Influence of crystallographic orientation and anisotropy ... · materials A and B, with peaks...

9
Influence of crystallographic orientation and anisotropy on Kapitza conductance via classical molecular dynamics simulations J. C. Duda, C. J. Kimmer, W. A. Soffa, X. W. Zhou, R. E. Jones et al. Citation: J. Appl. Phys. 112, 093515 (2012); doi: 10.1063/1.4764921 View online: http://dx.doi.org/10.1063/1.4764921 View Table of Contents: http://jap.aip.org/resource/1/JAPIAU/v112/i9 Published by the American Institute of Physics. Related Articles Physical insight into reduced surface roughness scattering in strained silicon inversion layers Appl. Phys. Lett. 101, 073504 (2012) Comment on “Twin symmetry texture of energetically condensed niobium thin films on sapphire substrates (a- plane Al2O3)” [J. Appl. Phys. 110, 033523 (2011)] J. Appl. Phys. 112, 016101 (2012) Molecular dynamics simulations of a hard sphere crystal and reaction-like mechanism for homogeneous melting J. Chem. Phys. 136, 234505 (2012) Physics of shell assembly: Line tension, hole implosion, and closure catastrophe J. Chem. Phys. 136, 184507 (2012) Temperature dependent magnetic structure of lithium delithiated LixFeSO4F (x=0, 1) by Mössbauer spectroscopy J. Appl. Phys. 111, 07E138 (2012) Additional information on J. Appl. Phys. Journal Homepage: http://jap.aip.org/ Journal Information: http://jap.aip.org/about/about_the_journal Top downloads: http://jap.aip.org/features/most_downloaded Information for Authors: http://jap.aip.org/authors Downloaded 28 Dec 2012 to 137.54.10.198. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions

Transcript of Influence of crystallographic orientation and anisotropy ... · materials A and B, with peaks...

Page 1: Influence of crystallographic orientation and anisotropy ... · materials A and B, with peaks centered at x 8:5 and 20. The phonon dispersion relations in selected directions of high

Influence of crystallographic orientation and anisotropy on Kapitzaconductance via classical molecular dynamics simulationsJ. C. Duda, C. J. Kimmer, W. A. Soffa, X. W. Zhou, R. E. Jones et al. Citation: J. Appl. Phys. 112, 093515 (2012); doi: 10.1063/1.4764921 View online: http://dx.doi.org/10.1063/1.4764921 View Table of Contents: http://jap.aip.org/resource/1/JAPIAU/v112/i9 Published by the American Institute of Physics. Related ArticlesPhysical insight into reduced surface roughness scattering in strained silicon inversion layers Appl. Phys. Lett. 101, 073504 (2012) Comment on “Twin symmetry texture of energetically condensed niobium thin films on sapphire substrates (a-plane Al2O3)” [J. Appl. Phys. 110, 033523 (2011)] J. Appl. Phys. 112, 016101 (2012) Molecular dynamics simulations of a hard sphere crystal and reaction-like mechanism for homogeneous melting J. Chem. Phys. 136, 234505 (2012) Physics of shell assembly: Line tension, hole implosion, and closure catastrophe J. Chem. Phys. 136, 184507 (2012) Temperature dependent magnetic structure of lithium delithiated LixFeSO4F (x=0, 1) by Mössbauerspectroscopy J. Appl. Phys. 111, 07E138 (2012) Additional information on J. Appl. Phys.Journal Homepage: http://jap.aip.org/ Journal Information: http://jap.aip.org/about/about_the_journal Top downloads: http://jap.aip.org/features/most_downloaded Information for Authors: http://jap.aip.org/authors

Downloaded 28 Dec 2012 to 137.54.10.198. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions

Page 2: Influence of crystallographic orientation and anisotropy ... · materials A and B, with peaks centered at x 8:5 and 20. The phonon dispersion relations in selected directions of high

Influence of crystallographic orientation and anisotropy on Kapitzaconductance via classical molecular dynamics simulations

J. C. Duda,1,2,a) C. J. Kimmer,3 W. A. Soffa,4 X. W. Zhou,5 R. E. Jones,5

and P. E. Hopkins1,b)

1Department of Mechanical and Aerospace Engineering, University of Virginia, Charlottesville,Virginia 22904, USA2Microscale Science and Technology Department, Sandia National Laboratories, Albuquerque,New Mexico 87185, USA3School of Natural Sciences, Indiana University Southeast, New Albany, Indiana 47150, USA4Department of Materials Science and Engineering, University of Virginia, Charlottesville,Virginia 22904, USA5Mechanics of Materials Department, Sandia National Laboratories, Livermore, California 94550, USA

(Received 20 July 2012; accepted 12 October 2012; published online 8 November 2012)

We investigate the influence of crystallographic orientation and anisotropy on local phonon density

of states, phonon transmissivity, and Kapitza conductance at interfaces between Lennard-Jones

solids via classical molecular dynamics simulations. In agreement with prior works, we find that

the Kapitza conductance at an interface between two face-centered cubic materials is independent

of crystallographic orientation. On the other hand, at an interface between a face-centered cubic

material and a tetragonal material, the Kapitza conductance is strongly dependent on the relative

orientation of the tetragonal material, albeit this dependence is subject to the overlap in vibrational

spectra of the cubic and tetragonal materials. Furthermore, we show that interactions between

acoustic phonons in the cubic material and optical phonons in the tetragonal material can lead to

the interface exhibiting greater “thermal anisotropy” as compared to that of the constituent

materials. Finally, it is noted that the relative match or mismatch between the Debye temperatures

of two materials comprising an interface does not serve an accurate gauge of the efficiency of

interfacial thermal transport when those materials have different crystal structures. VC 2012American Institute of Physics. [http://dx.doi.org/10.1063/1.4764921]

