methodology for determining the electronic thermal

9
This is a repository copy of Methodology for determining the electronic thermal conductivity of metals via direct nonequilibrium ab initio molecular dynamics . White Rose Research Online URL for this paper: http://eprints.whiterose.ac.uk/103356/ Version: Accepted Version Article: Yue, S-Y, Zhang, X, Stackhouse, S et al. (3 more authors) (2016) Methodology for determining the electronic thermal conductivity of metals via direct nonequilibrium ab initio molecular dynamics. Physical Review B, 94 (7). 075149. ISSN 2469-9950 https://doi.org/10.1103/PhysRevB.94.075149 © 2016, American Physical Society. This is an author produced version of a paper accepted for publication in Physical Review B. Uploaded in accordance with the publisher's self-archiving policy. [email protected] https://eprints.whiterose.ac.uk/ Reuse Unless indicated otherwise, fulltext items are protected by copyright with all rights reserved. The copyright exception in section 29 of the Copyright, Designs and Patents Act 1988 allows the making of a single copy solely for the purpose of non-commercial research or private study within the limits of fair dealing. The publisher or other rights-holder may allow further reproduction and re-use of this version - refer to the White Rose Research Online record for this item. Where records identify the publisher as the copyright holder, users can verify any specific terms of use on the publisher’s website. Takedown If you consider content in White Rose Research Online to be in breach of UK law, please notify us by emailing [email protected] including the URL of the record and the reason for the withdrawal request.

Upload: others

Post on 15-Jan-2022

12 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Methodology for determining the electronic thermal

This is a repository copy of Methodology for determining the electronic thermal conductivity of metals via direct nonequilibrium ab initio molecular dynamics.

White Rose Research Online URL for this paper:http://eprints.whiterose.ac.uk/103356/

Version: Accepted Version

Article:

Yue, S-Y, Zhang, X, Stackhouse, S et al. (3 more authors) (2016) Methodology for determining the electronic thermal conductivity of metals via direct nonequilibrium ab initio molecular dynamics. Physical Review B, 94 (7). 075149. ISSN 2469-9950

https://doi.org/10.1103/PhysRevB.94.075149

© 2016, American Physical Society. This is an author produced version of a paper accepted for publication in Physical Review B. Uploaded in accordance with the publisher's self-archiving policy.

[email protected]://eprints.whiterose.ac.uk/

Reuse

Unless indicated otherwise, fulltext items are protected by copyright with all rights reserved. The copyright exception in section 29 of the Copyright, Designs and Patents Act 1988 allows the making of a single copy solely for the purpose of non-commercial research or private study within the limits of fair dealing. The publisher or other rights-holder may allow further reproduction and re-use of this version - refer to the White Rose Research Online record for this item. Where records identify the publisher as the copyright holder, users can verify any specific terms of use on the publisher’s website.

Takedown

If you consider content in White Rose Research Online to be in breach of UK law, please notify us by emailing [email protected] including the URL of the record and the reason for the withdrawal request.

Page 2: Methodology for determining the electronic thermal

New methodology for determining the electronic thermal conductivity of metals via

direct non-equilibrium ab initio molecular dynamics

Sheng-Ying Yue,1 Xiaoliang Zhang,2 Stephen Stackhouse,3 Guangzhao Qin,2 Edoardo Di Napoli,1, 4 and Ming Hu1, 2, ∗

1Aachen Institute for Advanced Study in Computational Engineering Science (AICES),RWTH Aachen University, 52062 Aachen, Germany

2Institute of Mineral Engineering, Division of Materials Science and Engineering,Faculty of Georesources and Materials Engineering,RWTH Aachen University, 52064 Aachen, Germany

3School of Earth and Environment, University of Leeds, Leeds LS2 9JT, United Kingdom4Julich Supercomputing Centre, Forschungszentrum Julich and JARA–HPC, 52425 Julich, Germany

