You are on page 1of 10

PHYSICAL REVIEW B 69, 195316 共2004兲

Thermal conductivity modeling of periodic two-dimensional nanocomposites


Ronggui Yang and Gang Chen*
Mechanical Engineering Department, Massachusetts Institute Technology, Cambridge, Massachusetts 02139, USA
共Received 22 December 2003; revised manuscript received 11 March 2004; published 27 May 2004兲
In this paper, the phonon Boltzmann equation model is established to study the phonon thermal conductivity
of nanocomposites with nanowires embedded in a host semiconductor material. Special attention has been paid
to cell–cell interaction using periodic boundary conditions. The simulation shows that the temperature profiles
in nanocomposites are very different from those in conventional composites due to ballistic phonon transport
at nanoscale. Such temperature profiles cannot be captured by existing models in literature. The general
approach is applied to study silicon wire/germanium matrix nanocomposites. We predict the thermal conduc-
tivity dependence on the size of the nanowires and the volumetric fraction of the constituent materials. At
constant volumetric fraction the smaller the wire diameter, the smaller is the thermal conductivity of periodic
two-dimensional nanocomposites. For fixed silicon wire dimension, the lower the atomic percentage of ger-
manium, the lower the thermal conductivity of the nanocomposites. Comparison is also made with the thermal
conductivity of superlattices. The results of this study can be used to direct the development of high efficiency
thermoelectric materials.

DOI: 10.1103/PhysRevB.69.195316 PACS number共s兲: 68.65.⫺k, 44.10.⫹i, 63.22.⫹m

I. INTRODUCTION succeeded in fitting experimental values of several superlat-


tice systems in the thick period range. However, because the
The efficiency and energy density of thermoelectric de- wave nature of phonons is ignored, these models fail to ex-
vices are determined by the dimensionless thermoelectric plain the thermal conductivity recovery in the short period
figure of merit ZT⫽S 2 ␴ T/k, where S is the Seebeck coeffi- limit that is observed in some superlattices. More recently,
cient, ␴ is the electrical conductivity, k is the thermal con- models9,18 and direct simulations19 combining the above
ductivity, and T is the absolute temperature.1 Significant ad- two pictures have been presented. These models and simula-
vances for increasing ZT have been made based on new ideas tions further confirm the importance of diffuse interface
to engineer electron and phonon transport.2 One particularly scattering for thermal conductivity reduction. The diffuse in-
fruitful and exciting approach has been the use of nanostruc- terface scattering cannot only reduce the phonon mean
tures, so that the electron performance can be improved or free path 共MFP兲 but also destroy the coherence of phonons.
maintained concurrently with a reduction of phonon thermal Due to the loss of coherence, the phonon dispersion change
conductivity.3–5 Nanostructure-based materials such as in nanostructures predicted by ideal lattice dynamics
Bi2 Te3 /Sb2 Te3 superlattices and PbTe/PbSeTe quantum dot approximation cannot be realized. Previous studies on the
superlattices have shown significant increases in ZT values thermal conductivity of superlattices demonstrated that clas-
compared to their bulk counterparts6,7 due to mainly reduced sical size effect models are expected to be applicable to a
phonon thermal conductivity of these structures. Nanocom- wide range of nanostructures. This has been further
posites may realize a similar thermal conductivity reduction confirmed by a recent study on developing the classical size
and provide a pathway to scale-up the nanoscale effects ob- effect model for phonon transport in nanowires and superlat-
served in superlattices to thermoelectric material in bulk tice nanowires.20 The model, assuming gray and diffuse
form. phonon scattering at interfaces and side walls, has success-
Most of the previous studies on thermal transport in nano- fully explained the thermal conductivity reduction effect
structures have focused on thin films, semiconductor super- measured from nanowires. Based on previous studies, we
lattices, and nanowires. One key question in modeling of the will apply in this work the phonon Boltzmann equation to
thermal conductivity in nanostructures is when the wave ef- study the classical size effect on the thermal conductivity of
fect, i.e., the phonon dispersion change, should be nanocomposites.
considered.8,9 For example, models for phonon transport in Another field of related research is effective thermal con-
superlattices generally fall into two groups. One group ductivity of composites. The effective thermal conductivity
共‘‘wave models’’兲 assumes that phonons form superlattice of composites in macroscale has been studied since Maxwell
bands and calculates the modified phonon dispersion using 共for a review, see Milton21兲 and a variety of methods have
lattice dynamics or other methods.10–13 The other group of been proposed to estimate physical properties of heteroge-
models 共‘‘particle models’’兲 assumes that the major reason neous media. In most of these research works, the interfaces
for the thermal conductivity reduction is the sequential scat- between two heteromedia are treated as nonresistive to heat
tering of phonons at interfaces.14 –17 In these models, phonon flow. The interface thermal resistance, or Kapitza
transport falls in the totally incoherent regime and superlat- resistance,22 has been considered only recently. The first two
tices are treated as inhomogeneous multilayer structures. Dif- theoretical analyses that included the interface thermal resis-
fuse interface scattering is usually incorporated into the Bolt- tance were conducted by Hasselman and Johnson23 and by
zmann equation 共BE兲 as boundary conditions. These models Benvensite,24 respectively. Hasselman and Johnson extended

0163-1829/2004/69共19兲/195316共10兲/$22.50 69 195316-1 ©2004 The American Physical Society