I. INTRODUCTION

The thermal transport properties of modern devices are

now dictated more so by the interfaces between materials

than they are by the materials themselves.1 In response, a

significant effort has been put forth to both investigate and

explain the behavior of thermal transport across solid-solid

interfaces, the efficiency of which can be described by the

interface Kapitza conductance,2 or hK: For example, prior

works have demonstrated how the mismatch of phonon spec-

tra,3–7 interfacial structure,8–14 and interfacial chemistry15–21

can each impact conductance. However, despite the many

advances within this rapidly growing body of work, rela-

tively few studies have focused on the influence of crystallo-

graphic orientation and anisotropy on Kapitza conductance

at solid-solid interfaces.

In nanostructured material systems, thermal transport

properties can exhibit high degrees of anisotropy due to both

geometric and material factors. With regard to the former,

reduced characteristic lengths can lead to increased anisot-

ropy through boundary scattering rates that vary based on

direction22–26 and through changes in the phonon spec-

tra.24,27 As for the latter, material factors (i.e., thermal ani-

sotropy inherent to the materials comprising the interface)

have been less frequently studied. Theoretical28,29 and exper-

imental30 investigations of Kapitza conductance at metal:car-

bon-nanotube and metal:graphite interfaces have indicated

that conductance can vary by an order of magnitude depend-

ing on the orientation of the nanotube or graphite; this behav-

ior has been attributed to phonon group velocities that vary

with crystallographic direction. On the other hand, Costescu

et al.8 measured the conductance of epitaxially grown

TiN:MgO(001) and TiN:MgO(111) interfaces via time-

domain thermoreflectance (TDTR) and reported no depend-

ence on crystallographic orientation. Similarly, Hopkins

et al.31 measured Kapitza conductance at Al:Si(001) and

Al:Si(111) interfaces via TDTR and likewise reported no

orientational dependence. However, in that same study, con-

ductance at Al : Al2O3(0001) and Al : Al2O3(11�20) interfaces

differed by up to 40%. As with the earlier metal:nanotube and

metal:graphite studies, this effect was attributed to direction-

ally dependent phonon group velocities.

In this paper, we report on the influence of crystallo-

graphic orientation and anisotropy on Kapitza conductance

at interfaces between Lennard-Jones (LJ) solids via classical

molecular dynamics (MD) simulations. In agreement with

the aforementioned studies, we find that the Kapitza conduct-

ance at an interface between two cubic materials (face-

centered cubic or fcc type) is independent of crystallographic

orientation, while at an interface between a fcc material and

a tetragonal material (L10 type) the Kapitza conductance is

strongly dependent on the relative orientation of the L10 ma-

terial. In contrast with earlier works, this observed effect is

a)[email protected])[email protected].

0021-8979/2012/112(9)/093515/8/$30.00 VC 2012 American Institute of Physics112, 093515-1

JOURNAL OF APPLIED PHYSICS 112, 093515 (2012)

Downloaded 28 Dec 2012 to 137.54.10.198. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions

Page 3: Influence of crystallographic orientation and anisotropy ... · materials A and B, with peaks centered at x 8:5 and 20. The phonon dispersion relations in selected directions of high

not attributed to directionally dependent group velocities, as

the magnitudes of the group velocities in the L10 material do

not vary significantly with crystallographic direction. To

gain further insight into this observed behavior, we spatially

resolve the phonon density of states (DOS) near the interface

and determine frequency-dependent phonon transmissivities

via wave-packet simulations. Briefly, it is found that the

interactions between acoustic phonons in a fcc material and

the optical phonons in a L10 material can dramatically

enhance thermal anisotropy at the interface beyond that in-

herent to the constituent materials.

The remainder of this paper will be organized as fol-

lows: Sec. II will detail the Lennard-Jones solids considered,

including their bulk vibrational properties as calculated via

harmonic lattice dynamics (LD); Sec. III will report on the

thermal conductivities of the constituent materials and the

Kapitza conductances at interfaces between them as deter-

mined via non-equilibrium molecular dynamics (NEMD)

simulations; Sec. IV will discuss the results in the context of

local phonon density of states calculations and wave-packet

simulations at interfaces between fcc and L10 solids; finally,

Sec. V will provide a summary of major results.

II. PROPERTIES OF CONSTITUENT MATERIALS

We considered three different materials which we will

refer to as material A, material B, and the AB alloy. All intera-

tomic interactions were described by the 6–12 LJ potential,

UðrÞ ¼ 4e½ðr=rÞ12 � ðr=rÞ6�, where U is potential energy, r is

the interatomic separation, and r and e are the LJ length and

energy parameters, respectively. To reinforce the generality of

this work (our objective is not to extract quantitative data for

specific interfaces but rather to discover the dependence of

Kapitza conductance on interfacial configuration), we present

all relevant quantities in non-dimensional LJ units (see Table I

for the nondimensionalization scheme).5,18 The LJ potential

was parameterized identically for each material,32 and the cut-

off distance was set to rc ¼ 2:5r. Consequently, all materials

had the same zero-temperature cubic lattice constant,

alat ¼ 1:55, and thus differed only by atomic mass and

arrangement. Materials A and B were fcc solids comprised of

a single type of atom with a unique mass; mA ¼ 40 amu and

mB ¼ 120 amu. The AB alloy contained both types of atoms

in the layered-tetragonal L10 configuration (see Fig. 1). In this

structure, an A atom occupies the atomic site at (0, 0, 0) and a

B atom occupies the site at (0.5, 0.5, 0.5) relative to the primi-

tive vectors of the tetragonal unit cell. This atomic arrange-

ment forms a 1� 1 superlattice in the [001] crystallographic

direction, or along the c-axis.

