+ All Categories
Home > Documents > c 2008 Chia-Yi Chen - University of...

c 2008 Chia-Yi Chen - University of...

Date post: 30-Apr-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
111
THERMAL TRANSPORT IN NANOSTRUCTURED MATERIALS By CHIA-YI CHEN A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2008 1
Transcript

THERMAL TRANSPORT IN NANOSTRUCTURED MATERIALS

By

CHIA-YI CHEN

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOLOF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OFDOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2008

1

c© 2008 Chia-Yi Chen

2

To my Mother and in memory of my Father who sacrificed so much for me.

3

ACKNOWLEDGMENTS

I take this opportunity to state my deep sense of gratitude to my advisor, Dr. Dmitry

I. Kopelevich, for introducing me to field of modeling and numerical simulations. I am

deeply indebted to his patience teaching, guidance and constant encouragement out

of his best interest for me. I am thankful for his effort in advising three PhD students

at the same time, working with us late and being tireless to improve my technical

writing. I would also like to thank Dr. Antony Ladd for his encouragement when I

was his teaching assistant, suggestions for research method and his effort to set up the

computational cluster in our department. I also appreciate the members of my committee,

Dr. Jason Weaver, Dr. Ant Ural for their advice and availability. In addition, I express

my sincere thank Dr. Chauhan and Dr. Tseng for their advise for my future career path;

encouragement and inspiration in both my professional and personal life during the last

stage of my PhD study.

I am grateful to the people I have chance to work with during this 5 years, including

Gunjan Mohan, Ashish Gupta, Benjamin James, Young-Min Ban, Chris Cook and Young-

Nam Ahn. Especially Gunjan Mohan for assisting me on the numerical method and

programming techniques. I would like to express my great appreciation for my roommate,

Han Chang, for her emotional support and friendship in the last year of my study. I thank

Jamie Wang, Pastor Steve Pettit, Ella Pettit, and Patti Buckelew Bryant for their loving

words and faithful prayers at all time.

This acknowledgement would not be complete without Dr. Keesling, Pei-Hsun Wu,

our friendly staff, faculties in Chemical Engineering department and all my friends in

US and Taiwan for making my study abroad experience in Gainesville an unforgeable

experience. Also I acknowledge the funding from University of Florida and computational

resources of UF High Performance Computing Center.

For my family in Taiwan: my Mother, sisters and brother, I am extremely grateful for

their love, understanding, comfort and constant support for me through numerous phone

4

calls during these five years. Last but certainly not least, I thank God for His unfailing

love through Jesus Christ in all things of my life.

5

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

CHAPTER

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2 NUMERICAL METHODS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.1 Molecular Dynamics Simulations . . . . . . . . . . . . . . . . . . . . . . . 182.2 Temperature Coupling in Molecular Dynamics . . . . . . . . . . . . . . . . 18

2.2.1 Berendsen Thermostat . . . . . . . . . . . . . . . . . . . . . . . . . 192.2.2 Nose-Hoover Thermostat . . . . . . . . . . . . . . . . . . . . . . . . 192.2.3 Langevin Thermostat . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.3 Steady State Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212.4 Stability Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3 NONLINEAR LATTICE VIBRATION MODES IN MODEL SYSTEMS . . . . 24

3.1 One-Dimensional FPU system . . . . . . . . . . . . . . . . . . . . . . . . . 243.1.1 Thermal Relaxation Simulation . . . . . . . . . . . . . . . . . . . . 243.1.2 Steady State Solutions . . . . . . . . . . . . . . . . . . . . . . . . . 26

3.1.2.1 FPU-β model . . . . . . . . . . . . . . . . . . . . . . . . . 263.1.2.2 FPU-α model . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.2 Lattice Model Systems In Higher Dimensions . . . . . . . . . . . . . . . . . 303.2.1 Single Chain System in Two Dimensional System . . . . . . . . . . 313.2.2 Two Coupled FPU Chains . . . . . . . . . . . . . . . . . . . . . . . 323.2.3 Body Centered Cubic Structure . . . . . . . . . . . . . . . . . . . . 34

3.3 Hexagonal Tube Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363.3.1 Dispersion Relationship . . . . . . . . . . . . . . . . . . . . . . . . . 383.3.2 Steady State Solutions . . . . . . . . . . . . . . . . . . . . . . . . . 403.3.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

4 NONLINEAR LATTICE VIBRATIONAL MODES IN CARBON NANOTUBES 46

4.1 System Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 474.2 Non-Equilibrium MD Simulation . . . . . . . . . . . . . . . . . . . . . . . 514.3 Steady States Solutions in CNTs . . . . . . . . . . . . . . . . . . . . . . . 514.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

4.4.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

6

5 EFFECTS OF SORBATE MOLECULES ON THERMAL TRANSPORT INZEOLITES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

5.1 Model Details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 655.2 Simulation details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685.3 Normal modes of sodalite crystal . . . . . . . . . . . . . . . . . . . . . . . 705.4 Nonlinear phonon and sorbate dynamics . . . . . . . . . . . . . . . . . . . 735.5 Phonon statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

5.5.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

6 MODEL DEVELOPMENT FOR SORBATE MOLECULES IN 1D SYSTEM . 88

6.1 Thermal Conductivity from NEMD Simulations . . . . . . . . . . . . . . . 886.2 Sorbate in a Harmonic Lattice . . . . . . . . . . . . . . . . . . . . . . . . . 91

6.2.1 Scattering of a Phonon Wavepacket . . . . . . . . . . . . . . . . . . 926.2.2 Scattering of a Plane Wave . . . . . . . . . . . . . . . . . . . . . . . 966.2.3 Possible Theoretical Approach: Multi-scale Expansion . . . . . . . . 986.2.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

7 CONCLUSIONS AND POSSIBLE DIRECTIONS OF FUTURE RESEARCH . 102

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

BIOGRAPHICAL SKETCH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

7

LIST OF TABLES

Table page

3-1 The neighbor list and equilibrium bond length for the particles in hexagonalsystem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

4-1 Parameters for Brenner-Tersoff interaction potential. . . . . . . . . . . . . . . . 49

5-1 Parameters for the lattice potential energy model Eq. 5–1. . . . . . . . . . . . . 68

5-2 Lennard-Jones parameters for sorbate-sorbate and sorbate-lattice interactions. . 68

5-3 Normalized averages of phonon amplitudes, ∆Qjk =< Qjk > /σjk. . . . . . . . . 79

8

LIST OF FIGURES

Figure page

3-1 Local energy evolution and detail of breather interaction in a FPU chain usingthe protocol of Reigada et al.[38] for k = β = 1/2, α = 0. . . . . . . . . . . . . . 25

3-2 Floquet multipliesr and the effect of unstable perturbation of breather solutions. 28

3-3 Families of breathers with different configurations. . . . . . . . . . . . . . . . . . 28

3-4 A family of breather solutions with ST mode configuration and their stabilityanalysis for a FPU-α system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3-5 Results of the stability analysis of the ST mode breathers for a range of α, ω,and fixed β = 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3-6 Nonlinear vibration mode and its stability analysis for a single FPU chain intwo-dimensional system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3-7 The configuration for two coupled FPU chains. . . . . . . . . . . . . . . . . . . 33

3-8 Nonlinear vibration mode and its stability analysis for two coupled FPU chainsin two-dimensional system with kFPU = kcoupling = 1, α = β = 0. . . . . . . . . . 33

3-9 Nonlinear vibration mode and its stability analysis for two coupled FPU chainsin two-dimensional system with (kFPU , kcoupling, βFPU) = (1, 1, 1). . . . . . . . . . 34

3-10 The configuration for body-centered cubic structure system. . . . . . . . . . . . 34

3-11 Continuation curves and configuration of nonlinear solutions corresponding todifferent amplitudes of the phonon modes. . . . . . . . . . . . . . . . . . . . . . 36

3-12 Nonlinear vibration mode of two-dimensional body-centered cubic structure for(k, β)coupling = (1.10, 1.0) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

3-13 Structure of a unit cell of the model hexagonal system in three dimensions. . . . 38

3-14 The dispersion curves for three-dimensional hexagonal tube system. . . . . . . 39

3-15 One of the phonon modes of the hexagonal system and its stability analysis with(k,β)coupling/FPU = (1, 0). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

3-16 Initial guess for the Newton’s method for 3D hexagonal system. . . . . . . . . . 41

3-17 Dependence of amplitude and frequency of a steady-state mode on the strengthof coupling between chains in hexagonal system. . . . . . . . . . . . . . . . . . . 41

3-18 Configuration of the nonlinear solutions with (k, β)coupling = (1, 1). . . . . . . . . 42

3-19 The comparison of solutions with and without RWA. . . . . . . . . . . . . . . . 43

9

3-20 Comparison of nonlinear modes of hexagonal system for three different couplingstrength. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

3-21 Stability analysis for the solutions with (without) RWA approximation. Equivalenteigenvectors are observed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3-22 Stability analysis for the solutions with (without) RWA approximation. Differenteigenvectors are observed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

4-1 Preparation of a nanotube by rolling the graphite sheet in a direction specifiedby the chiral vector Ch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4-2 Comparison of ratio of the magnitude of the nonlinear and linear force of CNTand FPU system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

4-3 Temperature profile and local energy evolution of NEMD simulations of a segmentof a 100 unit-cell (5,0) CNT. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4-4 Phonon dispersion curves of the (5,0) carbon nanotube. . . . . . . . . . . . . . . 53

4-5 The summary of the numerical procedure to obtain Fourier series expansion fora complex potential. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

4-6 Comparison between a linear phonon mode of the (5,0) carbon nanotube andthe nonlinear mode obtained from this mode by the continuation method. . . . 56

4-7 Dependence of the mode energy and frequency of the Taylor solutions on ε forsolutions shown in Figure 4-6. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

4-8 The solutions of (5,0) CNT in two unit cells system starting with the amplitudecorresponding to three different temperature. . . . . . . . . . . . . . . . . . . . 58

4-9 Dependence of the mode energy and frequency of the Taylor solutions startingwith three initial thermal energy on ε for solutions shown in figure 4-8. . . . . . 59

4-10 The comparison the displacement for the 7th atom in each unit cell for ε = 1nonlinear mode and starting phonon mode in a 24 unit cells system with thesimplified potential under RWA approximation. . . . . . . . . . . . . . . . . . . 60

4-11 The nonlinear vibration modes with simplified potential for the systems up to24 unit cells. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

4-12 Configuration of a linearly stable nonlinear solution based on simplified potentialfunction for a four unit cells system. . . . . . . . . . . . . . . . . . . . . . . . . 61

4-13 The nonlinear solution based on simplified potential function for a four unit cellssystem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

4-14 The eigenvector corresponding the Floquet multiplier shown as the open circlein Figure 4-13B. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

10

5-1 A block of 2× 2× 2 sodalite unit cells containing nine sodalite cages. . . . . . . 66

5-2 Dispersion relationships for sodalite. . . . . . . . . . . . . . . . . . . . . . . . . 72

5-3 Examples of autocorrelation functions Cjk(τ) of phonons in a sorbate-free sodalitecrystal. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

5-4 Power spectra Sjk(ω) corresponding to the phonon autocorrelation functionsshown in Figure 5-3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

5-5 Power spectra SV (ω) of sorbate velocities in rigid zeolite cages. . . . . . . . . . 78

5-6 Phonon lifetimes τjk in sorbate-free sodalite lattice. . . . . . . . . . . . . . . . 80

5-7 Effects of sorbates on phonon lifetimes. . . . . . . . . . . . . . . . . . . . . . . . 83

5-8 Relative differences δωjk between the anharmonic (ωajk) and harmonic (ωh

jk) frequenciesin a sorbate-free sodalite crystal. . . . . . . . . . . . . . . . . . . . . . . . . . . 84

5-9 Relative differences δωsjk between phonon frequencies in sodalite with encapsulated

sorbates (ωsjk) and in the sorbate-free sodalite (ωa

jk). . . . . . . . . . . . . . . . . 84

5-10 Magnitudes of correlation coefficients ρ(jk, j′k′) between phonon modes jk andj′k′ in a sorbate-free sodalite crystal. . . . . . . . . . . . . . . . . . . . . . . . . 85

5-11 Effect of sorbates on phonon-phonon correlations. . . . . . . . . . . . . . . . . . 86

6-1 Established temperature profile and heat flux in NEMD simulations for 1D sorbate-free lattice system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

6-2 Dependence of thermal conductivities on sorbate-lattice interaction parameters. 90

6-3 Steady state temperature profile in NEMD simulation with imposed temperaturegradient dT

dx= 0.002. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

6-4 Simulation snapshots of a wavepacket scattering simulation for wavenumber centeredaround q0 = 0.1π, (ε, σ) = (2, 0.44) and Ms = 1. . . . . . . . . . . . . . . . . . . 93

6-5 Dependence of transmission ratios on sorbate mass for three different wave numberswavepackets. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

6-6 Scattering of wavepackets with wave numbers q = 0.475π by a sorbate withfixed sigma = 0.44 and (Ms,epsilon)=(0.5,5) and (1,1). . . . . . . . . . . . . . . 95

6-7 Evolution of mode energy for incident plane wave with wavenumber q = 0.245πinteracting with different sorbate molecules. . . . . . . . . . . . . . . . . . . . . 96

6-8 Fourier transform of the sorbate displacement for the planar wave scatteringsimulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

11

6-9 Dispersion of incident mode energy (q = 0.245π) among other phonon modesunder different values of Ms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

6-10 Comparison of Fk and Gk between model prediction and simulation results. . . . 100

6-11 Comparison of frequencies obtained from degenerate perturbation method andregular eigenvalue solver. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

12

Abstract of Dissertation Presented to the Graduate Schoolof the University of Florida in Partial Fulfillment of theRequirements for the Degree of Doctor of Philosophy

THERMAL TRANSPORT IN NANOSTRUCTURED MATERIALS

By

Chia-Yi Chen

August 2008

Chair: Dmitry I. KopelevichMajor: Chemical Engineering

Thermal transport in nanostructured materials often exhibits significant deviations

from predictions of the classical Fourier’s law for thermal conductivity. The deviations

occur because the length of the mean free path of heat-carrying phonons is comparable

with characteristic length-scale of these materials. Therefore, it is necessary to develop

a theory for thermal transport applicable to nanomaterials. In this study we investigate

thermal conductivity in two classes of nanomaterials, namely quasi-one-dimensional

materials and nanoporous materials with adsorbed guest molecules. For quasi-one-

dimensional (Q1D) materials, we aim to understand nonlinear dynamics involved in heat

transfer using a combination of molecular dynamics simulations and bifurcation theory.

In non-equilibrium molecular dynamics simulations, we observe ballistic propagation of

energy packets in model Q1D systems as well as in carbon nanotubes, which suggests

the significance of ballistic heat transfer mechanism. To decipher structure of the

waves propagating in the lattice, we obtain nonlinear lattice vibration modes by solving

fundamental equations of motion numerically, without ignoring any structural details. We

focus on localized nonlinear vibration modes, and investigate their properties and stability

on the structured details of the lattice, and potential energy of interaction between lattice

atoms.

In the second part of this work, we investigate the effect of sorbate-lattice interaction

in thermal transport for nanoporous materials. There is increasing evidence that thermal

13

conductivity of nanoporous materials can be significantly affected by adsorption of guest

molecules. These molecules serve as moving defects and provide additional scattering

centers for heat-carrying phonons. In order to understand the sorbate-phonon interactions,

we first perform molecular dynamics simulations of a realistic system, namely sodalite

zeolite with small molecules (argon, xenon, and methane) encapsulated in its cages.

We observe that the phonon lifetime often increases upon encapsulation of a sorbate

into the zeolite which suggests that the sorbate-phonon interactions are qualitatively

different from phonon scattering by point defects fixed in the lattice. We then proceed

to develop a model for the sorbate-lattice interaction. For simplicity, we consider a one-

dimensional lattice system. We investigate the role of the sorbate in the energy exchange

between lattice modes and observe that even a weak interaction between the sorbate and

lattice induces dispersion of energy over a wide spectrum of normal modes.

14

CHAPTER 1INTRODUCTION

Advances in materials science and manufacturing technology have enabled integration

of nanostructured materials in electronics, energy conversion systems and sensors. For

example, it is believed that CMOS (complementary metal-oxide semiconductor) will

decrease to 22-nm within next 10 years [1]. To improve footprint of elements of integrated

circuits, non-volatile memory devices using 1D structures such as nanowires and nanotubes

have drawn significant attention. In order to ensure the performance and stability of

incandesces, it is necessary to assess thermal properties of new nanoscale materials [2, 3].

The classical Fourier’s law of thermal conductivity, states that the heat flux J is linearly

proportional to the temperature gradient ∇T ,

J = −κ∇T, (1–1)

where the thermal conductivity κ can be approximated as [4]:

κ ∝ Cvl. (1–2)

Here, C, v, and l are the heat capacity, the group velocity and the mean free path of the

phonons respectively. The assumptions behind equations (1–1) and (1–2) are that (i)

the phonon mean free path l is much smaller than the characteristic size of the material

and (ii) the temperature gradient is sufficiently small so that collisions between phonons

maintain local equilibrium.

Both of these assumptions are likely to fail in nanosomia systems. For example, the

mean free path of heat-carrying phonons in silicon at 300 K is 300 nm whereas the current

dimension of a thin silicon film is under 100 nm [2] and the characteristic size of a hotspot

in a transistor can be as small as 10 nm [5]. Therefore, it is necessary to consider the

effect of boundary scattering and phonon confinement [6–8] explicitly. These issues are

15

currently being addressed by development and solution of the Boltzmann equation for the

phonon scattering due to phonon-phonon and phonon-boundary interactions.

The second assumption is also likely to fail in nanosomia materials, where thermal

gradients can be significant. In this case, the Boltzmann equation for phonon modes might

be invalid since the role of nonlinearities of the lattice vibration will not be limited to

energy exchange during collisions of linear phonon modes. In fact, recent investigations

have shown that different type of lattice vibration modes, known as the intrinsic nonlinear

localized modes or breathers [9–11], make a strong contribution to the thermal energy

transfer and lead to qualitatively different thermal properties in low dimensional materials.

In fact, it has been shown both theoretically, using the self-consistent mode coupling

theory [12, 13], and by numerical simulations [13–15] that the bulk heat conductivity

exhibits anomalous dependence on the lattice size for one- and two-dimensional lattices.

In the current work, we investigate the process of thermal conductivity in one- and

quasi-one-dimensional systems such as carbon nanotubes (CNTs) [16]. Due to their

unique mechanical and electronic properties, CNTs have drawn significant attention

for a wide variety of potential applications, [17] and are currently being integrated into

electronic devices such as high-performance ballistic field-effect transistors, nanotube

random access memory (NRAM) and assembling integrated logic circuit on an individual

carbon nanotube [18–20]. Currently there is only incomplete understanding regarding

thermal transport for CNTs, which is extremely important for heat management issue.

Recent experiments [21, 22] and molecular dynamics simulations [23–26] report anomalous

thermal conductivity of the nanotubes. However, there are significant discrepancies

between results of the experimental measurements and the simulations. Moreover,

discrepancies between results obtained using different simulation techniques confirm that

the conventional thermal conductivity theory is invalid for this system and there is a need

for a better description of thermal transport phenomena in these Q1D structures. In the

current work, we use the following two approaches to investigate the thermal transport in

16

Q1D materials. (i) numerical solution of the fundamental equations of motion to obtain

steady state nonlinear lattice vibration modes and (ii) non-equilibrium molecular dynamics

(MD) simulations to investigate the dynamics of a heat propagation process.

The second part of this thesis is devoted to the investigation of thermal transport in

nanoporous materials. Efficient thermoelectric materials received significant attention in

development of energy conversion technologies. Systems with low thermal conductivity

and good electrical conductance are considered as promising candidates for solid-state

thermoelectric refrigerators and power generators. It has been shown that addition of

sorbate molecules to voids within a nanoporous material such as skutterudite [27, 28]

leads to an order of magnitude decrease of its thermal conductivity, and that in turn

increases its thermoelectric figure of merit. A series of experimental and computational

studies indicate that oscillations (”rattling”) of guest molecules (sorbates or cations)

within zeolites and skutterudites have significant impact on thermal conductivity of these

crystals. Murashov et al. [29] performed MD simulations to show that presence of cations

can lead to either an increase or a decrease of thermal conductivity, depending on their

mass. Available data suggests complex dependence of the thermal conductivity on the

nature of a guest molecule and a host crystal. Hence fundamental understanding of the

effect of guest molecules on host matrix is necessary in order to engineer heat conductance

by the use of host-guest interactions.

17

CHAPTER 2NUMERICAL METHODS

In this chapter, we briefly review molecular simulation tools and numerical methods

used in this work.

2.1 Molecular Dynamics Simulations

The heat transfer phenomena can be modeled by molecular dynamics simulations.

Consider a general lattice system with the Hamiltonian

H =∑

n

u2n

2+ V ({un}), (2–1)

where un =√

mn(rn − reqn ) is the mass-normalized displacement of the n-th atom of mass

mn from its equilibrium position reqn , rn is the coordinate of the n-th atom, and V ({un})

is the potential energy of interaction between the atoms. Molecular dynamics simulations

solve Newton’s equations of motion for each particle,

un = Fn({un}), (2–2)

where forces Fn are given by

Fn({u}) = −∂V ({un})∂un

. (2–3)

In order to obtain the evolution of the entire molecular system, we integrate Eq. 2–2 using

the velocity Verlet algorithm [30],

r(t + ∆t) = r(t) + v(t)∆t + 12F(t)∆t2

v(t + ∆t) = v(t) + 12(F(t) + F(t + ∆t))∆t,

(2–4)

where r(t) and v(t) are the atomic position and velocity, respectively.