RONGGUI YANG AND GANG CHEN PHYSICAL REVIEW B 69, 195316 共2004兲

the classical work of Maxwell and Rayleigh to derive nanocomposites and serve as a prototype to be further devel-
Maxwell–Garnett-type effective medium approximation oped to characterize more complex nanocomposites, such as
共EMA兲 for calculating effective thermal conductivity of sim- nanoparticle composites. Some fundamental questions,
ply spherical particulate and cylindrical fiber reinforced ma- such as 共a兲 how is phonon transport in nanocomposites
trix composites in which interface effect and particle size different from that in macroscale composites, 共b兲 what is the
effect are included. Benvensite and Miloh24 developed a gen- behavior of the size effect in nanowire-embeded composites,
eral framework incorporating thermal boundary resistance by and 共c兲 can nanocomposites have lower effective thermal
averaging all pertinent variables like heat flux and intensity conductivity than a superlattice, will be addressed in the
over the composite medium viewed as a continuum consist- investigation.
ing of a matrix with inclusions. Every et al.25 refined the
effective medium theory and presented an asymmetric
II. THEORETICAL MODEL
Bruggeman type EMA, corrected for Kapitza resistance, as a
simple solution for high volumetric fraction of inclusion, Based on the reasoning that the phonon particle model
based on Bruggeman’s integration-embedding principle. A can be a good predictive tool for thermal conductivity of
noticeable work was by Nan et al.,26 who adopted the mul- nanocomposites, we focus our work on the phonon transport
tiple scattering theory27 to develop a more general EMA in nanowire-embedded composites for the case where the
formulation for the effective thermal conductivity of arbi- heat-flow direction is perpendicular to the wire axis. As
trary particulate composites with interfacial thermal resis- shown in Fig. 1共a兲, there is no heat flow along the wire axis
tance. They considered the properties of the matrix and rein- thus the problem is simplified to two-dimensional although
forcement, particle size and size distribution, volume the nature of phonon transport is three-dimensional in nature
fraction, interface resistance, and the effect of shape. Other as shown in Fig. 1共c兲. The unit cell to be simulated is shown
models have also been proposed including the bounding in Fig. 1共b兲. The detail about the interface and boundary
model28,29 and thermal resistance network theory. Numerical condition will be presented in a later section. To make a
simulation of thermal conductivity of composites can comparison, we also calculated the cross-plane thermal con-
also be found in literature.30–33 However, these macroscopic ductivity of a simple one-dimensional Si–Ge layered struc-
models are developed based on Fourier heat conduction ture, which is often called superlattices when the thickness of
theory that is not valid at nanoscale due to the ballistic each layer is tens of nanometers, as shown in Fig. 1共d兲. To
phonon transport. establish the phonon particle model the following assump-
So far, there are not many theoretical studies on the ther- tions are made: 共1兲 the phonon wave effect can be excluded;
mal conductivity of nanocomposites despite their importance 共2兲 the frequency-dependent scattering rate in the bulk me-
in both thermoelectrics and thermal management of electron- dium is approximate by an average phonon MFP; and 共3兲 the
ics 共especially the development of thermal interface materi- interface scattering is diffuse.
als兲. Closely related works are done by Khitun et al.34 and
Balandin and Lazarenkova35 to explain the ZT enhancement
of Ge quantum dot structures 共where Ge quantum dots are A. Phonon Boltzmann equation
⬃4 nm and can be thought of as nanoparticles兲. The model With the above assumptions, the phonon BE can be
by Khitun et al. calculates the reduced thermal conductivity expressed in terms of the total phonon intensity defined
through the relaxation time change due to the nanoparticles as38,39
embedded using the Mathiessen rule.36 They used the Mie
theory for acoustic wave scattering to calculate the scattering
cross section of a single particle and thus the additional re-
laxation time due to single particle scattering. Their method I i⫽
1
4␲ 兺m 冕
0
␻ max
兩 ␯ mi 兩 f ប ␻ D mi 共 ␻ 兲 d ␻ , 共1兲
is valid if there is no inelastic scattering inside the particle
and the interface scattering must be specular.37 This approach
does not recover to bulk material properties of composites where subscript i 共⫽1,2兲 denotes properties of the host and
since the thermal conductivity of nanoparticle material does wire material 共or alternative material of a 1D layered
not get into the final picture. Similar to the wave model of structure, in this work i⫽1 is for germanium and i⫽2
superlattices, Balandin and Lazarenkova35 assumed that a for silicon兲, D( ␻ ) is the density of states per unit volume, f
new homogeneous material is formed and calculates the new the phonon distribution function, ប the Planck constant, 兩 v mi 兩
electron and phonon dispersion relation thus the group ve- the magnitude of the phonon group velocity, ␻ the phonon
locity reduction. This approach requires phonon coherence frequency, and ␻ max is the maximum frequency of each
over several unit cells and does not apply to diffuse interface polarization. The summation index m is over the three
scattering as observed in previous studies on superlattices. phonon polarizations. The purpose of introducing the phonon
The motivation of this work is to develop a microscopic intensity concept38 is for the mathematical convenience be-
framework for thermal conductivity of nanocomposites in cause the concept of photon intensity is widely used in ther-
terms of the phonon particle transport model. For simplicity, mal radiation literature and many results in radiative transfer
we will study the thermal conductivity of periodic two- might be used for phonon transport studies.40,41
dimensional 共2D兲 nanocomposites. The model will help to The 2D phonon BE under the single mode relaxation time
understand some basic physical phenomena that exist in approximation can be written as

195316-2
THERMAL CONDUCTIVITY MODELING OF PERIODIC . . . PHYSICAL REVIEW B 69, 195316 共2004兲

FIG. 1. 共a兲 Heat flow across periodic 2D composite with silicon wires embedded in germanium host, 共b兲 the unit cell to be simulated, 共c兲
local coordinate used in phonon Boltzmann equation simulation, and 共d兲 heat flow across 1D Si–Ge layered structure 共superlattice兲.