The bulk phonon DOS and phonon dispersion relations

of materials A, B, and the AB alloy were determined via har-

monic lattice dynamics calculations. These calculations were

carried out with the General Utility Lattice Program

(GULP),33 the results of which are shown in Fig. 2. In the

top panel, the phonon DOS is plotted as a function of angular

frequency, x. The DOS of materials A and B exhibits the

characteristic shape associated with fcc materials, and their

cutoff frequencies differ by a factor offfiffiffiffiffiffiffiffiffiffiffiffiffiffimB=mA

ffiffiffi3p

. The

shape of the DOS of the AB alloy differs greatly from that of

materials A and B, with peaks centered at x � 8:5 and 20.

The phonon dispersion relations in selected directions of

high symmetry of materials A, B, and the AB alloy at zero

temperature are plotted in the lower three panels of Fig. 2. In

TABLE I. Non-dimensional (LJ-reduced) scheme. All quantities were nor-

malized such that m ¼ mA, or 40 amu. Throughout the text, the asterisks are

dropped and non-dimensional quantities are implied.

Quantity Non-dimensional quantity

Length, r r� ¼ r � r�1

Temperature, T T� ¼ T � kB e�1

Angular frequency, x x� ¼ x� r m0:5 e�0:5

Thermal flux, Q Q� ¼ Q� r3 m0:5 e�1:5

Thermal conductivity, j j� ¼ j� r2 m0:5 k�1B e�0:5

Kapitza conductance, hK h�K ¼ hK � r3 m0:5 k�1B e�0:5

FIG. 1. Schematics of the (a) tetragonal primitive unit cell and (b) cubic

conventional unit cell of the layered-tetragonal L10 AB alloy.

FIG. 2. Bulk DOS (top) and phonon dispersion relations (bottom) in selected

directions of high symmetry of materials A, B, and the AB alloy as calcu-

lated via harmonic lattice dynamics. The wavevector has been normalized to

2p=alat, where alat is the lattice constant of the cubic conventional cell. Note

that the slopes of the dispersion curves, i.e., phonon group velocities, of the

AB alloy in the in-plane and cross-plane directions do not differ

significantly.

093515-2 Duda et al. J. Appl. Phys. 112, 093515 (2012)

Downloaded 28 Dec 2012 to 137.54.10.198. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions

Page 4: Influence of crystallographic orientation and anisotropy ... · materials A and B, with peaks centered at x 8:5 and 20. The phonon dispersion relations in selected directions of high

all three panels, the wavevector has been normalized by

2p=alat. Note that the maximum wavevector of the AB alloy

in the cross-plane direction (cp) is half of that along the in-

plane direction (ip, or [0.5, 0.5, 0] relative to the primitive

vectors of the tetragonal unit cell). Both the DOS and disper-

sion relations demonstrate that while the vibrational spectra

of material A and the AB alloy better match one another in

terms of their cutoff frequencies (and hence, Debye tempera-

tures), material B and the AB alloy exhibit better spectral

overlap at frequencies below x ¼ 10. In addition, it is

important to note that the phonon group velocities,

vg ¼ j@x=@kj, do not differ significantly from the in-plane to

cross-plane direction in the AB alloy.

III. NON-EQUILIBRIUM MOLECULAR DYNAMICS

Classical molecular dynamics simulations are well

suited to study thermal transport across interfaces in semi-

conductor systems as the time-evolution of the positions and

energies of atoms around an interface are exactly known. For

example, a standard steady-state NEMD simulation can be

set up to investigate either the thermal conductivity of a ma-

terial, j, or the Kapitza conductance at the interface between

two different materials, hK. In either case, a thermal flux, Q,

is applied across a computational domain in order to estab-

lish a steady-state spatial temperature gradient, rT. If the

thermal conductivity of a single material is sought, the

observed spatial temperature gradient can be related to the

thermal conductivity by invoking the Fourier law, Q ¼ jrT.

As for Kapitza conductance, differences in vibrational prop-

erties will lead to a temperature discontinuity at the interface

between two materials when a flux is applied. This disconti-

nuity, DT, is related to conductance through the relationship

Q ¼ hKDT.

In the present study, the thermal conductivities of mate-

rials A, B, and the AB alloy, as well as the Kapitza conduc-

tances at various interfaces between them, were calculated

via NEMD simulations at temperatures from 0.09 to 0.42

(the melt temperature of LJ solids is Tm � 0:5 (Ref. 5)). The

computational domains implemented when calculating con-

ductivities and conductances were rectangular cuboids elon-

gated in the z-direction with periodic boundary conditions

applied in the x- and y-directions, thus taking on the general

characteristics associated with many previous NEMD simu-

lations.5,19,34–36 The four outermost monolayers in the

positive and negative z-directions formed rigid walls. The

eight monolayers immediately inside these walls were bath

atoms, to which energy would be added or removed to estab-

lish a temperature gradient. Further specifics regarding the

domains will be discussed in subsequent sections.