2.2 Temperature Coupling in Molecular Dynamics

In MD simulations the instantaneous value of the temperature is related to the total

kinetic energy of all atoms within the system as follows:

N∑i=1

1

2miv

2i =

kBT

2Nd. (2–5)

18

Here, N is the total number of atoms and d is the dimensionality of the system. Some of

the simulations performed in this work are performed in the canonical (NVT) ensemble.

In order to maintain constant temperature in such simulations, we bring the system

into thermal contact with a large thermal bath. In this work, this is accomplished by

employing one of the following thermostats : Berendsen [31], Nose-Hoover [32],or Langevin

thermostats.

2.2.1 Berendsen Thermostat

Deviation of the current temperature (T ) from a prescribed temperature (T0) is

corrected by rescaling atom velocities by a factor

λ =[1 +

∆t

τT

(T0

T− 1)

]1/2

, (2–6)

where ∆t is the step size for the time integration scheme and parameter λ specifies the

strength of temperature coupling between the system and thermal bath. The velocity

rescaling leads to an exponential decay of temperature deviation with time constant τT ,

dT

dt=

T0 − T

τT

. (2–7)

2.2.2 Nose-Hoover Thermostat

While Berendsen thermostat is extremely efficient for relaxing a system to the

target temperature, rescaling the velocity creates discontinuity in momentum. Nose-

Hoover thermostat provides a better approximation of canonical ensemble. The idea

of the method proposed by Nose is to reduce the effect of the heat reservoir to an

additional artificial degree of freedom: with coordinate s and mass Q. The magnitude

of Q determines the coupling strength between the reservoir and the real system and also

influences the temperature fluctuations as well. The Hamiltonian in the extended system

H =N∑

i=1

p2i

2mi

+ V (q) +Qs2

2+ gkBT ln(s), (2–8)

19

is conserved. Here, q and p are the generalized coordinates and momenta of the system,

g is the number of degrees of freedom in the system and V is the potential energy. The

corresponding equations of motion for the extended system are

dqi

dt=

pi

mi

, (2–9)

dpi

dt= −∂V

∂qi

− ζpi, (2–10)

∂ln(s)

∂t= ζ, (2–11)

dt=

1

Q(

N∑i=1

p2i

2ms− gkBT ). (2–12)

Note that the value of Q should be the inverse of the characteristic time scale of the

system in order to avoid inefficient thermostat scheme (too large value of Q) or high

frequency temperature oscillations (too small values of Q) in the system [33].

2.2.3 Langevin Thermostat

In Langevin thermostat, the contact of the system with thermal bath is performed by

random collisions between atoms of the system and the thermal bath atoms, and thus the

equations of motion is

un = Fn − γun + η(t), (2–13)

where γ is the friction coefficient and η(t) is the random force due to the particle

collisions. The random force has a zero mean and its autocorrelation function satisfies

the fluctuation-dissipation theorem,

< η(t)η(t′) >= 2γkBTδ(t− t′) (2–14)

Numerical method of solutions of a stochastic differential equation is different from

that for an ordinary deterministic equation of motion. The numerical integration of

Eq. 2–13 implemented in this work follows the method of Ermak et. al [34].

20

2.3 Steady State Solutions

In order to obtain the time-periodic lattice vibration modes for the systems, we

solve fundamental equations of motion based on Eq. 2–2. The corresponding steady

state solutions can be thought of as periodic orbits in the 2Nd-dimensional phase space.

The coordinates of a phase point Y in this phase space are the normalized positions

and velocities of the atoms, i.e. Y = {u1, . . . ,uN , u1, . . . , uN}, where un and un are

d-dimensional vectors. The numerical method used to obtain the steady state solutions

of Eq. 2–2 is based on the approximation of the periodic solutions by a truncated Fourier

series expansion,

un(t) =M∑

j=−M

xn,jeiωjt, (2–15)

where ω is the frequency of the mode and xn,j are time-independent coefficients.

Substituting Eq. 2–15 into Eq. 2–2, we obtain the equations for xn,j

F(x) = 0. (2–16)

We then solve Eq. 2–16 using Newton’s method to obtain x. Iterations of Newton’s

method are given by xn+1 = xn + δx, where δx is a solution of the following linear system

of equations:

Jxn · δx = −F(xn) (2–17)

and Jx0 is the Jacobian matrix of Eq. 2–16.

The lattice systems are translationally invariant leading to singularity in the Jacobian

matrix. In order to remove these singularities, we use the singular value decomposition

(SVD) method to solve the system of linear Eq. 2–17. In this method matrix J is

decomposed as

J = U ·w ·VT , (2–18)

where T denotes matrix transpose, U and V are column-orthogonal matrices, and w

is a diagonal matrix with positive or zero elements. This decomposition allows one to

21

obtain the nullspace and the range of matrix J. Specifically, the columns of U in which

same-numbered elements ωj are nonzero represent an orthonormal basis of the range; the

columns of V in which same-numbered elements ωj are zero represent an orthonormal

basis for the nullspace. If the right-hand side of Eq. 2–17 lies in the range of J, the system

of Eq. 2–17 has more than one solutions, since any vector in the nullspace of J can be

added to a solution x. Physically, this corresponds to an arbitrary translation and/or

rotation of the entire molecular system. SVD enables us to find the solutions with the

smallest norm, i.e. prevents the rotation and translation of the system by ignoring the

nullspace basis.

The inverse of matrix J is

J−1 = V · [diag(1/wj)] ·UT (2–19)

and unknown δx is

δx = V · [diag(1/wj)] ·UT · (−F) (2–20)

Here we replace 1wj

by zero if wj = 0 so that the nullspace vectors are ignored in the

solutions.

2.4 Stability Analysis

In order to assess the role of the obtained steady-state vibration modes in thermal

conductivity, it is necessary to analyze their stability to perturbations. An unstable mode

will be quickly destroyed by perturbations, and therefore is unlikely to contribute to

the heat transfer process. We examine the linear stability (i.e. stability to infinitesimal

perturbations) of the time-periodic nonlinear modes using the Floquet theory [35].

Consider a steady-state periodic solution Y(0)(t) = {u(0)1 (t), . . . ,u

(0)N (t), u

(0)1 (t), . . . , u

(0)N (t)}

of Eq. 2–2; the period of this solution is T = 2π/ω. For convenience, let us rewrite the

equations of motion Eq. 2–2 as a first-order system of equations,

Yn(t) = fn(t), (2–21)

22

where

fn(t) = un(t) = YN+n(t), n = 1, . . . , N, (2–22)

fN+n(t) = un(t) = − ∂V

∂Yn

, n = 1, . . . , N. (2–23)

Now consider a perturbation of Y(0)(t) of the form Y(t) = Y(0)(t) + εY(1)(t), ε ¿ 1. The

equations linearized near the steady-state solution are

Y(1)n =

∑m

Lnm(Y(0)(t))Y(1)m , n = 1, . . . 2N, (2–24)

where the time-periodic d× d matrices Lnm are given by

Lnm(Y(0)) =∂fn(Y(0))

∂Y(0)m

. (2–25)

In order to apply the Floquet analysis, we obtain a 2Nd × 2Nd matrix F (t) of

fundamental solutions of the linearized equations (2–24). The j-th column of this matrix

corresponds to the solution with the initial condition of the form Y(1,j)(t = 0) =

(0, 0, . . . , 0, 1, 0, . . . , 0) with the only non-zero element of the vector Y(1,j)(t = 0) located

at the j-th position. These solutions are obtained by the numerical integration of the

system of linearized equations (2–24). Once the fundamental solutions are obtained, we

compute the Floquet multipliers, i.e. the eigenvalues µj (j = 1, . . . , 2Nd) of the matrix

F (t) at time t = T . The steady-state solution is stable only if all of the Floquet multipliers

have magnitude less than or equal to 1. Since the system considered in this work is

Hamiltonian, for each eigenvalue µj the numbers µ∗j and 1/µj are also eigenvalues [11]

(here, asterisk denotes complex conjugation). Thus the neutral stability of the steady-state

solution requires that all the multipliers lie on the unit circle of the complex plane.

23

CHAPTER 3NONLINEAR LATTICE VIBRATION MODES IN MODEL SYSTEMS

3.1 One-Dimensional FPU system

In this section, we illustrate the methods employed in this work by application to a

one-dimensional Fermi-Pasta-Ulam (FPU) [36] system, i.e. a linear anharmonic chain of

atoms with the following Hamiltonian:

H =∑

n

[u2

n

2+

k

2(un − un−1)

2 +α

3(un − un−1)

3 +β

4(un − un−1)

4

]. (3–1)

Here, un is the mass-normalized displacement of the n-th atom from its equilibrium

position and k, α, and β are the harmonic, cubic, and quartic force constants, respectively.

The considered chain consists of N = 50 atoms with imposed periodic boundary

conditions, i.e. u1 = uN+1 and u1 = uN+1. In the analysis of the FPU system, we fix

the value of the harmonic spring constant k = 1 and vary the anharmonic spring constants

α and β. In literature, the system with α = 0 is referred to as FPU-β model and the

system with α 6= 0 is referred to as FPU-α model.

3.1.1 Thermal Relaxation Simulation

It has been shown in the literature that in addition to phonons, energy in FPU

system may be transmitted by intrinsic localized modes, also known as discrete breathers

(DBs) [9, 11, 37], DBs are intrinsic localized nonlinear vibration modes that are

qualitatively different from the phonon modes and can be observed in thermal relaxation

of the lattice. MD simulations of thermal relaxation of FPU system have been performed

by Reigada et al. [38]. They prepared a 30-site lattice initially thermalized at T = 0.5

by Langevin thermostat and after the system reached equilibrium, they disconnected the

thermalized lattice system from the heat bath and connected the ends of the chain via a

friction term γ to dissipate the energy into the thermal reservoir of zero temperature. Here

we show the results of our simulations following their protocol for a N = 50 lattice with

k = β = 1/2, α = 0. The evolution of the local energy of this system during relaxation

24

A B

10 20 30 40 50−0.5

0

0.5

Atom

Dis

plac

emen

t

Time=1754.5

C

10 20 30 40 50−0.5

0

0.5

Atom

Dis

plac

emen

t

Time=1809.5

D

10 20 30 40 50−0.5

0

0.5

AtomD

ispl

acem

ent

Time=1846.5

E

Figure 3-1. (A) The breathers in a FPU chain obtained using the protocol of Reigada etal.[38] for k = β = 1/2, α = 0. The gray scale in this figure represents the localenergy magnitude, with darker shading corresponding to more energeticregions. The horizontal axis indicates the position along the chain and thevertical axis corresponds to time. The energy density is shown by a gray scalefrom 0 (white) to the maximum energy recorded during the simulation (black).The energy localizes into narrow breathers which are seen as the black lines onthis plot. The noisy phonon modes correspond to randomly shaded areas. B),Details of the energy localization and breather interaction. C), D) and E) aresnapshots for the process of breather-breather corresponding to three differenttimes in relaxation simulation indicated by the dashed lines in plot B.

process is plotted in Figure 3-1 A), where the darker region represents higher energy. We

observe the ballistic energy transfer clearly after the dissipations of phonon modes.

The breathers are seen to move with essentially constant speed and appear to be

stable with respect to the low-energy noisy phonon modes. Details of breather-breather

interactions are shown in plot B and snapshots for this process are shown in plots C-E.

25

Elastic interaction is observed between breathers. It is clear that, after initial transient,

the energy becomes highly localized and in addition to phonons, the system contains

localized nonlinear structures (breathers).

3.1.2 Steady State Solutions

Steady-state solutions are approximated by rotating wave approximation (RWA) [39],

in which coefficients with j ≥ 2 in Eq. 2–15 are neglected. Therefore, the nonlinear terms

in the equations of motion are approximated as follows:

cos2(ωt) ' 12, cos3(ωt) ' 3

4cos(ωt) (3–2)

We compare each of the Fourier coefficients in Eq. 2–2 to set up the equations for

Newton’s method. Since quadratic force term is approximated by a constant under

RWA approximation, it leads to non-zero average displacements xn,0 of atoms.

3.1.2.1 FPU-β model

In this case, there is no quadratic nonlinear force and we seek the solution only with

j = 1 term:

un = xncos(ωt) (3–3)

Based on the solution configuration and the researchers who proposed the analytical

solution form of nonlinear solutions, DBs with asymmetric configuration in Figure 3-3 A)

(lower inset plot) is referred to as P mode (abbreviation for Page [40]). The DBs with

symmetric configuration in the lower inset plot of Figure 3-3 B) are referred to as ST

mode (abbreviation for Sievers and Takeno [41]). The initial guess for Newton’s method

are two of the high frequency phonon modes with wave vector k ≈ πa, shown in the

upper insets in Figure 3-3 A) and Figure 3-3 B). We perform continuation of solution by

gradually increasing β and ω by ∆β = 0.05 and ∆ω = 3× 10−3 until β = 1. The frequency

of DBs should be outside of the phonon frequency band to avoid resonance with phonon

solutions [9]. These two degenerate normal modes lead to two family of P mode and ST

mode breathers. The dependence of the mode energy on β are shown in Figure 3-3. Linear

26

stability of the breather modes to perturbations is examined using the Floquet analysis

described in Section 2.4. Recall that a vibration mode is stable only if all of its Floquet

multipliers µj lie on the unit circle of the complex plane. The closed and open circles

in these two figures represent the stable and unstable solutions, respectively. Results of

Floquet analysis for a stable ST mode DB with β = 1 is shown in Figure 3-2 A), where

µj are located on the unit circle. However, P modes are linearly unstable with unstable

multipliers being pure real numbers as shown in Figure 3-2 B). Figure 3-2 C) shows the

eigenvector corresponding to multiplier indicated by an open circle in Figure 3-2 B). In

order to understand the effect of this perturbation vector, we perform MD simulation with

initial condition as P mode breather with perturbation vector. Two snapshots are shown

in plot D and E. Initially, this perturbation results in the DB moving to the right and

eventually leading to a transition between P mode and ST mode configurations.

3.1.2.2 FPU-α model

In order to assess the role of the cubic term in Hamiltonian (Eq. 2–2), we turn to

the FPU-α model. We seek solutions with the RWA approximation, Eq. 3–2. Now, the

solutions will contain a static component,

un = xn,0 + xn,1cos(ωt). (3–4)

We take the DBs of the FPU-β systems as an initial guess for Newton’s method and

increase the cubic coefficient α. An example of an obtained steady-state P mode is shown

in Figure 3-4 A). We observe that the breather amplitude A increases with increase of the

magnitude of the cubic nonlinearity α. The dependence of A on α for a fixed frequency

ω = 2.89 and quartic term β = 1 is shown in Figure 3-4 B). We obtained breather families

for a range of ω between 2.2 and 2.93 (while keeping β = 1), and observed the increasing

of A while increasing α value as well. Interestingly, we observe that introduction of the

cubic nonlinearity destabilizes the breather modes. The stable and unstable breathers are

shown respectively by solid and open circles in Figure 3-4 B). To summarize the results of

27

−1 0 1

−1

0

1

Re(µ)

Im(µ

)

A

−1 0 1

−1

0

1

Re(µ)

Im(µ

)

B

20 30

−0.5

0

0.5

Particle Position

Dis

plac

emen

t

DBe

x

ev

C

10 20 30 40 50

−0.6

0

0.6

Time=0.025

Atom

Dis

plac

emen

t

D

10 20 30 40 50

−0.6

0

0.6

Time=2.2

Atom

Dis

plac

emen

t

E

Figure 3-2. Floquet multipliers DBs: A) ST mode and B) P mode solutions. (C) shows theDB configuration as well as the unstable displacement (ex) and velocity (ev)perturbation corresponding to the open circle in plot B The snapshots of MDsimulations for perturbed P mode breathers. (D) Initial configuration (E)shows the transition from P mode to ST mode.

0 0.2 0.4 0.6 0.8 1

1

2

3

4

5

6

7

β

Ene

rgy

−0.3

0

0.3β = 0

−0.8

0

0.8β = 1

A

0 0.2 0.4 0.6 0.8 1

1

2

3

4

5

6

7

β

Ene

rgy

−0.3

0

0.3β = 0

−0.8

0

0.8β = 1

B

Figure 3-3. Two family of (A) ST modes and (B) P mode DBs. The closed (open) circlesrepresent the linearly stable (unstable) modes.

28

10 20 30 40 50−0.8

−0.4

0

0.4

0.8ω=2.87, β=1, α=0.2

n, atom number

Xn,

j

Static (j=0)Dynamic (j=1)

A

0 0.2 0.4 0.6 0.8 10.38

0.4

0.42

0.44

0.46

0.48

0.5

0.52

α

A

ω=2.3061, β = 1

−1 1−1

1

α = 0.24

−1 1−1

1

α = 0.78

B

0 0.2 0.4 0.6 0.8 1

0.8

0.85

0.9

α

A

0 0.2 0.4 0.6 0.8 10

0.01

0.02

α

ν

C

Figure 3-4. (A)An ST mode for α = 0.2, β = 1, ω = 2.87. Both the static (•) and dynamic(¤) atom displacements are shown (B) A P mode breather family forω = 2.89, β = 1, with varying α. (C) Results of the stability analysis of thebreathers of the family shown in plot (B). The stability parameter ν is definedby Eq. 3–5. In these plots, the closed (open) circles denote the linearly stable(unstable) modes.

the Floquet analysis we introduce the quantity

ν = maxj

||µj| − 1| (3–5)

so that a steady-state vibration mode is stable if ν = 0 and is unstable if ν > 0. Figure 3-

5 B) shows the dependence of ν on the value of cubic nonlinearity α for ω = 2.89 and

β = 1. It is clear that the breather mode becomes unstable as α increases. We extended

this analysis to breathers corresponding to other values of frequency ω. The results are

summarized in Figure 3-5 C). The breather modes are observed to be unstable for all

values of α if ω ≤ 2.85, whereas for larger values of ω the modes become more stable for

at least some range of α. In general, we destabilizing effects of the cubic nonlinearity on

the breather modes. However, the detailed stability properties of the breathers are more

complex. For example, there are “islands” of stability at larger values of α for ω = 2.88

and ω = 2.92. This agrees with published report that shows cubic anharmonicity reduces

the thermal conductivity [42]. The instability of localized breathers in FPU-α indicates

the short life time of ballistic energy packet during MD simulation, which implies the

reduction of thermal conductivity [42].

29

0 0.2 0.4 0.6 0.8 1

2.84

2.88

2.92

α

ω

Figure 3-5. Results of the stability analysis of the ST mode breathers for a range of α, ω,and fixed β = 1. The closed (open) circles denote the stable (unstable) modes.

3.2 Lattice Model Systems In Higher Dimensions

Breathers in one-dimensional model lattices have been studied both theoretically and

numerically [10, 43–53]. The breathers have also been observed in models of real physical

system, such as a molecular model for a row of atoms in a semiconductor crystal GaN

[54]. In addition to periodic atomic chains, breathers have been observed in simulations of

disordered systems [44, 55], which has important implications for thermal conductivity in

polymers and biological systems. Moreover, several experimental studies indicate existence

of breathers in molecular systems. For example, spectroscopic studies of laser-induced

vibrations in a quasi-one-dimensional chain of halogen-bridged mixed Pt complex [56, 57]

report a Raman spectrum characteristic of localized nonlinear vibration modes. The

intrinsic localized modes have also been experimentally observed in myoglobin [58].

In the previous section we discussed FPU model with atoms allowed to move only

along the chain. It is more realistic to allow vibrations in more than one direction and

so below we investigate the existence and stability analysis for nonlinear lattice vibration

modes in two and three-dimensional lattice systems with high aspect ratio. We perform

the numerical calculation for few different model systems: a single FPU chain in two-

dimensional system; two coupled FPU chains in two dimensional system, body-centered

cubic structure in two-dimensional system, and a three-dimensional hexagonal tube formed

30

by six coupled FPU chains. In all these systems, periodic boundary condition is imposed

only in the axial direction to mimic quasi-one-dimensionality.

The general Hamiltonian for the model systems can be written in the following form

H =∑

n

{u2

n

2+

1

2

[l∑

j=1

knj

2(rnj − dj)

2 +βnj

4(rnj − dj)

4

]}, (3–6)

where un is the mass-normalized displacement of the n-th atom of mass mn from its

equilibrium position; rjn and djn are the instantaneous distance and equilibrium bond

length between particle n and its j − th neighbor; knj, βnj represent the linear and

nonlinear force coefficients, respectively. The corresponding equations of motion are:

un = −l∑

j=1

[k(rnj − dj) + β(rnj − dj)3]

rnj

rnj

, (3–7)

3.2.1 Single Chain System in Two Dimensional System

We first examine the stability of DBs presented in section 3.1.2 We consider one FPU

chain which extends along the axial x direction with particle vibrations in both x and

radial y directions. Figure 3-6 A) shows the vibration mode with breather configuration

for x direction displacement, which satisfies the equations of motion. However, the

stability analysis plotted in Figure 3-6 B) reveals that this solution is linearly unstable.