⳵Ii ⳵Ii I i ⫺I oi group velocity, and thus account for interactions among
sin ␪ i cos ␾ i ⫹cos ␪ i ⫽⫺ , 共2兲 phonons of different frequencies. This will relax the assump-
⳵y ⳵x ⌳i
tion 共2兲. However, it requires solution of the phonon BE for
where ␪ and ␾, as shown in Fig. 1共c兲, are the polar and many different frequencies. Previous works show that an av-
the azimuthal angles, respectively, and ⌳ is the average erage MFP is a good approximation for thermal conductivity
phonon MFP. I oi is determined by the Bose-Einstein distri- modeling across the interfaces 共cross-plane transport兲.16 For
bution of phonons and depends on the local equilibrium transport parallel to the interface, frequency dependent relax-
temperature. In nanostructures, however, local equilibrium ation time gives a better solution.14 However, the existing
cannot be established and thus the temperature obtained theories for the frequency dependence of relaxation time
should not be treated the same as the case of local thermal contain large uncertainties because they are based on many
equilibrium. An energy balance shows that I oi can be calcu- approximations and rely on the fitting parameters from ex-
lated from16,42 perimental data.43 Therefore we will use a frequency inde-
pendent phonon MFP for simplicity. Often the phonon MFP
I oi 共 x,y 兲 ⫽
1
4␲
冕 4␲
I i 共 r,⍀兲 d⍀
⌳ is estimated from the thermal conductivity, the specific
heat, and the speed of sound, according to the standard ki-
netic theory expression,


4
1

冕 冕
2␲

0 0

I i 共 x,y, ␪ , ␾ 兲 sin ␪ d ␪ d ␾ ) 共3兲 k⫽ 13 C v ⌳, 共4兲

and the corresponding temperature obtained is a measure of where C is the volumetric specific heat. This method, how-
the local energy density. We note the equations for the 1D ever, neglects the facts that at room temperature most acous-
layered structure are much simpler because the intensity does tic phonons are populated close to the zone boundary where
not depend on the ␾ direction and y position. the phonon group velocity is significantly smaller than the
A rigorous phonon transport simulation should incorpo- sound velocity and that the optical phonons contribute a sig-
rate the frequency dependence of phonon relaxation time and nificant portion of specific heat but little to thermal conduc-

195316-3
RONGGUI YANG AND GANG CHEN PHYSICAL REVIEW B 69, 195316 共2004兲

TABLE I. Room-temperature parameters used in the calculation.

C 共J/m3 K兲 v 共m/s兲 k 共W/m K兲 a 共nm兲

Silicon 0.93⫻106 1804 150 0.5431


Germanium 0.87⫻106 1042 60 0.5658

tivity due to their near zero group velocity. A better estima- transport of the whole structure can be represented by
tion of the phonon MFP and the group velocity can be that in the unit cell. Boundary conditions thus defined require
obtained by approximating the dispersion of the transverse an iterative scheme to solve the equation, which implies
and the longitudinal-acoustic phonons with simple sine func- that it is more complicated than the front-to-end scheme that
tions and subtracting the optical phonon contribution to the is usually adopted to solve the emitted temperature boundary
specific heat C in calculating the phonon MFP. This estima- condition problems. This periodic boundary condition has
tion leads to a mean free path in silicon of the order of been used in studying the cross-plane thermal conductivity
250–300 nm at room temperature. Experiments and model- of superlattices 共which is a 1D structure兲 before.16 It can be
ing from Goodson’s group also lead to similar numbers.44 proven that the temperature difference across the unit cell
The method gave a better fitting of the experimental thermal should be independent of y. That is,
conductivity data on superlattices than that based on bulk
specific heat for C and the speed of sound for v in Eq. C 1 v 1 关 T 共 0,y 兲 ⫺T 共 L Ge ,y 兲兴
I o 共 0,y 兲 ⫺I o 共 L Ge ,y 兲 ⫽ ⫽const.
共4兲.14,16 In our analysis, some parameters used are taken from 4␲
Ref. 16 as listed in Table I. 共8兲
In our simulation, we superimposed T(0,y)⫺T(L Ge ,y)
B. Boundary and interface conditions ⫽1 K in the above equation. If we do not superimpose such
Mathematically, the structure of the BE demands the a temperature difference in the program, Eq. 共7兲 will auto-
boundary conditions to be specified on those parts of bound- matically converge to a constant temperature difference
aries where the characteristics point into the domain.45 This T(0,y)⫺T(L Ge ,y) value. The converged value varies with
boundary condition is termed as emitted temperature at simulated structures. But the final results of thermal conduc-
wall.16 Most of the previous works on phonon transport tivity value do not depend on whether the temperature dif-
simulation used a specified emitted temperature boundary ference is superimposed. However, the calculation is much
condition. This can give an artificial temperature jump at the faster when the temperature difference value is superim-
boundaries due to the additional nonphysical scattering at the posed. We should note that superimposing the temperature
boundaries.46 A more physical boundary condition is difference across the unit cell is physically different from
desirable to simulate phonon transport in realistic structures. superimposing temperature 共either emitted or Fourier-limit
In this work, the periodic boundary condition is used based temperature兲 at each boundary.
on the underlying physics for phonon transport in periodic The interface scattering between the nanowire and the
structures. As shown in Fig. 1共b兲, heat is enforced to host material is assumed to diffuse. Ziman36 proposed the
flow in the x direction. The specular reflected boundary con- following expression for estimating the interface specularity
ditions are enforced at y⫽0 and y⫽L Ge boundaries due to parameter p:

冉 冊
symmetry,
16␲ 3 ␦ 2
I 共 x,L Ge , ␪ , ␾ 兲 ⫽I 共 x,L Ge , ␪ ,⫺ ␾ 兲 , 共5兲 p⫽exp ⫺ , 共9兲
␭2
I 共 x,0,␪ , ␾ 兲 ⫽I 共 x,0,␪ ,⫺ ␾ 兲 . 共6兲 where ␦ is the characteristic interface roughness and ␭ is the
characteristic phonon wavelength. At room temperature, the
In the x direction, we applied periodic boundary conditions. characteristic phonon wavelength ␭⫽h v /k B T is about 1 nm,
At equilibrium, the phonon distribution, i.e., the phonon where h is the Planck constant, k B is the Boltzmann constant,
intensity 关see Eq. 共1兲兴, is isotropic. If heat is enforced to and v is the sound velocity in the material. Clearly even one
flow, the phonon distribution/intensity will be distorted. monolayer roughness ␦ ⬃0.3 nm gives an interface specular-
Physically the periodic boundary condition means that the
ity parameter p⫽0, and allows for totally diffuse interface
distortion of phonon intensity in each direction at each point
assumption. Determining the phonon reflectivity and
(0,y) at the x⫽0 boundary is the same as the distortion in transmissivity at interface is difficult as in the treatment
the corresponding direction at the corresponding point of the classical thermal boundary resistance problem.47– 49
(L Ge ,y) at the x⫽L Ge boundary. The equation can be ex- One rather crude model is called the diffuse mismatch
pressed as model,49 which assumes that phonons emerging from the
I 共 L Ge ,y, ␪ , ␾ 兲 ⫺I 0 共 L Ge ,y 兲 ⫽I 共 0,y, ␪ , ␾ 兲 ⫺I 0 共 0,y 兲 . 共7兲 interface do not really bear any relationship to its origin, i.e.,
one cannot tell which side they come from. This assumption
With this boundary condition, the total surface heat flux is implies that
conserved in the x direction and the cell–cell interaction is
taken into account. Because of the periodicity, the phonon T d21⫽R d12⫽1⫺T d12 , 共10兲

195316-4
THERMAL CONDUCTIVITY MODELING OF PERIODIC . . . PHYSICAL REVIEW B 69, 195316 共2004兲

where the second equation comes from the energy conserva- tive transfer for a set of discrete directions spanning the en-
tion identity R d12⫹T d12⫽1. We remind again that subscript tire solid angle. The integrals over the solid angle are ap-
12 means from medium 1 into medium 2 and vice versa. proximated by numerical quadratures. For phonon transport
Dames and Chen20 obtained the following equation for in nanostructures, the challenge is to reduce the ‘‘ray effect,’’
T d12 , which is expected to be valid over a wide temperature which often happens similarly in thermal radiation in the
range: optical thin limit. In our previous work,46 double Gauss-
Legender quadratures have been used to replace the conven-
U 2v 2 tional S N quadratures for the discrete ordinate method and
T d12共 T 兲 ⫽ , 共11兲
U 1 共 T 兲v 1 ⫹U 2 共 T 兲v 2 demonstrated to successfully resolve the ray effect problem
of phonon transport simulation in nanostructures. We briefly
where U is the volumetric internal energy. discuss the calculation method here. The method separately
With the given interface transmissivity and reflectivity, we discretizes the integrating points in the ␮ ⫽cos ␪ 共the angle
can write down the phonon intensity at interfaces. As an ␪兲 and the angle ␾ using Gauss-Legender quadrature. To ob-
example, the phonon intensity for ␪ ⬎0 at the x⫽(L Ge tain high accuracy, ␮ is discretized into 120 points from ⫺1
⫺L Si)/2, (L Ge⫺L Si)/2⭐y⭐(L Ge⫹L Si)/2 interface can be to 1 and ␾ is discretized into 24 points for 0⬃ ␲ 共not 0
written from the energy balance equation. ⬃2 ␲ due to symmetry兲. Then Eq. 共3兲 can be written in dis-

冕 冕 冉 ␪ ␾冊
crete form as
2␲ ␲ /2 L Ge⫺L Si
I ,y, , cos ␪ •sin ␪ d ␪ d ␾ 2
0 0 2
I 0 共 x,y 兲 ⫽ 兺m 兺n I 共 x,y, ␮ n , ␾ m 兲 w n w m⬘ . 共14兲

冕 冕 冉 冊
4␲
2␲ ␲ L Ge⫺L Si
⫽R d21 I ,y, ␪ , ␾ cos ␪ •sin ␪ d ␪ d ␾ The weights satisfy 兺 m 兺 n w n w m ⬘ ⫽2 ␲ . In order to accurately
0 ␲ /2 2
capture the physics of the transport phenomena and mini-

冕 冕 冉 ⫹T d12
2␲

0 0
␲ /2
I
L Ge⫺L Si
2 冊
,y, ␪ , ␾ cos ␪ •sin ␪ d ␪ d ␾ .
mize the calculation time, a nonuniform grid system is used.
The step scheme is used for spatial discretization. Total spa-
tial grids of 102⫻102 are used in the calculation. Fewer
共12兲 points in both spatial and angular coordinates can be used for
faster calculation. The equation is solved by iteration over
Because the phonons are scattered diffusely at interfaces, the
the value of the equivalent equilibrium intensity I 0 (x,y)
phonons leaving an interface are isotropically distributed,
and Eq. 共11兲 can be written as ⬘ . When the relative error
⫽2/4␲ 兺 m 兺 n I(x,y, ␮ n , ␾ m )w n w m
of the calculated value of the equivalent equilibrium inten-

I 冉 L Ge⫺L Si
2
,y, ␪ , ␾ ⫽
R d21
␲ 冊 冕 冕 冉
0
2␲ ␲

␲ /2
I
L Ge⫺L Si
2
,y, ␪ , ␾ 冊 sity between two iteration steps is less than 2⫻10⫺6 , the
program is assumed to converge and the effective tempera-
ture and heat flux are calculated.
⫻cos ␪ •sin ␪ d ␪ d ␾ Although at nanoscale, temperature cannot be defined as a

冕 冕 冉 冊
measure of equilibrium, we can use the effective temperature
T d12 2␲ ␲ /2 L Ge⫺L Si to reflect the local energy density inside the medium.36,39
⫹ I ,y, ␪ , ␾
␲ 0 0 2 Assuming constant specific heat over a wide temperature
range, we can write the effective temperature as
⫻cos ␪ •sin ␪ d ␪ d ␾ for ␪ ⬎0. 共13兲
4 ␲ I 0 共 x,y 兲 1
Phonons leaving for ␪ ⬍0 and other interfaces can be simi- T 共 x,y 兲 ⫽
C兩v兩