Many physical properties of metals can be understood in terms of the free electron model, asproven by the Wiedemann-Franz law. According to this model, electronic thermal conductivity canbe inferred from the Boltzmann transport equation (BTE). However, the BTE does not performwell for some complex metals, such as Cu. Moreover, the BTE cannot clearly describe the originof the thermal energy carried by electrons or how this energy is transported in metals. The chargedistribution of conduction electrons in metals is known to reflect the electrostatic potential ofthe ion cores. Based on this premise, we develop a new methodology for evaluating electronicthermal conductivity of metals by combining the free electron model and non-equilibrium ab initiomolecular dynamics simulations. We confirm that the kinetic energy of thermally excited electronsoriginates from the energy of the spatial electrostatic potential oscillation, which is induced by thethermal motion of ion cores. This method directly predicts the electronic thermal conductivityof pure metals with a high degree of accuracy, without explicitly addressing any complicatedscattering processes of free electrons. Our methodology offers a new route to understand the physicsof heat transfer by electrons at the atomistic level. The methodology can be further extended tostudy similar electron involved problems in materials, such as electron-phonon coupling, which isunderway currently.

PACS numbers: 64.30.Ef, 63.20.dk,72.10.Bg,05.70.Ln

I. INTRODUCTION

The electronic thermal conductivity (κel) is one of themost important physical properties of metals. The ana-lytical solution of κel based on the Boltzmann transportequation (BTE) and free electron model can be expressedas [1, 2]

κel =π2nk2BTτel

3m, (1)

where n is the concentration of free electrons, m is theelectron mass, kB is the Boltzmann constant, T is the sys-tem temperature and τel is the collision time of free elec-trons, which is mainly determined by electron-electron,electron-hole and electron-phonon scattering. In prin-ciple, we can obtain an approximate value for τel fromMatthiessen’s rule. However, describing every scatteringprocess involved in the heat transfer by electrons of solidmetals is too complicated. Recently, there have been anumber of studies of the κel of solid metals, based onBTE methodology [3, 4]. However, it is well known thatthe BTE of electrons is based on single relaxation-timeapproximation which may not hold true for all metals. In

∗Electronic address: [email protected]

addition, several methods have been used to evaluate theκel of liquid-phase metals within the framework of den-sity functional theory (DFT), such as ab initio molecu-lar dynamics (AIMD), using the Kubo-Greenwood equa-tion [5–7]. In view of this, there remains a need for aneffective method to evaluate κel of solid metals.

In this paper, we develop a new methodology to de-scribe the electronic heat-transport process in solid met-als without explicitly addressing detailed scattering pro-cesses. From the second law of thermodynamics, we knowthat heat transfer in solids is driven by the temperaturegradient ∇T . It should be noted that, the temperaturein heat transfer describes the thermal motion of atoms,i.e. the kinetic energy of nuclei. In the meantime, thevibrations of ions can lead to spatial electrostatic poten-tial oscillation (EPO), as can be easily deduced from themathematical expression for the total Hamiltonian of sys-tem. It easily follows that the local variation of the elec-trostatic potential can induce the collective oscillations offree electrons, and those free electrons near the Fermi sur-face can be excited above the Fermi surface and obtainadditional thermal kinetic energy with respect to 0 K.These are called thermally excited electrons. Figs. 1(a, b)show two cartoons describing how the thermally excitedelectrons move in the vibrational lattice and the localEP field. Higher temperatures, which induce larger andfaster ionic vibrations, lead to stronger EPO. Thus, thethermally excited electrons in high-temperature regions

Page 3: Methodology for determining the electronic thermal

2

FIG. 1: Cartoons of free electrons in a metal moving in (a)the vibrating lattice and (b) electrostatic potential field.

have more kinetic energy than those in low-temperatureregions. Once a stable distribution of the thermal ki-netic energy of thermally excited electrons is establishedalong the direction of ∇T , then the heat flux carried bythermally excited electrons and κel can be calculated.