The eigenvectors corresponding to the unstable Floquet multipliers are the perturbations

along the radial direction. An example is presented in Figure 3-6 C). To understand the

cause of instability, a simple force analysis is done for three atoms x1, x2, x3 connected by

FPU interaction potential. We consider the case where only the center particle (x2) has

displacement (dx, 0) from its equilibrium position, hence the force acting on x2 along y

direction is expressed as

F2,y = −∂U(r12)

∂y2

− ∂U(r23)

∂y2

(3–8)

31

10 20 30 40 50−0.5

0

0.5

u i

ione chain

XY

A

−1 0 1−1

0

1

Re(µ)

Im(µ

)

B

10 20 30 40 50−0.5

0

0.5

ione chain

u i

XY

C

Figure 3-6. A) The nonlinear vibration mode for a single FPU chain in two-dimensionalsystem B) Stability analysis of this solution; C) The unstable eigenvectorcorresponding to the open circle in B), which indicates that the nonlinearsolution is unstable against the perturbation along radial y direction.

This displacement will be stable against a small perturbation along y direction y2 if the

stable criterion

y2 × F2,y < 0 (3–9)

is satisfied. We look at the condition where both dx and y2 are positive. We then

substitute explicit interaction expression U and Eq. 3–8 into Eq. 3–9 and note that

r12− req = req − r23 for this configuration. After algebraic calculation, we have the stability

criterion as

k(r12 − req)(1

r23

− 1

r12

) +β

6(r12 − req)

3(1

r23

− 1

r12

) < 0, (3–10)

where force coefficients k and β are positive. Since r23 < r12 and r12 − req > 0, Eq. 3–10

will not be satisfied for small values of y2, this indicates that the displacement of the

central particle in y direction always increase.

3.2.2 Two Coupled FPU Chains

As we saw in the previous section, localized modes in a single atom chain are unstable

to perturbations in the radial direction. In this section, we increase the restriction on the

chain movement in the radial direction by introducing another parallel FPU chain coupled

to the original chain by springs. The system under consideration is illustrated in Figure 3-

7. The force constants along each FPU chain are denoted with the subscript FPU

and the force constants between these two FPU chains are denoted with the subscript

32

Figure 3-7. The configuration for two coupled FPU chains.

1 50−0.1

0

0.1Displacement

X

1 50−0.1

0

0.1

Particle

Y

A

−1 0 1

−1

0

1

Re(µ)

Im(µ

)

B

Figure 3-8. A) Steady state solution for two coupled FPU chains withkFPU = kcoupling = 1, α = β = 0. B) Stability analysis of this solution.

coupling. In this system, we analyze the stability for both the linear (α = β = 0) and

the nonlinear coupling potential between the chains. Note that in the systems with higher

dimensionality, even the linear force function k(r − req) for the bonds between the chains

introduces an additional nonlinearity in the equations of motion. Therefore, it is necessary

to use Newton’s method to obtain a steady state solution corresponding to the system

with α = β = 0. The corresponding solution with (kFPU , kcoupling) = (1.0, 1.0) and its

stability analysis are shown in Figure 3-8 A) and B), respectively.

To examine the existence of nonlinear vibration modes with configuration similar to

DBs, we obtain the nonlinear solutions by using breathers with (k, α, β)FPU = (1, 0, 1)

as the initial guess in each FPU chain, and then gradually increase the values of the

coupling force constants. An example of such a solution is shown in Figure 3-9 along with

its stability analysis, which indicates that this solution is linearly unstable. We plot the

configuration of unstable perturbation eigenvector in Figure 3-9 C), which corresponds to

33

1 50−0.6

0

0.6Displacement

X

1 50−0.1

0

0.1

Y

Particle

A

−1 0 1

−1

0

1

Re(µ)

Im(µ

)

B

1 50−0.2

0

0.2

Particle

Dis

plac

emen

t

XY

C

Figure 3-9. A) Steady state solutions for two coupled FPU chains with(kFPU , kcoupling, βFPU) = (1, 1, 1) with the initial guess shown in Figure 3-6. B)Stability analysis for solution A) C) The unstable eigenvector corresponds tothe open circle in B), which indicates that the nonlinear solution is unstableagainst the perturbation along radial y direction.

Figure 3-10. The configuration for body-centered cubic structure system.

the open circle multiplier in Figure 3-9 B). It shows that the instability is still caused by a

perturbation in y direction.

3.2.3 Body Centered Cubic Structure

To further restrict the movement in the radial direction, we place one more particle

at the center of each unit cell and investigate nonlinear modes of the body-centered cubic

system shown in Figure 3-10. We refer to the particles in the upper and lower chains

as the first and second particles of the unit cell and to the center particle as the third

particle.

In section 3.2.2, the initial guesses for Newton’s method are localized vibration modes

for each FPU chain. We reached the final nonlinear solutions by gradually increasing

the coupling strength. However, this approach cannot be used for all the systems, such

as carbon nanotubes due to the complex structure of their unit cell. On the other hand,

34

phonon modes are always available for any given lattice system. Therefore, in this

section, we explore the process of finding the localized nonlinear solution starting from

phonon modes. We first obtain the nonlinear vibration mode with (k, β)FPU = (1, 0.98),

(k, β)coupling = (0.1, 0), using the phonon solution in Figure 3-8 as the initial guess for

the Newton’s method. Due to the degeneracy of the phonon modes in this lattice system,

we apply a degenerate perturbation method to this phonon to obtain the initial guess for

Newton’s method. In addition, different from the previous calculation where frequency

ω is a given and fixed parameter in Newton’s iteration, here we allow ω to be one of the

unknowns in the numerical process. Hence, we need to add one more equation in order to

utilize Newton’s method. This can be done by fixing the center of mass in the system, i.e.

∑xn,0 = 0. (3–11)

Furthermore, we use the phonon modes with different amplitudes as initial guesses in

Newton’s method in order to study the energy threshold to excite localized nonlinear

modes in this system. We obtain several families of nonlinear solutions corresponding

to different magnitude for the phonon mode used as an initial guess. The dependence

of ω on β is shown in Figure 3-11 A), and the configuration of full nonlinear solutions

(β = 1) from various initial amplitudes are shown in Figure 3-11 B). In this model,

we observed that degenerate perturbation with appropriate mode energy will allow

us to reach the localized solutions. We continue to increase the coupling strength to

(k, β)coupling = (1.1, 1.0), and perform the stability analysis of the obtained solution. In

order to examine the stabilization effect due to the interaction with centered particle, we

compare the solution configuration and Floquet multipliers for two sets of parameter. The

Floquet multipliers for the second set of parameters (k, β)coupling = (0.55, 0.55) is presented

in Figure 3-12 E). There is a significant decrease of the number and magnitude of unstable

eigenvalues in compared to Figure 3-12 D) due to stronger coupling with the central

particle. Moreover, we observe the disappearance of the highly unstable perturbation

35

0.2 0.4 0.6 0.8 1.02

2.1

2.2

βω

A=1.2

A=2.0

A=2.9

A

1 50−0.4

0

0.4

Atom

Dis

plac

emen

t

B

1 50−0.4

0

0.4

Atom

Dis

plac

emen

t

C

1 50−0.4

0

0.4

Atom

Dis

plac

emen

t

D

Figure 3-11. A) Families of nonlinear solutions corresponding to different amplitudes ofthe phonon modes. B) nonlinear vibration mode (ε = 1) for A = 1.2 C)nonlinear vibration mode (ε = 1) for A = 2.0 D) nonlinear vibration mode(ε = 1) for A = 2.9

with multiplier magnitude close to zero (shown as an open circle), which corresponds to

perturbation along axial direction.

3.3 Hexagonal Tube Model

In this section we study a hexagonal system in three dimensions which contains

6 FPU chains in order to mimic a nanotube or a nanowire. A unit cell of the model

structure containing 13 particles is shown in Figure 3-13. The particles located at the

corners of the hexagon are referred to as type A, the particles located at the face centers

are referred to as type B, and the particle located at the center of the hexagon is referred

to as type C. We assign each particle an index (i,Kj), which indicates the particle is

located in the i−th unit cell and belongs to j−th particle of type K (K =A, B or C;

j = 1, . . . , 6 if K = A or B and j = 1 if K = C). The hexagon sides and the bond

36

1 50

−0.1

0

0.1x j,0

X

Y

1 50−0.2

0

0.2

jth Unit Cell

x j,1

A

1 50

−0.1

0

0.1

x j,0

X

Y

1 50−0.2

0

0.2

jth Unit Cell

x j,1

B

1 50

−0.1

0

0.1

x j,0

X

Y

1 50

0

jth Unit Cell

x j,1

C

−2 −1 0 1 2

−1

0

1

Im(µ

)

Re(µ)

D

−2 −1 0 1 2

−1

0

1

Im(µ

)

Re(µ)

E

Figure 3-12. The nonlinear vibration mode of two-dimensional body-centered cubicstructure for (k, β)coupling = (1.10, 1.0). A), B) and C) show thedisplacement along the upper, lower, and central chains, respectively. D)Stability analysis for the solution with (k, β)coupling = (1.10, 1.0). E) Stabilityanalysis for the solution with (k, β)coupling = (0.55, 0.45).

37

Figure 3-13. Structure of a unit cell of the model hexagonal system in three dimensions. 6particles located at corners of the hexagon (shown by closed circles) arereferred to as type A, 6 particles located at the centers of face plane (shownby open circles) are referred to type B, and the particle at the center of thehexagon is referred to as C. see text for detail.

Table 3-1. The neighbor list and equilibrium bond length for the particles in hexagonalsystem.

Particle index Neighbor indexes Equilibrium length(i,C1), dAC=1

(i,Aj) (i,Bj), (i,Bj−1), (i-1,Bj), (i-1,Bj−1), dAB = 1√(2)

(i,Aj+1), (i,Aj−1), (i+1,Aj), (i-1,Aj) dAA = 1(i,C1), (i+1,C1), dBC = 1

(i,Bj) (i,Aj),(i,Aj+1), (i+1,Aj), (i+1,Aj+1), dAB = 1√(2)

(i,Bi+1), (i,Bi−1) dBB =

√(3)

2

(i,C1) (i,A1−6) dAC = 1(i,B1−6),(i-1,B1−6) dBC = 1

lengths in the axial z direction have a unit length. The neighbor list of each particle and

equilibrium lengths between the interacting neighbors are listed in Table 3-1. Since we are

interested in the difference and transition of DBs in 1D and Q1D latice system, each Aj

particle along the axial direction can be considered as one 1D FPU chain, with type B and

C particles serving as connectors to couple the independent FPU systems.

3.3.1 Dispersion Relationship

We consider a system of 50 unit cells with periodic boundary condition imposed in

the axial direction. The dispersion curve ωj(q) along q = [001] direction is shown in

Figure 3-14. There are 4 acoustic branches in this system; 2 doubly degenerate transverse

38

0 1 2 30

1

2

3

Wave vector k

Fre

quen

cy

Figure 3-14. The dispersion curves for three-dimensional hexagonal tube system.

acoustic (TA) modes, which have x and y vibrations perpendicular to the axial (z)

direction. Single highest energy acoustic mode is the longitudinal acoustic (LA) mode in

the axial direction. The fourth acoustic mode is related to a rotation around the axis,

which is called a twisting mode (TW).

39

1 50−0.03

0

0.03

Dis

plac

emen

t

Unit Cell

XYZ

A

−1 0 1−1

0

1

Re(µ)

Im(µ

)

B

Figure 3-15. A) One of the linear phonon modes of the hexagonal system with(k,β)coupling/FPU = (1, 0). B) Stability analysis for this solution.

3.3.2 Steady State Solutions

Similar to the two dimensional systems considered in section 3.2.3, we first obtain

the steady state solutions for linear system (α = β = 0), which is obtained via

Newton’s method with phonon mode solutions as initial guess. The solution and its

Floquet multipliers are plotted in Figure 3-15. Compared to the linear solutions in two

dimensional system, the linear solutions are linearly stable due to the presence of higher

constraint in both axial and radial directions. More interesting, the gap in this unit circle

is only observed in this system and it exists for three of the linear vibration modes we

computed. To obtain nonlinear localized solution, we assign the breather solutions with

(k, β)FPU = (1, 2) to the axial direction displacement in each chain. In the calculation,

we gradually increase the values of kcoupling and βcoupling with given increasing ω value.

The structure for initial guess is shown in Figure 3-16. The dependence of amplitude and

frequency on varying kcoupling (βcoupling) while keeping βcoupling (kcoupling) fixed is shown in

Figure 3-17. As it is shown in this plot, increasing kcoupling will decrease the localization

of the solutions and while increasing βcoupling supports localized vibration modes. This

relation can be illustrated in the comparison of the configurations of nonlinear modes

where (k, β)coupling = (0.1, 0.1), (1,0.1) and (1,1) are plotted in Figure 3-20 A), B) and C)

respectively. It is clear that increasing kcoupling weakens the localization of the solutions;

while solution with larger values of βcoupling supports highly localized structures.

40

20 27 34

−0.35

0.35

Unit CellZ

Dis

plac

emen

t

Figure 3-16. Initial guess for the Newton’s method for hexagonal system. This vibrationmode corresponds to breather solution in 1D FPU system with k = 1 andβ = 2.

0.5 1

0.2

0.25

0.3

Am

plitu

de

0.5 1

2.51

2.61

2.71

Fre

quen

cy

βcoupling

A

0.5 10.3

0.35

0.4

0.45

Am

plitu

de

0.5 12.62

2.68

Fre

quen

cy

kcoupling

B

Figure 3-17. Dependence of amplitude and frequency of a steady-state mode on thestrength of coupling between chains in hexagonal system. A) βcoupling isvaried; kcoupling = 1 is fixed B) kcoupling is varied with βcoupling = 1 fixed.

Detailed structure of one of the converged nonlinear solution corresponding to

(k, β)coupling = (1, 1) is shown in Figure 3-18 along with its stability analysis. It shows

that increasing coupling strength reduces the magnitude of unstable Floquet multipliers.

The stability analysis in plot D indicates that this mode is still linearly unstable solution.

Since we remove the cause of instability from system configuration, here we examine the

instability from RWA approximation. Therefore, we seek the solution without using RWA

approximation and compare these two solutions. In other words, in the system without

applying RWA approximations, we solve the system of equations for xn,1 and xn,3 using

41

20 27 34

−1.e−2

1.e−2

Unit Cell

X D

ispl

acem

ent

A

B

C

A

20 27 34

−1.e−2

1.e−2

Unit Cell

Y

B

20 27 34

−0.3

0.3

Unit Cell

Z

C

−1 0 1

−1

0

1

Re(µ)

Im(µ

)

D

Figure 3-18. One of the nonlinear solutions with (k, β)coupling = (1, 1). Plots A), B) andC) show the displacement in x, y and z direction respectively. D Stabilityanalysis for this solution.

the Newton’s method. The initial guess are the solutions under RWA approximation. The

results from both approaches are plotted in Figure 3-19. The comparison of xn,1 vector for

both solutions is shown in Figure 3-19 A) and the higher frequency term, xn,3 is plotted in

Figure 3-19 B).

Firstly, the magnitude of xn,3 is 2 orders smaller than xn,1, and quantitative similarity

for xn,1 in two solutions indicate the appropriate assumption of ignoring higher frequency

terms in this system. In order to ensure the validity of RWA approximation in higher

dimensional systems, we compare their Floquet multipliers in Figure 3-21 A), where

the multipliers for the solution using and not using RWA approximation are shown

by diamonds and pluses, respectively. The comparison of two pairs of eigenvectors

corresponding to close Floquet multipliers are shown in Figure 3-21 and Figure 3-22. It is

clear that the first pair correspond to the same perturbation vector. For the second pair,

42

20 27 34−0.4

0.4

Unit Cell

Z D

ispl

acem

ent

3w

w

A

20 27 34−6.e−3

6.e−3

Unit Cell

Z D

ispl

acem

ent

B

Figure 3-19. The comparison of solutions with and without RWA. A) shows the cos(ωt)vector B) the cos(3ωt) vector

20 27 34

−0.35

0.35

Unit Cell

Z D

ispl

acem

ent

A

B

C

A

20 27 34

−0.35

0.35

Unit Cell

Z D

ispl

acem

ent

B

20 27 34

−0.35

0.35

Unit Cell

Z D

ispl

acem

ent

C

Figure 3-20. Comparison of nonlinear modes of hexagonal system for three differentstrengths of coupling between chains of atoms: A) (k, β)coupling = (0.1, 0.1);B) (k, β)coupling = (1.0, 0.1); C) (k, β)coupling = (1.0, 1.0); The linear andnonlinear forces weaken and strengthen localization of the solution structure,respectively.

even it has qualitative different value but these two vectors have the same period, which

implies the RWA approximation should still be valid in this system.

3.3.3 Conclusions

We have investigated the existence and stability of nonlinear vibration modes in

Q1D lattice model systems with various dimensionality. The DBs in 1D FPU system

are obtained from continuation of the nonlinearity strength in the interaction potential.

The unstable eigenvectors correspond to exciting the steady state breathers into moving

breathers. We also observed the complex stability pattern while introducing the cubic

43

−1 0 1

−1

0

1

Re(µ)

Im(µ

)

A

25 50 75 100

−0.15

0

0.15

B

Figure 3-21. A) Stability analysis for the solutions with (without) RWA approximationby diamond (cross) markers. B) Comparison of the z-direction eigenvectorsfor two close eigenvalues shown as open triangle and circular markers.Equivalent vectors are observed.

−1 0 1

−1

0

1

Re(µ)

Im(µ

)

A

25 50 75 100

−0.04

0

0.04

B

Figure 3-22. A) Stability analysis for the solutions with (without) RWA approximationby diamond (cross) markers. B) Comparison of the z-direction eigenvectorsfor two close eigenvalues shown as open triangle and circular markers.Different perturbation vectors are observed.

potential function. Due to the high frequency value for breather solutions, they have

extremely long life time in NEMD simulations. In higher dimensions model system,

we showed the existence of localized nonlinear vibration modes in both two and three-

dimensional systems. However, there are no linearly stable nonlinear vibration modes

under the particular interaction parameters that we explored in this chapter. We

44

also present the cause for instability due to interaction potential and as well as system

configuration.

45

CHAPTER 4NONLINEAR LATTICE VIBRATIONAL MODES IN CARBON NANOTUBES

Investigations of thermal conductivity of carbon nanotubes, along with various

other properties of these materials, is a subject of active research. Review of the recent

literature reveals that the values of thermal conductivity obtained in the experimental

studies [21, 22] significantly differ from the results of the molecular dynamics simulations

[23–26, 59]. Moreover, different MD simulations techniques lead to different results.

These studies have used two complementary MD techniques: the non-equilibrium and

equilibrium molecular dynamics simulations (referred to as NEMD and EMD respectively).

The NEMD simulations consist of imposing the temperature gradient along the

nanotube axis by coupling the opposing ends of the nanotube to the thermal baths

at different temperatures. This coupling is typically implemented by using one of the

thermostat techniques [60]. The thermal flux between the nanotube and the thermal bath

is computed and the thermal conductivity is obtained from the Fourier’s law (1–1). The

EMD simulations are based on simulations in an equilibrium ensemble and the thermal

conductivity is computed from the fluctuations of the thermal flux in the system using the

Green-Kubo formula of the linear response theory [4]. Hence, both the NEMD and EMD

methods are derived in the framework of the linear response theory, which assumes that

the relationship between the perturbation (i.e. temperature gradient) and the response

(i.e. the heat flux) is linear. The discrepancy between the results of these simulations

suggests that the assumption of linear response is not valid in the case of the carbon

nanotubes and nonlinear effects play a significant role.

Recent analyses [61, 62] have shown that nonlinear lattice vibrations in carbon

nanotubes may lead to formation of strongly nonlinear localized waves (solitons). These

studies have considered lattice vibrations in the continuum limit and approximated

nanotubes by a one-dimensional chain of atoms, thus neglecting the details of the lattice

vibration in the directions normal to the nanotube axis. Under these assumptions, it

46

has been shown that the vibrations of the carbon nanotubes can be approximated by

the Korteweg-de Vries (KdV) equation. It is well known that the soliton solutions of the

KdV equation possess constant speed and their interaction can be modeled by elastic

collisions. Therefore, the energy transfer by these solitons is purely ballistic as opposed

to the diffusive heat transfer assumed by the Fourier’s law Eq. 1–1. We note that this

treatment is somewhat approximate since in real nanotubes the heat transport will be

due to a mix of the ballistic and diffusive effects. In particular, it is expected that the

nonlinear localized structures will differ from the idealized KdV solitons in that the

collisions between them will be non-elastic and thus will lead to the transfer of energy

between the localized vibration modes.

In this chapter, we present the investigation of the nonlinear localized vibration

modes in carbon nanotubes, while accounting for the molecular details of the nanotube

structure into account.

4.1 System Configuration

The structure of a nanotube is obtained by rolling a graphite sheet (see Figure 4-1)

into a cylinder [63]. The position r of each carbon atom on the unrolled graphite sheet can

be described by two basis vectors a1 and a2,

r = l1a1 + l2a2, (4–1)

defined in Figure 4-1, where l1, l2 are integers. The structure of a carbon nanotube is

specified by the vector (−→OA) which corresponds to a section of the nanotube perpendicular

to the nanotube axis (−−→OB). A carbon nanotube is constructed by rolling the graphite

sheet so that points O and A coincide. The vectors−→OA and

−−→OB define the chiral vector,

Ch and translation vector T in axial direction. This chiral vector can be expressed in

terms of the basis vectors a1 and a2,

Ch = na1 + ma2 ≡ (n,m) (4–2)

47

The values of n and m determine the chirality of the nanotube, which in particular affects

the electronic conductivity and other properties of the nanotube. A carbon nanotube is

metallic if the value (n −m) is divisible by three and is semiconducting otherwise. Since

we focus on the contribution of lattice vibration modes on thermal transport, we consider

a zigzag semiconductor (5,0) CNT (see Figure 4-1 B).

