen parameter. These values
However, these results should be compared with previous mod- can be used in a combinatorial approach with the Slack equation to
els to evaluate their performance. To the best of our knowledge, the obtain different values of j l (ha , c). We use a set of 42 materials to val-
most established and accurate model to predict j l using only har- idate our results and elucidate the best combination of parameters.
(2)
monic force constants has been developed by Madsen et al. [29]. Best results are obtained when ha is used, being the most accurate 
(2)
They used an approximated value for the phonon scattering time procedure to predict the acoustic Debye temperature. jl ha , c(3)
using the Debye temperature and the Grüneisen parameters. This is the most robust and accurate of all the combinations, obtain-
 
method jlM and our combinatorial approach have the same com- ing qualitatively and quantitatively better results than other models
putational cost, the calculation ofthe harmonic
 force constants. The in the literature with similar computational cost. The methodology
(2)
comparison between jlM and jl ha , c(3) is depicted in Fig. 2 (j). presented in this paper can be extremely useful because of its quan-
Both models show similar behaviors,
 although
 the statistical descrip- titative predictive power and its low computational cost compared
(2) to the calculation of the anharmonic force constants.
tors are slightly different. jl ha , c(3) presents a smaller RMSrD
than jlM (75% and 89% respectively) and a higher Pearson
 correlation

(2)
(0.99 and 0.98 respectively). We can also compare jl ha , c(3) with
Acknowledgments
less expensive methodologies such as AGL (see Fig. 2 (j)). Although
jlAGL captures some of the overall trends and can be used as a fast We would like to acknowledge support by DOD-ONR (N00014-
13-1-0635, N00014-11-1-0136, N00014-09-1-0921) and by DOE
(DE-AC02- 05CH11231), specifically the BES program under Grant
#EDCBEE. The AFLOW consortium would like to acknowledge the
Table 2
Root mean square relative deviation (RMSrD) for the material data set. Lowest RMSrD Duke University - Center for Materials Genomics and the CRAY
in bold. corporation for computational support.
(1) (2) (3) (4)
ha ha ha ha
Appendix A. Supplementary data
(1)
ca 237% 120% 355% 200%
(2)
ca 230% 90% 435% 200% Supplementary data to this article can be found online at http://
c(3) 112% 75% 269% 99%
dx.doi.org/10.1016/j.scriptamat.2016.09.034.
P. Nath et al. / Scripta Materialia 129 (2017) 88–93 93