During the simulation, the equations of motion for the

system were integrated using the Nordsieck fifth-order pre-

dictor-corrector algorithm37 with a time step of 0.002. The

systems were first equilibrated at a predefined temperature

via a velocity scaling routine and zero pressure, maintained

by the Berendsen barostat algorithm.38 In this routine, pres-

sure was calculated for all dynamic atoms in the system and

the volume scaling involved displacement of the rigid walls.

Subsequently, the NEMD procedure was implemented. The

addition of energy to/removal from the baths was performed

through a constant-flux approach.36,39 This routine slightly

changes the forces acting on a particular atom depending on

the amount of energy to be added or removed. The total force

acting on atom i is given by

Fi;total ¼ Fi þ nmivTi ; (1)

where mi is the mass of the atom, vTi is the thermal velocity

of the atom, and n is a scaling factor. This scaling factor is

expressed as

n ¼ Q1

2KT¼ DE

2DtKT; (2)

where Q is the flux across the domain, Dt is the time step

used in the MD integration of the equations of motion, DE is

the amount of energy to be added to or removed from the

bath per time step, and KT is the total thermal kinetic energy

of the bath. For thermal conductivity and Kapitza conduct-

ance calculations, the applied flux varied between 7:95�10�5 and 15:9� 10�5 depending on the overall thermal re-

sistivity of the particular domain of interest (in order to

make the change in temperature small but significant relative

to the noise).

During NEMD, the systems were divided into equally

sized bins such that the temperature of each monolayer could

be calculated, thus allowing for the construction of spatial

temperature profiles along the z-axis. Once in steady-state

(as determined by monitoring the temporal evolution of the

bath temperatures), time-averaged profiles were constructed

from 3500 system snapshots taken over 1:75� 106 time

steps. Linear least-squares fits of these temperature profiles

were then performed (separately for each half of the domain

in the Kapitza conductance simulations). In certain cases,

some non-linearity in the temperature profiles was observed

near the baths and/or the interface; these data were not

included in the fits. The slopes of these fits were used to cal-

culate the thermal conductivity; the discontinuities between

fits at the interfaces were used to calculate Kapitza conduct-

ance. When calculating Kapitza conductance, it was ensured

that the thermal conductivities of each half of the domain

were in agreement with the thermal conductivities of the re-

spective materials as determined from separate thermal con-

ductivity simulations of larger, homogenous domains in

which bulk properties could be reproduced. Finally, to

ensure that our results were independent of the direction of

applied heat flux, additional simulations were conducted

with this direction reversed; no systematic dependence was

observed.

A. Thermal conductivity

The thermal conductivities of materials A and B were

calculated along the [001] direction and the thermal con-

ductivity of the AB alloy was calculated in both the in-

plane and cross-plane directions (the thermal conductivity

tensors of materials A and B are isotropic due to their

cubic symmetry). Two pre-equilibrated domain sizes were

implemented to ensure size effects did not distort results;

093515-3 Duda et al. J. Appl. Phys. 112, 093515 (2012)

Downloaded 28 Dec 2012 to 137.54.10.198. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions

Page 5: Influence of crystallographic orientation and anisotropy ... · materials A and B, with peaks centered at x 8:5 and 20. The phonon dispersion relations in selected directions of high

12:4� 12:4� 93 and 12:4� 12:4� 124. These domain

lengths, Lz, correspond to 60 and 80 conventional unit cells,

respectively. The thermal conductivity of material A as cal-

culated in domains of each size is plotted in Fig. 3. As is

evident from the plot, this change in domain length did not

produce any systematic difference in the calculated values

of thermal conductivity, suggesting finite size effects did

not distort our results. Regardless, the Lz ¼ 124 length was

used in all subsequent simulations in order to generate

more data points over which to fit when calculating the spa-

tial temperature gradients.

Results of our thermal conductivity simulations are

plotted as a function of temperature in Fig. 3. Each data

point is the mean result of three independent simulations;

standard deviations were less than 6% of the mean value in

all cases. The thermal conductivities of the materials exhibit

the inverse temperature dependence attributed to phonon-

phonon scattering.27,40 This temperature dependence is

illustrated by the inverse power fits of the NEMD data rep-

resented by the dashed lines in Fig. 3. Across the entire tem-

perature range, the thermal conductivity of the AB alloy in

the in-plane direction is approximately a factor of two

higher than in the cross-plane direction. This difference is

due to the fact that the Brillouin zone of the AB alloy is nar-

rower by a factor of two in the cross-plane direction as com-

pared to the in-plane direction. As a result, the maximum

wavevector of phonons traveling along that axis is half that

of phonons traveling in the in-plane direction, and thus, the

likelihood of Umklapp scattering is higher in the cross-

plane direction as compared to that in the in-plane direc-

tion.41,42 While the formation of a superlattice typically

leads to band flattening and lower phonon group velocities,

the dispersion curves in Fig. 2 indicate that directionally de-

pendent group velocities alone cannot explain the conduc-

tivity trends of the AB alloy.