A

B

Figure 4-1. A) Preparation of a computational model for a nanotube by rolling thegraphite sheet in a direction specified by the chiral vector Ch(n, m) [63]; B)nanotube with chiral vector Ch(5, 0).

We have implemented the Brenner parametrization [64] of the Tersoff potential,

referred to as VB, to describe the interaction between carbon atoms [65, 66]. This

potential has been used in the past for thermal conductivity calculations, as well as

for investigations of other properties of CNTs [67–69]. The explicit potential function is

48

Table 4-1. Parameters for Brenner-Tersoff interaction potential between carbon atoms.

Parameter Value

Re 1.39 AD 6.0 eVβ 2.1 A−1

S 1.22δ 0.5

R1 1.7 AR2 2.0 Aa0 0.00020813c0 330d0 3.5

presented in Eq. 4–3 and the parameters are listed in Table. 4-1.

Vi =∑

j( 6=i)[f(rij)A(rij)−Bijf(rij)C(rij)],

f(rij) = 12[1 + cos(π

rij−R1

R2−R1)], R1 ≤ rij ≤ R2,

A(rij) = DS−1

exp(−β√

2S(rij −Re)),

C(rij) = DSS−1

exp(−β√

2/S(rij −Re)),

Bij = [1 +∑

k(6=i,j) G(θijk)f(rik)]−δ,

G(θijk) = a0[1 +c20d20− c20

d20+(1+cosθijk)2

].

(4–3)

In order to ensure that the equilibrium positions of the carbon atoms are consistent

with the implemented potential. We use the conjugate gradients method [70] to adjust

the equilibrium atom coordinates so that the total potential energy of the nanotube

is minimized. Since our calculation require gradually increase the nonlinearity in the

potential function, we approximate the Brenner potential by a Taylor series up to the

fourth power,

un = −∑m

∂2V

∂un∂um

um − ε

[1

2

m,l

∂3V

∂un∂um∂ul

umul +1

6

m,l,k

∂4V

∂un∂um∂ul∂uk

umuluk

],

(4–4)

so that we can gradually increase the nonlinearity ε for atomic interaction. The 2nd, 3rd,

and 4th derivatives of the potential energy function VB that are necessary in the analysis

are computed numerically by the central finite difference scheme.

49

In order to estimate the importance of the nonlinearity of interatomic interactions in

the nanotube, we compute the ratio |Fn|/|Fl| of the averaged magnitudes of the linear and

the nonlinear components of the force acting on the carbon atoms,

|Fl| =

∣∣∣∣∣∑m

∂2V

∂un∂um

um

∣∣∣∣∣ (4–5)

|Fn| =

∣∣∣∣∣1

2

m,l

∂3V

∂un∂um∂ul

umul +1

6

m,l,k

∂4V

∂un∂um∂ul∂uk

umuluk

∣∣∣∣∣ . (4–6)

Here, the atomic displacements {un} are taken to be those of the normal phonon modes

with the amplitude assigned, according to the Boltzmann distribution at the temperature

T = 300 K, i.e,

1

2m

∑n

un2 =

1

2kBT. (4–7)

. The force ratios |Fn|/|Fl| for the nanotube phonon modes within a single unit cell system

are shown in Figure 4-2. For comparison, we show similar ratios for the FPU system

discussed in chapter 3. It is clear that the magnitudes of nonlinearity of these two systems

are comparable and therefore we expect that the carbon nanotubes possess the nonlinear

vibration modes qualitatively different from the normal phonon modes.

0.7 0.8 0.9 1

0.08

0.12

0.16

ω/ωmax

|Fn|/|

Fl|

FPU

(5,0) Nanotube

Figure 4-2. Ratio of the magnitude of the nonlinear |Fn| and linear |Fl| forces acting onthe equilibrium linear phonon modes of the (5,0) nanotube (♦) and the FPUlattice at α = β = 1 (•). Here, ωmax is the maximum phonon frequency of thesystem.

50

4.2 Non-Equilibrium MD Simulation

In order to assess the propagating structures in the carbon nanotubes during

heat transfer, we perform NEMD simulation for a 100 unit-cell (43.2 nm) nanotube

with chirality (5,0) and impose periodic boundary condition along z-axis. We set up

a temperature gradient by imposing two thermostats of 300K and 280K on the 50th

and the first unit cell of the nanotube via Berendsen algorithm. We use velocity Verlet

time integration scheme with time-step dt = 0.1fs. After the system has reached

equilibrium, the temperature gradient is obtained by a linear fit of the average nanotube

cell temperatures which gives T ′(z) = 0.45 Knm

. The developed temperature profile is

shown in Figure 4-3 A); the evolution of local energy distribution is shown in the Figure 4-

3 B). The local energy is the sum of the kinetic and potential energy of an individual

unit cell. This reveals that a part of energy is transferred by localized packets and that

ballistic transport mechanism does play an important role in carbon nanotube systems. To

determine the corresponding heat carriers for this phenomena, we solve the equations of

motion to obtain the steady state solutions in carbon nanotubes.

4.3 Steady States Solutions in CNTs

As described in chapter 2, we obtain steady-state solutions using Newton’s method.

Since VB potential is closely approximated by VT potential at ε = 1, initial guesses for the

Brenner modes are taken to be the Taylor modes with ε = 1 and the latter are obtained

by perturbing the linear normal modes (ε = 0) with analytical approximations from

perturbation theory , and using these approximations as initial guess for potential VT ,

followed by a continuation of the solution by increasing ε until it reaches 1.

The normal modes and dispersion curve are obtained from the following eigenvalue

problem [71]:

ω2un =∑m

∂2V

∂un∂um

um ≡ −F(un) (4–8)

where ω is the vibration frequency. We computed the phonon dispersion curves for the

(5,5) carbon nanotubes, and observed that these are in good agreement with the results of

51

10 20 30 40280

290

300

Position [nm]

Tem

pera

ture

[K]

A

B

Figure 4-3. NEMD simulations of a segment of a (5,0) CNT. The system contains 100unit cells and contains a heat source at the center (in the 50-th unit cell) anda heat sink in the first cell: A) Temperature profile of the system ; B) Thelocal energy transport in a steady-state system

molecular dynamics simulations reported in Ref. [26]. The dispersion curve for the (5,0)

carbon nanotubes is presented in Figure 4-4.

52

2 4 6

100

200

300

Wave vector k [1/nm]

Fre

quen

cy [

1012

Hz

]

Figure 4-4. Phonon dispersion curves of the (5,0) carbon nanotube under Brenner-Tersoffpotential.

The unknowns for equations of motion are the atomic displacement, which are

expressed using the Fourier series as in Eq. 2–15. The frequency ω is a fixed and given

parameter during the iterations of the Newton’s method. We impose periodic boundary

conditions in the axial direction and also apply RWA approximation, i.e, the coefficients

xn,j with j ≥ 2 are neglected. The numerical solution of Eq. 2–2 requires a function

which, given the Fourier series expansions for the atom displacements ul(t), computes

the Fourier series expansions for the force Fn(u) acting on the atoms. For sufficiently

simple potentials Fn(u) can be obtained analytically by direct substitution of u into

an expansion for F. For more complex potentials, such as the Brenner potential, we

use the following procedure summarized in Figure 4-5. Given Fourier coefficients x for

atomic displacements u, we perform inverse Fourier transform to obtain values of u(t)

at evenly spaced times tj within one period of oscillations. Then we compute forces

fk(tj) = fk(u(tj)) acting on each atom and perform Fourier transform on the force vectors

at times of t1, . . . , tj. Practically, we solve the equations of motion set up for zeroth

and first Fourier mode coefficients. The nonlinear solutions x(ε) are obtained through

continuation by Newton’s method from normal modes solutions, which represent ε = 0

solutions. The Jacobian matrix in Eq. 2–17 is obtained from central finite difference

scheme. We will refer to nonlinear modes of CNTs obtained with potentials VT and VB as

Taylor and Brenner modes (solutions), respectively.

53

Figure 4-5. The summary of the numerical procedure to obtain Fourier series expansionfor a complex potential. Here, fft and ifft denote fast Fourier transform andinverse fast Fourier transform.

Application of Newton’s method to potential VB is extremely time consuming, mainly

due to the necessity of using a finite-difference scheme to compute Jacobian matrix J .

On the other hand, J for VT potential can be obtained analytically, which allows us to

obtain the solutions relatively fast even in large systems. Therefore, nonlinear solutions

for periods exceeding 2 are obtained for VT potential only. Moreover, for systems with

period exceeding 2, we neglect the cubic terms and approximate VT potential as a sum

of quadratic and quartic terms. The solution corresponding to such a potential does not

include a static displacement term (see Eq. 3–2), which reduces the number of unknowns.

This simplifications allows us to obtain the steady state solutions with periods up to 24

unit cells.

4.4 Results

In this section, we present the solutions for the Taylor and the Brenner potentials for

different number of unit cells. In these calculations, ω is a given parameter and the results

for different values of ω illustrate the role of this parameter. First we describe the results

for a single unit cell nanotube, which consists of two carbon rings. The result in Figure 4-6

is obtained using the phonon mode with frequency ω0 = 3.125 × 1014 Hz as the starting

54

point. The steady-state vibration mode corresponding to the nonlinear Brenner-Tersoff

potential was obtained using a series of solutions of Eq. 4–4 with gradually increasing

magnitude of nonlinearity ε. The step for ε is chosen to be ∆ε = 0.02. The initial

condition for the solution of the nonlinear system at ε = 0.02 was obtained from the linear

mode using the regular perturbation analysis. We obtained ∆ω(ε = 0.02) = 1.25× 1011 Hz

from this perturbation analysis. During the process of the continuation by parameter ε,

ω was changed at every step by the same value, ∆ω = 1.25 × 1011 Hz up to β = 0.68

and then is kept as constant to β = 1. We note that although for the first step the value

of ω was dictated by the perturbation method, the changes in ω in the consecutive steps

were somewhat arbitrary with the goal to obtain a solution with frequency sufficiently

different from that of phonon modes. Once the continuation reached the value of ε = 1,

we used the Newton’s method again to obtain the vibration mode corresponding to the

complete Brenner-Tersoff potential. The nonlinear mode obtained by this method is

shown in Figure 4-6 and is compared with the corresponding linear phonon mode in the

same figure. The displacements in the x and y directions are very similar for atoms of

the two different rings of the nanotube unit cell (both for linear and nonlinear modes),

and thus, we show these displacements for only one of the rings. These results indicate

that the system described by the complete non-linear carbon nanotube potential possesses

vibration modes qualitatively different from those of the linearized systems. In addition

to the large difference in the dynamic displacement of the modes, the nonlinear modes

exhibit large static displacement (see Figure 4-6 B,E) that is absent in the linear modes.

The corresponding mode energy along the continuation by parameter ε is also presented

in Figure 4-7. The left hand side and right hand side y axis represents mode energy and

the frequency respectively, x axis is the value of nonlinearity ε. Even ω value in this

calculation is given arbitrarily, from the initial drop of mode energy for solution with

ε = 0.02 and the constant mode energy while ω is kept constant for ε > 0.68, which implies

the dependence of mode energy on ω.

55

0 0.2

−0.2

0

0.2

Z(nm)

Y(n

m)

[x20]

A

0 0.2

−0.2

0

0.2

Z(nm)

Y(n

m)

[x20]

B

0 0.2

−0.2

0

0.2

Z(nm)

Y(n

m)

[x20]

C

−0.3 −0.2 −0.1 0 0.1 0.2 0.3

−0.2

−0.1

0

0.1

0.2

X (nm)

Y (

nm)

[x20]

D

−0.3 −0.2 −0.1 0 0.1 0.2 0.3

−0.2

−0.1

0

0.1

0.2

X (nm)

Y (

nm)

[x11]

E

−0.3 −0.2 −0.1 0 0.1 0.2 0.3

−0.2

−0.1

0

0.1

0.2

X (nm)

Y (

nm)

[x100]

F

Figure 4-6. Comparison between a linear phonon mode of the (5,0) carbon nanotube andthe nonlinear mode obtained from this mode by the continuation method. Thefirst row of plots [ A), B), C)] shows the displacement of atoms in the zdirection. For clarity, the atoms belonging to the same ring are plotted on thesame line. The second row of plots [ D), E), F)] shows the atom displacementin the x and y directions for one of the nanotube rings. The second ringexhibits similar displacements and hence is not shown. The first column ofplots [ A), D)] shows the linear solution; the second [ B), E)] and the third[ C), F)] columns show respectively the static and the dynamic displacementsof the nonlinear mode. The displacements are shown by arrows which forclarity are magnified by a factor of 20 [all plots except F)] or by a factor of100 [plot F)].

In the two unit cells system, we use ∆ω obtained from degenerate perturbation

for the first step in the continuation curve. During the process of the continuation by

parameter ε, ω is changed only at the first step. The nonlinear mode configuration shown

in Figure 4-8 is obtained using the phonon mode with frequency ω0 = 1.2 × 109 Hz as

the starting point with mode energy corresponding to 350K and ∆ω(ε = 0.02) = −3.1 ×108 Hz. The value of frequency is kept constant until ε reaches 1. In addition, to explore

the energy threshold of initial mode energy for exciting the localized nonlinear vibration

56

0 0.2 0.4 0.6 0.8 10

200

400

Mod

e E

nerg

y [K

J / M

ol]

β0 0.2 0.4 0.6 0.8 1

3.1

3.15

3.2

ω [1

014 H

z]

β

Figure 4-7. Dependence of the mode energy and frequency of the Taylor solutions on ε forsolutions shown in Figure 4-6. The mode energy is shown by diamonds. Thevalue of frequency ω is shown by pluses.

modes as we observed in section 3.2.3 in Q1D model system, we start with the normal

modes with thermal energy corresponding to 200K, 250K and 350K. Below we show

the displacement for the first carbon ring and only compare the dynamical displacement

between nonlinear and linear solutions for the three temperatures. The results for both

Taylor and Brenner solutions and the corresponding mode energy continuation curves

are shown in Figure 4-8. Plots A and D are the displacement for linear system; B and E

are the Taylor solutions for three different initial thermal energy; C and F are Brenner

solutions. This nonlinear vibration modes have configuration similar to the phonon mode

where we started the calculation. The results starting with different temperatures lead

to solutions with similar configuration and Brenner solution is very similar to Taylor

solutions, which justifies our numerical process. The mode energy for these three solution

branches along the continuation by parameter ε are presented in Figure 4-9. Because ωnl

is close to ω0 we obtain the vibration modes with configuration qualitatively similar to the

phonon modes.

For system with simplified potential, we obtain nonlinear vibration modes for 4, 8,

16 and 24 unit cells. In this potential function, we are able to reach nonlinear vibration

57

−0.2 0

−0.2

0

0.2

Phonon; ω=82.4532

[x80]

X (nm)

Y (

nm)

[x80][x80]

A

−0.2 0 0.2

−0.2

0

0.2

ε=1

[x300][x300][x300]

X (nm)

Y (

nm)

B

0 0.2

−0.2

0

0.2

[x80]

Z (nm)

Y (

nm)

[x80][x80]

(D)

C

0 0.2

−0.2

0

0.2

[x80]

Z (nm)

Y (

nm)

[x80][x80]

(D)(D)

D

0 0.2

−0.2

0

0.2

[x300][x300][x300]

Z (nm)

Y (

nm)

E

0 0.2

−0.2

0

0.2

[x300][x300][x300]

Z (nm)

Y (

nm)

F

Figure 4-8. The solutions of (5,0) CNT in two unit cells system starting with theamplitude corresponding to three different temperature. The first row of plots[ A), B), C)] shows the displacement of atoms x and y directions for one ofthe nanotube rings. to the same ring are plotted on the same line. The secondrow of plots [ D), E), F)] shows the atom displacement in the z direction. Forclarity, the atoms belonging The first column of plots [ A), D)] shows thelinear solution; the second [ B), E)] and the third [ C), F)] columns showrespectively the Taylor and Brenner solutions. The displacements are shownby solid lines which for clarity are magnified by a factor shown in the leftcorner in each subfigures.

modes for ε = 1 and ∆ω = 8.6 × 1011 Hz in one step of Newton’s method calculation.

An example of the nonlinear vibration modes with longest length, 24 unit cells is shown

in Figure 4-10, in which the solution is obtained using the phonon mode with frequency

ω0 = 1.58 × 1014 Hz. Since it is a long system, we present this nonlinear vibration mode

by plotting the dynamical displacement for the 7th atom in each unit cell. Plots A), B)

and C) are the atomic displacement for the nonlinear vibration mode along x, y and z

directions, respectively; and plots D), E) and F) are the dynamical displacement for

linear phonon mode along x, y and z directions, respectively. The nonlinear solution has

58

0 0.2 0.4 0.6 0.8 10

2

4

ε

Mod

e E

nerg

y [ K

J/M

ol ]

0 0.2 0.4 0.6 0.8 182.4526

82.4528

82.453

82.4532

0 0.2 0.4 0.6 0.8 182.453

82.4531

82.4532

0 0.2 0.4 0.6 0.8 182.4528

82.453

82.4532

Fre

quen

cy [

1012

Hz

]

ε

200K250K350K

Figure 4-9. Dependence of the mode energy and frequency of the Taylor solutions startingwith three initial thermal energy on ε for solutions shown in figure 4-8. Themode energy is shown by diamonds. The value of frequency ω is shown bypluses.

the qualitatively similar configuration compared to the initial linear phonon mode. We

observe that in this case the solution is qualitatively different from the initial phonon

mode. In order to summarize the nonlinear solutions for this potential function, we mark

the converged solutions on the dispersion curve in Figure 4-11. Plot A shows the solution

frequency, open (closed) marker represent the solutions with different (same) wave number

(periodicity) from the initial phonon modes. In order to quantify the difference between

nonlinear modes and phonon modes, we normalize the nonlinear solutions and compute

the norm of the difference between these two solutions, the results are plotted in plot B.

It is noted that large differences occurs for solutions for which the wave number equals π2a

,

where a is the length of one unit cell in z direction.

We also perform Floquet analysis for the nonlinear solution in the 4 unit-cell system.

We obtain both linearly stable and unstable vibration modes. The configuration and their

Floquet multipliers are shown in Figure 4-12 and Figure 4-13, respectively. To investigate

the cause for the instability, we plot the eigenvector corresponding to the open circle in

Figure 4-13B in Figure 4-14. We still need to analyze this eigenvector to fully understand

the effect of this unstable perturbation on the nonlinear vibration mode.

59

1 24−0.02

0

0.02

X

Unit Cell

A

1 24−0.04

0

0.04

X

Unit Cell

B

1 24−0.02

0

0.02

Y

Unit Cell

C

1 24−0.04

0

0.04

Y

Unit Cell

D

1 24−0.04

0

0.04

Z

Unit Cell

E

1 24

−0.06

0

0.06

Z

Unit Cell

F

Figure 4-10. The comparison the displacement for the 7th atom in each unit cell for ε = 1nonlinear mode and starting phonon mode in a 24 unit cells system with thesimplified potential under RWA approximation. The first row of plots [ A),B), C)] shows the displacement of atoms for solution. The second row ofplots [ D), E), F)] shows the atom displacement for the starting phononmode. The first, second and third column of plots shows the displacement inx, y and z direction. displacement for the first, fourth and 7th atom in eachunit cell.

4.4.1 Conclusions

We have investigated the thermal transport mechanism in a semiconductor carbon

nanotube. In non-equilibrium MD simulations, localized packets are observed, and this

indicates the important of ballistic transport mechanism in carbon nanotubes. We have

explored the nonlinear vibration modes in carbon nanotubes under Brenner-Tersoff

potential, as well as with simplified force coefficients. Nonlinear vibration modes with

qualitatively different configuration compared to linear phonon modes are obtained.

Among the solutions obtained, there are some stable nonlinear vibration modes. Even

though the structures are not localized, the roles of these nonlinear modes, including

their group velocity and phonon mean free path need further investigation in order to

describe the dynamics of vibration modes in carbon nanotubes using Boltzmann transport

equation.

60

0 1 2 3 4 5 6 70

5

10

15

20

25

30

Wave number [1/nm]

Fre

quen

cy [1

013 H

z]

A

0 1 2 3 4 5 6 70

5

10

15

20

25

30

Wave number [1/nm]

Fre

quen

cy [

1013

Hz]

B

Figure 4-11. The nonlinear vibration modes with simplified potential for the systems of 4,8, 16 and 24 unit cells, which are represented by circles, triangles, diamondsand squares respectively. Open and closed markers represent the nonlinearvibration modes having wave number different from the phonon modes westarted the calculation. A) location of ω value. B) norm of the differencebetween normalized ε = 1 and ε = 0 solutions.