II. THEORY AND EVIDENCE

To confirm this conjecture and quantify κel, we per-formed non-equilibrium ab initio molecular dynamics(NEAIMD) simulations [8, 9] by modifying the ViennaAb initio Simulation Package (VASP) [10, 11]. Theatomic heat flux was realized using the Muller-Plathe al-gorithm [12], in which the kinetic energies of the atoms inthe heat source and heat sink are exchanged (Supplemen-tary Information (SI) Sec. 1). With sufficient simulationtime, we can establish a stable temperature gradient inmetals. Figs. 2(a, b) present the Cu model and the cor-responding temperature profile, respectively. Simultane-ously, we can calculate the spatial distribution and thedynamical evolution of the EP, which is expressed as

U =

U(r) · ρtest · (|r −R|)d3r, (2)

where the test charge ρtest is the norm 1, and R repre-sents the ion position. Fig. 2(c) shows the theoreticalresults of the static distribution of the EP for a perfectCu lattice. In the rest of this paper, we confirm the rela-tionship between the spatial EPO and lattice vibrations,and that the EPO provides additional kinetic energy tothermally excited electrons. Following this, we show howto predict κel within our theoretical framework.To demonstrate the relationship between EPO and lat-

tice vibrations, we analyse the data from our AIMDsimulations using the power spectral density (PSD)method [13, 14]. For a stationary signal x(t), the PSD isdefined as

Sx(f) =

−∞

Rx(τ)e−2πifτdτ, (3)

whereRx(τ) = E[x(t)x(t+τ)] is the autocorrelation func-tion of x(t) [13, 14], and E[ · ] denotes the expectation

FIG. 2: Overview of the simulation model, temperature pro-file and EP field of copper. (a) Model of copper used inNEAIMD simulations and (b) the corresponding temperatureprofile. One unit cell length comprises two layers of atoms.We use fixed boundary conditions with the layers of fixedatoms and vacuum layers along the direction of ∇T . Periodicboundary conditions are adopted in the other two dimensions.(c) Theoretical EP field of a perfect copper structure (the testcharge number is norm 1).

value. Here, we consider four signals from an AIMDsimulation: atomic displacement Dion, atomic velocityVion, EP displacement Uion, and velocity of EPO (VEPO)∆Uion; these are used to calculate their respective spec-tral densities SD, SV [15], SU , S∆U (SI Sec. 2). SD andSV reflect the frequency-dependent lattice vibrations ata specific T . Analogously, SU and S∆U provide informa-tion regarding the EPO with respect to frequency. Weshow results for Al from a 10-ps equilibrium AIMD run at100.90K (Figs. 3(a, b)) and a 70-ps NEAIMD simulationat 299.46K (Figs. 3(c, d)). Fig. 3(a) clearly shows thatthe locations of the density peaks of SD and SU are con-sistent, demonstrating that the EPO is mainly caused bythe lattice vibration of ion cores. Fig. 3(b) confirms thisrelationship. Similar results are shown in Figs. 3(c, d):for most of the frequency ranges, the peaks of SD and SU ,SV and S∆U are consistent with each other. However, forsome specific frequencies in Figs. 3(c, d) (3.5 ∼ 5.0 THz),some discrepancies exist in the peaks’ magnitudes. It ispossible that this phenomenon could be due to the heatflux applied in the NEAIMD simulations. Nevertheless,Figs. 3(a ∼ d) provides an unambiguous physical pictureof the EPO being directly induced by the lattice vibra-tion of ions in metals.

To understand the dynamical evolution of spatial EPOintuitively, we present the representative case of Cu, cal-culated using NEAIMD at 298.49K, in Fig. 4(a). Lit-

Page 4: Methodology for determining the electronic thermal

3

FIG. 3: Overview of the relationship between EPO at ioncores and lattice vibrations. Blue and red lines representthe spectral density of (a) atom displacements (SD) and (c)the EP displacement (SU ), respectively, at specific ion cores.Green and orange lines represent the spectral density of (b)atom velocities (SV ) and (d) the EPO velocities (S∆U ), re-spectively, at specific ion cores. Data in (a) and (b) are from a10-ps equilibrium AIMD simulation of Al at 100.90K, whereas(c) and (d) present the same physical quantities from a 70-psNEAIMD simulation of Al at 299.46K.