C兩v兩 兺n 兺m 共 x,y, ␮ n , ␾ m 兲 w n w m⬘ .
larly written. 共15兲

C. Method of numerical solution The heat fluxes at every point can be accordingly written
as
Equation 共2兲 is similar to the photon radiative transport
equation 共RTE兲.42 The key is to solve for the intensity distri-
bution I(r,⍀). Once the intensity is found by solving Eq. q x 共 x,y 兲 ⫽ 兺m 兺n I 共 x,y, ␮ n , ␾ m 兲 ␮ n w n w m⬘ , 共16兲
共2兲, the heat flux q x (x,y), q y (x,y) and the effective tempera-
ture T(x,y) can be determined through the numerical inte-
gration over the whole solid angle 4␲. A variety of solution q y 共 x,y 兲 ⫽ 兺m 兺n I 共 x,y, ␮ n , ␾ m 兲 冑1⫺ ␮ 2n cos ␾ m •w n w m⬘ ,
method is available in the thermal radiative transfer
共17兲
literature.40,41 The most often used methods are the discrete
ordinates method 共DOM兲50 and the finite volume method.51 where q x and q y are heat flux in the x and y directions,
The discrete ordinate method is a tool to transform the equa- respectively. After local effective temperature distribution
tion of radiative transfer into a set of simultaneous partial and heat flux are obtained, the thermal conductivity calcula-
differential equations. This is based on a discrete representa- tion is straightforward, taking advantage of the unit cell con-
tion of directional variation of intensity. A solution to the cept. The surface heat flux in the x direction can be calcu-
transport problem is found by solving the equation of radia- lated as

195316-5
RONGGUI YANG AND GANG CHEN PHYSICAL REVIEW B 69, 195316 共2004兲

Q x 共 x 兲 ⫽L Z 冕
0
L Ge
q x 共 x,y 兲 dy, 共18兲
of the interface resistance to the effective thermal resistance
of the composite. When the nanowire dimension is much
smaller than the phonon MFP, say L Si⫽10 nm, the tempera-
where L z is the unit length in the z direction. Equations ture gradient along the x direction can be negative in some
共5兲 and 共6兲, i.e., the specular reflected boundary conditions local regions. The heat fluxes in the x direction, however, are
enforced at y⫽0 and y⫽L Ge boundaries have ensured always positive as shown in Figs. 3共a兲 and 3共b兲. This phe-
Q x (x)⫽const. The average energy density 共average tempera- nomenon was not observed before in macroscale composites
ture兲 at each y-z plane along the x direction can be written and cannot be predicted by Fourier heat conduction based
as theories. Another question that may raise is whether the re-
sults shown in Figs. 2共c兲 and 2共d兲 violate the second law of
T̄ 共 x 兲 ⫽
1
L Ge
冕 L Ge

0
T 共 x,y 兲 dy. 共19兲
thermodynamic because the temperature inside the cell is
larger than the cell boundary. To answer this question, we
should first remind, as pointed out before, that temperature as
Therefore the effective thermal conductivity k of the com- defined is not the same as the case of thermal equilibrium or
posite can be obtained as local thermal equilibrium. When ballistic transport domi-
nates, no local thermal equilibrium can be established and
Qx temperature calculated represents the local energy density.
k⫽ 共20兲 Figures 4共a兲 and 4共b兲 illustrate the mechanisms of the obser-
L z 关 T̄ 共 L Ge兲 ⫺T̄ 共 0 兲兴 vation in Fig. 2. When L Si is much smaller than the phonon
from MFP, the internal scattering in the medium 共both host mate-
rial and the wire兲 is negligible. We further assume that the
关 T̄ 共 L Ge兲 ⫺T̄ 共 0 兲兴 phonon reflectivity is unity at the wire and host material
Q x ⫽k 共 L z L Ge兲 . 共21兲 interface. Then the scenario can be simplified as thermal ra-
L Ge
diation in vacuum with opaque wire inclusions 共host
The following dimensionless parameters have been intro- material–vacuum, interface–solid wall, wire-opaque solid
duced to present temperature and heat flux results: body兲. Referring to Fig. 4共a兲, we are interested in knowing
the temperature distribution of points A – F when the heat is
q Qx x y enforced to flow in the x direction. We can qualitatively call
q *⫽ , x⫽
Q* , x *⫽ , y *⫽ . the left half of the region shown as the ‘‘hot’’ region and the
C 1v 1 L GeL Z C 1 v 1 L Ge L Ge
共22兲 right half as the ‘‘cold’’ region. As shown in the figure, points
D and F ‘‘see’’ the hot region and points A and C ‘‘see’’ the
cold region, thus points D and F locally receive higher en-
III. RESULTS AND DISCUSSIONS
ergy phonons and have higher effective temperature then
A. Nonequilibrium temperature and heat flux distribution points A and C. Moving from point A to F 共or from C to D兲,
Figures 2共a兲 and 2共c兲 show the effective temperature dis- more hot area is seen than cold area, thus the effective tem-
tribution in the composite structures with silicon wire dimen- perature increases. Comparing to D and F, point E has a
sion of L Si⫽268 nm and L Si⫽10 nm, respectively. The lower temperature due to a small view factor from the hot
atomic percentage is 20% for Si and 80% for Ge. Simple region. Similarly point B has a higher effective temperature
calculation gives the geometric ratio L Ge /L Si⫽2.35. The than A and C. Figure 2 shows the temperature distribution
only in one unit cell. To visualize the temperature distribu-
choosing of L Si⫽268 nm is based on the fact that the MFP
tion, one needs to stitch several periods of Fig. 2. Figure 4共b兲
value is around 268 nm calculated from Eq. 共4兲 by using
shows the temperature distribution along x at several y points
the parameters listed in Table I. Figure 2共a兲 is very close to
over three periods. The energy over those regions with even
the temperature we expect in macroscale composites
higher temperature than the unit cell boundary comes
with interface thermal resistance. Therefore, for a wire
from the much higher temperature region in their previous
dimension larger than 268 nm, the effective temperature dis-
cells. The results do not violate the second law of thermo-
tribution is expected to be similar to that plotted in Fig. 2共a兲.
dynamics.
The comparison of Figs. 2共a兲 and 2共c兲 shows that the
temperature or energy density distribution at nanoscale in B. Effect of wire dimension
periodic 2D composites can be very different from that at
macroscale due to the ballistic nature of phonons. To better To calculate the effective thermal conductivity of compos-
understand the effect of interface thermal resistance, we ites, the surface heat flux and the average temperature 共aver-
plot the temperature distribution along the x direction at cer- age energy density兲 at each y-z plane along the x direction is
tain fixed y positions of the two structures in Figs. 2共b兲 and calculated. As examples, Fig. 5 shows the dimensionless av-
2共d兲. Apparent temperature jumps at the wire–host material erage energy density distribution along the x direction in a
interfaces are clearly shown in Figs. 2共b兲 and 2共d兲. There are Si0.2 – Ge0.8 composite with a silicon wire dimension of L Si
also temperature jumps along the y direction as indicated in ⫽268 and 10 nm, respectively. The dimensionless surface
Figs. 2共a兲 and 2共c兲. The larger the wire dimension, the lower heat flux is conserved to Q x* (x)⫽0.088 for an L Si⫽268 nm
the temperature jump relative to the total temperature differ- composite and Q x* (x)⫽0.037 for an L Si⫽10 nm composite.
ence across the interface, and thus the lower the contribution Again, the surface heat flux is conserved and the temperature

195316-6
THERMAL CONDUCTIVITY MODELING OF PERIODIC . . . PHYSICAL REVIEW B 69, 195316 共2004兲

FIG. 2. Effective temperature (T⫺T ref) distribution in the unit cell of Si0.2 – Ge0.8 composites with T(0,y)⫺T(L Ge ,y)⫽1 K applied for
different wire dimensions: 共a兲 temperature contour for L Si⫽268 nm, 共b兲 the temperature distribution along x * at y * ⫽0.5, 0.7, and 0.85 for
L Si⫽268 nm, 共c兲 temperature contour for L Si⫽10 nm, and 共d兲 the temperature distribution along x * at y * ⫽0.5, 0.7, and 0.85 for L Si
⫽10 nm. The temperature discontinuity at the interface is clearly shown. The temperature distribution in L Si⫽10 nm nanocomposite is very
different from macroscale composites due to the ballistic phonon transport at nanoscale and cannot be captured by Fourier heat conduction
theory.

jump appears at the interface. The smaller the wire dimen- shows that at constant volumetric fraction 共or atomic
sion, the larger the average temperature jump and thus the percentage兲 the smaller the characteristic length of silicon
larger is the interface resistance contribution to the effective 共the silicon wire dimension in composites and the thickness
thermal resistance of the composite. We can expect of the silicon layer in superlattices兲, the smaller the thermal
that when the wire dimension is 2 to 3 times or even larger conductivity. The simple 1D layered structure has a lower
than the silicon MFP, the contribution of the interface thermal conductivity than periodic nanowire composites at
thermal resistance will be negligible and the results will this atomic percentage. We point out that the thermal
recover the Fourier limit. Figure 6 shows the thermal conductivity of superlattices calculated here is lower than the
conductivity of Si0.2 – Ge0.8 composites as a function of the experimental data because the interface scattering in super-
silicon wire dimension. To make a comparison, we include lattices may not fall into a totally diffuse scattering limit.14,16
the results of cross-plane 共perpendicular to the interfaces兲 The comparison shown in this paper is just for theoretical
thermal conductivity of a simple one-dimensional Si–Ge consistency.
layered structure 共superlattices兲. Simple calculation shows
C. Effect of atomic percentage
that the thickness of the germanium layer should be 4.52
times the thickness of the silicon layer in 1D stacks with a Some other questions of interests are: 共1兲 can the thermal
germanium atomic percentage of 80%. Figure 6 clearly conductivity of nanowire composites be lower than that of

195316-7
RONGGUI YANG AND GANG CHEN PHYSICAL REVIEW B 69, 195316 共2004兲

FIG. 3. The dimensionless heat flux distribution in the x direction q *x : 共a兲 L Si⫽268 nm composite, and 共b兲 L Si⫽10 nm composite. It
shows that the x-directional heat flux is always positive even in the localized negative temperature gradient region shown in Fig. 2.

simple 1D stacks 共superlattice兲? 共2兲 How the thermal con-


ductivity changes with the atomic percentage? Figure 7
shows the thermal conductivity of Si1⫺x Gex composites as a
function of atomic percentage x of germanium for wire di-
mensions L Si of 50 and 10 nm, respectively. By changing the
atomic percentage, the geometric ratio of the unit cell, i.e.,
the dimension of germanium, is changed in the numerical
simulation. It shows that for a fixed silicon wire dimension,
the lower the atomic percentage of germanium; the lower is
the thermal conductivity of the nanocomposites. This is very
different from macroscale composites and nanoparticle-filled
polymers, in which thermal conductivity of the composites
increases with the decreasing volumetric fraction of the low
thermal conductivity component. This is caused by the bal-
listic transport of phonons in both the host material and the
nanowires, and the interface resistance between the host ma-
terial and the nanowires. In polymer nanocomposites, the
thermal conductivity of the host polymer is usually very low
and the thermal transport in polymers is diffusive. Thus the
thermal conductivity of polymer nanocomposites increases
with the volumetric fraction of high thermal conductivity
nanoparticle fillers. Figure 7 also shows that the thermal con-
ductivity of the periodic 2D nanocomposites is lower than
that of the superlattice with corresponding characteristic
length when the atomic percentage x of germanium is lower
than 35%. Simple calculation shows that the geometric ratio
L Ge /L Si is around 1.182 when x⫽0.35. For a simple 1D
layered structure as shown in Fig. 8共a兲, phonons experience
cross-plane interface scattering in all the cross-sectional area
z-y when the heat is enforced to flow in the x direction.
Comparing Figs. 8共a兲 and 8共b兲, we know that phonons can
flow through a fraction of (L Ge⫺L Si)/L Ge open area without
experiencing cross-plane interface scattering in nanocompos-
ites. However, phonons must experience an additional frac-
tion of L Si /L Ge interface scattering parallel to the heat flow
direction 共in-plane scattering兲. When the thermal conductiv- FIG. 4. 共a兲 Illustration to show the mechanisms of negative tem-
ity of a simple 1D layered structure is the same as that of the perature gradient in the localized regions using thermal radiation
periodic 2D nanocomposites at x⫽0.35, we can approxi- analogy. 共b兲 The temperature distribution along x * at y * ⫽0.5, 0.7,
mately infer that a fraction of L Si /L Ge and 0.85 for L Si⫽10 nm over three periods.

195316-8
THERMAL CONDUCTIVITY MODELING OF PERIODIC . . . PHYSICAL REVIEW B 69, 195316 共2004兲

FIG. 5. The dimensionless average temperature distribution


FIG. 7. Thermal conductivity of Si1⫺x – Gex composites as a
along the x direction in Si0.2 – Ge0.8 composite with silicon wire
function of atomic percentage x of germanium. For fixed silicon
dimension of L Si⫽268 nm and L Si⫽10 nm, respectively.
wire dimension, the lower the atomic percentage of germanium, the
lower is the thermal conductivity of the nanocomposites. The result
in-plane interface scattering is equivalent to a fraction of is very different from the bulk material due to the ballistic nature of
(L Ge⫺L Si)/L Ge cross-interface scattering. In other words, the phonon transport at nanoscale and interface effect.
efficiency of cross-interface scattering to reduce thermal con-
ductivity is around five times as effective as the scattering
embedded in a host semiconductor material using the phonon
parallel to the interface. This result is consistent with
BE. Special attention has been paid to cell–cell interaction
previous experiments and modeling of the in-plane and
using periodic boundary conditions. The simulation
cross-plane thermal conductivity of superlattices.14,16 It also
shows that the temperature profiles in nanocomposites are
suggests that the anisotropic nanocomposites might be
very different from those in conventional composites due
more effective in reducing thermal conductivity of nanocom-
to ballistic phonon transport at nanoscale. Such temperature
posites.
profiles cannot be captured by existing models in literature.
We predict the thermal conductivity dependence on the
IV. CONCLUSIONS interface conditions, the size of the nanowires, and the
volumetric fraction of the constituent materials. The smaller
We studied theoretically the phonon thermal conductivity the wire diameter, the smaller is the thermal conductivity
of periodic two-dimensional nanocomposites with nanowires of periodic two-dimensional nanocomposites. The thermal
conductivity of 2D Si–Ge composites is predicted to be a
function of the atomic percentage of germanium for wire
dimensions of 50 and 10 nm. It shows that for fixed
silicon wire dimension, the lower the atomic percentage of
germanium, the lower the thermal conductivity of the

FIG. 8. Illustration to show that phonons experience less cross-


FIG. 6. Thermal conductivity of Si0.2 – Ge0.8 composites as a interface scattering in periodic 2D composites than that in 1D lay-
function of the silicon wire dimension or layer thickness. The ered structures but additional scattering parallel to the interface. The
smaller the characteristic length of silicon 共the silicon wire dimen- efficiency of cross-interface scattering to reduce thermal conductiv-
sion in composites and the thickness of the silicon layer in super- ity is around five times as effective as scattering parallel to the
lattices兲, the smaller is the thermal conductivity. interface.

195316-9
RONGGUI YANG AND GANG CHEN PHYSICAL REVIEW B 69, 195316 共2004兲

nanocomposites. This is very different from bulk composites face material with high thermal conductivity particle or wire
because the interface dominates the ballistic transport in inclusions.
nanocomposites. The periodic 2D nanocomposites can have
a lower thermal conductivity than their 1D counterpart when ACKNOWLEDGMENTS
the Ge atomic percentage is lower than 35%. This suggests The authors would like to thank Professor M.S. Dressel-
that the anisotropic nanocomposites might be more effective haus, W. L. Liu, and C. Dames for comments and useful
in reducing thermal conductivity of nanocomposites. Results discussions. R.Y. would like to acknowledge the help from
of this study can be used to direct the development of both Yong Li. The work was financially supported by DOE BES
high efficiency thermoelectric materials and thermal inter- 共DE-FG02-02ER45977兲 and NSF 共CTS-0129088兲.

*Electronic address: gchen2@mit.edu Nan, J. Appl. Phys. 76, 1155 共1994兲.