B. Kapitza conductance

Kapitza conductances were calculated at six different

interfaces via NEMD simulations; the specifics of these

interfaces and the computational domains are listed in

Table II. Note that the (111) orientation of sample II pre-

vented a square cross section and thus prescribed different

overall dimensions; still, the number of atoms contained in

that domain was within 4% of that in the others. The calcu-

lated conductances are plotted as a function of temperature

in Fig. 4. Each data point represents the mean result of five

independent simulations (with the exception of the hollow

symbols in Fig. 4(b), which each represent the result of a sin-

gle simulation) and error bars represent the standard devia-

tion of these five simulations, i.e., repeatability. In some

cases, the standard deviation is small enough that the error

bars are masked by the data point itself. In Figs. 4(b) and

4(c), the dashed lines are linear fits of the NEMD data.

Samples I and II contained interfaces between materials

A and B and differed in terms of their crystallographic orien-

tation; sample I was oriented along the [001] crystallographic

direction, whereas sample II was oriented along the [111]

direction. While their sizes do differ slightly, this discrepancy

does not affect our results, as the difference in size between

these domains is small relative to that between those which

were checked for finite size effects (see Sec. III A). The

Kapitza conductances at the interfaces contained within sam-

ples I and II are plotted as a function of temperature in Fig.

4(a). Our results support the conclusion of previous experi-

mental work that the Kapitza conductance at interfaces

between two cubic crystals does not vary with orientation.8,31

Unique to this data set is the fact that materials A and B are

lattice matched, and thus, the interfaces were atomically per-

fect. In reality, dislocation densities can vary with orientation

in systems that are not lattice matched. Still, prior experimen-

tal data at TiN/MgO (Ref. 8) and Al/Si (Ref. 31) interfaces ex-

hibit the same orientational independence observed here.

Though dislocation densities at those interfaces did likely

vary with orientation, they did not vary substantially enough

to affect Kapitza conductance (Ref. 11 suggests it can take

several order-of-magnitude changes in dislocation density to

vary Kapitza conductance by a factor of two).

Samples III through VI contained interfaces between

a cubic material oriented along the [001] direction (either

A or B) and the AB alloy oriented in the in-plane or

FIG. 3. Thermal conductivity of materials A, B, and the AB alloy as a func-

tion of temperature. Each data point represents the mean result of three inde-

pendent simulations. Dashed lines are inverse temperature fits of the NEMD

data; the fit for the AB in-plane data has been omitted for clarity. For mate-

rial A, open circles represent the calculated thermal conductivity of the

shorter domain and filled circles represent that of the longer domain; the

close agreement between the data sets suggests that size effects are not

affecting our results.

TABLE II. The pre-equilibrated dimensions, constituent materials, and crys-

tallographic orientations of the computational domains employed in the

Kapitza conductance simulations. In all cases, the cold bath was at the end

of material 1 and the hot bath was at the end of material 2, with the excep-

tion of the hollow symbols plotted in Fig. 4(b), for which the heat flux direc-

tion was reversed.

Sample Dimensions Material 1 Material 2

I 12:4� 12:4� 124 fcc A (001) fcc B (001)

II 12:0� 13:3� 123 fcc A (111) fcc B (111)

III 12:4� 12:4� 124 fcc A (001) L10 AB (ip)

IV 12:4� 12:4� 124 fcc A (001) L10 AB (cp)

V 12:4� 12:4� 124 fcc B (001) L10 AB (ip)

VI 12:4� 12:4� 124 fcc B (001) L10 AB (cp)

093515-4 Duda et al. J. Appl. Phys. 112, 093515 (2012)

Downloaded 28 Dec 2012 to 137.54.10.198. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions

Page 6: Influence of crystallographic orientation and anisotropy ... · materials A and B, with peaks centered at x 8:5 and 20. The phonon dispersion relations in selected directions of high

cross-plane direction; again, all interfaces were atomically

perfect. Schematics representing these domains are illustrated

in Fig. 5. From the data plotted in Figs. 4(b) and 4(c), it is

clear that Kapitza conductance is strongly dependent on the

relative orientation of the AB alloy. At interfaces between

material A and the AB alloy, Kapitza conductance varies by

upwards of a factor of five depending on AB alloy orienta-

tion, whereas at interfaces between B and the AB alloy, it

varies by a factor of two. To further elucidate this point, we

define a thermal anisotropy factor, c, as

c ¼ jip

jcpor

hK;ip

hK;cp; (3)

where the subscripts ip and cp refer to the orientation of the

AB alloy. In Fig. 6, both the conductivity of the AB alloy

and conductance of the B:AB interfaces exhibit nearly iden-

tical anisotropy factors; c � 2 across the entire temperature

range. On the other hand, the anisotropy factor is not only

much higher for the A:AB interfaces, but it also exhibits

greater temperature dependence.

IV. DISCUSSION

As detailed in the Introduction, prior works examining

the influence of crystallographic orientation and anisotropy on

Kapitza conductance have ascribed any observed dependence

on orientation to directionally dependent group velocities.

While this argument is viable in the case of materials exhibit-

ing large degrees of vibrational and thermal anisotropies, e.g.,

graphite or carbon nanotubes, the lattice dynamics calcula-

tions presented in Sec. II indicate that the phonon group veloc-

ities of the L10 AB alloy do not differ substantially with

crystallographic direction. In addition, directionally dependent

group velocities cannot explain the significant difference in

anisotropy between the A:AB and B:AB interfaces. In order

to investigate this phenomenon further, we calculate the local

phonon density of states and frequency-dependent phonon

transmissivity at interfaces corresponding to those contained

within samples III through VI.