tle variation occurs in the local electronic field betweenneighbouring atom layers, and the directions of these lo-cal fields continually change with time. The variationsof these local fields will drive the collective vibration offree electrons, as theoretically illustrated in Fig. 4(b).We see that only free electrons near the Fermi surfacecan be thermally excited. Because the direction of thelocal field continually changes with time, the vectors ofthe local momentum of the thermally excited electronsshould also continually change with time. Therefore, fora sufficiently long statistical time average, no net electriccurrent should arise during the thermal transport pro-cess of metals. This is consistent with the traditionalfree electron model [1].To confirm that the EPO provides additional kinetic

energy to thermally excited electrons in metals, we runa 100-ps equilibrium AIMD for Al at 329.40K and Li at283.97K (both with a 2×2×2 conventional-cell and 32total atoms). When T > 0K, the total energy of the freeelectron system can be written as [1, 2](SI Sec. 3)

Esys = E0 + ET = E0 +π2

4·N

(kBT )2

E0F

,

where E0 is the total energy of the free electron system at0K, ET is the thermally excited energy of the free elec-tron system obtained from the outside environment whenT > 0K, N is the total number of free electrons, and E0

F

is the Fermi energy at 0K. Because the Fermi energy

FIG. 4: (a) The variation of EPO in space over time (the testcharge number is 1). Data shown are from a 20-ps NEAIMDsimulation of Cu at 298.49K. (b) Schematic of the wholeFermi sphere oscillation as local electric field vibration alongthe ~z direction.

changes very little with temperature, here, we take theEF at room temperature as E0

F and adopt the experi-mental data[1]. We also calculate the energy providedby EPO using

EEPO = 2 · UEPO ·N · e,

where UEPO is the average effective EPO amplitude. ForAl, ET = 2.6293×10−21 J and EEPO = 2.7674×10−21 J.For Li, ET = 1.6049 × 10−21 J and EEPO = 1.5909 ×10−21 J. Based on these results, it is evident that

ET ≈ EEPO. (4)

This relation, although not a strict theoretical deriva-tion, confirms that lattice vibrations cause EPO in met-als, which, in turn, induces the collective vibration of freeelectrons. In fact, the energy of these collective vibra-tions provides additional kinetic energy to the thermallyexcited electrons. This is the core concept underlyingthis methodology.

III. METHOD AND RESULTS

Within this theoretical framework, higher tempera-tures strengthen the spatial EPO. To confirm this rela-tionship, we perform direct FFT of the relative displace-ment of EP Uion and VEPO ∆Uion. Uion and ∆Uion wereused to calculate SU and S∆U in Figs. 3(a − d). Uion de-scribes the strength of the EPO in space, whereas ∆Uion

Page 5: Methodology for determining the electronic thermal

4

FIG. 5: The direct fast Fourier transform (FFT) amplitudesof (a) the displacement of EP (DEP, Uion) and (b) VEPO(∆Uion) of ion cores at different temperatures. Data shownare from a 20-ps NEAIMD simulation of Cu at 298.49K. Thehigh and low temperatures correspond to 374.92 K and 198.39K, respectively.

reflects how fast the oscillation changes. Figs. 5(a, b)show the frequency-dependent FFT amplitudes of Uion

and ∆Uion, respectively. Clearly, the EPO is strongerand faster at higher temperatures.In Fig. 6(a), we present the positive and nega-

tive integrations of the total ∆Uion in the same atomlayer with simulation time, which can be written as∑Nal

j=1

t ∆Uj(t), where j is the index of the atom in thelayer and Nal is the total number of atoms per layer. Thefour quantities in Fig. 6(a) show perfect linear behaviourover time. From the absolute values, we have:

Nal∑

j=1

t

|∆Uj(t)hot| >

Nal∑

j=1

t

|∆Uj(t)cold|,

which is consistent with the evidence shown inFigs. 5(a, b). Notably, in the same temperature region,

the positive and negative accumulations of∑Nal

j=1 ∆Uj(t)

are almost the same. In other words,∑Nal

j=1

t ∆Uj(t) ≃0, and thus, there is no net electric field gradient alongthe heat flux direction for a sufficiently long statisticaltime. This result confirms the physical picture illustratedin Fig. 4(b). Fig. 6(b) presents the distribution of the av-erage effective amplitude of EPO in each atom layer alongthe heat flux direction, UEPO(l), where l is the index ofthe atom layers. Moreover, the amplitude distributionof EPO explains how the thermal kinetic energy of ther-mally excited electrons is divided in space. We calculateUEPO(l) using the RMS method [16]:

UEPO(l) =1

Nal

Nal∑

j=1

1

nsteps

nsteps∑

ti

(Uj(ti)− U j)2,

where nsteps is the total number of simulation time steps,Uj(ti) is the U value of atom j in a specific layer at time

FIG. 6: (a) Integration of positive and nega-

tive VEPO (∑Nal

j=1

t ∆Uj(t)) at different tem-

peratures. (b) The effective amplitude of EPO

( 1Nal

∑Nal

j=1

1nsteps

∑nsteps

i=1 (Uij − U j)2, average root mean

square (RMS) [16] of EPO, where Nal is the atom numberper layer) in atom layers along the ~z direction. Data shownare from a 20-ps NEAIMD simulation of Cu at 298.49K.

step ti, and U j is the average value of Uj(ti). Then,

we define the heat flux of electrons ~Jel according to thekinetic energy of thermally excited electrons between twoadjacent atom layers. Because of the isotropy of the freeelectron model (SI Sec. 4), we take half of the difference ofthe thermal kinetic energy of thermally excited electronsbetween the two layers as

~Jel = −1

2

n(e) · e

S · t

∂[2 · UEPO(l) · nsteps]

∂Nl

= −n(e) · e · nsteps

S · t

∂UEPO(l)

∂Nl

, (5)

where S is the cross-sectional area, t is the total simula-tion time, n(e) is the number of free electrons per atom

layer, and ∂UEPO(l)∂Nl

is the gradient of the average effective

amplitude value of EPO by linear fitting of UEPO(l) withthe atom layer index numberNl shown in Fig. 6(b). Here,a nonlinear phenomenon exists in the effective EPO am-plitude distribution along the heat flux direction in somemetals, such as Al, Be, and Mg. According to a casestudy of Be, we find that the nonlinear effect of UEPO(l)can be reduced by increasing the system size (SI Sec. 4.2).Because of the non-linear effect, when we calculate the~Jel of Al, Be and Mg, we fit the linear part only. For Cu

Page 6: Methodology for determining the electronic thermal

5

FIG. 7: Integration of the atomic kinetic energy flux withtime based on the NEAIMD (Muller-Plathe) simulations ofAl, Li, Be and Mg.

and Li, the UEPO(l) distributions exhibit perfect linearbehaviour along the heat current direction. Thus, we cancalculate κel based on Fourier’s law:

~J = −κ∇T. (6)

Combining Eqs. (5, 6), we obtain the expression for κel:

κel =n(e) · e · nsteps

∇T · S · t

∂UEPO(l)

∂Nl

, (7)

where ∇T is obtained by linear fitting the temperatureprofile with the representative case shown in Fig. 2(b).Note that, the induced temperature gradient is greaterthan that expected in a real system. Preliminary calcu-lations showed that both κel and κph were invariant totemperature gradient within the range of values studied.This can be seen in our final values for κ (Fig. 8(a), seedetailed computation parameters in Table 1 and Table 3in SI). It should be noted that each case listed in Ta-ble 3 in the SI has different temperature gradient, i.e.different heat flux. In some cases the temperature gra-dient differs by a factor of 1.81, yet both κel and κph

show no noticeable difference. This indicates that differ-ent temperature gradients result in almost the same κ.Using larger temperature gradients (within the linear re-sponse regime) can help to establish a stable temperatureprofile with a smaller associated uncertainty, since thetemperature difference between layers becomes greaterthan statistical fluctuations. It is also helpful to reducenon-linear effects of the VEPO in space, enabling us toobtain a stable value for the thermal conductivities fromboth electrons and phonons. It also helps to ensure thatwe have enough data for statistics. As the NEAIMDis realized by the Muller-Plathe method, the tempera-ture gradient relies on the interval time for the exchangeof atomic velocities between hot and cold baths. Thesmaller the interval time (meaning more frequent veloc-ity exchange), the larger the heat flux of atoms and thelarger the temperature gradient. This means that withlarger temperature gradients we have more valid datapoints to calculate the heat flux.Within this framework, we studied the κel of five met-

als (Li, Be, Mg, Al, and Cu) near room temperature. Ad-

FIG. 8: NEAIMD-EPO simulation results for metals. (a)Total thermal conductivities of metals from NEAIMD simu-lation (with error bars determined from the calculation of ∇T

and ∂UEPO(l)∂Nl

along the heat-transport direction) and exper-

imental data at 300K. (b) Pie graphs showing the electronicand phononic contributions to the total thermal conductivityof Al, Li, Be, and Mg at 300K.

ditionally, by integrating the Muller-Plathe [12] atomickinetic energy flux, as shown in Fig. 7, we predict the lat-tice (phonon) thermal conductivities of the metals (κph).Here we adopt the statistical physics approximation that,with sufficient simulation time, the time average of κel

and κph are equal to the ensemble average. Because of fi-nite size effects, our NEAIMD results underestimate κph,especially for Cu and Mg (SI Sec. 5). By summing κel

and κph from the NEAIMD simulations, we obtain thetotal thermal conductivities of the metals, as presentedin Fig. 8(a). The results demonstrate that the thermalconductivities of metals slowly decrease with tempera-ture near room temperature, which is consistent withtraditional theory and experimental data. The error es-timates in Fig. 8(a) are calculated from the expressionfor κel and error propagation theory [17]. They mainlystem from the calculation of the gradient of UEPO(l) and∇T . Here, we note that because the statistical temper-ature fluctuation (∆T )2 = kBT

2/Cv [18] of each atomlayer is large (because of the small number of atomsper layer), the conventional error estimate of ∇T willbe quite large. However, NEAIMD consistently yields astable temperature profile after a sufficiently long simu-lation time. Thus, we adopt the error in the linear fittingfor ∇T . We also note that the aforementioned nonlinearphenomenon of the gradient of UEPO(l) can also lead tolarge error bars. The details of the error-bar analysis can

Page 7: Methodology for determining the electronic thermal

6

FIG. 9: Bar graph comparing the thermal conductivities ofmetals calculated using the NEAIMD method (with errorbars), Boltzmann method and experimental data at 300K.

FIG. 10: Autocorrelation function of VEPO from NEAIMDsimulations for the case of (a) Cu and (b) Al. The exponentialfitting follows the function: y = Aexp(−t/τexp), where τexprepresents the exponential autocorrelation time of VEPO.

be found in SI Sec. 6. Meanwhile, in Fig. 8(b) we usepie graphs to show the electronic and phononic contribu-tions to the total thermal conductivity. Our results showthat κel indeed dominates the thermal transport processin metals.To the best of our knowledge, the BTE of electrons is

the only theory to be relatively successful evaluating theκel of solid metals. To compare our results with those ofthe traditional BTE method, we also utilize the Boltz-TraP software [3] (based on electron energy band the-ory) to calculate κel

τel. However, it is very difficult to ac-

curately and straightforwardly calculate the lifetime ofelectrons. Theoretical studies indicate that the magni-tude of the lifetime of electrons is around 1 × 10−14 sat room temperature[1, 3, 4, 22], and so, similar to pre-

vious studies[3, 4], we also use the constant relaxationtime approximation, with τel = 1 × 10−14 s. To avoidfinite size effects in the calculation of κph, we also eval-uate κph from the BTE method with interatomic forceconstants obtained from ab initio calculations [19, 20], asimplemented in the ShengBTE package [21]. Then, weobtain the total thermal conductivities of the metals viathe BTE method by summing κel from BoltzTraP andκph from ShengBTE. Our NEAIMD method, the tradi-tional BTE method and experimental data are comparedin Fig. 9. The results demonstrate that the BTE methodis unable to correctly describe κel for all metals, and ourmethod is superior to the traditional BTE method inpredicting the electronic thermal conductivities of met-als, especially for Be and Cu at room temperature.