1
H. J. Golsmid, Thermoelectric Refrigeration 共Plenum, New York, 28
P. A. Smith and S. Torquato, J. Appl. Phys. 65, 893 共1989兲; S.
1964兲. Torquato and M. D. Rintoul, Phys. Rev. Lett. 75, 4067 共1995兲.
2
Special Issues on Thermoelectrics, edited by T. M. Tritt, Semi- 29
R. Lipton and B. Vernescu, J. Appl. Phys. 79, 8964 共1996兲.
cond. Semimetals 69–71 共2001兲. 30
See, for example, M. Jiang, I. Jasiuk, and M. Ostoja-Starzewski,
3
M. S. Dresselhaus, Y. M. Lin, S. B. Cronin, O. Rabin, M. R. Comput. Mater. Sci. 25, 329 共2002兲; F. W. Jones and F. Pascal,
Black, G. Dresselhaus, and T. Koga, Semicond. Semimetals 71, Geophysics 60, 1038 共1995兲.
1 共2001兲. 31
S. Graham and D. L. McDowell, ASME J. Heat Transfer 125, 389
4
G. Chen, Semicond. Semimetals 71, 2003 共2001兲. 共2003兲.
5
R. Venkatasubramanian, Semicond. Semimetals 71, 175 共2001兲. 32
K. Ramani and A. Vaidyanathan, J. Compos. Mater. 29, 1725
6
T. C. Harman, P. J. Taylor, M. P. Walsh, and B. E. LaForge, 共1995兲.
Science 297, 2229 共2002兲. 33
M. R. Islam and A. Pramila, J. Compos. Mater. 33, 1699 共1999兲.
7 34
R. Venkatasubramanian, E. Silvona, T. Colpitts, and B. O’Quinn, A. Khitun, A. Balandini, J. L. Liu, and K. L. Wang, J. Appl. Phys.
Nature 共London兲 413, 597 共2001兲. 88, 696 共2000兲; Superlattices Microstruct. 30, 1 共2001兲; A. Khi-
8
G. Chen, D. Borca-Tasciuc, and R. G. Yang, in Encyclopedia of tun, K. L. Wang, and G. Chen, Nanotechnology 11, 327 共2000兲;
Nanoscience and Nanotechnology, edited by H. S. Nalwa J. L. Liu et al., Phys. Rev. B 67, 165333 共2003兲.
共American Scientific Publishers, Stevenson Ranch, CA, 2004兲, 35
O. L. Lazarenkova and A. A. Balandin, J. Appl. Phys. 89, 5509
Vol. 7, N68. 共2001兲; Phys. Rev. B 66, 245319 共2002兲; A. A. Balandin and O.
9
B. Yang and G. Chen, Phys. Rev. B 67, 195311 共2003兲. L. Lazarenkova, Appl. Phys. Lett. 82, 415 共2003兲.
10
P. Hyldgaard and G. D. Mahan, Phys. Rev. B 56, 10754 共1997兲. 36
J. M. Ziman, Electrons and Phonons 共Oxford University Press,
11
S. Tamura, Y. Tanaka, and H. J. Maris, Phys. Rev. B 60, 2627 London, 1985兲.
共1999兲. 37
R. Prasher, Appl. Phys. Lett. 83, 48 共2003兲; ASME J. Heat Trans-
12
W. E. Bies, R. J. Radtke, and H. Ehrenreich, J. Appl. Phys. 88, fer 125, 1156 共2003兲.
1498 共2000兲. 38
A. Majumdar, ASME J. Heat Transfer 115, 7 共1993兲.
13 39
B. Yang and G. Chen, Microscale Thermophys. Eng. 5, 107 G. Chen and C. L. Tien, J. Thermophys. Heat Transfer 7, 311
共2001兲. 共1993兲.
14
G. Chen, ASME J. Heat Transfer 119, 220 共1997兲. 40
R. Siegel and J. R. Howell, Thermal Radiation Heat Transfer, 4th
15
G. Chen and M. Neagu, Appl. Phys. Lett. 71, 2761 共1997兲. ed. 共Taylor & Francis, Washington, D.C., 2001兲.
16
G. Chen, Phys. Rev. B 57, 14958 共1998兲. 41
M. F. Modest, Radiative Heat Transfer, 2nd ed. 共Academic Press,
17
P. Hyldgaard and G. D. Mahan, Therm. Conduct. 23, 172 共1996兲. New York, 2003兲.
18
M. V. Simkin and G. D. Mahan, Phys. Rev. Lett. 84, 927 共2000兲. 42
A. A. Joshi and A. Majumdar, J. Appl. Phys. 74, 31 共1993兲.
19
B. Daly, H. Maris, K. Imamura, and S. Tamura, Phys. Rev. B 66, 43
P. G. Klemens, Solid State Phys. 7, 1 共1958兲.
024301 共2002兲. 44
K. E. Goodson and Y. S. Ju, Annu. Rev. Mater. Sci. 29, 261
20
C. Dames and G. Chen, J. Appl. Phys. 95, 682 共2004兲. 共1999兲.
21
G. W. Milton, The Theory of Composites 共Cambridge University 45
P. Markowich, C. Ringhofer, and C. Schmeiser, Semiconductor
Press, New York, 2002兲. Equations 共Springer, Wien, 1990兲.
22
P. L. Kapitza, J. Phys. 4, 181 共1941兲; for a review on thermal 46
R. G. Yang, G. Chen, M. Laroche, and Y. Taur, ASME J. Heat
boundary resistance, see Ref. 49. Transfer 共to be published兲.
23 47
D. P. H. Hasselman and L. F. Johnson, J. Compos. Mater. 21, 508 R. M. Costescu, M. A. Wall, and D. G. Cahill, Phys. Rev. B 67,
共1987兲. 054302 共2003兲.
24
Y. Benvensite, J. Appl. Phys. 61, 2840 共1987兲; Y. Benvensite and 48
D. G. Cahill, W. K. Ford, K. E. Goodson, G. D. Mahan, A. Ma-
T. Miloh, ibid. 69, 1337 共1991兲. jumdar, H. J. Maris, R. Merlin, and S. R. Phillpot, J. Appl. Phys.
25
A. G. Every, Y. Tzou, D. P. H. Hasselman, and R. Raj, Acta 93, 793 共2003兲.
Metall. Mater. 40, 123 共1992兲. 49
E. T. Swartz and R. O. Pohl, Rev. Mod. Phys. 61, 605 共1989兲.
26
C.-W. Nan, R. Birringer, D. R. Clarke, and H. Gleiter, J. Appl. 50
W. A. Fiveland, ASME J. Heat Transfer 109, 809 共1987兲; J. S.
Phys. 81, 6692 共1997兲. Truelove, ibid. 109, 1048 共1987兲.
27
C.-W. Nan and F. S. Jin, Phys. Rev. B 48, 8578 共1993兲; C.-W. 51
G. D. Raithby, Numer. Heat Transfer, Part B 35, 389 共1994兲.

195316-10

You might also like