A. Local phonon density of states

We calculated the local phonon DOS of the bilayer of

atoms within the AB alloy immediately adjacent to the inter-

face in samples III through VI. The DOS is proportional to

the Fourier transform (F) of the velocity correlation func-

tion (VACF) but in practice is calculated using standard

FIG. 4. Kapitza conductance plotted as a function of temperature. Each data

point represents the mean result of five independent simulations (with the

exception of the hollow symbols in (b) which each indicate the result of a

single simulation in which the heat flux direction was reversed) and error

bars represent the standard deviation of this average, i.e., repeatability.

Dashed lines in (b) and (c) are linear temperature-fits of the NEMD data

(distorted due to semi-log axes).

FIG. 5. Schematics of samples III through VI, where material A is blue and

material B is red. Samples III and V have the heat flow aligned with the in-

plane direction of the AB alloy, whereas samples IV and VI have it aligned

with the cross-plane direction.

FIG. 6. The anisotropy ratio defined by Eq. (3) plotted as a function of tem-

perature. While the anisotropy factor of the B:AB interfaces closely follows

that of the conductivity of the AB alloy, the anisotropy factor of the A:AB

interfaces is significantly higher while also exhibiting greater temperature

dependence.

093515-5 Duda et al. J. Appl. Phys. 112, 093515 (2012)

Downloaded 28 Dec 2012 to 137.54.10.198. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions

Page 7: Influence of crystallographic orientation and anisotropy ... · materials A and B, with peaks centered at x 8:5 and 20. The phonon dispersion relations in selected directions of high

estimation procedures for power spectral density.37 Within

each bilayer of interest, the velocity of 32 atoms is obtained

at each integration time step to give a velocity fluctuation

time series of 73 728 points. The Welch method of power

spectral density estimation is then applied by creating eight

50% overlapping segments of 16 384 points to give an angu-

lar frequency resolution of 0.19 based on our time step of

0.002. Each segment is then multiplied by a Hamming win-

dow and the fast Fourier transform is computed. The power

spectral density, proportional to DOS and equivalent to

FðVACFÞ, is then obtained by ensemble averaging the Fou-

rier transform magnitudes of each segment.

From Fig. 2, the optical-type phonons in the AB alloy

are “elastically inaccessible” to the acoustic phonons of ma-

terial B; i.e., no acoustic phonons within material B are at

frequencies as high as those of the optical phonons in the AB

alloy. On the other hand, the optical phonons in the AB alloy

are “elastically accessible” to the acoustic phonons of mate-

rial A. In Fig. 7, the calculated local phonon DOS of the

bilayer of atoms within the AB alloy immediately adjacent

to the interface in samples III through VI at T¼ 0.09 are

plotted (this low simulated temperature ensures distinct spec-

tral features are observable). Overall, none of the curves ex-

hibit an overwhelming deviation from that of the bulk (and

harmonic) phonon DOS plotted in the top panel of Fig. 2, de-

spite the fact that the sampled atoms used to generate these

curves were immediately adjacent to an interface. However,

in sample IV, the frequency range 18 < x < 21 does exhibit

a degree of mode depletion. While this could be expected

due to the overlap in vibrational spectra of materials A and

the AB alloy, it is interesting that these modes are only

depleted for one orientation of the interface. This unique

characteristic of the interface within sample IV suggests that

interactions between the acoustic phonons of material A and

the optical phonons of the AB alloy lead to increases in the

thermal anisotropy of A:AB interfaces beyond that of B:AB

interfaces, or the AB alloy itself.

B. Wave-packet simulations

To investigate the frequency dependence of phonon

transmissivity, we use the multiple wave-packet method.43

For a single wave-packet,44 linear combinations of normal

modes with different wavevectors k are used as initial condi-

tions for a molecular dynamics simulation. The normal mode

amplitudes for wave-packet i are chosen to be a Gaussian

centered on a given central wavevector ki0, yielding a wave-

packet localized in real space. Moreover, the phases of the

modes are chosen such that the packet is initially located at a

specified position Ri0 in real space.

In this work, the propagation direction is along the

z-axis and localization is accomplished by mixing modes

with different kz. Specifically, multiple wave-packets are

superposed with packet i containing wavevectors k ¼ ki0

þð00dkzÞ with normal mode amplitudes nikk given by

nikk ¼ Aie�g2

0dk2

z e�iðki0þdkzzÞ�Ri

0 : (4)

Every wave-packet from phonon branch k has the same

energy so that jAij2 varies inversely with xkðki0Þ

2and Ri

0 is

chosen so that it begins in the middle of the fcc half of the

bicrystal. Allowed wavevectors are reciprocal to the simula-

tion box for the system and the normal modes are from the

longitudinal acoustic (LA) branch. Wave-packets propagate

along the positive z-direction, reaching the interface to either

be transmitted to the AB alloy or reflected into the fcc mate-

rial. By calculating time-dependent normal mode amplitudes

at particular simulation times, the phonon transmissivity is

obtained. We define this transmissivity to be the ratio of the

energy transmitted through the interface in a given range of

frequency to the energy incident on the interface in the same

frequency range.