Moreover, we observe an interesting phenomenon whencalculating the spectral density of VEPO (S∆U ) inFigs. 3(b, d). We perform exponential decay fitting ofthe autocorrelation function of the VEPO using the for-mula y = Aexp(−t/τexp). Surprisingly, the exponentialautocorrelation time of VEPO τexp at room temperatureis on the same approximate order of magnitude as thetheoretical collision time of the free electrons [3, 4, 22].The results for Cu and Al are shown in Fig. 10. We alsoexamine other metals (Be, Li, and Mg) and obtain simi-lar results (SI Sec. 8). Therefore, we anticipate that somephysical mechanisms must drive this phenomenon, i.e., itis not a coincidence.

Before closing, we would like to point out that, in theBTE expression for κel, the interactions between elec-trons and nuclei are implicit in the electron scatteringtime τel, or equivalently, the presence of nuclei has a sig-nificant effect on the electron scattering time, which, inturn, affects κel. Thus, in principle, κel should dependon the vibration of nuclei. In this sense, our theory andmethodology is not in conflict with the traditional freegas model. However, in contrast to the traditional elec-tron BTE method, our direct non-equilibrium ab initiomolecular dynamics simulation based on EPO (we nameit NEAIMD-EPO method) can calculate electronic ther-mal conductivity directly by mimicking the real physicalpicture of the heat transfer in metals, without artificialmanipulation and input parameters. The NEAIMD-EPOmethod naturally, but implicitly, includes the compli-cated interactions between electrons and electron-phononcoupling. Our method is applicable to all solid metals,whereas the traditional electrons BTE method strugglesto evaluate κel for some elements. Our method also hassome limitations, at present, such as: 1) as our NEAIMD-EPO framework is built on the free electron gas model,so far, this method is limited to simulation of pure met-als; 2) this method cannot be directly used to simulatethermal transport of metals at low temperatures; 3) asthis method is realized in the ab initio molecular dy-namics simulation, the simulation results will depend onthe pseudopotential used and 4) the computation costsfor the NEAIMD simulations are much higher than thatof normal density functional theory (DFT) simulations.

Page 8: Methodology for determining the electronic thermal

7

However, with theory and computational capacity im-proving, the NEAIMD- EPO method shows potential forinvestigation of different kinds of electronic systems, i.e.,alloys, semiconductors, metal/non-metal interfaces, andeven directly simulating nano-devices in the future.

IV. CONCLUSIONS

In summary, we have developed a new methodologybased on the concept of electrostatic potential oscilla-tion to predict the electronic thermal conductivities ofmetals via direct non-equilibrium ab initio molecular dy-namics simulation. We provide a clear and new physicalpicture of the origin of the thermal energy carried byelectrons and reveal how this energy is transported inmetals. Without explicitly addressing any complicatedscattering processes of free electrons, our NEAIMD-EPOmethod provides better predictions of the electronic ther-mal conductivities of pure metals than the traditionalBTE method near room temperature. Our methodol-ogy offers a new route to understand the physics of heattransfer by electrons at the atomistic level. We expectthat this methodology will be helpful and useful for un-derstanding and studying the heat-transfer problems ofmetal systems in the future. Further extension to cope

with some presently challenging problems in materials,such as electron-phonon coupling, is also foreseen.

Acknowledgments

M.H. gratefully acknowledges Prof. David G. Cahill(University of Illinois at Urbana-Champaign) for valuablecomments on the manuscript. S.Y.Y. gratefully thanksDr. Yang Han (RWTH Aachen University) and Dr. TaoOuyang (Xiangtan University) for their helpful and fruit-ful discussions, and Dr. Shi-Ju Ran (ICFO, Spain) for hishelp in plotting Fig. 1(b) and Fig. 2(c). X.Z. greatly ac-knowledges Dr. Zhiwei Cui (Northwestern University)for providing the input script and MEAM potential filesfor the EMD simulations of Li. The authors gratefullyacknowledge the computing time granted by the Johnvon Neumann Institute for Computing (NIC) and pro-vided on the supercomputer JURECA at Julich Super-computing Centre (JSC) (Project ID: JHPC25). Theclassical NEMD and EMD simulations were performedwith computing resources granted by the Julich AachenResearch Alliance-High Performance Computing (JARA-HPC) from RWTH Aachen University under project No.jara0135. S.S. was supported by NERC grant numberNE/K006290/1.