References for High-Throughput Materials Science Calculations, Comp. Mat. Sci. Part A 108
(2015) 233–238.
[19] M.A. Blanco, E. Francisco, V. Luaña, GIBBS: isothermal-isobaric thermodynam-
[1] S. Curtarolo, G.L.W. Hart, M. Buongiorno Nardelli, N. Mingo, S. Sanvito, O. Levy,
ics of solids from energy curves using a quasi-harmonic Debye model, Comput.
The high-throughput highway to computational materials design, Nat. Mater.
Phys. Commun. 158 (2004) 57–72. 1.
12 (2013) 191–201.
[20] G.A. Slack, The Thermal Conductivity of Nonmetallic Crystals,H. Ehrenreich, F.
[2] M. Zebarjadi, K. Esfarjani, M.S. Dresselhaus, Z.F. Ren, G. Chen, Perspectives on
Seitz, D. Turnbull (Eds.), Solid State Phys. 34, Academic, New York, 1979, pp. 1.
thermoelectrics: from fundamentals to device applications, Energ. Environ. Sci.
[21] G. Leibfried, E. Schlömann, Warmleitund in elektrische isolierenden Kristallen,
5 (2012) 5147–5162.
Nach. Akad. Wiss. Gottingen, Math. Phyz. Klasse 4 (1954) 71.
[3] L.-T. Yeh, R.C. Chu, Thermal Management of Microloectronic Equipment: Heat
[22] C.L. Julian, Theory of Heat Conduction in Rare-Gas Crystals, Phys. Rev 137
Transfer Theory, Analysis Methods, and Design Practices, ASME Press. 2002.
(1965) A128–A137.
[4] C.D. Wright, L. Wang, P. Shah, M.M. Aziz, E. Varesi, R. Bez, M. Moroni,
[23] C. Toher, J.J. Plata, O. Levy, M. de Jong, M.D. Asta, M. Buongiorno Nardelli, S.
F. Cazzaniga, The design of rewritable ultrahigh density scanning-probe
Curtarolo, High-Throughput Computational Screening of thermal conductivity,
phase-change memories, Trans. Nanotechnol. 10 (5608505) (2011) 900–912.
Debye temperature and Grüneisen parameter using a quasi-harmonic Debye
[5] D.G. Cahill, P.V. Braun, G. Chen, D.R. Clarke, S. Fan, K.E. Goodson, P. Keblinski,
Model, Phys. Rev. B 90 (2014) 174107.
W.P. King, G.D. Mahan, A. Majumdar, H.J. Maris, S.R. Phillpot, E. Pop, L. Shi,
[24] L. Bjerg, B.B. Iversen, G.K.H. Madsen, Modeling the thermal conductivities of the
Nanoscale thermal transport. II. 2003–2012, Appl. Phys. Rev. 1 (2014) 011305.
zinc antimonides ZnSb and Zn4 Sb3 , Phys. Rev. B 89 (2014) 024304.
[6] J. Ziman, Electrons and Phonons: The Theory of Transport Phenomena in Solids,
[25] G. Deinzer, G. Birner, D. Strauch, Ab initio calculation of the linewidth of various
Clarendon. 1960.
phonon modes in germanium and silicon, Phys. Rev. B 67 (2003) 144304.
[7] J. Callaway, Model for Lattice Thermal Conductivity at Low Temperatures, Phys.
[26] S.L. Shindé, G.P. Srivastava (Eds.), Ab Initio Thermal Transport, Springer.2014,
Rev 113 (1959) 1046–1051.
[27] F. Zhou, W. Nielson, Y. Xia, V. Ozoliņš, Lattice Anharmonicity and Thermal Con-
[8] P.B. Allen, Zero-point and isotope shifts: Relation to thermal shifts, Phil. Mag. B
ductivity from Compressive Sensing of First-Principles Calculations, Phys. Rev.
70 (3) (1994) 527–534.
Lett. 113 (2014) 185501.
[9] D.A. Broido, M. Malorny, G. Birner, N. Mingo, D.A. Stewart, Intrinsic lattice
[28] D.T. Morelli, G.A. Slack, High lattice thermal conductivity solids, in: S.L. Shindé
thermal conductivity of semiconductors from first principles, Appl. Phys. Lett
(Ed.), High Thermal Conductivity Materials, Springer. 2006,
91 (2007) 231922.
[29] G.K.H. Madsen, A. Katre, C. Bera, Calculating the thermal conductivity of the
[10] Z. Wang, S. Wang, S. Obukhov, N. Vast, J. Sjakste, V. Tyuterev, N. Mingo, Ther-
silicon clathrates using the quasi-harmonic approximation, Phys. Stat. Solidi A
moelectric transport properties of silicon: Towards an ab initio approach, Phys.
213 (2016) 802–807.
Rev. B 83 (2011) 205208.
[30] M.A. Blanco, A.M. Pendás, E. Francisco, J.M. Recio, R. Franco, Thermodynamical
[11] J. Carrete, W. Li, N. Mingo, S. Wang, S. Curtarolo, Finding Unprecedent-
properties of solids from microscopic theory: Applications to MgF2 and Al2 O3 ,
edly Low-Thermal- Conductivity Half-Heusler Semiconductors via High-
J. Mol. Struct., Theochem 368 (1-3 SPEC. ISS.) (1996) 245–255.
Throughput Materials Modeling, Phys. Rev. X 4 (2014) 011019.
[31] J.-P. Poirier, Introduction to the Physics of the Earth’s Interior, 2nd edn. ed.,
[12] J. Carrete, N. Mingo, S. Wang, S. Curtarolo, Nanograined Half-Heusler Semicon-
Cambridge University Press. 2000.
ductors as Advanced Thermoelectrics: An Ab Initio High-Throughput Statistical
[32] G. Kresse, J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B
Study, Adv. Func. Mater. 24 (2014) 7427–7432.
47 (1993) 558–561.
[13] M.S. Green, Markoff random processes and the statistical mechanics of
[33] P.E. Blöchl, Projector augmented-wave method, Phys. Rev. B 50 (1994)
time-dependent phenomena. II. Irreversible processes in fluids, J. Chem. Phys.
17953–17979.
22 (3) (1954) 398–413.
[34] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made
[14] R. Kubo, Statistical-mechanical theory of irreversible processes. I. General
simple, Phys. Rev. Lett. 77 (1996) 3865–3868.
theory and simple applications to magnetic and conduction problems, J. Phys.
[35] P. Nath, J.J. Plata, D. Usanmaz, R. Al Rahal Al Orabi, M. Fornari, M. Buongiorno
Soc. Jpn. 12 (6) (1957) 570–586.
Nardelli, C. Toher, S. Curtarolo, Comput. Mater. Sci. 125 (2016) 82–91.
[15] S. Curtarolo, G. Ceder, Dynamics of an Inhomogeneously Coarse Grained Multi-
[36] Y. Wang, J.J. Wang, W.Y. Wang, Z.G. Mei, S.L. Shang, L.Q. Chen, Z.K. Liu, A
scale System, Phys. Rev. Lett. 88 (2002) 255504.
mixed-space approach to first-principles calculations of phonon frequencies
[16] S. Curtarolo, W. Setyawan, G.L.W. Hart, M. Jahnátek, R.V. Chepulskii, R.H. Taylor,
for polar materials, J. Phys.: Condens. Matter 22 (20) (2010) 202201.
S. Wang, J. Xue, K. Yang, O. Levy, M.J. Mehl, H.T. Stokes, D.O. Demchenko,
[37] G.A. Slack, S. Galginaitis, Thermal Conductivity and Phonon Scattering by
D. Morgan, AFLOW: an automatic framework for high-throughput materials
Magnetic Impurities in CdTe, Phys. Rev. 133 (1964) A253–A268.
discovery, Comp. Mat. Sci. 58 (2012) 218–226.
[38] X. Yang, J. Carrete, Z. Wang, Role of force-constant difference in phonon
[17] S. Curtarolo, W. Setyawan, S. Wang, J. Xue, K. Yang, R.H. Taylor, L.J.
scattering by nanoprecipitates in PbTe, J. Appl. Phys. 118 (2015) 085701.
Nelson, G.L.W. Hart, S. Sanvito, M. Buongiorno Nardelli, N. Mingo, O. Levy,
[39] A.M. Hofmeister, Thermal diffusivity and thermal conductivity of single-crystal
AFLOWLIB.ORG: A distributed materials properties repository from high-
MgO and Al2 O3 and related compounds as a function of temperature, Phys.
-throughput ab initio calculations, Comp. Mat. Sci. 58 (2012) 227–235.
Chem. Miner. 41 (2014) 361–371.
[18] C.E. Calderon, J.J. Plata, C. Toher, C. Oses, O. Levy, M. Fornari, A. Natan, M.J.
Mehl, G.L.W. Hart, M. Buongiorno Nardelli, S. Curtarolo, The AFLOW Standard

Das könnte Ihnen auch gefallen