These phonon transmissivities at the four interfaces of in-

terest are plotted as a function of angular frequency in Fig. 8.

As is evident from the figure, phonon transmissivity is largely

insensitive to the relative orientation of the AB alloy at fre-

quencies below x ¼ 12. On the other hand, phonon transmis-

sivity at A:AB interfaces is strongly dependent on the relative

orientation of the AB alloy at frequencies in the range

17 < x < 25, i.e., nearly the same frequency range noted dur-

ing the discussion of local phonon density of states in Sec.

IV A. Again, this is the frequency range in which acoustic pho-

nons within material A are at the same frequencies of optical

phonons within the AB alloy. Finally, we note that transmis-

sivity at B:AB interfaces is marginally higher than at A:AB

interfaces at frequencies below x ¼ 12, consistent with the

earlier observation that material B and the AB alloy exhibit

better spectral overlap across that same frequency range.

These results, in conjunction with the Kapitza conduct-

ance data presented in Sec. III B, demonstrate that the

observed thermal anisotropy at B:AB interfaces is due almost

entirely to that inherent to the AB alloy, whereas the anisot-

ropy at A:AB interfaces is enhanced beyond this baseline (i.e.,

FIG. 7. The local phonon density of states at T¼ 0.09 within the bilayer of

atoms in AB alloy immediately adjacent to the interface in samples (a) III

and IV and (b) V and VI; III and V (both ip) are represented by shaded

regions, whereas IV and VI (both cp) are represented by solid lines. The fre-

quency range 18 < x < 21 in sample IV exhibits significant mode depletion

(as indicated by the bold arrow) as compared to all other samples.

093515-6 Duda et al. J. Appl. Phys. 112, 093515 (2012)

Downloaded 28 Dec 2012 to 137.54.10.198. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions

Page 8: Influence of crystallographic orientation and anisotropy ... · materials A and B, with peaks centered at x 8:5 and 20. The phonon dispersion relations in selected directions of high

c > 2) due to the dependence of interfacial phonon-phonon

interactions on the relative orientation of the AB alloy. In other

words, the anisotropy factor defined by Eq. (3) is nearly identi-

cal for B:AB Kapitza conductance and AB alloy conductivity

(see Fig. 6), despite the fact that phonon transmissivity at

B:AB interfaces is independent of the relative orientation of

the AB alloy; in turn, the thermal anisotropy of the B:AB inter-

faces can be ascribed to the thermal anisotropy inherent to the

AB alloy. In contrast, the thermal anisotropy of the A:AB

interfaces is enhanced beyond that inherent to the AB alloy

due to the dependence of phonon transmissivity at A:AB inter-

faces on the orientation of the AB alloy.

V. CONCLUSION

We have studied the influence of crystallographic orien-

tation and anisotropy on Kapitza conductance, local phonon

density of states, and phonon transmissivity via classical mo-

lecular dynamics simulations. Non-equilibrium molecular

dynamics simulations have demonstrated that Kapitza con-

ductance at interfaces between fcc and L10 materials

depends on both the relative orientation of the L10 material

and the degree of vibrational overlap between the two mate-

rials (but not on phonon cutoff frequencies). Local phonon

density of states and phonon transmissivity calculations indi-

cate that interactions between acoustic phonons in the fcc

material and optical phonons in the L10 material can greatly

enhance thermal anisotropy. Furthermore, interfaces may

exhibit higher degrees of thermal anisotropy (i.e., a larger

dependence on crystallographic orientation) than their con-

stituent materials. Despite the classical nature of these simu-

lations, the results presented provide qualitative insight into

the mechanisms responsible for prior experimental observa-

tions that Kapitza conductance is critically dependent on

crystallographic orientation. Finally, we would like to note

that the relative match or mismatch between the cutoff fre-

quencies or Debye temperatures of two materials comprising

an interface does not serve an accurate gauge of the effi-

ciency of interfacial thermal transport when those materials

have different crystal structures.

ACKNOWLEDGMENTS

The authors acknowledge funding through the LDRD

program office at Sandia National Laboratories. Sandia

National Laboratories is a multi-program laboratory man-

aged and operated by Sandia Corporation, a wholly owned

subsidiary of Lockheed Martin Corporation, for the United

States Department of Energy’s National Nuclear Security

Administration under Contract No. DE-AC04-94AL85000.

1E. Pop, Nano Res. 3, 147 (2010).2P. L. Kapitza, Phys. Rev. 60, 354 (1941).3R. J. Stoner and H. J. Maris, Phys. Rev. B 48, 16373 (1993).4R. J. Stevens, A. N. Smith, and P. M. Norris, J. Heat Transfer 127, 315 (2005).5R. J. Stevens, L. V. Zhigilei, and P. M. Norris, Int. J. Heat Mass Transfer

50, 3977 (2007).6P. E. Hopkins, P. M. Norris, and R. J. Stevens, J. Heat Transfer 130,

022401 (2008).7P. M. Norris and P. E. Hopkins, J. Heat Transfer 131, 043207 (2009).8R. M. Costescu, M. A. Wall, and D. G. Cahill, Phys. Rev. B 67, 054302

(2003).9P. E. Hopkins, P. M. Norris, R. J. Stevens, T. E. Beechem, and S. Graham,

J. Heat Transfer 130, 062402 (2008).10Y. Xu, R. Kato, and M. Goto, J. Appl. Phys. 108, 104317 (2010).11P. E. Hopkins, J. C. Duda, S. P. Clark, C. P. Hains, T. J. Rotter, L. M.