[1] Kittel, C. Introduction to Solid States Physics (8th Ed.).Ch.6, (John.Wiley & Sons, Inc., USA, 2004).

[2] Jones, W. & March, N. H. Theoretical Solid StatePhysics. (Courier Dover Publications, 1985).

[3] Madsen, G. K. H. & Singh, D. J. BoltzTraP. A code forcalculating band-structure dependent quantities. Com-put. Phys. Commun. 175, 67 (2006).

[4] Chen, M. X. & Podloucky, R. Electronic thermal conduc-tivity as derived by density function theory. Phys. Rev.B 88, 045134 (2013).

[5] Desjarlais, M. P., Kress, J. D. & Collins, L. A. Electri-cal conductivity for warm, dense aluminum plasmas andliquids. Phys. Rev. E 66, 025401(R) (2002).

[6] Pozzo, M., Davies, C., Gubbins, D. & Alfe, D. Thermaland electrical conductivity of iron at Earth’s core condi-tions. Nature 485, 355 (2012).

[7] de Koker, N., Steinle-Neumann, G. & Vlcek, V. Electricalresistivity and thermal conductivity of liquid Fe alloys athigh P and T, and heat flux in Earth’s core. Proc. Natl.Acad. Sci. USA 109, 4070 (2012).

[8] Stackhouse, S., Stixrude, L. & Karki, B. B. Thermal Con-ductivity of Periclase (MgO) from First Principles. Phys.Rev. Lett. 104, 208501 (2010).

[9] Kim, H., Kim, M. H. & Kaviany, M. Lattice thermal con-ductivity of UO2 using ab-initio and classical moleculardynamics. J. Appl. Phys. 115, 123510 (2014).

[10] Kresse, G. & Furthmuller, J. Efficient iterative schemesfor ab initio total-energy calculations using a plane-wavebasis set. Phys. Rev. B. 54, 11169 (1996).

[11] Kresse, G. & Furthmuller, J. Efficiency of ab-initio totalenergy calculations for metals and semiconductors using

a plane-wave basis set. Comput. Mater. Sci. 6, 15 (1996).[12] Muller-Plathe, F. A simple nonequilibrium molecular dy-

namics method for calculating the thermal conductivity.J. Chem. Phys. 106, 6082 (1997).

[13] Dennis, W. R. Echo Signal Processing. (Springer, 2003).[14] Storch, H. V. & Zwiers, F. W. Statistical Analysis in

Climate Research. (Cambridge University Press, 2001).[15] Zhang, X. & Jiang J. Thermal conductivity of zeolitic

imidazolate framework-8: a molecular simulation study.J. Phys. Chem. C. 117, 18441 (2013).

[16] Taylor, J. R. An Introduction to Error Analysis: theStudy of Uncertainties in Physical Measurements (2ndEd.). Ch.4, (University Science Books, 1997).

[17] Ku, H. H. Notes on the use of propagation of error for-mulas. Journal of Research of the National Bureau ofStandards - C. Engineering and Instrumentation, 70C,No. 4 (1966).

[18] Landau, L. D. & Lifshitz, E. M. Statistical Physics, Part1 (3rd Ed.). Ch.2, (Pergamon Press, 1980).

[19] Broido, D. A., Malorny, M., Birner, G., Mingo, N. &Stewart, D. A. Intrinsic lattice thermal conductivity ofsemiconductors from first principles. Appl. Phys. Lett.91, 231922 (2007).

[20] Lee, S., Esfarjani, K., Mendoza, J., Dresselhaus, S. &Chen, G. Lattice thermal conductivity of Bi, Sb, and Bi-Sb alloy from first principles. Phys. Rev. B. 89, 085206(2014).

[21] Li, W., Carrete, J., Katcho, N. A. & Mingo, N. Sheng-BTE: A solver of the Boltzmann transport equation forphonons. Comput. Phys. Commun. 185, 1747 (2014).

[22] Campillo, I., Pitarke, J. M., Rubio, A., Zarate, E., &

Page 9: Methodology for determining the electronic thermal

8

Echenique, P. M. Inelastic lifetimes of hot electrons inreal metals. Phys. Rev. Lett. 83, 2230 (1999).