−0.2 0 0.2

−0.2

0

0.2

X (nm)

Y (

nm)

Mode59

[x200]

0 0.1 0.2 0.3−0.2

0

0.2

Z(nm)

Y(n

m)

[x400]

Nonlinear soln ; ω=943.2252

−0.2 0 0.2

−0.2

0

0.2

X (nm)

Y (

nm)

ω=943.2204

[x200]

0 0.1 0.2 0.3−0.2

0

0.2

Z(nm)

Y(n

m)

[x400]

Phonon Modes

A

−1 0 1−1

0

1

−1

0

Re(µ)

Im(µ

)

B

Figure 4-12. A linearly stable nonlinear solution based on simplified potential function fora four unit cells system. A) the solution configuration in the first carbon ringB) Floquet multipliers of this solution

61

−0.2 0 0.2

−0.2

0

0.2

X (nm)

Y (

nm)

Mode171

[x100]

0 0.1 0.2 0.3−0.2

0

0.2

Z(nm)

Y(n

m)

[x100]

Nonlinear soln ; ω=323.2077

−0.2 0 0.2

−0.2

0

0.2

X (nm)

Y (

nm)

ω=323.1938

[x100]

0 0.1 0.2 0.3−0.2

0

0.2

Z(nm)

Y(n

m)

[x100]

Phonon Modes

A

−1 0 1−1

0

1

−1

0

Re(µ)

Im(µ

)

B

Figure 4-13. The nonlinear solution based on simplified potential function for a four unitcells system. Figure A) shows the atomic displacement and its Floquetmultipliers are shown in plot B)

−0.2 0 0.2

−0.2

0

0.2

X (nm)

Y (

nm)

[x1]

A

0 0.1 0.2 0.3

−0.2

0

0.2

Z(nm)

Y(n

m)

[x950]

B

Figure 4-14. The eigenvector corresponding the Floquet multiplier shown as the opencircle in Figure 4-13B.

62

CHAPTER 5EFFECTS OF SORBATE MOLECULES ON THERMAL TRANSPORT IN ZEOLITES

In this and the following chapters we present the investigation of thermal transport

in nanoporous materials including the molecular dynamics simulations for encapsulated

sodalite zeolites and the model development for the sorbate molecule in a 1D model is

reported in chapter 6.

Zeolites are microporous alumino-silicate crystals with pore sizes comparable to

molecular dimensions. They are widely used as molecular sieves, sorbents, catalysts,

and ion exchangers. In addition, several possible applications of zeolites combining

adsorption of guest molecules with temperature control are emerging. Examples include

solar adsorption heat pump [72] and adsorption chillers for microelectronic devices [73].

A wide variety of available zeolite structures provides a large range of flexibility in

fine-tuning their thermal properties to a specific task. Zeolite thermal conductivity can

also be altered by introducing point defects into the crystal lattice, e.g., by substitution

of some of the silicon atoms by aluminum [74]. The point defects provide additional

scattering centers for phonons thus reducing the crystal thermal conductivity [75].

Moreover, the nanoporous structure of zeolites provides an additional opportunity to

control the zeolite thermal properties through introduction of off-framework guest species

(sorbates or cations) into the crystal. In fact, it is well known that strong interaction

between zeolite lattice vibrations and guest molecules significantly affect transport

properties of the guest molecules (see e.g. [76–78]) as well as the dynamics of oscillation

of sorbate molecules at their adsorption sites [79]. Evidence is accumulating that presence

of guest molecules within zeolites also affects the dynamics of zeolite lattice vibrations

leading to changes in thermal properties of zeolites [29, 80].

Similar host-guest interactions play a significant role in thermal conductivity of

other nanoporous materials of technological importance. For example, skutterudites are

promising candidates for development of efficient thermoelectric materials, i.e. materials

63

with high ratio of electrical and thermal conductivities. It has been shown [27, 28] that

addition of ions to voids in skutterudites leads to an order of magnitude decrease of their

thermal conductivity thus increasing the thermoelectric figure of merit.

Available data suggest complex dependence of the thermal conductivity on the

nature of a guest molecule and a host crystal. As discussed above, addition of ions to

skutterudites reduces their thermal conductivity. Similarly, encapsulation of an atom in

a Ge clathrate leads to an order of magnitude reduction in thermal conductivity [81]. In

addition, molecular dynamics (MD) simulations [29] show a drastic reduction of thermal

conductivity of zeolite LTA-SiO2 in the presence of heavy cations. These results seem to

suggest that “rattling” of the guest species inside a crystal leads to scattering of phonons

thus reducing their mean free path and leading to the decrease of thermal conductivity.

However, other observations contradict this picture. For example, MD simulations

of xenon in zeolite LTA-SiO2 indicate an increase of the zeolite thermal conductivity

due to the guest-host interactions [29]. In addition, experiments of Greenstein et al.

[80] show that the conductivity of zeolite MFI is substantially higher when an organic

template cation tetrapropylammonium (TPA) is present in it, as compared to a sample

with removed templates.

Therefore, the phonon-guest interactions may be qualitatively different from

the phonon scattering by point defects fixed in the lattice due to strongly nonlinear

oscillations of guest molecules inside the crystal pores. Interactions of guest molecules

with host lattice vibrations have been extensively modeled in recent years [78, 79, 82–84].

Typically, the goal of these studies is to understand effects of the lattice vibration on the

sorbate dynamics inside crystals and the lattice vibrations are frequently modeled as a

thermal bath. In the current work, we are aiming at understanding the reverse process,

i.e. effects of the sorbate “rattling” on the crystal lattice vibrations. In this chapter

we present results of our investigations of effects of sorbate molecules on dynamics of

individual phonons. It is expected that understanding of sorbate-phonon interactions

64

will lead to a better understanding of the sorbate effects on thermal conductivity of

nanoporous materials. We consider a relatively simple system, namely a sodalite zeolite

with small molecules (argon, xenon, or methane) trapped inside its cages. Forces between

these sorbates and zeolite lattice atoms are short-ranged and, when the sorbate size is

sufficiently small, the interaction between the sorbate and the phonons takes place only

during collisions between the sorbate and the zeolite wall. Therefore, the current model

allows us to focus on effects of the sorbate “rattling” on the phonon dynamics and our

observations are not obscured by other possible contributions of guest molecules to the

lattice dynamics, such as long-range electrostatic interactions between adsorbed charged

species and lattice ions.

5.1 Model Details

The crystal structure of silica sodalite Si12O24 is shown in Figure 5-1. This zeolite

possesses a cubic symmetry and its lattice parameter is 8.83 A, see [85]. A sodalite

unit cell consists of a cage shaped like a truncated cuboctahedron bounded by six 4-

ring windows (i.e. windows formed by four oxygen and four silicon atoms) and eight

6-ring windows. Diameter of the 4-ring windows is very small and these windows are

impermeable by the sorbates considered in our simulations. In addition, transport rates of

argon and larger molecules (methane and xenon) through the 6-ring windows are orders

of magnitude slower than the phonon-phonon and phonon-sorbate interactions [77] and a

sorbate remains inside a cage on the time-scale of interest.

Silica sodalite is usually synthesized by growing the crystal around organic template

molecules which become encapsulated in the sodalite cages after the synthesis is complete

[85, 86]. In the current work, we neglect the presence of encapsulated templates and

perform MD simulations of either bare silica sodalite containing only Si and O atoms in its

lattice structure or silica sodalite with encapsulated argon, methane, or xenon molecules.

The equilibrium lattice configuration and the potential model for zeolite lattice

vibrations used in this study are the same as those used by Kopelevich and Chang [77]

65

Figure 5-1. A block of 2× 2× 2 sodalite unit cells containing nine sodalite cages. Siliconand oxygen atoms are shown as larger and smaller spheres, respectively. InMD simulations of sorbate-lattice systems, the sorbates are located in eightcorner cages of this block.

in a study of sorbate transport through 6-ring windows. Zeolite lattice vibrations are

modeled by a truncated version of an anharmonic potential energy field proposed by

Nicholas et al. [87],

V (u) =∑O−Si

Vr(rO−Si) +

∑O−Si−O

Vα(αO−Si−O) +∑

Si−O−Si

[Vβ(βSi−O−Si) + VUB(rSi−Si)

], (5–1)

where

Vr(r) =1

2Kr(r − rO−Si

0 )2 (5–2)

is the potential energy of stretching of the O-Si bond r,

Vα =1

2Kα(α− α0)

2 (5–3)

66

is the potential energy of bending of the O-Si-O bond angle α,

Vβ =1

2

[K

(1)β (β − β0)

2 −K(2)β (β − β0)

3 + K(3)β (β − β0)

4]

(5–4)

is the potential energy of bending of the Si-O-Si bond angle β, and

VUB(r) =1

2KUB(r − rSi−Si

0 )2 (5–5)

is the Urey-Bradley term which represents lengthening of the Si-O bond as the Si-O-Si

angle becomes smaller. In Eq. 5–5, r denotes distance between two silicon atoms of a

Si-O-Si angle.

The potential model Eq. 5–1 neglects smaller contributions included in the original

model [87] such as torsion energy of the dihedral Si-O-Si-O angle, nonbonded Lennard-

Jones interaction, and electrostatic interaction between zeolite atoms due to the partial

charges of Si and O atoms. In principle, long-range electrostatic interactions may have

a significant effect on the lattice dynamics. However, it has been shown in [87] that

electrostatic interactions have little effect on the structure or dynamics of the silica

sodalite lattice due to high symmetry of this crystal and charge neutrality of each SiO2

group. Moreover, since the sorbates considered in the current work are electrically neutral,

electrostatic interactions with the partial charges of the lattice atoms are expected to have

negligible effects on the sorbate dynamics. In fact, this truncated model has been shown

to yields good agreement between computed and experimental values of transport rates

of inert gases in sodalite [77]. Since this transport process involves large deformations of

6-ring windows, it is expected that, despite the introduced approximations, the model

Eq. 5–1 provides an accurate description of anharmonic lattice dynamics.

The values of the force constants K as well as the values of the equilibrium distances

(rO−Si0 and rSi−Si

0 ) and equilibrium bond angles α0 and β0 are summarized in Table 5-

1. The force constants and the equilibrium bond angles were taken from the paper

of Nicholas et al. [87]. The equilibrium distances rSi−Si0 and rO−Si were obtained from

67

Table 5-1. Parameters for the lattice potential energy model Eq. 5–1.

Si-O Kr = 2500.1 kJ mol−1 A−2 rO−Si0 = 1.58 A

O-Si-O Kα = 578.1 kJ mol−1 rad−2 α0 = 109.5◦

Si-O-Si K(1)β = 45.4 kJ mol−1 rad−2 β0 = 149.5◦

K(2)β = 95.1 kJ mol−1 rad−3

K(3)β = 55.5 kJ mol−1 rad−4

Si-Si KUB = 228.5 kJ mol−1 A−2 rSi−Si0 = 3.1219 A

Table 5-2. Lennard-Jones parameters for sorbate-sorbate and sorbate-lattice interactions.

Interaction ε (J/mol) σ (A)Ar-Ar [88] 1183.0 3.350Ar-O [88] 1028.0 3.029

CH4-CH4 [89] 1139.0 3.882CH4-O [90] 1108.3 3.214Xe-Xe [88] 3437.0 3.849Xe-O [91] 1133.1 3.453

a requirement that the equilibrium crystal structure predicted by the potential field

coincides with the structure obtained experimentally by Richardson et al. [85]. The value

of rO−Si0 used in this work is somewhat different from that proposed by Nicholas et al. [87]

due to the differences in the potential as discussed in detail in Ref. [77].

In order to assess sorbate size effects on the phonon-sorbate interactions, we consider

three different sorbates, namely argon, methane, and xenon. All these sorbates are

modeled as spheres which interact with each other and the zeolite lattice atoms through

the Lennard-Jones potential. The values of the Lennard-Jones parameters ε and σ used

in our calculations are listed in Table 5-2. The sorbate-lattice interactions are modeled

using the common assumption [92] that the interaction between sorbates and lattice silicon

atoms can be neglected and the only contribution to the sorbate-lattice potential energy is

due to interaction between sorbates and lattice oxygen.

5.2 Simulation details

The simulations were performed at temperature T = 300 K for a 2 × 2 × 2 block

of sodalite unit cells satisfying the periodic boundary conditions. Initial configurations

68

for the simulations were prepared as follows. First, sorbate-free zeolite was considered.

Initial conditions were generated by placing zeolite atoms at their equilibrium positions

and sampling their velocities from the Maxwell distribution. The bare zeolite lattice was

then equilibrated for 10 ns using NVT simulations with Berendsen thermostat [31] with

the time constant 1 ps.

After this equilibration, one or more sorbate molecules were added to each of the

corner cage of the 2 × 2 × 2 sodalite block (see Figure 5-1) and the sorbate-lattice system

was equilibrated using the NVT simulations for an additional 2 ns. The initial locations

for the sorbates were taken to correspond to the minimum of the sorbate-zeolite potential

energy. Due to small size of the sodalite cages, only one xenon molecule and no more than

two argon or methane molecules can be placed into a single cage. Therefore, we consider

the following five sorbate-lattice systems: one sorbate (Ar, CH4, or Xe) per unit cell

and two sorbates (Ar or CH4) per unit cell. We will denote these systems as 1 Ar/cage,

1 CH4/cage, 1 Xe/cage, 2 Ar/cage, and 2 CH4/cage.

The equilibration was followed by a 5 ns production run of NVE simulations for each

of these five sorbate-lattice systems and the sorbate-free zeolite. Since one of the main

goals of this work is to assess chaotic nonlinear dynamics of phonons, we chose a fairly

small step size, ∆t = 0.1 fs, for the microcanonical simulations. This step size ensures that

the total energy fluctuations are on the order of 0.001%.

In order to demonstrate that the sorbate “rattling” inside the zeolite cage is

qualitatively different from harmonic or nearly harmonic oscillations of point defects

in the lattice, we performed additional simulations of sorbate dynamics in the absence

of the sorbate-phonon interactions. In these simulations, the zeolite atoms were fixed at

their equilibrium positions. The sorbates were initially placed at positions corresponding

to the minimum of the sorbate-zeolite potential energy and their velocity was sampled

from the Maxwell distribution. The system was then equilibrated for 2 ns using the NVT

simulations, which were followed by a 5 ns NVE production run. The parameters of these

69

NVT and NVE simulations are the same as those of the simulations of the flexible zeolite

systems.

5.3 Normal modes of sodalite crystal

Before presenting analysis of our MD simulations in sections 5.4 and 5.5, we briefly

review background information on harmonic lattice dynamics and calculation of sodalite

normal modes.

Consider a crystal modeled by a periodically repeated block of L1 × L2 × L3 unit cells.

Each unit cell contains N atoms and two vector sets {a1, a2, a3} and {b1,b2,b3} form

bases of the unit cell and the reciprocal lattice, respectively; ai · bj = 2πδij.

Let integer vector l = (l1, l2, l3) specify the unit cell with coordinates

r(l) = l1a1 + l2a2 + l3a3, li = 0, . . . , (Li − 1). (5–6)

In addition, let r(lκ) denote the coordinates of the κ-th atom within the l-th unit cell and

u(lκ) = r(lκ) − req(lκ) denote displacement of this atom from its equilibrium position

req(lκ). Then the normal mode coordinates Qjk are defined as the projection of the

Fourier transform of u(lκ) on eigenvectors e(jk) of the Fourier transform D(k) of the

dynamical matrix [71]. Here, j is a number of the eigenmode and k is a wavevector,

k =h1

L1

b1 +h2

L2

b2 +h3

L3

b3, hi = 0, . . . , Li − 1, i = 1, 2, 3. (5–7)

In what follows, we will use crystallographic notation for this vector, i.e. the right-hand

side of equation (5–7) will be written as k = [h1h2h3].

Elements of the matrix D(k) are given by

Dαβ(κκ′;k) = (mκmκ′)−1/2 ∑

l′

(∂2V

∂rα(lκ)∂rβ(l′κ′)

)∣∣∣r=req

e−ik·r(l−l′),

α, β = 1, 2, 3, κ, κ′ = 1, . . . , N.

(5–8)

70

Solution of the eigenvalue problem for this 3N × 3N matrix,

N∑

κ′=1

3∑

β=1

Dαβ(κκ′;k)eβ(κ′; jk) = ω2jkeα(κ; jk), j = 1, . . . , 3N, (5–9)

yields frequencies ωjk of the normal mode vibrations.

The normal mode coordinates can now be defined as follows [71]

Qj(k) = (L1L2L3)−1/2

l

N∑κ=1

3∑α=1

m1/2κ uα(lκ)e∗α(κ; jk)e−ik·r(l). (5–10)

In the case of a harmonic and sorbate-free lattice, the normal modes are independent

of each other and the Hamiltonian of each mode is

Hjk =1

2|Qjk|2 +

1

2

(ωh

jk

)2 |Qjk|2. (5–11)

Here and in the remainder of the chapter, we use superscript h to distinguish frequency of

a normal mode in a harmonic lattice from that in an anharmonic lattice.

The sodalite unit cell contains N = 36 atoms. Therefore, there are 108 normal

modes corresponding to each wavevector k. Since in the current work we consider a

2 × 2 × 2 block of unit cells and sodalite possesses a cubic symmetry, there are only

four independent wavevectors, k = [000], [100], [110], and [111], in our MD system. For

reference, the dispersion relationships ωhj (k) for sodalite for wavevectors k pointing in

directions [100], [110], and [111] are shown in Figure 5-2. Normal modes accessible by the

MD simulations correspond to the smallest and the largest values of |k| in each of these

three plots.

In general, encapsulation of a sorbate inside a zeolite cage may affect the linearized

lattice dynamics and lead to changes in the eigenvectors of the dynamical matrix, which

would require one to use different normal modes in the analysis of zeolite lattice vibration

in the presence of sorbates. However, this effect is significant only in the case of strong

interaction between phonons and a sorbate located at an equilibrium adsorption site.

This situation would occur if the sorbate fits tightly within a zeolite cage or if there are

71

0 0.05 0.1 0.15 0.2 0.25 0.3 0.350

20

40

60

80

100

120

140

160

180

200

220

|k|, (A−1)

ωh

(ps−

1)

A

0 0.1 0.2 0.3 0.4 0.50

20

40

60

80

100

120

140

160

180

200

220

|k|, (A−1)

ωh

(ps−

1)

B

0 0.1 0.2 0.3 0.4 0.5 0.60

20

40

60

80

100

120

140

160

180

200

220

|k|, (A−1)

ωh

(ps−

1)

C

Figure 5-2. Dispersion relationships for sodalite in directions (A) k = [100], (B) k = [110],and (C) k = [111].

long-range electrostatic interactions between the sorbate and the zeolite lattice. It has

been shown in [84] that coupling between the phonon modes and small electrically neutral

sorbates, such as those considered in the current work, is negligible when the sorbates are

located at adsorption sites inside sodalite cages. For these systems, the sorbate-phonon

coupling becomes significant only when the sorbate approaches the zeolite wall. Therefore,

the presence of the sorbates does not affect the dynamical matrix of the crystal.

The nonlinear effects of the sorbates, such as the change in phonon lifetime and the

nonlinear corrections to the phonon frequency are analyzed in the next two sections.

The insensitivity of the harmonic approximation to the lattice dynamics to the sorbates

considered in the current work allows us to use the normal modes of the sorbate-free

72

zeolite to analyze the lattice dynamics in the presence of the sorbates. In the remainder

of the chapter, we use harmonic phonon frequencies ωhjk to parameterize various plots of

properties of individual modes. This allows us to consistently identify a phonon mode even

if its frequency is changed upon addition of various sorbates.

5.4 Nonlinear phonon and sorbate dynamics

Nonlinear phonon dynamics is usually modeled by the Boltzmann transport equation

BTE [75]

vjk · ∇T∂njk

∂T=

(∂njk

∂t

)

collision

, (5–12)

where T is the temperature and vjk and njk are the group velocity and the occupation

number of the phonon mode jk. The occupation number njk is proportional to the

harmonic potential energy of the mode jk, see, e.g., [93]. Phonon-phonon interactions due

to lattice anharmonicity, defects, and other factors are modeled by the collision integral

in the right-hand-side of Eq. 5–12. Calculation of the collision integral in exact form is

very challenging and it is usually approximated by various models, such as the single mode

relaxation time (SMRT) approximation,

(∂njk

∂t

)

collision

=neq

jk − njk

τjk

. (5–13)

Here, neqjk is the equilibrium phonon distribution and τjk is the relaxation time of the

phonon mode jk, which is assumed to coincide with the phonon lifetime. The relaxation

time can be estimated from MD simulations by computing relaxation times of the

occupation number [93] or the harmonic energy (5–11) of the normal mode [94].

The analysis of the phonon dynamics based on BTE approach has several drawbacks.

The relaxation time approximation may not be adequate to model phonon dynamics

in complex materials. Although SMRT approximation (5–13) can be extended to

account for multiple relaxation times due to different phonon scattering mechanisms,

these extensions require the phonon autocorrelation function to be a sum of multiple

exponentials. However, as will be shown below, some of the sodalite phonon modes do not

73

satisfy this requirement. In this case, a straightforward extension of the relaxation time

model does not seem to be possible. In addition, the relaxation time approximation may

not be appropriate to account for phonon-sorbate interactions.

In order to attain more flexibility in modeling phonon dynamics, we choose an

alternative model based on Langevin equation [12],

Qjk + γjkQjk = (ωajk)

2 (Qjk − 〈Qjk〉) + Γjk(t). (5–14)

Here, 〈Qjk〉 is the mean value of the normal mode coordinate Qjk, which may differ from

zero due to anharmonicity of the system, ωajk is the anharmonic frequency, γjk is the

friction coefficient, and Γjk(t) is the random force with zero mean which is related to the

friction coefficient by the fluctuation-dissipation theorem

〈Γjk(t)Γjk(t + τ) = 2kBTγjkδ(τ). (5–15)

We assume that the deterministic force in this Langevin model is still harmonic and the

anharmonicity of the oscillations can be adequately captured by shifts in the phonon

frequency and average normal mode coordinates, as well as friction and stochastic forces.

The autocorrelation function of normal mode Qjk satisfying equation (5–14) is [95]

Cjk(τ) = 〈δQjk(t)δQ∗jk(t + τ)〉 = e−τ/τjk cos ωjkτ. (5–16)

Here,

δQjk = Qjk − 〈Qjk〉 (5–17)

is the deviation of the phonon mode from its average,

τjk = −2/γjk (5–18)

is the phonon lifetime, and

ωjk =√

(ωajk)

2 − 1/τ 2jk. (5–19)

74

is the apparent phonon frequency. Eq. 5–16 demonstrates that, similarly, to BTE with

SMRT approximation, the Langevin model (5–14) predicts an exponential decay of

the normal mode autocorrelation function. However, the Langevin equation provides

more flexibility and allows one to perform relatively simple adjustments of the model to

fit observed phonon dynamics. This can be done by including an explicit anharmonic

term into the equation or modifying statistics of the random force. For example, non-

exponential behavior of the autocorrelation function can be modeled by a non-Markovian

random force [96].

Examples of phonon autocorrelation functions obtained from our MD simulations

of sorbate-free sodalite are shown in Figure 5-3. Most of these functions, such as that

shown in Figure 5-3a, are in good agreement with Eq. 5–16. However, some modes

exhibit significant deviations from the predictions of this Markovian Langevin equation.

Autocorrelation functions of all these non-Markovian modes are qualitatively similar

to one of the autocorrelation functions shown in Figure 5-3b-d. These modes possess

secondary (slow) oscillations which are qualitatively similar to oscillations of the

autocorrelation function of the energy of an entire sodalite cage observed by McGaughey

and Kaviany [97]. These secondary oscillations were interpreted as corresponding to

localization of energy in sodalite cages. Our analysis of individual phonon modes shows

that only a fraction of optical phonons modes possesses this secondary time-scale.

In the current work we assume that the Markovian Langevin equation (5–14) provides

an adequate model for the phonon dynamics and leave development of its extensions

to acount for secondary oscillations to future studies. Therefore, the lifetimes and

the frequencies of all modes are obtained by applying Eq. 5–16 to analysis of their

autocorrelation function

In particular, Eq. 5–16 implies that the power spectrum Sjk(ω) of the phonon mode

jk attains its maximum at ω = ωjk. Therefore, we define the apparent frequency of

vibration, ωjk, as the location of the maximum of Sjk(ω) for all modes, including those

75

0 2 4 6−1

−0.5

0

0.5

1

τ (ps)

Cjk

A j = 7

0 2 4 6−1

−0.5

0

0.5

1

τ (ps)

Cjk

B j = 23

0 2 4 6−1

−0.5

0

0.5

1

τ (ps)

Cjk

C j = 64

0 20 40−1

−0.5

0

0.5

1

τ (ps)C

jk

D j = 104

Figure 5-3. Examples of autocorrelation functions Cjk(τ) of phonons in a sorbate-freesodalite crystal. The normal mode numbers j are shown in the correspondingplots; the wavevector is k = [000] in all four examples. In plots A) and B),envelopes ±Ejk(τ) of the autocorrelation functions are shown by dashed lines.

exhibiting significant deviations from Eq. 5–16. Typical examples of power spectra of the

normal modes are shown in Figure 5-4. We observe that for most modes, even those not

satisfying the Langevin model (5–14), the maximum of Sjk corresponds to the highest

frequency of oscillations which we associate with the apparent phonon frequency. In a

few cases, such as that shown in Figure 5-4c, the maximum of Sjk corresponds to slower

secondary oscillations. In these cases, the apparent phonon frequency was defined as the

frequency of a local maximum of Sjk(ω) with the largest value of ω.

In order to estimate the lifetime of a phonon mode, we define an envelope Ejk(t) of

its autocorrelation function Cjk(t) as a line connecting local maxima of Cjk(t), as shown

by dashed lines in Figure 5-3a,b. The function Ejk is then fitted to an exponential. In the

cases of non-exponential decay of Ejk, the phonon lifetime is obtained by fitting its initial

(quickly decaying) segment to an exponential.

76

0 50 100 1500

50

100

150

200

ω (ps−1 )

Sjk

A j = 7

0 50 100 1500

50

100

150

200

ω (ps−1 )

Sjk

B j = 23

0 50 100 1500

50

100

ω (ps−1 )

Sjk

C j = 64

0 100 200 3000

500

1000

ω (ps−1 )

Sjk

D j = 104

Figure 5-4. Power spectra Sjk(ω) corresponding to the phonon autocorrelation functionsshown in Figure 5-3. The normal mode numbers j are shown in thecorresponding plots and the wavevector is k = [000] in all four examples.

Once the apparent phonon frequency ωjk and the lifetime τjk are obtained, the

anharmonic phonon frequency ωajk is computed from Eq. 5–19. This equation is correct

only for phonon normal modes that satisfy the Markovian Langevin equation (5–14).

However, the differnce between ωjk and ωajk is negligible if ωjk >> 1/τjk. As will be shown

in section 5.5, phonon lifetimes range from 0.4 to 30 ps. Therefore, the correction of the

apparent frequency (5–19) is significant only for phonons with very small frequencies.

Our results indicate that dynamics of the low-frequency modes are in good agreement

with the predictions of Eq. 5–14, see e.g. Figure 5-3a. The correction of the apparent

frequency is accurate for these modes. On the other hand, this correction is negligible for

the high-frequency modes which do not satisfy the Markovian model (5–14).

Autocorrelation functions Cjk of a phonon mode jk in a sorbate-free sodalite and a

sodalite containing sorbates are qualitatively similar for the same values of j and k. This

provides an additional confirmation that small neutral sorbates do no alter the phonon

eigenmodes (see discussion at the end of section 5.3). The sorbates affect such phonon

77

0 10 20 30 40 50 60 700

10

20

30

40

50

60

70

80

90

100

ω (ps−1)

SV

1 Ar/cage1 CH

4/cage

2 Ar/cage2 CH

4/cage

Figure 5-5. Power spectra SV (ω) of sorbate velocities in rigid zeolite cages.

properties as their lifetime and frequency. It will be shown in the next section that some

of these changes, namely an increase of lifetime of some phonon modes, are qualitatively

different from those expected from the simple phonon scattering picture. This implies

that the sorbate dynamics is qualitatively different from that of point defects coupled to

the lattice by a nearly harmonic potential. In fact, the sorbate dynamics is chaotic even

in the absence of the thermal interaction with the lattice vibration. This is confirmed by

the power spectra SV (ω) of the sorbate velocities in the rigid zeolite shown in Figure 5-5.

These power spectra are rather wide, implying chaotic sorbate dynamics due to strongly

nonlinear interactions between the sorbates and the zeolite walls. The nonlinear effects are

even stronger for systems with two sorbates per cage, as indicated by the wider sorbate

velocity spectra in these systems.

5.5 Phonon statistics

Anharmonicity of lattice vibrations and sorbate-phonon interactions lead to non-

zero mean values of some normal mode coordinates Qjk. The normal modes Qjk with

78

Table 5-3. Normalized averages of phonon amplitudes, ∆Qjk =< Qjk > /σjk. Only modeswith sufficiently large average deviations from zero, ∆Qjk ≥ 0.04, are shown.Normal modes are listed in the order of descending ∆Qjk. Harmonic phononfrequencies ωh

jk corresponding to the listed modes are also shown.

∆Qjk

jk ωhjk (ps−1) Sorbates/cage

Bare lattice 1 Ar 1 CH4 1 Xe 2 Ar 2 CH4

7 [000] 9.6 4.00 3.98 4.03 3.95 4.00 4.3363 [000] 98.4 -3.14 -3.21 -3.18 -3.18 -3.23 -3.6948 [000] 80.6 – – – – 0.26 0.8360 [000] 93.5 – – – – -0.048 -0.1521 [110] 4.3 – – – – – 0.04878 [000] 142.0 – – – – – -0.043

sufficiently large relative deviation of their mean from zero,

∆Qjk =〈Qjk〉σjk

≥ 0.04, (5–20)

are listed in Table 5-3. In Eq. 5–20, σjk denotes the standard deviation of the normal

mode fluctuations.

In the sorbate-free lattice, modes Qjk with j = 7 and j = 63 and k = [000] deviate

from zero by 4 and 3 standard deviations, respectively. All other modes have much smaller

deviations, ∆Qjk < 0.02. Addition of one sorbate per cage essentially does not change

the values of ∆Qjk. However, addition of two sorbates per cage leads to a substantial

shift of average values of several additional modes, implying that the equilibrium lattice

configuration slightly changes due to the presence of the sorbates. This change is larger

in the case of 2 CH4/cage. Addition of 2 CH4/cage shifts averages of several normal

modes away from zero as well as further increases ∆Qjk of the modes j = 7 and j = 63

(k = [000]) that were already shifted in the sorbate-free lattice. Nevertheless, even the

strongest sorbate effects on the mean normal coordinates seen in the case of 2 CH4/cage

are significantly weaker than effects of introduction of anharmonicity to a harmonic

sodalite lattice.

79

0 50 100 150 200 2500

5

10

15

20

25

30

ωhjk (ps−1)

τjk

(ps)

Figure 5-6. Phonon lifetimes τjk in sorbate-free sodalite lattice.

It is interesting to note that all substantial shifts of 〈Qjk〉 take place for optical lattice

modes with k = [000]. The largest relative displacement of an acoustic mode is rather

small (∆Qjk = 0.048) and is observed for mode j = 1,k = [110] when 2 CH4/cage are

added.

Phonon lifetimes τjk in sorbate-free anharmonic lattice range from 0.4 to 30 ps, as

shown in Figure 5-6. The modes with intermediate frequencies, 70 ps−1 ≤ ωhjk ≤ 150 ps−1,

possess short lifetimes. The phonons with high frequencies, ωhjk > 150 ps−1, correspond

mostly to fast vibrations of individual bonds. Interactions between these modes and other

modes in the system are weak leading to large phonon lifetimes for the high-frequency

modes. Some modes with low frequencies, ωhjk < 70 ps−1, also possess long lifetimes.

However, most of the low-frequency modes possess relatively short lifetimes indicating that

they are strongly coupled with other modes in the system.

For comparison, analysis of McGaughey and Kaviany [97] based on a decomposition

of the heat current autocorrelation function predicts decay time for heat transfer

associated with long-range acoustic modes in sodalite to be 1.67 ps. In addition,

80

Greenstein et al. [80] have estimated phonon relaxation time in MFI zeolite to be 9.2 ps

by fitting a theoretical expression to experimental thermal conductivity data. This

estimate is based on an assumption that the relaxation time is the same for all phonon

modes. Both of these estimations are within the range of the phonon lifetimes observed in

the current work.

Relative changes of phonon lifetimes,

δτjk =τ sjk − τa

jk

τajk

, (5–21)

upon encapsulation of sorbates into the zeolite cages are shown in Figure 5-7. In Eq. 5–

21 and elsewhere in this chapter superscript s refers to a property related to a zeolite

with encapsulated sorbates. The distributions of δτjk shown in Figure 5-7a are almost

identical for all three systems with 1 sorbate/cage. These distributions are symmetric with

respect to δτ = 0, which implies that the phonon-sorbate interactions are equally likely

to decrease as well as increase the phonon lifetime. The increase of the phonon lifetime

contradicts a simple picture of phonon scattering by sorbates and implies that a more

complex sorbate-phonon interaction is in play.

When two sorbates per cage are added to the system, the distribution of δτjk

becomes skewed towards average decrease of the phonon lifetime. This trend is especially

pronounced in the case of a larger sorbate, namely methane. This can be explained,

in part, by a tighter fit of the sorbates within the cages leading to a smaller amplitude

of the sorbate oscillations, which suggests more similarities between sorbates in these

systems with point defects. However, as Figure 5-5 indicates, the sorbate dynamics in the

2 sorbate/cage systems is more chaotic than in the 1 sorbate/cage systems. Therefore,

the analogy between the 2 sorbates/cage systems and crystals with point defects is not

complete. Indeed, Figure 5-7 shows that some of the phonon modes in the 2 sorbates/cage

systems undergo a significant (on the order of 100%) increase of their lifetime and hence

the scattering model is still not applicable to this case.

81

In order to assess which of the modes undergo increase or decrease of their lifetime,

in Figure 5-7b we plot δτjk for every phonon mode. For clarity, only two extreme cases

are shown: 1 Ar/cage and 2 CH4/cage. For 1 Ar/cage, the changes of phonon lifetimes are

evenly distributed among different frequencies. In the 2 CH4/cage system, the modes lying

in the small and large frequency regions experience, on average, larger change of their

lifetimes than the modes with intermediate frequencies.

Effects of anharmonicity on the phonon frequency in a sorbate-free sodalite crystal are

summarized in Figure 5-8 which shows relative differences,

δωjk =ωa

jk − ωhjk

ωhjk

, (5–22)

between the anharmonic and harmonic phonon frequencies. With few exceptions,

anharmonicity leads to an increase of the phonon frequency. This increase is especially

large for low frequency modes.

Addition of sorbates to zeolite cages typically leads to a further frequency increase, as

can be seen from the relative differences

δωsjk =

ωsjk − ωa

jk

ωajk

(5–23)

between phonon frequencies in sodalite with encapsulated sorbates (ωsjk) and in the

sorbate-free sodalite (ωajk), see Figure 5-9. The frequency changes δωs

jk are substantially

larger when more than one sorbate per cage is introduced. Similarly to δωjk, δωsjk tends

to increase with the decrease of the phonon frequency. However, the changes of phonon

frequencies due to addition of sorbates are smaller than the changes due to introduction of

anharmonicity to a sorbate-free harmonic lattice.

Up to this point, we have investigated effects of anharmonicity and sorbate-phonon

interactions on individual phonon modes. To complete the picture, we now consider

correlations between different phonon modes. It is expected that this information will

be helpful in developing a more precise form of the Langevin model (5–14) for phonon

82

−1 −0.5 0 0.5 10

0.5

1

1.5

2

2.5

3

δτjk

P(δ

τjk)

1 Ar/cage

1 CH4/cage

1 Xe/cage

2 Ar/cage

2 CH4/cage

A

0 50 100 150 200−1

−0.5

0

0.5

1

1.5

ωhjk (ps−1)

δτjk

1 Ar/cage

2 CH4/cage

B

Figure 5-7. Effects of sorbates on phonon lifetimes. A) Distributions P (δτjk) of therelative differences δτjk between the phonon lifetimes τ s

jk in sodalite withencapsulated sorbates and the phonon lifetimes τa

jk in a sorbate-free sodalite.B) Relative changes δτjk of lifetimes of individual modes jk for the cases of1 Ar/cage and 2 CH4/cage.

83

0 50 100 150 200 250−0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

ωhjk (ps−1)

δω

jk

Figure 5-8. Relative differences δωjk between the anharmonic (ωajk) and harmonic (ωh

jk)frequencies in a sorbate-free sodalite crystal.

0 20 40 60 80 100−0.25

−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

0.2

ωhjk (ps−1)

δω

s jk

1 Ar/cage1 CH

4/cage

1 Xe/cage2 Ar/cage2 CH

4/cage

Figure 5-9. Relative differences δωsjk between phonon frequencies in sodalite with

encapsulated sorbates (ωsjk) and in the sorbate-free sodalite (ωa

jk). For clarity,only δωs

jk with magnitudes greater than 10−2 are shown.

84

0 50 100 150 2000

50

100

150

200

ωhjk (ps−1)

ωh j′k

′(p

s−1)

0.08

0.16

0.24

0.31

0.39

0.47

Figure 5-10. Magnitudes of correlation coefficients ρ(jk, j′k′) between phonon modes jkand j′k′ in a sorbate-free sodalite crystal. Darker symbols correspond tostronger correlations. For clarity, correlation coefficients for jk = j′k′ are notshown and only correlations of magnitude greater than 0.1 are plotted.

dynamics since it will allow us to assess which phonons make dominant contributions to

stochastic forces acting on each of the phonon modes. The magnitudes of the correlation

coefficients

ρ(jk, j′k′) =

∣∣∣∣〈δQjkδQ

∗j′k′〉

σjkσj′k′

∣∣∣∣ (5–24)

between modes jk and j′k′ in a sorbate-free lattice are shown in Figure 5-10. For clarity,

only correlations of magnitude exceeding 0.1 are plotted. Most correlation coefficients

between individual phonons are rather weak. Correlations between low- and mid-

frequency modes constitute a notable exceptions: some correlations of low-frequency

modes with modes of frequency ωhjk ≈ 100 ps−1 have correlation coefficient as large as

0.5. These observations are consistent with the measured phonon lifetimes, see Figure 5-6.

Specifically, phonon-phonon correlations involving the high-frequency modes are very

weak, which is manifested in long lifetimes of these modes. The modes with intermediate

frequencies, 70 ps−1 ≤ ωhjk ≤ 150 ps−1, are strongly correlated with some of the low

frequency modes, which leads to small lifetime of these modes.

85

0 50 100 150 2000

50

100

150

200

ωhjk (ps−1)

ωh j′k

′(p

s−1)

2 Ar/cage

2 CH4/cage

0.08

0.16

0.24

0.31

0.39

0.47

Figure 5-11. Effect of sorbates on phonon-phonon correlations. Only those pairs (jk, j′k′)are shown for which addition of sorbates creates a correlation with coefficientgreater than 0.1 or changes the magnitude of an existing correlation by morethan 0.1. Magnitudes of the changed correlation coefficients are indicated bycolor.

Effect of the sorbate presence on the phonon-phonon correlations is shown in

Figure 5-11. Addition of a single sorbate per cage leads to relatively small changes in

phonon-phonon correlations, with no phonon pairs with correlation coefficient greater than

0.1 being affected. Addition of 2 Ar/cage creates one new phonon-phonon correlation with

magnitude slightly larger than 0.1. In contrast, addition of 2 CH4/cage creates several

strong correlations and magnifies some correlations which were present in the sorbate-free

lattice. The largest changes of the correlation coefficients upon addition of 2 CH4/cage are

between the low-frequency acoustic modes with wavevector k = [110] and other phonons.

The change of the correlation coefficient in these cases is ∆ρ ≈0.27.

5.5.1 Conclusions

We have investigated phonon dynamics in sodalite for five different sorbate-zeolite

systems: 1 Ar/cage, 1 CH4/cage, 1 Xe/cage, 2 Ar/cage, and 2 CH4/cage, as well as

86

a sorbate-free zeolite. We have observed that the Markovian Lanvegin equation (5–

14) provides an adequate model for dynamics of most of the phonon modes. However,

dynamics of some phonon modes does not agree with predictions of Eq. 5–14. These

modes exhibit secondary oscillations (see Figure 5-3b-d) which are most likely related to

the total energy fluctuations inside a single sodalite cage [97].

Encapsulation of sorbates in zeolite cages does not lead to qualitative changes of

phonon dynamics, as can be concluded from phonon autocorrelation functions. However,

depending on the type and number of sorbates inside sodalite cage, some or all of the

following aspects of phonon dynamics are changed: (i) the mean values of the normal

mode coordinates, (ii) phonon lifetimes, (iii) phonon frequencies, and (iv) phonon-phonon

correlations. The largest changes of phonon dynamics are observed upon encapsulation of

two methane molecules per sodalite cage.

The strongest effect of all considered sorbates on lattice dynamics is a significant

change of the phonon lifetimes. From the scattering picture of phonon-sorbate interactions

it is expected that these interactions will decrease phonon lifetimes. However, we observe

that encapsulation of sorbates leads to an increase of lifetimes of a large fraction of

phonons. Therefore, development of a more detailed model is necessary to understand the

complex nonlinear sorbate-phonon interactions.

87

CHAPTER 6MODEL DEVELOPMENT FOR SORBATE MOLECULES IN 1D SYSTEM

In this chapter, we present results of our efforts to develop a mathematical model

for the “rattling” effect of sorbate molecules on the lattice vibration of a host matrix.

Existing models for phonon scattering cannot be directly applied because they are limited

to prediction of phonon interaction with static impurities or defects [98]. To investigate

the influence of sorbates on lattice system and to develop a heat transfer model for

nanoporous materials, we consider a simple 1D system. First, we investigate dependence of

thermal conductivity of this system on sorbate properties using full scale non-equilibrium

MD simulations. To get insight into the sorbate-lattice dynamics, we investigate phonon-

sorbate dynamics in a harmonic lattice, which allows us to focus on the role of the

sorbate-lattice interactions in the energy exchange between phonon modes. In conclusion,

we discuss possible approaches to development of a theoretical model.

6.1 Thermal Conductivity from NEMD Simulations

We consider FPU system (see section 3.1) consisting of N = 1024 particles with force

constants k = 10, α = −50, and β = 180. The sorbate is located between lattice sites 256

and 257 and interacts with the lattice atoms through Lennard-Jones (LJ) potential:

VLJ = 4ε

[(σ

r

)12

−(σ

r

)6]

, (6–1)

where r is the distance between the sorbate and a lattice atom. Periodic boundary

conditions are imposed in the system. Hot and cold regions with temperatures Th = 1.2

and Tc = 0.8 each containing 100 particles are set up at the edge (atoms 1, . . . , 100) and

at the center (atoms 512, . . . , 612) of the chain using the Nose-Hoover thermostat (see

section 2.2.2). We use velocity Verlet algorithm with time step ∆t = 5 × 10−4 time unit.

The simulation is carried out for 225000 time units. After the system reaches a steady

state, we compute the heat flux

J =1

2

m6=n

vnfnm, (6–2)

88

1 500 1000

0.8

1

1.2

Position

Tem

pera

ture

A

1 500 1000−1.5

0

1.5

Position

Hea

t Flu

x

B

Figure 6-1. A) Established temperature profile in non-equilibrium MD simulations. Thesorbate is located between lattice sites and sorbate-lattice interactions aremodeled by LJ potential. B) Established heat flux in non-equilibrium MDsimulations.

where vn is the velocity of the n-th atom and fnm is the force with which the m-th atom

is acting on the n-th atom. The thermal conductivity κ is then calculated using Fourier’s

law, Eq. 1–1. We perform a series of NEMD simulations with sorbates of different mass

Ms and Lennard-Jones parameter ε of sorbate-lattice interaction. In all these simulations,

the effective LJ diameter is σ = 0.23. Examples of established temperature and heat

flux profiles are shown in Figure 6-1. Comparison between thermal conductivities of a

sorbate-free lattice and lattice containing various sorbate molecules is shown in Figure 6-2.

Even in this simple system, we observe a complex pattern of dependence of κ on the

sorbate-lattice interaction parameters. The presence of the sorbate may either increase

or decrease the thermal conductivity. In addition, sufficiently large temperature gradient

leads to development of a discontinuity in the temperature profile across the unit cell

containing the sorbate, as shown in Figure 6-3. This phenomenon is similar to Kapitza

thermal boundary resistance at an interface between two dissimilar materials [99]. In the

latter case, a temperature discontinuity is developed to maintain continuity of the heat

flux through the interface.

89

0 0.5 1 1.5 2 2.55.2

5.4

5.6

5.8

6

6.2

6.4

6.6

6.8

Ms/ε

The

rmal

Con

duct

ivity

[ κ

]

With Sorbate

pure lattice

Figure 6-2. Comparison of thermal conductivities, κ, obtained from NEMD simulationsfor a system containing 1024 atoms. Temperature gradient in all simulations isdTdx

= 0.001. Solid circles represent the values of κ in the presence of a sorbatemolecule with mass Ms and LJ interaction parameters ε. The solid linecorresponds to the thermal conductivity of lattice with no sorbate molecules.

1 250 500

0.9

1.0

1.1

Particle

Tem

pera

ture

Figure 6-3. Steady state temperature profile in NEMD simulation with imposedtemperature gradient dT

dx= 0.002 The sorbate is located between lattice

particles 128 and 129. A discontinuity in the temperature profile is developedacross the unit cell containing the sorbate.

90

6.2 Sorbate in a Harmonic Lattice

In order to isolate the phonon-phonon energy exchange facilitated by the sorbate from

the energy exchange due to lattice anharmonicity, in the remainder of this chapter we

consider dynamics of a sorbate in a harmonic lattice. The Hamiltonian of the system is

H =∑

q

Hq + Hs + U({Qj}, xs), (6–3)

where

Hq =1

2

(ω2

q |Qq|2 + |Qq|2)

, (6–4)

is the Hamiltonian of a phonon mode Qq with wavenumber q,

Hs =Msv

2s

2+ U0(xs) (6–5)

is the Hamiltonian of a sorbate in a rigid lattice, and U({Qj}, xs) is the potential energy

of interactions between the phonons and the sorbate. In Eq. (6–4), ωq is the frequency of

the phonon with wavenumber q, which is given by the following dispersion relationship,

ωq =

√4k

m

∣∣∣∣sin qa

2

∣∣∣∣ (6–6)

(a is the lattice constant, m is the mass of lattice atoms). In Eq. (6–5), Ms, xs, and vs

are the sorbate mass, position, and velocities, respectively, and U0(xs) is the potential of

interaction between the sorbate and the lattice atoms when the latter are fixed at their

equilibrium positions.

In the absence of the sorbate-lattice coupling, i.e. when U({Qj}, xs) ≡ 0, the system

is integrable; the phonons and the sorbate undergo periodic motion. Introduction of a non-

zero potential U({Qj}, xs) perturbs their periodic trajectories. It is known from the theory

of dynamical systems [100] that a small perturbation of an integrable systems may lead

to a complete destruction of the periodic trajectories. In this case, the system trajectory

becomes chaotic, which facilitates fast energy exchange between the system degrees of

freedom. Such chaotic behavior will take place if the unperturbed system satisfies the

91

resonance condition, i.e. if the ratio of frequencies of some of the unperturbed periodic

trajectories is a rational number.

6.2.1 Scattering of a Phonon Wavepacket

In this section we discuss simulations of scattering of a phonon wavepacket by a

single sorbate located in a harmonic lattice. Similar simulations have been performed in

Ref [101] for scattering of a wavepacket by an interface between two different crystals.

We consider one-dimensional harmonic lattice containing N = 2001 atoms. The

sorbate is located between lattice sites 1000 and 1001. The lattice spring constant is

k = 156 and the sorbate-lattice interactions are modeled by LJ potential, Eq. 6–1. The

mass of sorbate molecule, Ms, ranges from 1 to 6 and ε ranges from 1 to 10, while effective

LJ diameter is held fixed at σ = 0.44.

The sorbate does not move until a collision with a wavepacket. The simulations are

initialized with the sorbate placed at its equilibrium position within a unit cell and its

initial velocity set to zero. Initial lattice configuration consists of a phonon wavepacket

centered at wavenumber q0. This wavepacket is a linear combination of the normal modes

with wave numbers sufficiently close to q0. Therefore, the displacement ul of the l-th

lattice atom is

ul = Aeıq0(xl−x0)e−η2(xl−x0)2 , (6–7)

where A is the amplitude of the wave and

uq = ei(qx−ωqt) (6–8)

is a normal mode corresponding to wavenumber q.

This procedure generates a wavepacket localized in space around x0 with spatial

extent ∼ 1/η. The value of the parameter η is selected so that the range of wave numbers

in the generated wavepacket is sufficiently narrow. This ensures that the wavepacket

maintains its configuration until its collision with the sorbate. The initial velocities of the

92

1000 2000

−0.1

0.1

Position

Dis

plac

emen

tTime=1

A

1000 2000

−0.1

0.1

Position

Dis

plac

emen

t

Time=271

B

1000 2000

−0.1

0.1

Position

Dis

plac

emen

t

Time=500

C

Figure 6-4. Scattering of a wavepacket centered around wavenumber q0 = 0.1π by asorbate with LJ parameters (ε, σ) = (2, 0.44) and mass Ms = 1. Location ofthe unit cell containing the sorbate molecule is shown by the dashed line.

1 2 3 4 5 60

0.2

0.4

0.6

0.8

1

Mass of sorbate

Tra

nmis

sion

Rat

io

q0=0.2π ; ω=7.7155

q0=0.5π ; ω=17.6566

q0=0.8π ; ω=23.7525

Figure 6-5. Transmission ratios for scattering of different wavepackets by sorbatemolecules with different masses. LJ potential parameters are (ε, σ) = (2, 0.44)in all shown simulations.

lattice atoms are obtained as follows. We decompose ul into the normal modes and then,

using the normal mode frequencies, obtain the time derivatives of the atom displacements.

A typical example of the wavepacket scattering is shown in Figure 6-4. As can be

seen, a part of the wavepacket is transmitted through the unit cell containing the sorbate

molecule and part of the wavepacket is reflected. The frequency of the normal modes

contained in these wavepackets remain essentially unchanged.

Transmission ratios (i.e., ratios of the amplitude of the transmitted and the incident

wavepackets) for a series of the scattering simulations are summarized in Figure 6-5. In

93

these simulations, LJ parameters were fixed, σ = 1.0 and ε = 0.44, and the sorbate

mass Ms and wave numbers q0 of the incident wavepackets were varied. Depending on q0,

the transmission ratio may either monotonously increase, monotonously decrease as Ms

is increased. We also observe a sudden drop of the transmission ratio to an almost full

reflection for q0 = 0.5π and Ms = 2. These parameters correspond to a resonance between

the sorbate and the wavepacket.

In order to further study the resonant sorbate-lattice interactions, we perform

two scattering simulations using the same wavepacket with different parameters of the

sorbate molecule. We estimate the sorbate frequency from the linear approximation of LJ

potential,

VLJ ' 1

2kLJ(r − r0)

2, (6–9)

ωs '√

kLJ

Ms

. (6–10)

The wavepacket is centered around normal modes with wavenumber q0 = 0.475π and

ω = 16.9 with fixed effective LJ parameter σ = 0.44. We use two different values of

Ms and ε for the sorbate molecule: (0.5,5) and (1,1). We refer to these two interaction

parameters as system α and β, respectively. From Eq. 6–9, the sorbate frequency is

ωs = 16.8 and 5.36, respectively. Hence the sorbate frequency in system β is closer to the

resonance frequency with the wavepacket.

The initial configuration of the wavepacket is shown in Figure 6-6 A). The system

configurations after the wavepacket pass the sorbate molecule in system α and β are

shown in Figure 6-6 B) and C), respectively. There is a qualitative difference between

scattering results for these two wavepackets. Under resonance condition (system β),

wavepacket changes its shape, which suggests that additional modes are excited due

to interaction with this wavepacket. In order to obtain a better view of the scattering

process, we perform windowed-Fourier transform (WFT) of the lattice configuration.

This transform allows us to obtain local wave numbers of the lattice modes for different

94

1000 2000−0.05

0.05

Particle Position

Dis

plac

emen

t

A

1000 2000−0.05

0.05

Particle Position

Dis

plac

emen

t

B

1000 2000−0.05

0.05

Particle Position

Dis

plac

emen

t

C

D E F

Figure 6-6. Scattering of wavepackets with wave numbers with q = 0.475π with fixedsigma = 0.44 and (Ms,epsilon)=(0.5,5) and (1,1). A) Initial wavepacketconfiguration for both simulations.; B) System configuration after scatteringwith sorbate for (Ms,epsilon)=(0.5,5); C) System configuration afterscattering with sorbate for (Ms,epsilon)=(1,1); D) to F) are WindowedFourier transforms of configurations shown in A) to C); x axis corresponds tolattice sites and y axis corresponds to local wave numbers of latticeoscillations.

locations within the system. The results of WFTs for Figure 6-6 A)- C) are shown in

Figure 6-6 D)- F), respectively. In the system α (plot E), the wavepacket structure and

wavenumber remain the same as in the initial wave packet (plot D). This is expected

as this value of the initial wave vector is sufficiently far from the resonant condition.

On the other hand, in the system β we observe that the symmetric structure of the

wavepacket is destroyed and nearby wave modes are excited during the collision. This

quick redistribution of energy between different modes is characteristic of a resonant

system.

95

0 200 400 6000

10

20

30

Time

Ene

rgy

Ms=0.1, ε=2

Ms=1, ε=0.5

Ms=1, ε=1.5

Figure 6-7. Evolution of mode energy for incident plane wave with wavenumberq = 0.245π interacting with different sorbate molecules. Ms = 1, σ = 0.23 andthree ε values of 0.5, 1.5, 2.

6.2.2 Scattering of a Plane Wave

The wavepacket scattering simulations reveal the dependence of energy transport on

interaction in lattice system. However, due to the short time span of the sorbate-phonon

interactions in these simulations, we cannot obtain information regarding long-term energy

dispersion among the normal modes.

In order to get better insight of the phonon-sorbate interaction dynamics, we perform

scattering simulations of a plane wave. The initial conditions for these simulations are

plane waves consisting of a single normal modes. The lattice spring constant in these

simulations is k = 78. The simulations are performed for time length of 600 time units.

We perform a series of MD simulations with planar wave consisting of single phonon

mode, ωin = 6.6 Hz. We present the analysis of the results including incident mode

energy evolution, sorbate frequency and energy dispersion from three sets of interaction

parameters: σ is fixed to be 0.23 and (Ms, ε) = (0.1, 2), (1, 0.5) and (1, 1.5). These three

systems are referred to as system A, B and C respectively. The incident mode energy

evolution in Figure 6-7 shows that the energy change in both systems A and C are in

the form of periodic oscillation whereas rapid decaying behavior is observed in system

B. Note that in this harmonic system the energy exchange between different normal

96

5 10 150

0.05

0.1

ωs

| Fou

rier

Coe

ffici

ent |

A

5 10 150

0.02

ωs

| Fou

rier

Coe

ffici

ent |

B

5 10 150

0.02

0.04

0.06

ωs

| Fou

rier

Coe

ffici

ent |

C

Figure 6-8. Fourier transform of the sorbate displacement with Ms = 1, σ = 0.23 andε =0.5 A), 1.5 B) and 2.0 C) respectively.

0 5 10 150

0.5

1

1.5

2

ωin

∆ E

Ms=0.1d0, ε=2.d0

Ms=1.d0, ε=0.5d0

Ms=1.d0, ε=1.5d0

Figure 6-9. Dispersion of incident mode energy (q = 0.245π) among other phonon modesunder different values of sorbate mass. Ms = 1, σ = 0.23 and three ε values of0.5, 1.5, 2.

modes only occurs through sorbate-phonon interaction. Thus, rapidly decaying behavior

indicates the sorbate-phonon interaction allows significant dispersion of incident mode

energy among the other normal modes. The condition for occurrence of energy change

between two degrees of freedom is the resonance condition. Hence we present the sorbate

frequency spectrum in Figure 6-8. As expected, we observe a wide frequency spectrum for

system B. Moreover, in order to identify the dispersion of energy among the other phonon

modes, we plot the gained energy of the other phonon modes except the incident mode

in Figure 6-9. In system A, there is almost no energy transfer taking place, implying

weak interaction between sorbate and other phonon modes. This is also verified by the

97

very distinct sorbate frequency spectrum shown in Figure 6-8A. On the other hand, the

excitation of most phonon modes is observed for system B (closed circles). Note that

though in Figure 6-8B the magnitude at high sorbate frequency is not as significant

as that for lower frequency, the higher frequency phonon modes are excited due to the

satisfaction of resonance condition

ω = mωs. (6–11)

We also observe the excitation of modes with 2ωin = 13.30 Hz in the system C.

From these results we conclude that the energy exchange between phonon modes

through rattling of sorbate molecules depends on the frequency spectrum of sorbate

molecule. Sorbate molecule with wide frequency spectrum excites most normal modes

and on the other hand, sorbate molecule with distinct frequency is considered as a small

perturbation to the dynamics of incident phonon mode and hence only excites modes with

close frequencies. In order to understand the dependence of sorbate frequency spectrum

on sorbate-lattice interaction parameters and further quantify the energy exchange, we

present the possible approaches to develop a mathematical model using both dynamical

system approach and multi-scale expansion.

6.2.3 Possible Theoretical Approach: Multi-scale Expansion

In this section we initiate development of an analytical model for the phonon

scattering by sorbate molecules. The first step in this development is to describe the

above scattering process. We consider a simplified linear lattice-sorbate interaction as

in this simple model one can obtain an analytical solution which fully describes this

system. However, our eventual goal is to develop an approach that may be extended to

nonlinear interactions between the lattice and the sorbate. In the absence of the sorbate,

displacement of lattice atoms can be fully described by orthonormal eigenvectors φk(x)

and ±ıωk eigenvalues in a linear system. We also consider the presence of sorbate as

a perturbation to the linear system with new set of eigenvector ψk(x) and eigenvalue

98

±ıνk. Since φk and ψk are complete set of basis vectors, the displacement throughout the

scattering process can be expanded as follows:

f(x, t) =∑

k[Fk(t)eıωkt + Gk(t)e

−ıωkt]φk(x)

=∑

j[Ajeıνjt + Bje

−ıνjt]ψj(x).(6–12)

In addition, these two sets of eigenvectors are connected by constant matrix Γ given by

−→Ψ = Γ

−→Φ . (6–13)

Substituting Eq. 6–13 into Eq. 6–12, we obtain the relation

Fk(t)eıωkt + Gk(t)e

−ıωkt =∑

j

γjk(Ajeıνjt + Bje

−ıνjt) (6–14)

In a weakly perturbed system

νk = ωk + ενk, ε → 0, (6–15)

and thus the time-dependent coefficients, Fk(t) and Gk(t) will be slow-varying function.

Here another slow time variable, τ is introduced;

Fk(t) ≡ Fk(τ)

Gk(t) ≡ Gk(τ),(6–16)

where τ = εt. Under the weak perturbation assumption, Eq. 6–15, the relation between

these two sets of coefficients, Eq. 6–13, becomes

Fk(τ)eıωkt + Gk(τ)e−ıωkt '∑

j

γjk(Ajeıωkteıνjkτ + Bje

−ıωkte−ıνjkτ ) (6–17)

Comparing both sides of Eq. 6–17,

Fk(τ) =∑

j γjkAjeıνjkτ

Gk(τ) =∑

j γjkBje−ıνjkτ ,

(6–18)

99

0 10 200

2.e−3

4.e−3

Period

|Fk|

A

0 10 20

4.e−3

5.e−3

Period

|Gk|

B

Figure 6-10. A) Comparison of linear model prediction of Fk from Eq. 6–18 andsimulation result B) Comparison of linear model prediction of Gk fromEq. 6–18 and simulation result

where the mode j has to satisfy Eq. 6–15. The result from Eq. 6–18 along with the

MD simulation result is shown in Figure 6-10. The evolution of Fk and Gk coefficients

are captured well. In order to complete the above derivation, we need the perturbed

eigenvalue and eigenvectors, which is achieved through degenerate perturbation method.

The new sets of eigen-frequency obtained from degenerate perturbation method

as well as the normal eigen-problem solving function are shown in Figure 6-11 and

represented by solid circles and open circles respectively. The two curves match well

except near zero frequency, where the perturbation assumption is not valid.

Extending this analysis to nonlinear system

d2f

dt2= H0(f) + εN(f), (6–19)

100

0 0.5 1 1.50

10

20

Fre

quen

cy

Wave Vector, q

Figure 6-11. Frequencies obtained from degenerate perturbation method and regulareigenvalue solver are shown using closed and open circles respectively.

where N is a nonlinear operator. The first expansion of Eq. 6–12 still holds. In addition,

we perform multi-time scales (t and τ) analysis for time derivative. In other words,

f(t) → f(t; τ),

f(t, τ) = f0(t; τ) + εf1(t; τ) + ε2f2(t; τ),

ddt

→ ∂∂t

+ ε ∂∂τ

.

(6–20)

Substituting Eq. 6–20 into Eq. 6–19 and compare different ε order term, we expect to

develop a model with better description than the current theory for perturbation caused

by the static defect atoms.

6.2.4 Conclusions

We have investigated the dependence of sorbate-lattice interaction potential on the

effect of lattice dynamics in a 1D model system. The results in NEMD simulations show

a complex dependence ( both increasing and decreasing) of the LJ potential parameters

on thermal conductivity. . In the EMD simulations of one single phonon mode and the

sorbate molecule, we show that the energy dispersion through sorbate-lattice interaction

depends on the frequency spectrum of sorbate molecule. We have shown the development

of mathematical model for the sorbate-lattice interaction from KAM theorem and multi-

scale expansion.

101

CHAPTER 7CONCLUSIONS AND POSSIBLE DIRECTIONS OF FUTURE RESEARCH

In this work, we have examined the existence and stability of nonlinear lattice

vibrational modes in model system with different dimensionality and carbon nanotubes.

In model system, we have shown the nonlinear modes with localized structure. However,

these modes are linearly unstable localized solutions due to the force coefficients under

consideration as well as the structure of the system. Hence we need to investigate the

parameter as well as system structure in which system the stable nonlinear modes are

supported.

For Q1D (5,0) semi-conductive carbon nanotubes, we reported the nonlinear modes

with configuration qualitatively different from linear phonon modes for Brenner-Tersoff

potential in short carbon nanotubes as well as the modes up to periodicity of 24 unit

cells while ignoring cubic force coefficients. Thus we need to include the cubic term in

simplified potential function and in turn to obtain the vibration modes for Brenner-Tersoff

function in larger system. In order to reach larger time and length scales in the thermal

transport of CNT system, we need to develop a model based on BTE, Eq. 5–12.

In nanoporous materials, we have demonstrated the anharmonic effect on phonon

interaction induced by absorbed sorbate molecules (gas molecules) in zeolite molecules via

molecular dynamics simulations. We observe both increasing and decreasing phonon life

time caused by the presence of sorbate molecule. In order to develop a model to describe

the effect of sorbate molecule on phonon dynamics, we investigate a simple 1D system.

From NEMD simulations we observed complex dependence of thermal conductivity on

sorbate-lattice interaction parameters. Based on resonance condition, the incident mode

energy can be dispersed among the other normal modes through sorbate-lattice interaction

in a harmonic lattice system. To continue this work, we need to compute Arnold diffusion

and use generalized Langevin equation to describe the dynamics of lattice modes due to

the presence of sorbate molecule.

102

REFERENCES

[1] J.A. Hutchby, R. Cavin, V. Zhirnov, J.E. Brewer, and G. Bourianoff, “EmergingNanoscale Memory and Logic Devices: A Critical Assessment,” Computer, vol. 41,no. 5, pp. 28–32, 2008.

[2] P. Heino, “Thermal conduction simulations in the nanoscale,” Journal ofComputational and Theoretical Nanoscience, vol. 4, pp. 896–927, 2007.

[3] D.G. Cahill, W.K. Ford, K.E. Goodson, G.D. Mahan, A. Majumdar, H.J. Maris,R. Merlin, and S.R. Phillpot, “Nanoscale thermal transport,” Journal of AppliedPhysics, vol. 93, no. 2, 2003.

[4] D.A. McQuarrie, Statistical Mechanics, Harper & Row Publishers, New York, 1976.

[5] P. G. Sverdrup, S. Sinha, M. Asheghi, S. Uma, and K. E. Goodson, “Measurementof ballistic phonon conduction near hotspots in silicon,” Appl. Phys. Lett., vol. 78,pp. 3331–3333, 2001.

[6] P. G. Sverdrup, Y. S. Ju, and K. E. Goodson, “Sub-continuum simulations of heatconduction in silicon-on-insulator transistors,” J. Heat Transfer – Transactions ofthe ASME, vol. 123, pp. 130–137, 2001.

[7] S. K. Kim and I. A. Daniel, “Gradient method for inverse heat conduction problemin nanoscale,” Int. J. Numerical Methods in Engineering, vol. 60, pp. 2165–2181,2004.

[8] S.V.J. Narumanchi, J.Y. Murthy, and C.H. Amon, “Boltzmann transport equation-based thermal modeling approaches for hotspots in microelectronics,” Heat and MassTransfer, vol. 42, no. 6, pp. 478–491, 2006.

[9] S. Flach and C.R. Willis, “Discrete breathers,” Phys. Rep., vol. 295, pp. 181–264,1998.

[10] R. Reigada, A. Sarmiento, and K. Lindenberg, “Asymptotic dynamics of breathersin Fermi-Pasta-Ulam chains,” Phys. Rev. E, vol. 66, pp. 046607, 2002.

[11] J. L. Marin, S. Aubry, and L. M. Floria, “Intrinsic localized modes: Discretebreathers. existence and linear stability,” Physica D, vol. 113, pp. 283–292, 1998.

[12] S. Lepri, R. Livi, and A. Politi, “Thermal conduction in classical low-dimensionallattices,” Phys. Rep., vol. 377, pp. 1–80, 2003.

[13] A. Lippi and R. Livi, “Heat conduction in two-dimensional nonlinear lattices,” J.Stat. Phys., vol. 100, pp. 1147, 2000.

[14] T. Hatano, “Heat conduction in the diatomic Toda lattice revisited,” Phys. Rev. E,vol. 59, pp. R1–R4, 1999.

103

[15] B. Hu, B. Li, and H. Zhao, “Heat conduction in one-dimensional chains,” Phys. Rev.E, vol. 57, pp. 2992–2995, 1998.

[16] S. Iijima and T. Ichihashi, “Single-shell carbon nanotubes of 1-nm diameter,”Nature, vol. 363, pp. 603–605, 1993.

[17] R. H. Baughman, A. A. Zakhidov, and W. A. de Heer, “Carbon nanotubes – theroute toward applications,” Science, vol. 297, pp. 787–792, 2002.

[18] A. Javey, J. Guo, Q. Wang, and M. Lundstrom, “Ballistic carbon nanotube field-effect transistors,” Nature, vol. 424, pp. 654–657, 2003.

[19] Z. H. Chen, J. Appenzeller, Y. M. Lin, J. Sippel-Oakley, A. G. Rinzler, J. Y. Tang,S. J. Wind, P. M. Solomon, and P. Avouris, “An integrated logic circuit assembledon a single carbon nanotube,” Science, vol. 311, no. 5768, pp. 1735–1735, 2006.

[20] V. Meunier, S. V. Kalinin, and B. G. Sumpter, “Nonvolatile memory elements basedon the intercalation of organic molecules inside carbon nanotubes,” Physical ReviewLetters, vol. 98, no. 5, 2007, Meunier, Vincent Kalinin, Sergei V. Sumpter, Bobby G.

[21] P. Kim, L. Shi, A. Majumdar, and P. L. McEuen, “Thermal transport measurementsof individual multiwalled nanotubes,” Phys. Rev. Lett., vol. 87, pp. 215502, 2001.

[22] J. Hone, M. Whitney, C. Piskoti, and A. Zettl, “Thermal conductivity of single-walled carbon nanotubes,” Phys. Rev. B, vol. 59, pp. R2514, 1999.

[23] M. A. Osman and D. Srivastava, “Temperature dependence of the thermalconductivity of single-wall carbon nanotubes,” Nanotechnology, vol. 12, pp. 21–24, 2001.

[24] S. Berber, Y.-K. Kwon, and David Tomanek, “Unusually high thermal conductivityof carbon nanotubes,” Phys. Rev. Lett., vol. 84, pp. 4613–4616, 2000.

[25] J. Che, T. Cagin, and W. A. Goddard, “Thermal conductivity of carbonnanotubes,” Nanotechnology, vol. 11, pp. 65–69, 2000.

[26] S. Maruyama, “A molecular dynamics simulation of heat conduction of a finitelength single-walled carbon nanotube,” Microscale Thermophysical Engineering, vol.7, pp. 41–50, 2003.

[27] G. S. Nolas, G. A. Slack, D. T. Morelli, T. M. Tritt, and A. C. Ehrlich, “The effectof rare-earth filling on the lattice thermal conductivity of skutterudites,” Journal ofApplied Physics, vol. 79, no. 8, pp. 4002–4008, 1996, Part 1.

[28] G. P. Meisner, D. T. Morelli, S. Hu, J. Yang, and C. Uher, “Structure and latticethermal conductivity of fractionally filled skutterudites: Solid solutions of fully filledand unfilled end members,” Physical Review Letters, vol. 80, no. 16, pp. 3551–3554,1998.

104

[29] V. V. Murashov, “Thermal conductivity of model zeolites: Molecular dynamicssimulation study,” J. Phys.: Condens. Matter, vol. 11, pp. 1261–1271, 1999.

[30] W. C. Swope, H. C. Andersen, P. H. Berens, and K. R. Wilson, “A computersimulation method for the calculation of equilibrium constants for the formation ofphysical clusters of molecules – application to small water clusters,” J. Chem. Phys.,vol. 76, pp. 637, 1982.

[31] H. J. C. Berendsen, J. P. M. Postma, W. F. van Gunsteren, A. DiNola, and J. R.Haak, “Molecular dynamics with coupling to an external bath,” J. Chem. Phys., vol.81, pp. 3684–3690, 1984.

[32] S. Nose, “Constant temperature molecular dynamics methods,” Prog. Theor. Phys.Supp., vol. 103, pp. 1, 1991.

[33] DM Bylander and L. Kleinman, “Energy fluctuations induced by the Nosethermostat,” Physical Review B, vol. 46, no. 21, pp. 13756–13761, 1992.

[34] D. L. Ermak and H. Buckholz, “Numerical integration of the Langevin equation:Monte Carlo simulation,” J. Comput. Phys., vol. 35, pp. 169, 1980.

[35] F. Verhulst, Nonlinear Differential Equations and Dynamical Systems, Springer-Verlag, Berlin Heidelberg, 1990.

[36] E. Fermi, J. Pasta, and S. Ulam, “Los Alamos Report No. LA-1940 (1955),unpublished,” in Collected Papers of Enrico Fermi, E. Segre, Ed., vol. 2, p. 978.University of Chicago Press, Chicago, 1965.

[37] S. Aubry, “Breathers in nonlinear lattices: Existence, linear stability andquantization,” Physica D, vol. 103, pp. 201–250, 1997.

[38] R. Reigada, A. Sarmiento, and K. Lindenberg, “Energy relaxation in nonlinearone-dimensional lattices,” Phys. Rev. E, vol. 64, pp. 066608, 2001.

[39] S. R. Bickham and J. L. Feldman, “Calculation of vibrational lifetimes in amorphoussilicon using molecular dynamics simulation,” Phys. Rev. B, vol. 57, pp. 12234, 1998.

[40] J.B. Page, “Asymptotic solutions for localized vibrational modes in stronglyanharmonic periodic systems,” Phy. Rev. B, vol. 41, no. 11, pp. 7835, 1990.

[41] A.J. Sievers and S. Takeno, “Intrinsic localized modes in anharmonic crystals,”Phys. Rev. Lett., vol. 61, pp. 970–937, 1998.

[42] R.H.H. Poetzsch and H. Bottger, “Interplay of disorder and anharmonicity in heatconduction: Molecular-dynamics study,” Physical Review B, vol. 50, no. 21, pp.15757–15763, 1994.

105

[43] K. Ullmann, A. J. Lichtenberg, and G. Corso, “Energy equipartition starting fromhigh-frequency modes in the Fermi-Pasta-Ulam β-oscillator chain,” Phys. Rev. E,vol. 61, pp. 2471, 2000.

[44] G. Kopidakis and S. Aubry, “Intraband discrete breathers in disordered nonlinearsystems. II. Localization.,” Physica D, vol. 139, pp. 247, 2000.

[45] P. G. Kevrekidis, K. Ø. Rasmussen, and A. R. Bishop, “Two-dimensional discretebreathers: Construction, stability, and bifurcations,” Phys. Rev. E, vol. 61, pp. 2006,2000.

[46] P. G. Kevrekidis, K. Ø. Rasmussen, and A. R. Bishop, “Comparison of one-dimensional and two-dimensional discrete breathers,” Mathematics and Computersin Simulation, vol. 55, pp. 449, 2001.

[47] M. Johansson and S. Aubry, “Growth and decay of discrete nonlinear Schrodingerbreathers interacting with internal modes or standing-wave phonons,” Phys. Rev. E,vol. 61, pp. 5864, 2000.

[48] F. Zhang, D. J. Isbister, and D. J. Evans, “Nonequilibrium molecular dynamicssimulations of heat flow in one-dimensional lattices,” Phys. Rev. E, vol. 61, pp. 3541,2000.

[49] I. Bena, A. Saxena, G. P. Tsironis, M. Ibanes, and J. M. Sancho, “Confinement ofdiscrete breathers in inhomogeneously profiled nonlinear chains,” Phys. Rev. E, vol.67, pp. 037601, 2003.

[50] F. Piazza, S. Lepri, and R. Livi, “Cooling nonlinear lattices toward energylocalization,” Chaos, vol. 13, pp. 637, 2003.

[51] G. P. Tsironis and S. Aubry, “Slow relaxation phenomena induced by breathers innonlinear lattices,” Phys. Rev. Lett., vol. 77, pp. 5225, 1996.

[52] M. Peyrard, “The pathway to energy localization in nonlinear lattices,” Physica D,vol. 119, pp. 184, 1998.

[53] T. Cretegny, T. Dauxois, S. Ruffo, and A. Torcino, “Localization and equipartitionof energy in the β-FPU chain: Chaotic breathers,” Physica D, vol. 121, pp. 109,1998.

[54] A. Franchini, V. Bortolani, and R. F. Wallis, “Surface and gap intrinsic localizedmodes in one-dimensional III-IV semiconductors,” J. Phys.: Condens. Matter, vol.12, pp. 1, 2000.

[55] G. Kopidakis and S. Aubry, “Intraband discrete breathers in disordered nonlinearsystems. I. Delocalization.,” Physica D, vol. 130, pp. 155, 1999.

106

[56] B. I. Swanson, J. A. Brozik, S. P. Love, G. F. Strouse, A. P. Shreve, A. R. Bishop,W.-Z. Wang, and M. I. Salkola, “Observation of intrinsically localized modes in adiscrete low-dimensional material,” Phys. Rev. Lett., vol. 82, pp. 3288, 1999.

[57] K. Kisoda, N. Kimura, H. Harima, K. Takenouchi, and M. Nakajima, “Intrinsiclocalized vibrational modes in a highly nonlinear halogen-bridged metal,” J.Luminescence, vol. 94, pp. 743, 2001.

[58] A. Xie, L. van der Meer, W. Hoff, and R. H. Austin, “Long-lived Amide I vibrationalmodes in myoglobin,” Phys. Rev. Lett., vol. 84, pp. 5435, 2000.

[59] J. F. Moreland, J. B. Freund, and G. Chen, “The disparate thermal conductivityof carbon nanotubes and dimond nanowires studied by atomistic simulation,”Microscale Therm. Eng., vol. 8, pp. 61–69, 2004.

[60] M. P. Allen and D. J. Tildesley, Computer Simulation of Liquids, Oxford UniversityPress, New York, 1987.

[61] A. V. Savin and O. I. Savina, “Nonlinear dynamics of carbon molecularlattices: Soliton plane waves in graphite layers and supersonic acoustic solitonsin nanotubes,” Phys. Solid State, vol. 46, pp. 383–391, 2004.

[62] T. Yu. Astakhova, M. Menon, and G. A. Vinogradov, “Solitons in inharmoniclattices: Application to achiral carbon nanotubes,” Phys. Rev. B, vol. 70, pp.125409, 2004.

[63] R. Saito, G. Dresselhaus, and M. S. Dresselhaus, Physical Properties of CarbonNanotubes, Imperial College Press, London, 1998.

[64] D.W. Brenner, “Empirical potential for hydrocarbons for use in simulating thechemical vapor deposition of diamond films,” Phys. Rev. B, vol. 42, pp. 9458–9471,1990.

[65] J. Tersoff, “Empirical interatomic potential for carbon, with applications toamorphous carbon,” Phys. Rev. Lett., vol. 61, pp. 2879–2882, 1988.

[66] J. Tersoff, “New empirical approach for the structure and energy of covalentsystems,” Phys. Rev. B, vol. 37, pp. 6991–7000, 1988.

[67] KM Liew, CH Wong, and MJ Tan, “Tensile and compressive properties of carbonnanotube bundles,” Acta Materialia, vol. 54, no. 1, pp. 225–231, 2006.

[68] H. RAFII-TABAR, “Computational modelling of thermo-mechanical and transportproperties of carbon nanotubes,” Physics reports, vol. 390, no. 4-5, pp. 235–452,2004.

[69] A. Kalra, S. Garde, and G. Hummer, “From The Cover: Osmotic water transportthrough carbon nanotube membranes,” Proceedings of the National Academy ofSciences, vol. 100, no. 18, pp. 10175, 2003.

107

[70] W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery, NumericalRecipes in Fortran. The Art of Scientific Computing, Cambridge University Press,2nd edition, 1992.

[71] A. A. Maradudin, E. W. Montroll, G. H. Weiss, and I. P. Ipatova, Theory of LatticeDynamics in the Harmonic Approximation, Academic Press, New York, 1971.

[72] M. Tatlier and A. Erdem-Senatalar, “The performance analysis of a solar adsorptionheat pump utilizing zeolite coatings on metal supports,” Chemical EngineeringCommunications, vol. 180, pp. 169–185, 2000.

[73] J. M. Gordon, K. C. Ng, H. T. Chua, and A. Chakraborty, “The electro-adsorptionchiller: a miniaturized cooling cycle with applications to micro-electronics,”International Journal of Refrigeration-Revue Internationale Du Froid, vol. 25,no. 8, pp. 1025–1033, 2002.

[74] Y. Hudiono, A. Greenstein, C. Saha-Kuete, B. Olson, S. Graham, and S. Nair,“Effects of composition and phonon scattering mechanisms on thermal transport inmfi zeolite films,” Journal of Applied Physics, vol. 102, 2007.

[75] G. P. Srivastava, The Physics of Phonons, Adam Hilger, Bristol, 1990.

[76] T. R. Forester and W. Smith, “Bluemoon simulations of benzene in silicalite-1.Predictions of free energies and diffusion coefficients,” J. Chem. Soc., FaradayTrans., vol. 93, pp. 3249–3257, 1997.

[77] D. I. Kopelevich and H.-C. Chang, “Diffusion of inert gases in silica sodalite:Importance of lattice flexibility,” J. Chem. Phis., vol. 115, pp. 9519, 2001.

[78] S.C. Turaga and S. M. Auerbach, “Calculating free energies for diffusion in tight-fitting zeolite-guest systems: Local normal-mode Monte Carlo,” J. Chem. Phys., vol.118, pp. 6512–6517, 2003.

[79] F. Jousse, D. P. Vercauteren, and S. M. Auerbach, “How does benzene in NaYzeolite couple to the framework vibrations?,” J. Phys. Chem. B, vol. 104, pp.8768–8778, 2000.

[80] A. M. Greenstein, S. Graham, Y. C. Hudiono, and S. Nair, “Thermal properties andlattice dynamics of polycrystalline MFI zeolite films,” Nanoscale and MicroscaleThermophysical Engineering, vol. 10, pp. 321–331, 2006.

[81] J. J. Dong, O. F. Sankey, and C. W. Myles, “Theoretical study of the lattice thermalconductivity in ge framework semiconductors,” Physical Review Letters, vol. 86, no.11, pp. 2361–2364, 2001.

[82] R. Tsekov and E. Ruckenstein, “Stochastic dynamics of a subsystem interacting witha solid body with application to diffusive processes in solids,” J. Chem. Phys., vol.100, pp. 1450–1455, 1994.

108

[83] P. Demontis and G. B. Suffritti, “Structure and dynamics of zeolites investigated bymolecular dynamics,” Chem. Rev., vol. 97, pp. 2845–2878, 1997.

[84] D. I. Kopelevich and H.-C. Chang, “Does lattice vibration drive diffusion inzeolites?,” J. Chem. Phys., vol. 114, pp. 3776–3789, 2001.

[85] J. W. Richardson, J. J. Pluth, J. V. Smith, W. J. Dytrych, and D. M. Bibby,“Conformation of ethylene-glycol and phase-change in silica sodalite,” J. Phys.Chem., vol. 92, pp. 243, 1988.

[86] K. Knorr, C. M. Braunbarth, G. van de Goor, P. Behrens, C. Griewatsch, andW. Depmeier, “High-pressure study on dioxolane silica sodalite (C3H6O2)2[Si12O24] –neutron and X-ray powder diffraction experiments,” Solid State Commun., vol. 113,pp. 503, 2000.

[87] J. B. Nicholas, A. J. Hopfinger, F. R. Trouw, and L. E. Iton, “Molecular modeling ofzeolite structure. 2. Structure and dynamics of silica sodalite and silicate force-field,”J. Am. Chem. Soc., vol. 113, pp. 4792, 1991.

[88] S. El Amrani, F. Vigne, and B. Bigot, “Self-diffusion of rare-gases in silicalitestudied by molecular dynamics,” J. Phys. Chem., vol. 96, pp. 9417, 1992.

[89] D. Keffer, A. V. McCormick, and H. T. Davis, “Unidirectional and single-filediffusion in AlPO4-5: Molecular dynamics investigations,” Mol. Phys., vol. 87, pp.367, 1996.

[90] S. J. Goodbody, K. Watanabe, D. MacGowan, J. Walton, and N. Quirke, “Molecularsimulation of methane and butane in silicalite,” J. Chem. Soc. Faraday Trans., vol.87, pp. 1951, 1991.

[91] R. L. June, A. T. Bell, and D. N. Theodorou, “Transition-state studies of xenon andSF6 diffusion in silicalite,” J. Phys. Chem., vol. 95, pp. 8866, 1991.

[92] A. V. Kiselev, A. A. Lopatkin, and A. A. Shulga, “Molecular statistical calculationof gas adsorption by silicalite,” Zeolites, vol. 5, pp. 261, 1985.

[93] A. J. C. Ladd, B. Moran, and W. G. Hoover, “Lattice thermal conductivity: Acomparison of molecular dynamics and anharmonic lattice dynamics,” Phys. Rev. B,vol. 34, pp. 5058, 1986.

[94] A. J. H. McGaughey and M. Kaviany, “Quantitative validation of the boltzmanntransport equation phonon thermal conductivity model under the single-moderelaxation time approximation,” Physical Review B, vol. 69, no. 9, 2004.

[95] G. W. Gardiner, Handbook of Stochastic Methods for Physics, Chemistry, and theNatural Sciences, Springer-Verlag, Berlin, 1983.

[96] J. Luczka, “Non-markovian stochastic processes: Colored noise,” Chaos, vol. 15, no.2, 2005.

109

[97] A. J. H. McGaughey and M. Kaviany, “Thermal conductivity decomposition andanalysis using molecular dynamic simulations. Part II. Complex silica structures,”Int. J. Heat Mass Transfer, vol. 47, pp. 1799–1816, 2004.

[98] P. G. Klemens, “The scattering of low-frequency lattice waves by staticimperfections,” Proceedings of the Physical Society of London Section A, vol.68, no. 12, pp. 1113–1128, 1955, 16.

[99] WA Little, “THE TRANSPORT OF HEAT BETWEEN DISSIMILAR SOLIDSAT LOW TEMPERATURES,” Canadian Journal of Physics, vol. 37, no. 3, pp.334–349, 1959.

[100] AJ Lichtenberg and Lieberman MA, Regular and Chaotic Dynamics, Springer, 1992.

[101] P. K. Schelling, S. R. Phillpot, and P. Keblinski, “Phonon wave-packet dynamicsat semiconductor interfaces by molecular-dynamics simulation,” Applied PhysicsLetters, vol. 80, no. 14, pp. 2484–2486, 2002, 12.

110

BIOGRAPHICAL SKETCH

Chia-Yi Chen was born in Taiwan on September 4th, 1981. She received the

Bachelor of Science from National Taiwan University, Taipei, Taiwan in June 2003. After

completing her bachelor’s degree she joined the Department of Chemical Engineering,

University of Florida, in August 2003. Her research interests are modeling of thermal

transport in nanostuctured materials.

111


Recommended