Phinney, and G. Balakrishnan, Appl. Phys. Lett. 98, 161913 (2011).12P. E. Hopkins, J. C. Duda, C. W. Petz, and J. A. Floro, Phys. Rev. B 84,

035438 (2011).13T. S. English, J. C. Duda, J. L. Smoyer, D. A. Jordan, P. M. Norris, and L.

V. Zhigilei, Phys. Rev. B 85, 035438 (2012).14J. C. Duda and P. E. Hopkins, Appl. Phys. Lett. 100, 111602 (2012).15M. Hu, P. Keblinski, and P. K. Schelling, Phys. Rev. B 79, 104305

(2009).16K. C. Collins, S. Chen, and G. Chen, Appl. Phys. Lett. 97, 083102 (2010).17Z.-Y. Ong and E. Pop, Phys. Rev. B 81, 155408 (2010).18Y. Wang and P. Keblinski, Appl. Phys. Lett. 99, 073112 (2011).19J. C. Duda, T. S. English, E. S. Piekos, W. A. Soffa, L. V. Zhigilei, and P.

E. Hopkins, Phys. Rev. B 84, 193301 (2011).20M. Shen, W. J. Evans, D. Cahill, and P. Keblinski, Phys. Rev. B 84,

195432 (2011).21P. E. Hopkins, M. Baraket, E. V. Barnat, T. E. Beechem, S. P. Kearney, J.

C. Duda, J. T. Robinson, and S. G. Walton, Nano Lett. 12, 590 (2012).22G. Chen, J. Heat Transfer 119, 220 (1997).23G. Chen, Phys. Rev. B 57, 14958 (1998).24J. E. Turney, A. J. H. McGaughey, and C. H. Amon, J. Appl. Phys. 107,

024317 (2010).25D. P. Sellan, J. E. Turney, A. J. H. McGaughey, and C. H. Amon, J. Appl.

Phys. 108, 113524 (2010).26A. J. H. McGaughey, E. S. Landry, D. P. Sellan, and C. H. Amon, Appl.

Phys. Lett. 99, 131904 (2011).27N. Mingo, Phys. Rev. B 68, 113308 (2003).28R. Prasher, Phys. Rev. B 77, 075424 (2008).29J. C. Duda, J. L. Smoyer, P. M. Norris, and P. E. Hopkins, Appl. Phys.

Lett. 95, 031912 (2009).30J. Hirotani, T. Ikuta, T. Nishiyama, and K. Takahashi, Nanotechnology 22,

315702 (2011).31P. E. Hopkins, T. Beechem, J. C. Duda, K. Hattar, J. F. Ihlefeld, M. A.

Rodriguez, and E. S. Piekos, Phys. Rev. B 84, 125408 (2011).

FIG. 8. Phonon transmissivities at (a) A:AB interfaces and (b) B:AB interfa-

ces plotted as a function of angular frequency. Both sets of data show little

dependence of phonon transmissivity on the relative orientation of the AB

alloy at frequencies below x ¼ 12. However, at A:AB interfaces, phonon

transmissivity is strongly dependent on AB alloy orientation at frequencies

in the range 17 < x < 25. Note that the cutoff frequency for material B is

xc � 15 at which point we expect the B:AB(ip) transmissivity to go to zero.

093515-7 Duda et al. J. Appl. Phys. 112, 093515 (2012)

Downloaded 28 Dec 2012 to 137.54.10.198. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions

Page 9: Influence of crystallographic orientation and anisotropy ... · materials A and B, with peaks centered at x 8:5 and 20. The phonon dispersion relations in selected directions of high

32D. V. Matyushov and R. Schmid, J. Chem. Phys. 104, 8627 (1996).33J. D. Gale and A. L. Rohl, Mol. Simul. 29, 291 (2003).34E. S. Landry, M. I. Hussein, and A. J. H. McGaughey, Phys. Rev. B 77,

184302 (2008).35E. S. Landry and A. J. H. McGaughey, Phys. Rev. B 79, 075316

(2009).36J. C. Duda, T. S. English, D. A. Jordan, P. M. Norris, and W. A. Soffa,

J. Phys.: Condens. Matter 23, 205401 (2011).37M. P. Allen and D. J. Tildesley, Computer Simulation of Liquids (Claren-

don, Oxford, 1990).

38H. J. C. Berendsen, J. P. M. Postma, W. F. van Gunsteren, A. DiNola, and

J. R. Haak, J. Chem. Phys. 81, 3684 (1984).39D. S. Ivanov and L. V. Zhigilei, Phys. Rev. B 68, 064114 (2003).40A. Ward and D. A. Broido, Phys. Rev. B 81, 085205 (2010).41S. Y. Ren and J. D. Dow, Phys. Rev. B 25, 3750 (1982).42M. V. Simkin and G. D. Mahan, Phys. Rev. Lett. 84, 927 (2000).43C. Kimmer, S. Aubry, A. Skye, and P. K. Schelling, Phys. Rev. B 75,

144105 (2007).44P. K. Schelling, S. R. Phillpot, and P. Keblinski, Appl. Phys. Lett. 80,

2484 (2002).

093515-8 Duda et al. J. Appl. Phys. 112, 093515 (2012)

Downloaded 28 Dec 2012 to 137.54.10.198. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions