effects of gas interactions on the transport roperties

180
The Pennsylvania State University The Graduate School Department of Physics EFFECTS OF GAS INTERACTIONS ON THE TRANSPORT PROPERTIES OF SINGLE-WALLED CARBON NANOTUBES A Thesis in Physics by Hugo E. Romero © 2004 Hugo E. Romero Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy May 2004

Upload: others

Post on 21-Dec-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

The Pennsylvania State University

The Graduate School

Department of Physics

EFFECTS OF GAS INTERACTIONS ON THE TRANSPORT PROPERTIES

OF SINGLE-WALLED CARBON NANOTUBES

A Thesis in

Physics

by

Hugo E. Romero

© 2004 Hugo E. Romero

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

May 2004

The thesis of Hugo E. Romero has been reviewed and approved* by the following:

Peter C. Eklund Professor of Physics and Professor of Materials Science and Engineering Thesis Adviser Chair of Committee Gerald D. Mahan Distinguished Professor of Physics Vincent H. Crespi Downsbrough Professor of Physics and Professor of Materials Science and Engineering James H. Adair Professor of Materials Science and Engineering Jayanth R. Banavar Professor of Physics Head of the Department of Physics * Signatures are on file in the Graduate School

ABSTRACT

The work presented in this thesis discusses a series of in situ transport properties

measurements (thermoelectric power S and electrical resistance R) on networks of randomly

oriented single-walled carbon nanotube (SWNT) bundles (e.g., thin films, mats, and

buckypapers), in contact with various gases and chemical vapors. Results are presented on the

effects of gases that chemisorb and undergo weak charge transfer reactions with the carbon

nanotubes (e.g., O2 and NH3), gases and chemical vapors that physisorb on the tube wall (e.g., H2,

alcohols, water and cyclic hydrocarbons), and gases and small molecules that undergo collisions

with the carbon nanotube walls (e.g., inert gases, N2, CH4). The strong, systematic effects on the

transport properties of SWNTs due to exposure to six-membered ring and polar molecules

(alcohol and water) are found to increase with the quantity Ea/A, where Ea is the adsorption

energy and A is the molecular projection area. The magnitudes of the remarkable effects of

collisions of inert gases (He, Ne, Ar, Kr, and Xe) and small molecules (N2 and CH4) on the

transport properties of SWNTs are found to be proportional to ~ M1/3, where M is the mass of the

colliding species. This is approximately the same mass dependence exhibited by the maximum

deformation of the tube wall and the energy exchanged between the tube wall and the colliding

atoms as a result of this collision.

A model is proposed to explain the unusual behavior of the thermoelectric power in

SWNTs, wherein the metallic tubes provide the dominant contribution to this physical quantity

and the observed peak at ~ 100 K is attributed to the phonon drag effect. In addition, the details

of our transport model for the behavior of the carbon nanotubes in the presence of gases and

chemical vapors are presented, incorporating the effects of a new scattering channel for the

iii

charge carriers, associated with the adsorbed (or colliding) atoms and molecules. The model is

found to explain qualitatively the various transport phenomena observed.

iv

TABLE OF CONTENTS

LIST OF TABLES..................................................................................................................... VIII

LIST OF FIGURES ...................................................................................................................... IX

CHAPTER 1. INTRODUCTION .............................................................................................. 1

1.1. Structure of Carbon Nanotubes............................................................................... 3

1.2. Electronic Structure of Nanotubes .......................................................................... 8

1.3. Motivations for this Work..................................................................................... 13

CHAPTER 2. EXPERIMENTAL TECHNIQUES.................................................................. 15

2.1. Sample Preparation ............................................................................................... 15

2.1.1. The Arc-Discharge Method ......................................................................... 19

2.2. Thermoelectric Power Measurements................................................................... 20

2.2.1. The Analog Subtraction Circuit ................................................................... 21

2.2.2. The Thermopower Probe ............................................................................. 23

2.2.3. Experimental Setup...................................................................................... 26

2.2.4. The Thermopower Program......................................................................... 28

2.2.5. Calibration of the Thermocouples ............................................................... 30

2.3. Four-Probe Resistance Measurements.................................................................. 33

2.4. Gas/Chemical Adsorption Measurements............................................................. 37

CHAPTER 3. THERMOELECTRIC POWER OF SINGLE-WALLED CARBON

NANOTUBES ........................................................................................................................... 39

3.1. Seebeck Effect: Theory......................................................................................... 39

3.2. Thermoelectric Power of Carbon Nanotubes: Background.................................. 43

v

3.2.1. Parallel Heterogeneous Model of Metallic and Semiconducting Pathways 45

3.2.2. Variable-Range Hopping ............................................................................. 48

3.2.3. Electron-Phonon Enhancement.................................................................... 50

3.2.4. Fluctuation-Assisted Tunneling................................................................... 50

3.2.5. Kondo Effect................................................................................................ 51

3.2.6. Thermoelectric Power of Oxidized SWNT Networks ................................. 52

3.3. Thermoelectric Power of SWNT Films ................................................................ 55

3.3.1. Role of Contact Barriers on the Transport Properties of SWNTs ............... 56

3.3.2. Effect of Oxygen Doping on the Thermoelectric Power of SWNTs ........... 58

3.3.3. Compensating Doping and Defect Chemistry ............................................. 63

3.3.4. Model Calculations of the Thermoelectric Power of SWNTs ..................... 65

3.3.5. Thermopower from Enhanced D(EF) due to Impurities............................... 69

CHAPTER 4. PHONON DRAG THERMOELECTRIC POWER OF SINGLE-WALLED

CARBON NANOTUBES............................................................................................................. 70

4.1. Introduction........................................................................................................... 70

4.2. Phonon Drag Model.............................................................................................. 73

4.2.1. Phonon Lifetimes ......................................................................................... 74

4.3. Baylin Formalism Applied to Metallic Carbon Nanotubes .................................. 76

CHAPTER 5. CARBON NANOTUBES: A THERMOELECTRIC NANO-NOSE .............. 82

5.1. Introduction........................................................................................................... 82

5.2. Effects of Gas Adsorption on the Electrical Transport Properties of SWNTs ..... 85

5.3. Thermoelectric Power from Multiple Scattering Processes.................................. 89

vi

CHAPTER 6. EFFECTS OF MOLECULAR PHYSISORPTION ON THE TRANSPORT

PROPERTIES OF CARBON NANOTUBES.............................................................................. 95

6.1. Introduction........................................................................................................... 95

6.2. Effects of Adsorption of Six-Membered Ring Molecules .................................... 96

6.3. Effects of Adsorption of Polar Molecules .......................................................... 103

CHAPTER 7. EFFECTS OF GAS COLLISIONS ON THE TRANSPORT PROPERTIES OF

CARBON NANOTUBES........................................................................................................... 115

7.1. Introduction......................................................................................................... 115

7.2. Collision-Induced Electrical Transport of Carbon Nanotubes............................ 116

1.1. Molecular Dynamics Simulations....................................................................... 122

CHAPTER 8. CONCLUSIONS AND FUTURE WORK ..................................................... 130

APPENDIX A: DERIVATION OF THE MOTT RELATION.................................................. 136

APPENDIX B: DERIVATION OF THE PHONON DRAG THERMOPOWER ..................... 138

BIBLIOGRAPHY....................................................................................................................... 141

vii

List of Tables

Table 4–1. Best fit parameter values achieved with Eq. (4.12) ........................................ 79

Table 6–1. Comparison of the T = 40 ºC thermoelectric power and resistive responses of

a SWNT thin film to adsorbed C6H2n molecules. The vapor pressure at 24 ºC

and the adsorption energy Ea of the corresponding molecule (measured on

graphitic surfaces) are also listed. S0 and R0 refer to the degassed film before

exposure to C6H2n molecules. ........................................................................ 98

Table 6–2. Comparison of the T = 40 ºC thermoelectric power and resistive responses of

a SWNT thin film to adsorbed water and CnH2n+1OH; n = 1-4. The vapor

pressure p at 24 ºC, the molecular area A, the static dipole moment µ, and the

adsorption energy Ea of the corresponding molecule (measured on graphitc

surfaces) are also listed. S0 and R0 refer to the degassed film before exposure

to water and alcohols. An increase in vapor pressure did not change the

values of ∆Smax or ∆Rmax; see text................................................................ 106

Table 6–3. Adsorption time constants for thermoelectric ( Sτ ) and resistive ( Rτ ) response

of a SWNT thin film to adsorbed water and alcohol molecules.................. 108

viii

List of Figures

Figure 1.1. Stable forms of carbon clusters: (a) a piece of graphene sheet, (b) the fullerene

C60, and (c) a model for a carbon nanotube (Adapted from Dresselhaus11). ... 2

Figure 1.2. (a) The unrolled honeycomb lattice of a nanotube. When the lattice sites O

and A, and sites B and B’ are connected, an (n,m) = (4,2) nanotube can be

constructed.50 (b) STM image of a SWNT exposed at the surface of a rope. A

portion of a 2D graphene sheet is overlaid to highlight the atomic structure.51

......................................................................................................................... 4

Figure 1.3. (a) Computer-generated images of (10,10) armchair, (10,0) zigzag, and chiral

type SWNTs. The numbers in parenthesis are the chiral indices. (b) A 2D

graphene sheet showing the schematic of the indexing used for SWNTs. The

large dots denote metallic tubes while the small dots are for semiconducting

tubes.50 ............................................................................................................. 5

Figure 1.4. X-ray diffraction patterns at low angle of a SWNT sample obtained by (a) the

arc-discharge technique by Journet et al.43 and (b) the laser ablation

technique by Thess et al.42 The graphite peak, due to remaining graphitic

particles, has been removed for clarity; its position is shown by an asterisk.

The inset shows a single SWNT rope made up of ~100 SWNTs as it bends

through the image plane of the microscope, showing uniform diameter and

triangular packing of the tubes within the rope.42............................................ 7

ix

Figure 1.5. Graphene π band structure in the first Brillouin zone, constructed using Eq.

(1.4). The conduction and valence bands touch at the six Fermi points K

indicated at E = 0. ............................................................................................ 9

Figure 1.6. Examples of the allowed 1D subbands for (a) a (5,5) armchair, (b) a (5,0)

zigzag, and (c) a (7,1) chiral carbon nanotube. The hexagon defines the first

Brillouin zone of graphene and the dots in the corners are the graphene K

points.............................................................................................................. 10

Figure 1.7. One-dimensional energy dispersion relations for (a) armchair (10,10) tubes,

(b) zigzag (10,0) tubes, and (c) chiral (6,4) tubes, computed using the zone-

folded tight-binding dispersion relations described in the text...................... 11

Figure 1.8. Electronic 1D density of states per unit cell for a series of metallic tubes,

showing discrete peaks at the positions of the 1D band maxima or minima

(Adapted from Dresselhaus55). ...................................................................... 12

Figure 2.1. Room-temperature Raman spectrum for unpurified arc-derived SWNTs

excited at 514.5 nm........................................................................................ 17

Figure 2.2. Right: SEM image of a SWNT film showing entangled ropes synthesized by

the arc-discharge method. Left: High-resolution TEM image of an end view

of the SWNT bundles, showing the 2D hexagonal lattice arrangement of the

tubes............................................................................................................... 18

Figure 2.3. TEM images of SWNT bundles before (right) and after (left) purification.

Dark spots are catalyst clusters...................................................................... 18

Figure 2.4. Block diagram of the analog subtraction circuit to measure the thermoelectric

power. ............................................................................................................ 21

x

Figure 2.5. Schematic diagram of the thermopower and resistance measurements probe

suitable for the temperature range 4-500 K. .................................................. 24

Figure 2.6. Schematic diagram of the sample holder for the thermoelectric power and

four-probe resistance measurements.............................................................. 25

Figure 2.7. Block diagram of the system for thermoelectric power and four-probe

resistance measurements instrumentation...................................................... 27

Figure 2.8. The program “Thermopower Auto.vi”, showing the time evolution of the

thermoelectric power of a SWNT film during vacuum-degassing at 500 K. 28

Figure 2.9. The program “Thermopower Manual.vi”, showing the temperature

dependence of the thermoelectric power of constantan................................. 30

Figure 2.10. Schematic diagram of the connections to A2 amplifier to measure the sample

temperature (top) and the equivalent circuit (bottom). .................................. 31

Figure 2.11. The output voltage of amplifier A2 as a function of the sample temperature

(left) and the temperature dependence of the relative thermoelectric power of

a chromel-Au:Fe thermocouple pair (right). ................................................. 32

Figure 2.12. Temperature dependence of the thermoelectric power of chromel and

gold:iron alloy with respect to copper. The solid lines represent polynomial

fits to the data................................................................................................. 33

Figure 2.13. The program “DC 4-Probe Resistance.vi” showing the resistance as a

function of temperature for a SWNT mat...................................................... 35

Figure 2.14. Schematic diagram of the gas handling system for gas/chemical adsorption

experiments.................................................................................................... 37

xi

Figure 3.1. (a) Basic thermoelectric open circuit that displays the Seebeck effect. (b) The

Seebeck effect: A temperature gradient along a conductor gives rise to a

potential difference. ....................................................................................... 40

Figure 3.2. Temperature dependence of the thermoelectric power for an air-saturated

SWNT mat. The solid line is a guide to the eye. The dashed lines (a) and (b)

represent the ways in which metallic behavior could be incorporated in the

thermoelectric behavior of SWNTs. .............................................................. 44

Figure 3.3. Illustration of the combination of thermoelectric powers for conductors in

parallel (also applicable to the two-band model)........................................... 46

Figure 3.4. Fits to measured thermoelectric power of a SWNT mat using a parallel

heterogeneous model of semiconducting and metallic tubes. The solid line

represents a fit to the data using Eq.(3.13). Fitting parameters extracted from

our fit are also shown in the figure. ............................................................... 47

Figure 3.5. Fits to measured thermoelectric power of a SWNT mat using a parallel

heterogeneous model of disordered semiconducting and metallic tubes. The

solid line represents a fit to the data using Eq. (3.16). Fitting parameters

extracted from the fit are also shown in the figure. ....................................... 49

Figure 3.6. The temperature dependence of the thermoelectric power for an “as-prepared”

SWNT mat in its air-saturated and degassed states. The solid lines are guides

to the eye........................................................................................................ 53

Figure 3.7. Sketch of crystalline SWNT ropes, where fibrillar carbon nanotubes are

separated by disordered regions (Adapted from Kaiser et al.107) .................. 57

xii

Figure 3.8. Uniaxial pressure dependence of (a) the normalized room temperature

resistance R/R0 and (b) the thermopower S for two different “as-prepared”

SWNT mats. The inset shows the experimental geometry where the applied

force F is perpendicular to the sample.108...................................................... 58

Figure 3.9. Thermopower response to vacuum and O2 (1 atm) at T = 500 K. (A → C):

Vacuum-degassing of a sample initially O2-doped under ambient conditions

for several days. (C → D): Exposure of the degassed sample to 1 atm of O2

established at C.108 ......................................................................................... 59

Figure 3.10. Temperature dependence of the thermopower S for a SWNT thin film after

successively longer periods of O2 degassing at T = 500 K in vacuum. The

labels A, B, and C refer to a vacuum-degassing interval indicated in Figure

3.9. Curve D is for the same sample exposed to 1 atm O2 at T = 500 K for

about 4 h after being fully degassed to point C.108 ........................................ 62

Figure 3.11. Calculated thermoelectric power of a (10,10) carbon nanotube as a function

of the Fermi level position............................................................................. 68

Figure 4.1. Sketch of the thermoelectric power of a simple quasi-free electron pure metal

as a function of temperature. A: Electron diffusion component of

thermoelectric power approximately proportional to T. B: Phonon drag

component with magnitude increasing as T 3 at very low temperatures (T <<

TD), and decaying as 1/T at “high” temperatures (T > TD) (Adapted from

MacDonald69). ............................................................................................... 72

Figure 4.2. Temperature dependence of the thermoelectric power for a purified SWNT

thin film after successively longer periods of O2 degassing at 500 K in

xiii

vacuum. Curve 1 corresponds to the same sample exposed to 1 atm O2 at 500

K for about 4 h, after being fully degassed (curve 4). The solid lines in the

figure represent the fits to the data using Eq. (4.12)...................................... 77

Figure 4.3. Temperature dependence of the thermoelectric power for SWNT mats

prepared using different catalysts. The samples were not purified and

contained ~ 5 at% residual catalyst. The data were measured by Grigorian et

al.76 The solid lines represent the best fits to the data using Eq. (4.12). ....... 80

Figure 4.4. Fits to the measured thermoelectric power data (curve 1 in Figure 4.2) using a

model involving diffusion and phonon drag contributions to the

thermoelectric power. The solid curve represents a fit to the data using Eq.

(4.12). The dashed lines represent the contributions from Sd [Eq. (3.3)] and Sg

[Eq. (4.10)]..................................................................................................... 81

Figure 5.1. Schematic structure of a SWNT bundle showing the sites available for gas

adsorption. The dashed line indicates the nuclear skeleton of the nanotubes.

Binding energies EB and specific surface area contributions σ for hydrogen

adsorption on these sites are indicated.133...................................................... 84

Figure 5.2. The time dependence of the thermoelectric power response of a SWNT mat to

1 atm overpressure of He gas (filled circles), and to the subsequent

application of a vacuum over the sample (open circles). The dashed lines are

exponential fits of the data (see text).136 ........................................................ 86

Figure 5.3. In situ thermoelectric power versus time after exposure of a vacuum-degassed

SWNT mat to 1 atm overpressure of H2 at T = 500 K (solid symbols). The

response of the H2-loaded SWNT sample to a vacuum is also represented

xiv

(open symbols). The dashed lines are fits to the data using exponential

functions (see text)......................................................................................... 88

Figure 5.4. In situ thermoelectric power as a function of time after exposure of degassed

SWNT mats to a 1 atm overpressure of H2 at T = 500 K (solid symbols). The

open symbols are the response of the H2 loaded SWNT system to a vacuum.

Data are shown for three samples: not purified (bottom), HCl reflux for 4 h

(middle), HCl reflux for 24 h (top). The dashed lines are guides to the eye.

The catalyst residue in at% is indicated......................................................... 89

Figure 5.5. Nordheim-Gorter plots showing the effect of gas adsorption on the electrical

transport properties of a SWNT mat. The amount of gas stored in the bundles

increases to the right, tracking the increase in ρ. For the H2 data, the open

circles are from the time dependent response to 1 atm of H2 at T = 500 K and

the closed circles are from a pressure study at the same temperature. The

inset shows the Nordheim-Gorter plots for O2 (electron acceptor) and NH3

(electron donor). Note that the data in the inset, as opposed to that in the main

plot, is non-linear. The non-linearity is consistent with charge transfer and

Fermi energy shifts. ....................................................................................... 92

Figure 6.1. In situ (a) thermoelectric power and (b) resistance responses at 40 ºC as a

function of time during successive exposure of a degassed SWNT thin film to

vapors of six-membered ring molecules C6H2n; n = 3-6. The dashed lines are

guides to the eye. The vapor pressure was ~ 12 kPa. .................................... 97

Figure 6.2. Maximum change of the thermoelectric power of a SWNT film as a function

of the adsorption energy of the adsorbed molecule. The dashed line is a guide

xv

to the eye........................................................................................................ 99

Figure 6.3. S vs. ∆R/R0 plots during exposure to C6H2n (n = 3-6). The dashed curve is a fit

to the data using a quadratic function. ......................................................... 100

Figure 6.4. Temperature dependence of the thermoelectric power of the degassed SWNT

after saturation coverage of the various C6H2n molecules. The dashed lines

are guides to the eye. ................................................................................... 102

Figure 6.5. Time dependence of the (a) thermoelectric power and (b) normalized four-

probe resistance responses to vapors of water and alcohol molecules

(CnH2n+1OH; n = 1-4) at 40 ºC. The dashed lines are fit to S(t) and R(t) data

using an exponential function. The inset shows a simple schematic of the

measurement apparatus. The liquid temperature T2 establishes the vapor

pressure in the sample chamber which is at a temperature T1 > T2. The system

is evacuated through V2. After degassing, V2 is closed and V1 is opened. The

responses of S and R are then measured simultaneously. ............................ 105

Figure 6.6. S vs. ∆R/R0 plots during exposure of degassed SWNT bundles to water and

CnH2n+1OH (n = 1-4). The solid lines are linear fits to the data until saturation

is established................................................................................................ 110

Figure 6.7. Maximum thermoelectric power change ∆Smax of a SWNT thin film

successively exposed to vapors of water and alcohol molecules (CnH2n+1OH;

n = 1-4) as a function of the quantity AEaβ , where Ea and A are,

respectively, the molecular adsorption energy and the projection area. The

solid and dashed lines are guides to the eye. ............................................... 113

xvi

Figure 7.1. Time dependence of the thermoelectric power response of (a) PLV

buckypaper and (b) arc-derived thin film exposed to 1 atm of inert gas

(closed symbols), and to subsequent application of vacuum over the sample

(open symbols) at T = 500 K. The different values of S0 in (a) and (b) reflect

differences in defect densities in the PLV and the arc-derived material (see

Chapter 5). ................................................................................................... 118

Figure 7.2. S vs. ∆R/R0 plots showing the effect of inert gases on the transport properties

of a SWNT buckypaper prepared from PLV material. The closed symbols are

from the time evolution of S and R to 1 atm of gas at T = 500 K and the open

symbols are from a pressure study at the same temperature, where the

maximum response of S and R to a given pressure was measured. The inset

shows the pressure dependence of the maximum change of thermopower for

the same sample. .......................................................................................... 120

Figure 7.3. Computed power spectra of the radial motion of a C-atom nearest the point of

contact in a (10,0) carbon nanotube at 0 K. The figure shows the phonons

induced during (a) the first 5 ps of the collision (and includes the gas-tube

impact) and (b) the second 5 ps after the collision. The inset to (a) shows the

side view of a collision between a Xe atom (θi = 0º, Ei = 13 kcal/mol) and a

nanotube. The inset to (b) shows the schematic representation of the tube wall

deformation in response to an atom collision. ............................................. 124

Figure 7.4. Maximum thermoelectric power change ∆Smax of two SWNT samples

exposed to gases indicated (ARC: open circles and PLV: closed circles; data

from Figure 7.1), calculated total energy gained by a (10,0) nanotube upon

xvii

collision with a gas atom (θi = 0º, Ei = 3.97 kcal/mol, squares), and maximum

radial displacement ∆Dmax of the tube C-atom immediately after impact with

a gas atom (θi = 45º, Ei = 1.99 kcal/mol, triangles) as a function of the mass

of the colliding inert gas. The lines are power law fits to the data of the forms

35.0max 08.3 MS =∆ , 39.091.0 ME =∆ , and 35.0

max 04.0 MD =∆ . ................... 126

Figure 7.5. Dipole polarizability α as a function of the mass of the inert atom or small

molecule....................................................................................................... 128

xviii

DEDICATION

To my beloved parents,

Guillermo and Sergia,

who taught me the values I treasure,

who gave me the freedom of choice.

For their dedication and commitment to

furnish their children with the best possible future.

Without their sacrifice none of this would be possible.

To my brothers and sisters,

Zulay, Alba, Guillermo, John, Ana, Zulma, and Celeste,

my best friends.

I have been blessed with the good fortune and

privilege of having such wonderful people in my life.

To my wife,

Francelys,

for her love and support,

for her tolerance and patience,

for the joy and happiness she has brought to my life.

Gracias. Los amos a todos.

xix

Chapter 1.

Introduction

Carbon nanotubes have aroused worldwide excitement since their discovery by Sumio

Iijima in 1991.1 In retrospect, it is quite likely that such fascinating materials were produced as

early as the 1970s during research on carbon fibers by Morinobu Endo.2 The discovery of carbon

nanotubes was stimulated, in part, by the discovery in 1985 of fullerene C60 by groups led by

Harold Kroto at Sussex University and Richard Smalley at Rice University. C60 is a nearly

spherical molecule made of 60 identical carbon atoms bonded in hexagonal and pentagonal rings

[Figure 1.1(b)]. The pentagonal rings are necessary to close the structure. Exactly 12 pentagonal

rings are needed, as can be proven using Euler’s polyhedron theorem.3 Carbon nanotubes, on the

other hand [Figure 1.1(c)], do not require any pentagonal ring in the curved cylindrical surface.

The ends of the nanotube can be closed by a hemispherical fullerene molecule. By Euler’s

theorem, each end cap has exactly 6 pentagonal rings. It is clear that a nanotube can be

considered to be a graphene sheet [Figure 1.1(a)] rolled into a seamless cylinder. However, it is

not clear that this can be done in so many ways to produce a variety of chiral tubular structures.

Carbon nanotubes may be one of the key materials for nanoscale technology. It is hoped

that nanotube electronics may lead to progress in miniaturization of computing and power

devices. Small-diameter carbon nanotubes are attractive materials for nanoelectronics because

they provide a remarkable one-dimensional (1D) system, i.e., their electronic and phonon states

are described by a wave vector along the tube axis. They do not have a Fermi surface but exhibit

only two Fermi wave vectors ± kF. Because of the nearly 1D electronic structure, electronic

transport in carbon nanotubes can occur ballistically (i.e., without scattering) at low temperatures

-1-

and over long nanotube lengths, enabling them to carry high currents with essentially no heat

dissipation.4-7 Phonons also propagate easily along nanotubes; the measured room temperature

thermal conductivity of an individual nanotube (> 3000 W/m·K) is greater than that of natural

diamond and the basal plane of graphite (both 2000 W/m·K).8 Whether one considers phonon or

electron scattering, the interesting point is the limited number of final states into which these

excitations can scatter. This is the benefit from a crystalline 1D material. Small-diameter

nanotubes are also quite stiff in tension and exceptionally strong with Young’s modulus of 1.28

TPa and high tensile strength of 28.5 GPa, exceeding those of steel and SiC.9,10

Figure 1.1. Stable forms of carbon clusters: (a) a piece of graphene sheet, (b) the fullerene C60, and (c) a model for a carbon nanotube (Adapted from Dresselhaus11).

Among the potential applications12,13 proposed for carbon nanotubes are conductive and

high-strength composites,14,15 energy storage and energy conversion devices,16,17 chemical18 and

-2-

gas19 sensors, electron field emission displays20-22 and radiation sources,23-25 hydrogen storage

media,26-31 nanoprobes for AFM and STM tips,32-34 electronic interconnects35,36 and

semiconductor devices (e.g., field effect transistors,37,38 logic gates,39 etc.)

Iijima actually observed multi-walled carbon nanotubes (MWNTs) in his electron

microscope images. They showed tubular filaments consisting of multiple concentric shells.

Approximately two years after the discovery of MWNTs, single-walled nanotubes (SWNTs)

consisting of only one shell of carbon atoms were discovered independently by groups led by

Iijima at the NEC Fundamental Research Laboratory40 and Bethune at IBM’s Almaden Research

Center in California.41 Later work by Richard Smalley and his co-workers at Rice University

enabled the bulk production (i.e., 10s of mg) of ~ 1 nm diameter SWNTs.42 The bulk production,

increased by the arc-discharge approach,43 has led to a vast array of experiments on these

materials to explore their unique and remarkable physical properties, which span a wide range–

from structural to electronic. Here, we will concentrate on the electrical transport properties of

bundles of SWNTs. The literature contains some good reviews on this subject.44-46

1.1. Structure of Carbon Nanotubes

Carbon nanotubes can be described as cylindrical molecules. They have been produced in

the laboratory with diameters as small as ~ 0.4 nm47,48 and lengths up to several millimeters.49

They consist only of carbon atoms and can essentially be thought as a single atomic layer of

graphite (graphene) that has been wrapped into a seamless hollow cylinder; the ends of which

can be open or “capped” with half a fullerene molecule.3,50 A graphene sheet, depicted in Figure

1.2(a), is an sp2 bonded network of carbon atoms arranged in a hexagonal lattice with two atoms

-3-

per unit cell. The experimental verification of the honeycomb structure of a carbon nanotube

became possible via the scanning tunneling microscope (STM) images. A typical atomically

resolved image of the tube’s hexagonal lattice is shown in Figure 1.2(b).

Figure 1.2. (a) The unrolled honeycomb lattice of a nanotube. When the lattice sites O and A, and sites B and B’ are connected, an (n,m) = (4,2) nanotube can be constructed.50 (b) STM image of a SWNT exposed at the surface of a rope. A portion of a 2D graphene sheet is overlaid to highlight the atomic structure.51

The nanotube is uniquely characterized by the so-called chiral vector Ch, defined by

B

B’

O

A

θ

y

x

a1

a2

Ch

(a) (b)

B

B’

O

A

θ

y

x

a1

a2

Ch

(a) (b)

( ) , ,mnmnh ≡+= 21 aaC (1.1)

where a1 and a2 are the unit vectors in the two-dimensional (2D) hexagonal lattice, while n and m

are integers. As shown in Figure 1.2(a), the vector Ch connects two crystallographically

equivalent sites O and A on a 2D graphene sheet, where a carbon atom is located at each vertex

of the hexagonal structure. The chiral angle θ is defined as the angle between the vectors Ch and

a1.

-4-

(10,10) (10,0) (6,4)

d = 13.75 Å d = 7.94 Å d = 6.83 Å

armchair

zigzag

a2

a1

(0,0) (1,0) (2,0) (11,0)(3,0) (4,0) (5,0) (6,0) (7,0) (8,0) (9,0) (10,0)

(1,1) (2,1) (3,1) (4,1) (5,1) (6,1) (7,1) (8,1) (9,1) (10,1)

(2,2) (3,2) (4,2) (5,2) (6,2) (7,2) (8,2) (9,2) (10,2)

(3,3) (4,3) (5,3) (6,3) (7,3) (8,3) (9,3)

(4,4) (5,4) (6,4) (7,4) (8,4) (9,4)

(5,5) (6,5) (7,5) (8,5)

(6,6) (7,6) (8,6)

armchair

zigzag

a2

a1

(0,0) (1,0) (2,0) (11,0)(3,0) (4,0) (5,0) (6,0) (7,0) (8,0) (9,0) (10,0)

(1,1) (2,1) (3,1) (4,1) (5,1) (6,1) (7,1) (8,1) (9,1) (10,1)

(2,2) (3,2) (4,2) (5,2) (6,2) (7,2) (8,2) (9,2) (10,2)

(3,3) (4,3) (5,3) (6,3) (7,3) (8,3) (9,3)

(4,4) (5,4) (6,4) (7,4) (8,4) (9,4)

(5,5) (6,5) (7,5) (8,5)

(6,6) (7,6) (8,6)

(a)

(b)

(10,10) (10,0) (6,4)

d = 13.75 Å d = 7.94 Å d = 6.83 Å

armchair

zigzag

a2

a1

(0,0) (1,0) (2,0) (11,0)(3,0) (4,0) (5,0) (6,0) (7,0) (8,0) (9,0) (10,0)

(1,1) (2,1) (3,1) (4,1) (5,1) (6,1) (7,1) (8,1) (9,1) (10,1)

(2,2) (3,2) (4,2) (5,2) (6,2) (7,2) (8,2) (9,2) (10,2)

(3,3) (4,3) (5,3) (6,3) (7,3) (8,3) (9,3)

(4,4) (5,4) (6,4) (7,4) (8,4) (9,4)

(5,5) (6,5) (7,5) (8,5)

(6,6) (7,6) (8,6)

armchair

zigzag

a2

a1

(0,0) (1,0) (2,0) (11,0)(3,0) (4,0) (5,0) (6,0) (7,0) (8,0) (9,0) (10,0)

(1,1) (2,1) (3,1) (4,1) (5,1) (6,1) (7,1) (8,1) (9,1) (10,1)

(2,2) (3,2) (4,2) (5,2) (6,2) (7,2) (8,2) (9,2) (10,2)

(3,3) (4,3) (5,3) (6,3) (7,3) (8,3) (9,3)

(4,4) (5,4) (6,4) (7,4) (8,4) (9,4)

(5,5) (6,5) (7,5) (8,5)

(6,6) (7,6) (8,6)

(a)

(b)

Figure 1.3. (a) Computer-generated images of (10,10) armchair, (10,0) zigzag, and chiral type SWNTs. The numbers in parenthesis are the chiral indices. (b) A 2D graphene sheet showing the schematic of the indexing used for SWNTs. The large dots denote metallic tubes while the small dots are for semiconducting tubes.50

When the graphene sheet is “rolled up” to form the cylindrical part of the nanotube, the

ends OA of the chiral vector meet each other and the cylinder joint is made by joining the line

AB’ to the parallel line OB in Figure 1.2(a). The chiral vector thus forms the circumference of the

nanotube’s circular cross-section. In terms of the integers (n,m), the nanotube diameter d is given

-5-

by the relation

, 3 22C-C nmnmad h ++

π=

π=

C (1.2)

where aC-C is the nearest-neighbor carbon-carbon distance (1.421 Å in graphite). SWNT

diameters are typically found in the range ~ 0.4 nm < d < 3 nm. For example [Eq. (1.2)], the

diameter of a (10,10) armchair nanotube, shown in Figure 1.3(a), is ~ 13.75 Å. In MWNTs, the

outer tube can be as large as 30-50 nm.

Every pair of integers (n,m) leads to different nanotube structures [Figure 1.3(a)]:

armchair (n,n), zigzag (n,0) and chiral (n,m) nanotubes. Many of the possible vectors specified

by the pairs of integers (n,m) are shown in Figure 1.3(b), which define different ways of rolling

the graphene sheet to form the carbon nanotube with a specific chirality. Because of the point

group symmetry of the honeycomb lattice, several different integers (n,m) will give rise to

equivalent nanotubes. To define each nanotube once, and only once, it is only necessary to

consider the nanotubes arising in the 30º wedge of the 2D Bravais lattice shown in Figure 1.3(b).

The physical properties of nanotubes are determined by their diameter and chiral angle,

both of which depend on n and m. Typically, SWNT samples have a distribution of diameters

and chiral angles.

One interesting characteristic of the growth of the carbon nanotubes is the tendency for

large numbers of nanotubes to grow nearly parallel to each other, forming crystalline-like

bundles or ropes of nanotubes of about 10-50 nm in diameter. These bundles contain from tens to

hundreds of carbon nanotubes of nearly uniform diameter, self-organized in a close-packed

triangular lattice with a typical lattice constant a = 17 Å through van der Waals inter-tube

bonding. Thus, a raw macroscopic SWNT sample consists of a collection of bundles of different

size, with their axes isotropically distributed over all possible orientations.

-6-

Figure 1.4. X-ray diffraction patterns at low angle of a SWNT sample obtained by (a) the arc-discharge technique by Journet et al.43 and (b) the laser ablation technique by Thess et al.42 The graphite peak, due to remaining graphitic particles, has been removed for clarity; its position is shown by an asterisk. The inset shows a single SWNT rope made up of ~100 SWNTs as it bends through the image plane of the microscope, showing uniform diameter and triangular packing of the tubes within the rope.42

Figure 1.4 shows the X-ray diffraction patterns for SWNT samples obtained by the arc-

discharge43 and the laser ablation techniques.42 In an electron microscope, the nanotube material

produced by either of these methods looks like a mat of ropes or bundles of SWNTs. The ropes

are between 10 and 20 nm across and up to 100 µm long.42 The strong discrete peak near Q =

0.44 Å-1, as well as the four weaker peaks up to Q = 1.8 Å-1 in Figure 1.4, indicates the existence

of a 2D triangular lattice of SWNTs organized in bundles.42 The X-ray diffraction also shows

that the diameters of SWNTs in the bundles have a narrow distribution with a strong peak.

-7-

1.2. Electronic Structure of Nanotubes

The remarkable variety of electrical properties of SWNTs stems from the unusual

electronic structure of “graphene”–the 2D material from which they are made. Calculations for

the electronic structure of SWNTs show that carbon nanotubes can be either metallic or

semiconducting, depending on the choice of (n,m). It can be shown that metallic conduction in a

(n,m) carbon nanotube is achieved when

, 3qmn =− (1.3)

where q is an integer. Equation (1.3) shows that all armchair carbon nanotubes are metallic but

only one third of the possible zigzag and chiral nanotubes are metallic. Therefore, from Figure

1.3(b), about 1/3 of nanotubes are metallic and 2/3 are semiconducting.

In the simplest possible model, the band structure of nanotubes can be derived directly

from the 2D band structure of graphene, whose π bands are constructed from the overlapping pz

orbitals of adjacent carbon atoms. The simplest analytical form of the 2D dispersion relation for

the π bands of a single graphene sheet can be expressed in the nearest-neighbor tight-binding

approximation:52

( ) , 2

cos42

cos2

3cos41,

21

020002

⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛+⎟⎟

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛+γ±=

akakakkkE yyx

yxDg (1.4)

where 342.10 ×=a Å is the lattice constant for a 2D graphene sheet and γ0 is the nearest-

neighbor carbon-carbon overlap integral. Currently, the value γ0 ~ 2.9 eV is used to fit optical

data.53 We have used the tight-binding scheme [Eq. (1.4)] to compute the π bands of graphene in

the first Brillouin zone and the results are shown in Figure 1.5. The π* antibonding band and the

π bonding band, respectively, form the conduction and the valence bands of graphene. Since

-8-

there are two atoms per unit cell in a graphene sheet, the valence band is completely filled. Only

the π electrons contribute to the graphene electrical conduction. Note that the conduction and the

valence bands touch at the six corners (K points) of the hexagonal Brillouin zone, where the

Fermi energy EF = 0.

Figure 1.5. Graphene π band structure in the first Brillouin zone, constructed using Eq. (1.4). The conduction and valence bands touch at the six Fermi points K indicated at E = 0.

Using Eq. (1.4), 1D dispersion relations for carbon nanotube (n,m) can be calculated

based on a simple zone folding consideration, i.e., by imposing a periodic boundary condition

around the waist of a SWNT. The allowed wave vectors k in the direction parallel to the chiral

vector, resulting from radial confinement, follow from

, 2 qh π=⋅ kC (1.5)

where q is an integer. The 1D energy dispersion curves of a nanotube correspond to the cross-

section of the 2D energy dispersion surface shown in Figure 1.5, where the cuts are made on

-9-

parallel lines corresponding to the particular set of allowed states.3 In Figure 1.6 several cutting

lines, representing the allowed subbands of a nanotube, are shown. On the basis of this simple

scheme, if one of the allowed wave vectors passes through a Fermi point of the graphene sheet,

the SWNT should be metallic with a nonzero density of states at the Fermi level [Figure 1.6(a)].

When the K point of the 2D Brillouin zone [Figure 1.6(b)] is located between two cutting lines,

the K point is always located in a position one-third of the distance between two adjacent lines

and thus a semiconducting nanotube with a finite energy gap appears. It is important to note that

the states near the Fermi energy in both metallic and semiconducting tubes result from states

near the K point, and hence their transport and other properties are related to the properties of the

states on the allowed lines.

(5,5) (5,0) (7,1)

K

KK

K

(5,5) (5,0) (7,1)

K

KK

K

(a) (b) (c)

Figure 1.6. Examples of the allowed 1D subbands for (a) a (5,5) armchair, (b) a (5,0) zigzag, and (c) a (7,1) chiral carbon nanotube. The hexagon defines the first Brillouin zone of graphene and the dots in the corners are the graphene K points.

The resulting 1D energy dispersion relations of a (n,m) nanotube are given by,

( )

( ) ( nqka

kakan

qE nn

,....,1 , :nanotubesarmchair for 2

cos42

coscos41

0

2/1020

0,

=π<<π−

⎥⎦

⎤⎢⎣

⎡⎟⎠⎞

⎜⎝⎛+⎟

⎠⎞

⎜⎝⎛

⎟⎠⎞

⎜⎝⎛ π

±γ±=

) (1.6)

-10-

( )

( )nqka

qkan

qE n

,....,1 ,33

:nanotubes zigzagfor

2cos4

23

coscos41

0

2/1

2000,

=⎟⎟⎠

⎞⎜⎜⎝

⎛ π<<

π−

⎥⎥⎦

⎢⎢⎣

⎡⎟⎠⎞

⎜⎝⎛ π

+⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎠⎞

⎜⎝⎛ π

±γ±= (1.7)

( )

( ). :nanotubes chiralfor 2

cos42

cos2

cos41

0

2/10200

0,

π<<π−

⎥⎦

⎤⎢⎣

⎡⎟⎠⎞

⎜⎝⎛+⎟

⎠⎞

⎜⎝⎛

⎟⎠⎞

⎜⎝⎛ −

π±γ±=

ka

kakan

mkan

qE mn (1.8)

We have constructed the band structures for metallic nanotubes (n,m) = (10,10) and for

semiconducting nanotubes (10,0) and (6,4) as shown in Figure 1.7. Note that only two of the 1D

subbands cross the Fermi energy in metallic nanotubes.

-3

-2

-1

0

1

2

3

E(k)

/γ0

k

(a)

0 000aπ− 03aπ− 03aπ−

-3

-2

-1

0

1

2

3

k

(b)

-3

-2

-1

0

1

2

3

k

(c)

Figure 1.7. One-dimensional energy dispersion relations for (a) armchair (10,10) tubes, (b) zigzag (10,0) tubes, and (c) chiral (6,4) tubes, computed using the zone-folded tight-binding dispersion relations described in the text.

The results for the 1D electronic density of states (DOS) show sharp peaks associated

with the van Hove singularities about each subband edge (Figure 1.8). At a band edge,

-11-

( ) 0' →∇ Ek ( is the gradient with respect to k) and singularities arise in the DOS at E’ (the

energy of the particular band maxima or minima). Resonances in Raman scattering experiments

have provided evidence for such sharp peaks in the DOS of nanotubes.

k∇

54 The electronic DOS

also shows that the metallic nanotubes have a small, but non-vanishing 1D density of states at the

Fermi level. In contrast, the DOS for the semiconducting nanotubes is zero throughout the

bandgap.

Den

sity

of S

tate

s

Energy/γ0

(8,8)

(9,9)

(10,10)

(11,11)

Den

sity

of S

tate

s

Energy/γ0

(8,8)

(9,9)

(10,10)

(11,11)

Figure 1.8. Electronic 1D density of states per unit cell for a series of metallic tubes, showing discrete peaks at the positions of the 1D band maxima or minima (Adapted from Dresselhaus55).

-12-

1.3. Motivations for this Work

Research and knowledge about carbon nanotubes have been developing at a very fast

pace. Although a number of basic features in the electron transport through nanotubes were

discovered, many challenges and questions remain before most of the proposed applications can

be realized. For example, how molecules interact with carbon nanotubes and affect their physical

properties is of fundamental interest. This knowledge may have important implications for their

production and growth as well as their applications. The nanometer-scale spaces inside and

among the SWNTs in a bundle should provide large gas-adsorption capacities,56 which are

especially exciting when we consider, for example, methane and hydrogen adsorption.

Adsorption and storage of hydrogen on nanotubes have been studied extensively due to the

potential application in the next generation of energy sources, e.g., fuel cells.26-31 However,

experimental reports of high storage capacities are so controversial that it is impossible to assess

the potential applications.13 Numerous claims of high hydrogen storage levels have been shown

to be incorrect; other reports of room temperature capacities above 6.5 wt% (a U.S Department

of Energy benchmark) await confirmation.57 Adsorption phenomena are also of interest from a

fundamental point of view because gases adsorbed on SWNTs provide an excellent model

system to study the effect on conduction electrons.

There is also the possibility for the development of high-sensitivity gas/chemical sensors

based on carbon nanotubes.18,19 In addition, the control of the electronic properties of nanotube

devices using vapor phase chemical doping was shown to be crucial to the design and tuning of

these devices.58 Indeed, the modulation doping of a semiconducting SWNT along its length can

lead to intramolecular wire electronic devices.59 Another motivation for doping experiments has

-13-

been the search for superconductivity in carbon nanotubes.

Production and growth of carbon nanotubes often take place in inert gas environments at

elevated temperatures.58 It is also expected that the collision dynamics between the gas and the

outside of the nanotube affect the growth. However, the detailed collision dynamics between the

gas molecules and the nanotube, the diffusion of the adsorbed atom along the nanotube, and its

incorporation in the nanotube are not well understood, nor are the effects of these impacts on the

nanotube conductance. The latter is studied in this Ph.D. thesis.

From a broader perspective, SWNTs provide a unique opportunity to study the

interaction of molecules with a conducting surface. This stems from the unique structure of the

nanotube. The electron and phonon states of this unique “all-surface” solid state system are, by

comparison to many other solids, relatively simple, thereby allowing fundamental calculations

addressing the experimental observations presented here to be carried out.

In this Ph.D. research, we have sought a greater understanding of gas-SWNT interactions.

A series of electrical transport measurements (thermoelectric power and electrical resistance)

will be discussed in the remainder of this thesis. Chapter 2 discusses the experimental methods

used to study transport in SWNTs. Chapters 3 and 4 review previous treatments for the

thermoelectric power of SWNTs, before presenting a new model of the thermoelectric power in

these materials. In Chapter 3, only the diffusion thermopower is considered, while Chapter 4 is

devoted to a formulation of the phonon drag effect problem in metallic SWNTs. Chapter 5 gives

some details on the effects of adsorption of small gas molecules on the thermopower and the

electrical resistance of SWNTs. Chapter 6 deals with the effects of gaseous chemicals adsorption

on the transport properties of SWNTs. Chapter 7 discusses the effects of inert gas collisions on

the thermoelectric power and the electrical resistance of SWNTs.

-14-

Chapter 2.

Experimental Techniques

2.1. Sample Preparation

The single-walled carbon nanotubes studied in our experiments were in the form of thin

films, thin pellets or mats of tangled ropes. In most cases, the SWNT material was obtained from

CarboLex, Inc. and produced by the arc-discharge method using a Ni-Y catalyst. The

approximate volumetric yield was estimated on the basis of Raman scattering to be ~ 50-70 vol%

carbon as SWNT. The SWNT material was removed from the growth chamber and handled in

ambient conditions.

Figure 2.1 shows a typical Raman spectrum (514.5 nm excitation) of arc-derived SWNT

bundles at room temperature. The SWNT material was always found to exhibit the characteristic

Raman spectrum published previously,54 including the radial breathing mode band at 186 cm-1

and the stronger tangential mode band at 1593 cm-1. The high-frequency bands can be

decomposed into two main peaks around 1593 and 1567 cm-1 with shoulders at 1550 and 1526

cm-1. These features have previously been assigned to a splitting of the E2g mode of graphite.60

The peaks in the frequency range 300-1200 cm-1 can be mostly identified as overtones and

combinations of lower-frequency modes. The low-frequency domain shows at least two

components at 141 and 186 cm-1. According to earlier calculations,54 these modes are expected to

be of A1g symmetry, and are identified with the radial breathing modes.

For an isolated nanotube of any chirality (n,m), the radial breathing mode has been 0RBMω

-15-

shown theoretically to exhibit a simple inverse diameter relationship, i.e., d2240RBM ≈ω for

in cm0RBMω -1 and d in nm, where the proportionality factor is somewhat sensitive to the details

of the calculation.61 This simple relationship for must be corrected for weak inter-tube

interactions within a bundle to obtain the measured mode frequency . Theoretical

calculations have predicted that these interactions are responsible for a 6-21 cm

0RBMω

RBMω

-1 frequency

upshift, depending on the details of the calculation.62-65 The expression linking the radial

breathing to the nanotube diameter d has been reported to be well approximated by the

expression

RBMω

( ) ( ) , 10,100

10,10RBM0RBMRBMRBM ddω+ω∆=ω+ω∆=ω (2.1)

where is the radial breathing mode frequency for an isolated SWNT, is a frequency

upshift which is a constant for nanotube diameters near to that of a (10,10) armchair nanotube

, and is the radial breathing mode frequency of an isolated (10,10) nanotube. The

calculated values of these parameters reported by various research groups vary from one another,

but some typical values are ( ,

0RBMω RBMω∆

)10,10(d 0)10,10(ω

RBMω∆ )10,10(0

)10,10( dω ) = (14 cm-1, 224 cm-1nm),62 (6.5 cm-1, 232 cm-

1nm),65 and (6 cm-1, 214 cm-1nm).64

Using Eq. (2.1), we found that the average diameter of the tubes from the arc-derived

material was therefore close to that of a (10,10) tube. The tube diameter distribution in this

material was mainly confined to the range 1.2 < d < 1.6 nm, based on the Raman spectra of the

radial breathing modes collected at six different excitation wavelengths.

Typical high-resolution transmission electron microscopy images (see Figure 2.2)

showed that the nanotubes were present in the form of bundles. The bundle diameter for arc-

derived material was in the range 10-15 nm, i.e., the bundles contained ~ 100-200 tubes.

-16-

Ram

an in

tens

ity (a

.u)

2000160012008004000Frequency (cm-1)

186

1567

1593

13471102

1550

1526

Figure 2.1. Room-temperature Raman spectrum for unpurified arc-derived SWNTs excited at 514.5 nm.

Some of our samples were prepared from “as-grown” SWNT material, i.e., without any

post-synthesis chemical or thermal treatment. Others were prepared from purified SWNT

material. Purification of our SWNT material was done first by a selective oxidation step at 425

ºC in dry air for ~ 20 min to remove amorphous carbon and weaken the carbon shell covering the

metal catalyst. This treatment was followed by an acid reflux for 24 h in 4.0 M HCl to remove

the metal residue. The material was then vacuum-annealed at ∼ 10-7 Torr and ~ 1000-1200 ºC for

24 h. The final metal content after this purification process, as determined by ash analysis

(combustion in dry air) in an IGA thermogravimetric analyzer (Hiden Analytical, Inc.), yielded a

value of 0.2 at% metal. Figure 2.3 shows TEM images of SWNT bundles before and after the

purification process.

-17-

Figure 2.2. Right: SEM image of a SWNT film showing entangled ropes synthesized by the arc-discharge method. Left: High-resolution TEM image of an end view of the SWNT bundles, showing the 2D hexagonal lattice arrangement of the tubes.

Figure 2.3. TEM images of SWNT bundles before (right) and after (left) purification. Dark spots are catalyst clusters.

The SWNT films were prepared by placing drops of an alcohol solution containing

SWNTs onto thin (0.25 mm), ground and polished clear quartz substrates (Chemglass Scientific,

Inc.) The alcohol solution was mildly sonicated (medium power) before the sample preparation

using a microtip horn connected to a Misonix Sonicator™ Ultrasonic Cell Disruptor/Processor

XL2020. The substrates were cleaned first in a boiling bath of isopropanol, followed by a

-18-

refluxing vapor of the same alcohol.

Nanotube paper or “buckypaper” was produced using the conventional method,66 i.e., by

filtering SWNTs dispersed in a liquid and peeling the resulting sheet from the filter after washing

and drying. Finally, the sheet was vacuum-annealed at ~ 1200 ºC for 12 h to remove volatile

impurities and repair tube wall damage incurred during the purification.

2.1.1. The Arc-Discharge Method

Carbon nanotubes can be synthesized through various process routes. The carbon arc-

discharge method, used initially for producing C60 fullerenes, is perhaps the most common and

easiest way to produce carbon nanotubes. The method became popular for the production of

carbon nanotubes after a group of researchers at the University of Montpellier in France

demonstrated that this technique can produce high yields of SWNTs.43

The arc-discharge method synthesizes nanotubes through the arc-vaporization of carbon

from the ends of two electrodes separated by approximately 1 mm. A direct current of 50 to 100

A driven by approximately 20 V creates a high temperature discharge between the two electrodes.

The discharge vaporizes one of the carbon electrodes and forms a small rod shaped deposit on

the other electrode. Both the anode and the cathode are made of graphite rods (purity ~ 99.99%),

and only the anode is loaded with 2-4 at% metal for synthesizing SWNTs. Production of

nanotubes takes place inside a stainless steel chamber filled with helium gas at low pressure (~

500 Torr). The electrodes may be positioned manually, or automatically, based on the measured

voltage between them. The electrodes and the chamber are cooled by a flow of low pressure

water. The gas pressure is controlled via a He flow system, assisted by a mechanical vacuum

-19-

pump. An electronic flow/pressure controller is used to regulate added gas.

Large-scale production of carbon nanotubes depends on many factors including the

uniformity and stability of the plasma arc, the stability of the temperature distribution, gas

pressure, etc. One interesting and useful characteristic of the growth of the carbon nanotubes by

the arc-discharge method is the tendency for large numbers of nanotubes to grow parallel to each

other, forming bundles or ropes of nanotubes, which consist of 10-100 tubes. This is thought to

be a curious multifilament outcome of vapor-liquid-solid (VLS) growth where a metal

nanoparticle is thought to act like a solvent for carbon, and the nanotube is viewed as growing

from the surface of a carbon saturated particle. The precipitation of carbon from the saturated

metal particle leads to the formation of tubular carbon solids in a sp2 structure. Tubule formation

is favored over other forms of carbons because a nanotube contains no dangling bonds and

therefore is in a low energy form. To maximize van der Waals contact and lower their free

energy, individual SWNTs align themselves with each other to form ropes growing over large

metal particles (> 10 nm diameter).

2.2. Thermoelectric Power Measurements

In essence, the experiment to measure the thermoelectric power consists of generating a

thermal gradient along a conductor and measuring the resultant open-circuit voltage. In this study,

the thermoelectric power was measured using a heat-pulse method developed by Eklund and co-

workers67,68 and which employs a simple analog subtraction circuit.

-20-

2.2.1. The Analog Subtraction Circuit

The analog subtraction circuit, shown in Figure 2.4, allows simultaneous measurements

of the temperature difference and the Seebeck voltage, using two thermocouples electrically

connected to the sample. Figure 2.4 also identifies the primary thermoelectric voltages ∆Vi (i =

1,2,3) used to determine the absolute thermoelectric power SU of the sample.

The absolute thermoelectric power S is conveniently defined as the potential difference

developed per unit temperature difference, i.e.,69

. dTdVS = (2.2)

Thus, given the Seebeck coefficient S(T) for a homogeneous material, the voltage difference

between two points where the temperatures are T1 and T2, is given as

. (2.3) 2

1

∫=∆T

T

SdTV

Figure 2.4. Block diagram of the analog subtraction circuit to measure the thermoelectric power.

The voltages in Figure 2.4 can therefore be written as

-21-

(2.4)

( ) ( ) ,

0

0

1

∫ ∫ ∫∆+

∆+

∆+

∆−=−=

++=∆

TT

T

AUAU

T

T

TT

T

T

TT

AUA

TSSdTSS

dTSdTSdTSV

(2.5)

( ) ( , 0

0

0

2

TVdTSS

dTSdTSV

BA

T

T

AB

T

T

T

T

BA

=−=

+=∆

∫ ∫

)

(2.6)

( ) ( ) ,

0

0

3

∫ ∫ ∫∆+

∆+

∆+

∆−=−=

++=∆

TT

T

BUBU

T

T

TT

T

T

TT

BUB

TSSdTSS

dTSdTSdTSV

where ∆T is the small temperature difference between the two thermocouple junctions attached

to the sample and T0 is the reference junction temperature (~ 300 K). The thermocouples in

Figure 2.4 are thermally but not electrically anchored to the temperature reservoir at T0.

In these experiments, the voltages ∆Vi are generated by applying a small heat pulse to one

of the two ends of the sample, which establishes a time-dependent temperature difference. The

amplifiers Ai (i =1,2,3,4) respond as indicated in Figure 2.4. For small temperature differences, a

straight line should be obtained when plotting the output of A1 (or A3) versus the output of A4.

The slope is related to the sample thermopower measured relative to the conducting leads.

Taking into account the actual gains (g, G) of the amplifiers, we find

( )( ) , slope AABU SSSgS +−= (2.7)

or

( )( ) . slope BABU SSSgS +−= (2.8)

-22-

The sample temperature T is determined from an independent measurement of VAB(T). For this,

the output of A2 is polled just before the heat pulse is started, and just after it is terminated to

determine the average temperature of the sample during the measurement. The temperature

dependence of the relative thermopower of the thermocouple pair VAB(T) is previously measured

by attaching the thermocouple pair to the surface of a silicon-diode thermometer, which

determines the temperature while measuring the output of A2.

2.2.2. The Thermopower Probe

A schematic of the thermopower probe is shown in Figure 2.5. The probe consists of a

header with a hermetic multipin connector for electrical input/output, a vacuum valve, a vent

valve and a sample stage. The sample holder is fastened, using Teflon® screws, to a stainless

steel stage at the end of a 0.635 cm outer diameter, thin-walled stainless steel tube attached to the

header (the supporting tube). The supporting tube is also a gas vent line with an opening at the

lowest part of the probe. The overall probe length is 160 cm, which allows its insertion into an

ordinary liquid-helium-storage container or a tube furnace. An O-ring seals the vacuum jacket to

the header.

-23-

Figure 2.5. Schematic diagram of the thermopower and resistance measurements probe suitable for the temperature range 4-500 K.

-24-

The sample holder, shown in Figure 2.6, consists of a rectangular piece made of Macor®,

which is a white machineable ceramic. This material can be used continuously up to 1000 ºC, is

vacuum compatible (no outgassing) and provides good electrical and thermal insulation. Eight

equally spaced screws are used to provide the electrical connections. Twisted copper leads

connect the sample heater to the multipin electrical connector on the header. Similarly, twisted

thermocouple leads (0.003” diameter, Omega Engineering, Inc.) carry the thermoelectric

response to the multipin connector and from there, via copper leads, to the analog subtraction

amplifiers. Two additional copper leads (0.003” diameter, Omega Engineering, Inc.) are used to

measure the voltage during four-probe resistance measurements, as explained later in this section.

A platinum resistor (type H2104, Omega Engineering, Inc.) is thermally clamped at one end of

the sample holder and serves as the heat source.

Figure 2.6. Schematic diagram of the sample holder for the thermoelectric power and four-probe resistance measurements.

Three types of differential thermocouples could be used in our experiments: chromel-

-25-

alumel (type K), copper-constantan (type T) and chromel-gold (7 at% Fe). The thermocouple

wires (0.003” diameter, Omega Engineering, Inc.) are bonded together using a spark-bonding

technique. This is done with a device consisting of two tweezers connected, through copper

wires, to opposite polarities of a power supply set to 120 volts. The thermocouple wires are

picked up together at their bare ends with one of the tweezers, and momentarily touched with the

other tweezers. If properly held, the wires spark-bond at the junction and the sections of the

wires being held by the tweezers burn off.

The sample is mounted onto the copper heater clamp (Figure 2.6) by cementing one of

the sample ends with silver paint. Thermocouples and voltage leads for four-probe resistance

measurements also make contact with the sample via silver paint, which provides reasonably low

contact resistances especially after thermal annealing at 100 ºC. We have tried to use silver-

loaded epoxy resin, which can withstand higher temperatures and exhibits better adhesion than

silver paint, but have found that it is susceptible to cracking upon cooling. Both silver paint and

silver-loaded epoxy exhibit excellent electrical and thermal conductivity as well as

environmental resistance.

2.2.3. Experimental Setup

A schematic of the experimental setup is shown in Figure 2.7, including the computer-

interfaced system. The output from the analog subtraction circuit is sent to independent pre-

amplifiers before being collected by the computer. An IEEE-488.2 interface card and an analog-

to-digital converter (A/D) card DAS8 (Keithley MetraByte) are used for the data acquisition with

LabVIEW (National Instruments Corp.) programs. The heat-pulse generator with variable pulse

-26-

width and height is built using a micro controller (PIC 16C56, Microchip Technology, Inc.),

which can be triggered by an external TTL signal. The pulse height and duration are adjustable

in the ranges of 0-10 V and 1-20 s, respectively.

Figure 2.7. Block diagram of the system for thermoelectric power and four-probe resistance measurements instrumentation.

When the sample is at the desired stable temperature, the computer sends a TTL pulse via

one of the digital output lines of the A/D card to trigger the pulse generator. As a result, a voltage

pulse with the appropriate width and height is applied to the heater, which causes a temperature

gradient to develop and relax with time along the sample. Depending on the thermal mass of the

sample and heater block, a temperature gradient of about 0.5 K is typically developed and

relaxed over an interval of 5-20 s. After additional amplification, the thermopower data are

collected via the A/D card as ∆T increases and relaxes.

201

5

10

15

Power

ON

OFF

PULSE GENERATOR

20 0

5

1015

Voltage AOutput

BOutput

COutput

ANALOG SUBTRACTION CIRCUIT

100 1 100 1 100 1 100 1 100 1 100 1100 1100 1

Power

DC Pre-amplifiers

Pulse Generator Analog SubtractionCircuit

SourceMeter

TTL triggering signal

DAS8 A/D Board

IEEE-488.2 InterfaceBoard

4-Wire Sense

Heater

1

2

3

4

201

5

10

15

Power

ON

OFF

PULSE GENERATOR

20 0

5

1015

Voltage AOutput

BOutput

COutput

ANALOG SUBTRACTION CIRCUIT

100 1 100 1 100 1 100 1 100 1 100 1100 1100 1

Power

DC Pre-amplifiers

Pulse Generator Analog SubtractionCircuit

SourceMeter

TTL triggering signal

DAS8 A/D Board

IEEE-488.2 InterfaceBoard

4-Wire Sense

Heater

1

2

3

4

-27-

2.2.4. The Thermopower Program

Figure 2.8. The program “Thermopower Auto.vi”, showing the time evolution of the thermoelectric power of a SWNT film during vacuum-degassing at 500 K.

To measure the thermoelectric power, we created two LabVIEW (National Instruments

Corporation) programs entitled “Thermopower Auto” and “Thermopower Manual”. The

principles of operation of both programs are the same except for the fact that the former allows

us to collect data continuously, at regular intervals of time without further intervention by the

operator.

Figure 2.8 shows a sample set of thermoelectric power data as it was being taken with

“Thermopower Auto”. The program collects thermopower data at every interval of time

specified by the parameter “Data Recording”. Before each data collection, the program may be

-28-

instructed on the gain of the pre-amplifiers (Output Gain), the type of thermocouple pair used

(Thermocouple Selection), and the file where data is going to be saved (Filename). The variable

“Front Panel Connections” specifies whether the output of amplifier A1 or A3 in Figure 2.4 is

used to measure the thermopower.

A selected number of the outputs of the amplifiers A4 and A1 (or A3) are continuously

collected at a sampling rate specified by the parameter “Scan time”. When these two voltages are

plotted against each other, a straight line should be generated, which retraces itself as the

temperature difference relaxes to zero (provided the thermocouples are in good thermal contact

with the sample and are properly heat stationed), as shown in the lower left chart in Figure 2.8.

At the end of the data collection, the same data are plotted in the lower right graph, together with

a linear-least-square fitting curve. As we have discussed [Eq. (2.7)], the slope of this line is used

to deduce the sample thermoelectric power.

The upper graph in Figure 2.8 shows the thermoelectric power of a SWNT film as a

function of time, as the sample was vacuum-degassed at 500 K. The same graph in Figure 2.9

shows the thermoelectric power as a function of the temperature of a small piece of constantan,

measured according to the aforementioned method using the program “Thermopower Manual”.

The data are in good agreement with the tabulated values.

-29-

Figure 2.9. The program “Thermopower Manual.vi”, showing the temperature dependence of the thermoelectric power of constantan.

2.2.5. Calibration of the Thermocouples

The sample temperature, as well as the quantities AB SS − , , and in Eqs. (2.7) and

(2.8), is known from calibration experiments which is checked frequently. The sample

temperature is known by simply measuring the temperature dependence of the thermocouple

voltage V

AS BS

AB(T). This is done by attaching the thermocouple pair to the surface of a silicon-diode

thermometer and measuring the output voltage of amplifier A2 as a function of the temperature

determined by the silicon-diode thermometer. These data are then fitted with a polynomial

function.

-30-

The reference junction temperature T0 is needed for the calculation of the sample

temperature T. Rather than using the cumbersome ice bath (T0 = 0 ºC), T0 is measured by

thermally anchoring a type-K thermocouple to two pins on the hermetic connector of the

thermopower probe. A schematic of the connections is shown in Figure 2.10.

Figure 2.10. Schematic diagram of the connections to A2 amplifier to measure the sample temperature (top) and the equivalent circuit (bottom).

The junctions J2 and J3 and the thermocouple (or thermistor) are all assumed to be at the

same temperature T0. We can easily show, using Eq. (2.2), that the output voltage ∆V2 is

proportional to the temperature difference (T – T0). Usage of an ice bath at the reference junction

allows one to determine the temperature directly from a ∆V2 versus T calibration curve. If T

needs to be known to higher accuracy, perhaps a secondary thermometer such as a silicon diode

-31-

should be used. Although slow temperature drifts in room temperature T0 could cause some error

in absolute temperature, they are too slow to affect measurements of the thermopower, because

each data point is collected during a short period of time (~ 20 s).

-1.25

-1.00

-0.75

-0.50

-0.25

0.00

0.25

g T∆V 2 (

V)

300250200150100500T (K)

30

25

20

15

10

5

0

(SK

P – S

Au:

Fe) (

µV/K

)300250200150100500

T (K)

Figure 2.11. The output voltage of amplifier A2 as a function of the sample temperature (left) and the temperature dependence of the relative thermoelectric power of a chromel-Au:Fe thermocouple pair (right).

According to Eq. (2.5), the temperature-dependent relative thermopower of the

thermocouple pair is given by the relation

. dT

dVSS BA

AB =− (2.9)

Therefore, by simply evaluating the derivative of the ∆V2 versus T calibration curve, we can

determine the temperature dependence of AB SS − . Figure 2.11 shows the temperature

dependence of the output voltage ∆V2 of the thermocouple amplifier A2 and the temperature

derivative of that calibration curve for a chromel-Au:Fe thermocouple pair ( ) . Fe:AuKP SS −

The absolute thermopower of an unknown sample is obtained from either Eq. (2.7) or

(2.8) by using calibration data for SA(T) or SB(T) with respect to copper. These quantities are

determined by simply using a piece of high-purity copper as the sample. Figure 2.12 shows the

-32-

relative thermopower data of chromel ( )CuKP SS − and Au:Fe ( )Fe:AuCu SS − with respect to

copper, measured according to the aforementioned method. Finally, tabulated values for the

thermopower of copper are used to obtain absolute thermoelectric power values.

20

15

10

5

0

S (µ

V/K

)

300250200150100500T (K)

SKP – SCu

SCu – SAu:Fe

Figure 2.12. Temperature dependence of the thermoelectric power of chromel and gold:iron alloy with respect to copper. The solid lines represent polynomial fits to the data.

2.3. Four-Probe Resistance Measurements

To measure the resistance of our samples the four-probe method was employed. This is a

very versatile means, used widely in physics, for the investigation of electrical phenomena.

Resistance measurements in the normal range (> 1 kΩ) are generally made using the two-probe

-33-

method. Here, a test current is forced through two test leads and the resistance being measured.

Then, the voltmeter measures the voltage across the resistance through the same set of test leads

and computes the resistance accordingly. The main problem with the two-probe method,

particularly for samples of low resistance, is that one inadvertently also measures the contact

resistance of the wires to the sample. Since the test current causes a small, but significant,

voltage drop across the contact resistances, the voltage measured by the meter will not be exactly

the same as the voltage directly across the test resistance and considerable error can result. For

example, one ohm of cable and contact resistance in a conventional two-probe circuit adds a

0.1% error to the 1-kΩ measurement. When one is measuring a very small resistance, especially

under variable temperature conditions, the contact resistance can dominate and completely

obscure changes in the resistance of the sample itself. This is the situation that exists for SWNT

networks.

The effect of cabling and contact resistances, as well as other series resistance errors, can

be eliminated with the use of a four-probe method, often called a Kelvin measurement. A

schematic of a four-probe connection, as well as the experimental arrangement, is shown in

Figure 2.7. In this method, four wires attached to the sample are used for resistance

measurements. A constant current is made to flow along the length of the sample through one set

of test leads (probes labeled 1 and 4 in the figure), while the voltage across the sample is

measured through a second set of leads called sense leads (probes labeled 2 and 3 in the figure).

In our case, one arm of each thermocouple is used as the current lead.

Although some small current may flow through the sense leads, it is usually negligible

(typically pA or less) and can generally be ignored for all practical purposes. Since the voltage

drop across the sense leads is negligible, the voltage measured by the meter is essentially the

-34-

same as the voltage across the sample. Consequently, the resistance value can be determined

much more accurately than with the two-probe method.

Figure 2.13. The program “DC 4-Probe Resistance.vi” showing the resistance as a function of temperature for a SWNT mat.

A SourceMeter 2400 (Keithley Instruments, Inc.) was used to provide a constant current

to the experiment and measure the voltage. This instrument combines a precise, low-noise,

highly stable DC power supply with a low-noise, highly repeatable, high-impedance multimeter.

To cancel thermal emf, the current reversal algorithm is employed. That is, two measurements

with currents of opposite polarity are programmed. Then, the two measurements are combined to

eliminate unwanted offsets.

Data acquisition is carried out remotely via an IEEE-488.2 connection with a computer as

shown in Figure 2.7. For this purpose, we created a LabVIEW program entitled “DC 4-Probe

-35-

Resistance”, shown in Figure 2.13. The program collects sample and thermocouple voltages

either continuously at a specified rate (Scan rate) or at the user command from the front panel.

After colleting a specified number of data points (No. Samples), the program plots the calculated

sample’s resistance versus temperature or time (Type of Measurement).

The variables under “Control Settings” specify the methods used to measure resistance

with the Keithley SourceMeter 2400. The program allows four-probe and two-probe resistance

measurements. The offset-compensation method cancels out unwanted offset in the voltage and

current readings by measuring resistance at the specific source level, and then subtracts a

resistance measurement made with the source set to zero. Temperature changes across

components within the instrument can cause the reference and zero values for the A/D converter

of the SourceMeter to drift due to thermoelectric effects. Auto zero acts to negate the effects of

drift in order to maintain measurement accuracy over time. Without auto zero enabled,

measurements can drift and become erroneous. Note that auto zero and offset-compensation

measurements are additional corrections to the one obtained by using the current reversal

algorithm, but the use of all these correction procedures simultaneously will decrease the

measurement speed. Control display command is used to enable and disable the front panel

display circuitry of the SourceMeter. When disabled, the instrument operates at a higher speed.

The variables under “Power Source Parameters” allow the configuration of the current

source. The operator can set the current level (I-Source), the limit voltage (V-compliance),

auto/manual range for voltage measurements (V-AutoRange), the range for manual current (I-

range) and voltage (V-range) measurements.

Before each data collection, the program may also be instructed on the type of

thermocouple pair used to measure the temperature (Thermocouple Selection) and the file where

-36-

data is going to be saved (Filename).

2.4. Gas/Chemical Adsorption Measurements

Figure 2.14. Schematic diagram of the gas handling system for gas/chemical adsorption experiments.

For gas adsorption measurements, a gas handling system is attached to the vent valve of

the thermopower probe. The basic components are shown in Figure 2.14. This system consists of

an oxygen/moisture trap (OT-4-SS, R&D Separations, Inc.), capable of reducing the oxygen and

moisture content of a gas stream to less than 20 ppb; a general purpose differential capacitance

manometer (Baratron 221, MKS Instruments, Inc.), a vacuum valve and a vent valve. Before

admitting the gas into the thermopower probe, the gas handling system is properly evacuated and

leak checked through the vacuum valve. An ultra-high purity grade gas cylinder is connected to

the vent valve of the gas handling system for gas adsorption experiments. If needed, a side arm

attached to a glass bulb containing a spectral grade liquid chemical (Sigma-Aldrich, Co.) is

-37-

connected to the vent valve of the gas handling system for chemical adsorption studies.

-38-

Chapter 3.

Thermoelectric Power of Single-Walled Carbon Nanotubes

3.1. Seebeck Effect: Theory

The Seebeck effect, depicted schematically in Figure 3.1, is the open-circuit (zero

current) voltage response to a temperature gradient in a material. This phenomenon was

discovered in 1821 by the German physicist Thomas Johann Seebeck who observed that a

voltage (electromotive force, emf) was developed in a loop containing two dissimilar metals,

labeled A and B in Figure 3.1(a), provided that the two junctions c and d were maintained at

different temperatures.69,70 The voltage across the loop was found to depend on the type of

metals used and the temperature of the junctions.

Physically, the phenomenon arises in a single material [Figure 3.1(b)] because the

electrons at the hot end of such a conductor can find states of lower energy at the cold end,

towards which they diffuse. This diffusion current is accompanied by the accumulation of extra

electrons at the cold end, setting up an electric field or a potential difference between the two

ends of the material. The electric field builds up until a state of dynamic equilibrium is

established between electrons rushing down the temperature gradient and those moving against

the gradient due to the electrostatic field. This thermoelectric effect is sometimes referred to as

the “diffusion thermoelectric power”.

-39-

Figure 3.1. (a) Basic thermoelectric open circuit that displays the Seebeck effect. (b) The Seebeck effect: A temperature gradient along a conductor gives rise to a potential difference.

The Seebeck coefficient, thermoelectric power or simply thermopower S is the ratio of

the open-circuit voltage developed ∆V to the temperature difference ∆T:

. lim0 T

VST ∆

∆=

→∆ (3.1)

Almost all the theoretical treatments of thermoelectric power in macroscopic systems are

based on the Boltzmann transport equation, from which the following expression for the

thermopower can be derived:71

[ ]

,)()()(

)()()()(1

)(

)(

∫∫

ε=ε

ε=ε

⎟⎠⎞

⎜⎝⎛

ε∂∂

τ

⎟⎠⎞

⎜⎝⎛

ε∂∂

−ετ

=

k

k

kvkvkk

kkvkvkk

fd

fEd

eTS

F

(3.2)

where τ is the electron relaxation time, v is the electron group velocity, e is the electronic charge,

-40-

f is the equilibrium distribution function of the electrons, ε is the energy of the electron relative

to the chemical potential EF, and the integral is taken over all momentum states k. Note that the

thermoelectric power is related to the energy carried by the electrons per unit charge, which is a

function of the relative contribution of the electron to the total conduction. The sign of the

thermopower is determined by whether the dominant conduction takes place in states above, or

below, the chemical potential EF.

For a degenerate metallic conductor, a calculation based on Eq. (3.2) (see Appendix A)

leads to the following expression for the diffusion thermopower:

( ) , ln3

F

22

ATdE

Ede

TkSEE

B ≡⎟⎠⎞

⎜⎝⎛ σπ

−==

(3.3)

where σ(E) is the conductivity that would be found in a metal for electrons of energy E, given by

the well known relation

( ) ( ) ( ) ( ). 22 EENEeE τυ=σ (3.4)

Here N(E) is the density of states, and υ(E) and τ(E) are, respectively, the free carrier velocity

and relaxation time at energy E. For an electron-like band, the thermopower is negative, while

for a hole-like band the thermopower is positive.

The energy dependence of the relaxation time is often written as

, (3.5) )( mEE β=τ

where β is a constant and m a number that depends on the type of scattering that is dominant.

Consequently, it is easily shown that in the free-electron approximation Eq. (3.3) reduces to

, 23

3 F

22

⎟⎠⎞

⎜⎝⎛ +

π−= m

EeTkS B (3.6)

which indicates that S is expected to vary linearly with temperature. Elementary calculations71

-41-

predict that m = 3/2 in the temperature range where the relaxation time is limited primarily by

electron-phonon scattering, and m = – 1/2 at very low temperatures where τ(E) is limited by

impurity scattering. Hence, a change in the slope dTdS as the temperature is reduced would be

expected in the free-electron approximation, i.e.

region) resistance (residual low very 22

TEe

TkSF

Bπ−= (3.7)

region) scattering(phonon high 3

22

TEeTkSF

Bπ−= (3.8)

The thermopower of semiconductors with relatively few conduction electrons (non-

degenerate semiconductors) shows temperature dependence different from that found in metals.

In the case of semiconductors, one has to replace the Fermi-Dirac statistics in Eq. (3.2) by the

Boltzmann statistics, in which case one obtains

, for Fcc

B

cB EEATkEE

ek

S >⎥⎦

⎤⎢⎣

⎡+

−= (3.9)

, for Fvv

B

vB EEATkEE

ek

S <⎥⎦

⎤⎢⎣

⎡+

−= (3.10)

where Ac and Av are temperature-independent constants. In general, the thermoelectric power for

a non-degenerate semiconductor is of the form

, ⎥⎦

⎤⎢⎣

⎡β+

λ=

TekS B (3.11)

where λ is the gap temperature measured from the midgap to the band edge and β is a constant. β

= 3 when both the density of states and the mobility increase linearly with E. β = 1 for constant

density of states and mobility.72 Using the energy dependence of the relaxation time in Eq. (3.5),

m−=β 25 .73

-42-

Thus, in contrast to the linear temperature dependence observed in metals, a

semiconductor should exhibit a thermoelectric power which is proportional to the reciprocal of

temperature. In addition, the thermoelectric power of a semiconductor is usually large at room

temperature (in the mV/K range). Metals, on the other hand, usually display a small

thermoelectric power (in the order of a few µV/K).69

In recent years, there has been increasing interest in the thermoelectric phenomena in

carbon nanotubes. Thermoelectricity has been shown to be a valuable and effective probe of

electronic structures, and a suitable tool for understanding the scattering dynamics of electrons

and phonons and the electron-phonon interactions in solids.

3.2. Thermoelectric Power of Carbon Nanotubes: Background

Figure 3.2 shows a typical temperature dependence of the thermoelectric power for an

“as-prepared” SWNT mat. The thermoelectric power data is in reasonably good agreement with

the earliest results,74-76 which yielded surprising results: at high temperatures the thermoelectric

power has a large positive value (i.e., S ~ 40-60 µV/K at 300 K, depending on the sample

history) that decreases monotonically with decreasing temperature, while at low temperatures the

thermoelectric power is quasi-linear in temperature and rapidly approaches zero as T → 0. There

is also a strong non-linearity in the range ~ 80-100 K, either in the form of a superimposed

“knee” (change of slope) or a more pronounced “bump” or peak.

The temperature dependence of the thermoelectric power of SWNT bundles is unusual; it

does not correspond to that of a simple metal or semiconductor. However, a metallic behavior

could be incorporated at low temperatures (below ~ 100 K), or by considering an extension of

-43-

the linear section at very low temperature (below 50 K) as shown, respectively, by the curves (a)

and (b) in Figure 3.2.

This behavior is in sharp contrast with the thermoelectric power observed for the basal

plane of graphite which exhibits a small value for the thermoelectric power (S = – 4 µV/K at 300

K) and nearly linear temperature dependence.77 Graphite has a pair of weakly overlapping

electron and hole π bands with near mirror symmetry about the Fermi level. Approximately

equal numbers of electrons and holes in these symmetric π bands are consistent with the small

(negative) linear thermoelectric power observed below room temperature.

40

30

20

10

0

S (µ

V/K

)

300250200150100500T (K)

(a)

(b)

Figure 3.2. Temperature dependence of the thermoelectric power for an air-saturated SWNT mat. The solid line is a guide to the eye. The dashed lines (a) and (b) represent the ways in which metallic behavior could be incorporated in the thermoelectric behavior of SWNTs.

-44-

Based on the fact that the sign of the thermoelectric power is always positive in the

measuring temperature range, it was implied that the majority charge carriers in the carbon

nanotubes should be p-type, and the hole carriers are dominant. The large magnitude of the

thermoelectric power is surprising because metallic tubes are predicted to have electron-hole

symmetry and hence, a thermoelectric power close to zero. The obtained temperature

dependence of the thermoelectric power is also unusual, and cannot be simply explained by a

single-band model for a metal [Eq. (3.3)] or a non-degenerate semiconductor [Eq. (3.11)], over

the studied temperature range. Numerous models, however, have been proposed to explain this

complicated behavior, including parallel metallic and semiconducting pathways,74,78 variable-

range hopping,78 electron-phonon enhancement,78 fluctuation-induced tunneling,79 and Kondo

effect.76 More complicated heterogeneous models have also been considered.80 We discuss next

the merits and difficulties of some of these approaches.

3.2.1. Parallel Heterogeneous Model of Metallic and Semiconducting Pathways

The observed temperature dependence of the thermoelectric power for SWNT mats

suggests a parallel heterogeneous model of metallic and semiconducting nanotubes within the

ropes.74,78 In principle, since a nanotube bundle consists of various nanotubes, the total

thermoelectric power should result from the combination of the contributions from all nanotubes

with different (n,m). For two types of parallel conductors (metallic and semiconducting), we can

write the total thermopower S as a weighted average

,ss

mm S

GGS

GGS += (3.12)

where Sm and Ss are the thermopower of the metallic and semiconducting tubes, respectively, Gm

-45-

and Gs are the corresponding electrical conductances, and G is the total conductance of the rope.

To derive Eq. (3.12) let us consider the situation shown in Figure 3.3. From Eq. (3.1), the

open-circuit potential difference produced across conductor a alone would be given by

, and similarly TSV aa ∆=∆ TSV bb ∆=∆ for conductor b acting alone. It then follows from

simple circuit theory that the resultant open-circuit voltage produced by the two conductors in

parallel is given by an expression similar to Eq. (3.12). This circuit theorem is sometimes known

as the “ladder” theorem.

Figure 3.3. Illustration of the combination of thermoelectric powers for conductors in parallel (also applicable to the two-band model).

The conductance from the metallic tubes is expected to be proportional to the reciprocal

of temperature, i.e., TGm 1∝ , at least at high temperatures. Assuming an activated form for the

semiconducting conductance, ( )TGs λ−∝ exp , and that the total conductance is dominated by

the metallic tubes, then from Eqs. (3.3) and (3.11), the total thermopower is74

,exp)( ⎟⎠⎞

⎜⎝⎛ λ−+λ+=

TCTBATS (3.13)

where A, B, and C are constants. The solid line in Figure 3.4 represents our fit of Eq. (3.13) to

our data set shown in Figure 3.2. Even though this model provides a reasonably good fit to the

-46-

thermoelectric power data, the fitting parameters are less than satisfactory in several respects.74

The fitted thermoelectric power for a mat of SWNTs is seen to be dominated by the first

term in Eq. (3.13), where the positive value of the temperature coefficient A implies a

thermoelectric power contributed by the holes of metallic nanotubes, with a room temperature

magnitude of S ~ 80 µV/K. In general, the model predicts an unphysically large magnitude of

the diffusion thermoelectric power for the metallic tubes (Sm ~ 80-200 µV/K at 300 K).78

40

30

20

10

0

S (µ

V/K

)

300250200150100500T (K)

A = 0.249 µV/K2

B = -0.668 µV/K2

C = 0.486 µV/K2

λ = 521 K

Figure 3.4. Fits to measured thermoelectric power of a SWNT mat using a parallel heterogeneous model of semiconducting and metallic tubes. The solid line represents a fit to the data using Eq.(3.13). Fitting parameters extracted from our fit are also shown in the figure.

The negative term would presumably be contributed by the electrons of the conduction

band for the semiconducting nanotubes. Therefore, within this model, the measured magnitude

-47-

of the temperature-dependent thermoelectric power is due to a near cancellation of very large

metallic and semiconducting thermopowers of opposite sign; a situation regarded as unlikely.

The value of semiconducting tube energy gap, obtained from the parameter λ, is ~ 0.09

eV. Previous studies have found values of 10-20 meV, which are significantly smaller than

values expected for the energy gap. In general, the energy gap of a semiconducting nanotube is

expected to vary from 0.5 eV to a few eV in the diameter range 1 nm < d < 2 nm, depending

upon the geometry.51,81 The mean tube diameter measured for the sample whose data appear in

Figure 3.4 is ~ 1.4 nm. This model consequently predicts very small semiconducting gaps and

conductances that are considerably larger than typical observations.82,83 Moreover, the model

with metallic and semiconducting conduction in parallel cannot by itself account for the

resistivity behavior, because the semiconductor contribution to conductivity is frozen out at

lower temperatures where it is observed experimentally.

3.2.2. Variable-Range Hopping

An alternative parallel tube model was proposed where metallic conduction is in parallel

with disordered semiconductor conduction via 3D variable-range hopping.78 For conduction by

variable-range hopping, the expected thermopower (which is positive for hole conduction) is

given by84

( ) ( ) , ln2

F

21

0

2

EE

B

dEENdTT

ek

S=

⎥⎦⎤

⎢⎣⎡= (3.14)

where N(E) is the density of states at the Fermi level and T0 is the parameter appearing in the

Mott variable-range hopping law of conduction:

-48-

, exp4

1

00

⎥⎥⎦

⎢⎢⎣

⎡⎟⎠⎞

⎜⎝⎛−=

TT

GG (3.15)

G0 is usually considered to be weakly dependent on temperature.

Using Eq. (3.12) one now obtains

. exp4

1

21 0

⎥⎥⎦

⎢⎢⎣

⎡⎟⎠⎞

⎜⎝⎛−+=

TT

BTATS (3.16)

40

30

20

10

0

S (µ

V/K

)

300250200150100500T (K)

A = -0.423 µV/K2

B = 31.54 µV/K3/2

T0 = 636 K

Figure 3.5. Fits to measured thermoelectric power of a SWNT mat using a parallel heterogeneous model of disordered semiconducting and metallic tubes. The solid line represents a fit to the data using Eq. (3.16). Fitting parameters extracted from the fit are also shown in the figure.

As shown in Figure 3.5, this expression does not provide a reasonable fit to the observed

thermoelectric power (Figure 3.2), as the predicted change in slope is smoother than that

-49-

predicted. The model also gives an even larger metallic contribution when fitted to the data. On

the other hand, the simple combination of metallic and hopping conduction regions suggested by

the resistivity behavior of SWNTs fails to provide reasonable fitting to both the resistivity and

the thermopower using the same set of parameters; mainly T0.

3.2.3. Electron-Phonon Enhancement

The quasi-linear temperature dependence of the thermoelectric power in a metal can be

understood by the electron-phonon enhancement effect.78,85 The enhancement of the linear

temperature dependence of metallic thermoelectric power can be expressed as

[ ] ,)(1 ATTS Sλ+= (3.17)

where λS(T) denotes the temperature dependent enhancement from the coupling of the electrons

and phonons and A is the coefficient of diffusive thermoelectric power.

The fits to the thermoelectric power data using Eq. (3.17) give the reduction of slope

above ~ 100 K. However, the values of the effective enhancement λS(0) are as large as those

expected for good superconductors.78 Since nanotubes have not been observed to superconduct at

temperatures as low as 1.5 K, a very large value for λ seems inappropriate.

3.2.4. Fluctuation-Assisted Tunneling

A SWNT network can be assumed to contain many metallic regions separated by small

insulating barriers (for example, due to poor connectivity between individual nanotubes). The

contact resistance of these barriers can dominate the overall resistance of the system. Such

-50-

systems can be described phenomenologically by Sheng’s theory of fluctuation-induced

tunneling.86 According to this theory, the electrical conductivity is given by the relation

. exp0

10 ⎟⎟

⎞⎜⎜⎝

⎛+

−σ=σTT

T (3.18)

The thermoelectric power can also be calculated from this theory.79 The contribution to

the thermoelectric power would be weighted by the fraction of the temperature difference

appearing across the barriers compared to that appearing across the metallic portions.

This model can describe the thermoelectric power data of SWNTs reasonably well and a

knee, such as that occurring in the thermopower data at ~ 100 K, is predicted by this theory.

However, additional contributions to the thermoelectric power must be considered in order to

achieve the positive curvature seen in the data for T ≥ 100 K.

3.2.5. Kondo Effect

An anomalously large peak superimposed on the metallic thermopower has been

observed in the range 80-100 K in some transition-metal doped SWNT mats, and has been

tentatively attributed to a Kondo anomaly associated with the residual (magnetic) catalyst

residing as particles on the bundles or rope surface, or trapped as atoms or small clusters within

the bundles.76 In other materials, the interaction between the magnetic moments of the impurity

atoms and the spin of the conduction electrons has been shown to lead to a new spin-dependent

scattering mechanism and a narrow hybridization peak or “Kondo resonance” in the electronic

density of states positioned near the Fermi level.87

The Kondo mechanism was also suggested because the particular catalyst caused a

-51-

concomitant change in the strength of the upturn in the electrical resistivity with decreasing

temperature below ~ 100 K. Further support for this argument is that the chemical treatment of

the SWNT samples with iodine, significantly reduces the Kondo contribution.76 The drawback in

the Kondo proposition to explain the thermoelectric power of SWNTs is that the thermoelectric

power peak should occur at BK kETT ∆≈= , where ∆E is the width of the Kondo resonance

near the Fermi level. One might expect a stronger variation in TK with various magnetic

impurities (e.g., Fe, Ni, Co), unless the effect is associated with Y, which is a common element

in all of the catalysts considered in that study.

3.2.6. Thermoelectric Power of Oxidized SWNT Networks

Recent experimental studies88,89 have reported that the measured SWNT electronic

properties (including the thermoelectric power and the electrical resistance) are extremely

sensitive to the presence of molecular oxygen in the nanotube. Specifically, it has been found

that, at room temperature, the resistance changes by 10-15% when SWNT mats are cycled

between vacuum and air exposure. In addition, after exposure for ~ 2-3 h under ambient

conditions to room air, SWNTs were found to acquire thermopower values unusually large

compared to those of ordinary metals and graphite. Most strikingly, it was reported that some

small-gap semiconducting nanotubes exhibit metallic behavior when they are exposed to

oxygen.90

Our in situ measurements while heating these O2-doped samples in a ~ 10-7 Torr vacuum

at ~ 500 K (i.e., to remove the adsorbed O2) showed that the large positive thermopower

identified with O2 doping first decreases with time and then changes sign, with the fully degassed

-52-

sample finally exhibiting a large negative thermopower after 10-15 h. These results were

interpreted as due to the formation of a charge transfer complex . For the

semiconducting tubes, the position of the Fermi level must shift toward the valence band. This

will be discussed in detail in the following section.

δ−δ+ − 2OC p

-40

-30

-20

-10

0

10

20

30

40

S (µ

V/K

)

300250200150100500T (K)

Air-saturated

Degassed @ 500 Kin vacuum

Figure 3.6. The temperature dependence of the thermoelectric power for an “as-prepared” SWNT mat in its air-saturated and degassed states. The solid lines are guides to the eye.

Figure 3.6 shows our S(T) data for the same SWNT sample for two cases: air-saturated

and degassed states. For the air-saturated sample, the thermoelectric power is always positive

and approaches smoothly S = 0 as T → 0, from a large value S ~ 40 µV/K at 300 K. For the

sample degassed at 500 K for 24 h in a ∼ 10-7 Torr vacuum, the thermoelectric power is negative

over the entire temperature range 4 K < T < 300 K, with a room temperature magnitude

-53-

comparable to that measured under air-saturated condition. As can be seen, the functional forms

of S(T) for air-saturated and degassed samples appear to be almost “mirror images” of each other,

reflected about the horizontal temperature axis. The results in Figure 3.6 demonstrate that the

previously published large positive thermoelectric power data should not be considered as an

intrinsic SWNT behavior, but rather the result of various degrees of oxygen doping.

The interactions of O2 with carbon nanotubes have been investigated in several recent

theoretical studies.91-103 According to the results, there are three different pathways for oxygen

adsorption on SWNTs:

(a) Molecular oxygen (in its spin-triplet ground state) physisorbs on the outer surface of a perfect

tube wall with relatively small binding energies up to 0.25 eV, accompanied by a small charge

transfer of ≤ 0.1 electron from the carbon nanotube to the O2 molecule. There may be no barrier

for O2 physisorption onto perfect carbon nanotubes.101 On the basis of the calculated density of

states, it was also suggested that the adsorbed O2 molecules can dope the semiconducting

nanotubes with hole carriers and that conducting states are present near the band gap.92

(b) Molecular oxygen (in its spin-singlet excited state) exothermically chemisorbs to the tube

wall through an intermediate cycloaddition step to give two spatially well-separated epoxide

groups or two single C−O bonds, with significant charge transfer from the nanotubes to the O2

molecules.94,97,99,103 The adsorption is high (~ 0.34 eV) and the activation barrier can be as low as

~ 0.6 eV, which is accessible at room temperature.97 O2 molecules can be excited into a more

reactive spin-singlet state by UV-light,100 sunlight, photosynthesizers (e.g., fullerene

impurities),97 or by topological defects.101 In addition, adsorption at defect and impurity sites

may help overcome the kinetic activation barrier.100,101

(c) Molecular oxygen diffuses to the edge of open tubes where it dissociatively chemisorbs with

-54-

an adsorption energy of up to 8.4 eV to form strong C−O bonds, even without any activation

barrier.93

3.3. Thermoelectric Power of SWNT Films

Here we present the results of a systematic study on a purified thin film of tangled SWNT

ropes, which show that the thermoelectric power is determined by coordinated effects in both the

semiconducting and the metallic tubes. However, the thermoelectric power will be shown to be

dominated by the metallic tubes in the ropes. The sign and magnitude of the thermoelectric

power will be shown to be determined by the relative concentrations of (acceptor state; δ =

fractional charge) and an unidentified donor state in the semiconducting tubes, possibly due to

wall defects. In fact, we show that a fully compensated sample can exhibit a thermopower S ~ 0

over a wide temperature range (4 < T < 500 K). In the compensated case, we will argue that the

Fermi level for a rope containing semiconducting and metallic tubes is very near the intrinsic

position. Furthermore, in this case, the mirror symmetry of the metallic band structure makes

offsetting contributions from electrons and holes. In contrast to this view, Avouris and co-

workers

δ−2O

104-106 have demonstrated a similar behavior using SWNT field effect transistors (CNT-

FET), i.e., the p-type electronic character can be turned into a fully n-type one by simply

annealing in vacuum. Based on the observation that oxygen treatment has no effect on the

threshold voltage for turning on a CNT-FET, they argued that barriers at the metal-

semiconducting contact control the carrier injection even though bulk doping may take place.

According to Avouris and co-workers,104-106 the dependence of these barriers on oxygen

determines the electrical character of the CNT-FET. It may be that the nanotubes in the CNT-

-55-

FET studied were defect-free and contained only weakly interacting (physisorbed) oxygen. In

contrast, our samples contained sufficient defect density that could bind strongly with O2

involving charge transfer.

Despite the numerous theoretical and experimental investigations, the microscopic

mechanism responsible for the observed changes in the electronic-transport properties of SWNTs

is still discussed controversially. The sensitivity of carbon nanotubes to O2 exposure is an

important issue, as it raises questions about the stability of devices made of carbon nanotubes

upon air exposure. On the other hand, the observed effect of oxygen exposure on the properties

of carbon nanotubes raises the possibility that unintentional oxygen contamination during

preparation of nanotubes samples might have led to incorrect analysis of the experimental data.90

3.3.1. Role of Contact Barriers on the Transport Properties of SWNTs

It has been argued79,80 that the measured thermoelectric power of a SWNT network may

be affected by rope-rope contacts and other barriers (e.g., defects, tube-tube contacts, etc.) in the

SWNT mat or film and their random orientations relative to the thermal gradient. This is a

reasonable concern, particularly if we note that a SWNT network consists of ropes, themselves

containing metallic and semiconducting tubes that may be loosely touching each other through

semiconducting tubes and/or amorphous carbon on the rope surface, not eliminated by the

purification process (Figure 3.7).

-56-

Figure 3.7. Sketch of crystalline SWNT ropes, where fibrillar carbon nanotubes are separated by disordered regions (Adapted from Kaiser et al.107)

In this picture, the thermoelectric power could be the result of a pathway of

semiconducting tubes broken by series-connected inter-tube barriers, where the thermopower

due to the insulating barriers could be described by either an activated hopping-like conduction

model80 or a fluctuation-induced tunneling model.79 We investigated the possible influence of

rope-rope contacts on the four-probe resistance [Figure 3.8(a)] and the thermoelectric power

[Figure 3.8(b)] on film or mat samples by observing the change in these quantities under the

action of uniaxial stress.

The experimental geometry is shown schematically in the inset to Figure 3.8(b). The

measurements were made in air at T ~ 300 K as a function of the loading force F applied normal

to the substrate supporting the SWNT mat. Thermocouples [TC(1), TC(2)] and voltage leads

[V(1), V(2)] made contact with the SWNT mat via silver epoxy outside the region of applied

stress. The pressure was calculated directly form the cross-sectional surface area of the insulating

rod and the weights placed on top of it.

-57-

1.00

0.95

0.90

0.85

0.80

R/R 0

1.81.51.20.90.60.30.0Pressure (MPa)

(a)70

60

50

40

30

S (µ

V/K

)

1.81.51.20.90.60.30.0Pressure (MPa)

F

V(1) V(2)

TC(1) TC(2)

F

V(1) V(2)

TC(1) TC(2)

(b)

Figure 3.8. Uniaxial pressure dependence of (a) the normalized room temperature resistance R/R0 and (b) the thermopower S for two different “as-prepared” SWNT mats. The inset shows the experimental geometry where the applied force F is perpendicular to the sample.108

As can be seen in Figure 3.8, the applied force (stress) impacts the four-probe resistance

but not the thermopower. It should be noted that the data are for two O2-doped samples under

ambient conditions. If the contact regions among the ropes dominated the thermopower, when

these regions become better heat conductors under the applied stress, the thermopower might be

expected to decrease. However, as shown in Figure 3.8(b), the pressure has little effect on the

thermopower. The insensitivity of S to the improved rope-rope contact resistance, observed via

the decreasing R in Figure 3.8(a), is taken as direct evidence that the contact barrier between

ropes is not significantly involved in the thermoelectric power of the SWNT sample. This is

consistent with the measurement of the thermoelectric power under “open-circuit” (zero current)

conditions.

3.3.2. Effect of Oxygen Doping on the Thermoelectric Power of SWNTs

Figure 3.9 shows the time evolution of the thermoelectric power at T = 500 K for a

-58-

typical purified SWNT thin film sample under vacuum. The SWNT sample was previously

exposed to air under ambient conditions for several days, then mounted in the measurement

apparatus, evacuated to ~ 10-7 Torr and heated from 300 to 500 K. At point A, the sample is still

nearly air-saturated which was accomplished under ambient conditions.

30

20

10

0

-10

S (µ

V/K

)

121086420Time (h)

O2 desorption

O2 adsorption

T = 500 K

(A)

(B)

(D)

(C)

Figure 3.9. Thermopower response to vacuum and O2 (1 atm) at T = 500 K. (A → C): Vacuum-degassing of a sample initially O2-doped under ambient conditions for several days. (C → D): Exposure of the degassed sample to 1 atm of O2 established at C.108

As the sample was degassed at T = 500 K in a vacuum of ∼ 10-7 Torr, the thermopower

was observed to decrease slowly from an initial value S = 8 µV/K, change sign at B (fully

compensated state), and then gradually approach a constant value of S0 = – 10 µV/K near C

(fully degassed state). The observed thermoelectric behavior at T = 500 K in vacuum is in

agreement with previously reported results on similar samples.88,89 However, very recently,

-59-

Goldoni et al.109 have used high resolution core-level photoemission spectroscopy to study the

interaction between O2 and SWNTs at low temperature (150 K). A strong interaction with O2

was found for samples contaminated with traces of Na (mainly chemical residues of the

purification, dispersion, and filtration processes) due to charge transfer from the tube to the

Na−O complex, whereas weak interaction with O2 was observed when dosing the Na-free sample.

Thus, Goldoni et al.109 suggested that O2 molecules have no effect on the transport properties of

SWNTs if impurities (i.e., catalyst particles, contaminants and defects coming from the chemical

treatments) are carefully removed from the nanotube samples. Note that, in our purification

procedure, we do not use surfactants or NaOH, which might leave residual Na in the SWNTs.

Besides, as mentioned in Ref. 109, the high-T annealing at ultrahigh-vacuum completely

removes any Na contamination and strongly reduces the number of defects introduced by the

purification treatments, restoring the nanotube structure and the bundle network. Our samples

were annealed at ~ 1200 ºC in a ∼ 10-7 Torr vacuum for 24 h. We also note that the experiments

in Ref. 109 have been carried out by exposing nanotubes to O2 at 150 K. At this low temperature,

we expect O2 to interact weakly with SWNTs through a physisorption process only. Due to the

negligible charge transfer between physisorbed oxygen and SWNTs, such O2 species are not

expected to facilitate the doping responsible for the observed change in the transport properties

of SWNTs.110 As suggested by Ulbricht et al.,110 based on thermal desorption experiments and

molecular mechanics calculations, it seems likely that the observed effect of O2 on the transport

properties of SWNTs is predominantly due to charge transfer by minority oxygen species,

weakly bound either at defect sites on the SWNT bundles or at tube-metal contacts in electronic

devices.104-106

Exposure of the fully degassed film to 1 atm overpressure of pure oxygen at T = 500 K

-60-

irreversibly changes the thermopower to large, positive value S = 25 µV/K, as indicated by the

point labeled D (high-T O2-doped state) in Figure 3.9. This indicates that O2 exposure at T = 500

K results in a more strongly bound oxygen acceptor, possibly a C–O bond at a wall defect. When

vacuum was applied at D, we were unable to change the thermoelectric power.

Note that the sample was fully degassed for ~ 8 h. Such a long equilibration time taken to

attain the negative value of the thermoelectric power representative of the “degassed state” of

purified SWNT films exposed to ambient air suggests that some of the O2 must reside in the

interstitial channels and/or within the central pore of the opened SWNTs. Fujiwara et al.111 have

used adsorption isotherms and X-ray diffraction at 77 K to investigate the gas adsorption

properties of bundled carbon nanotubes and have concluded that O2 molecules are adsorbed

preferentially inside the bundles, and then mostly in the interstitial channels. Single-file diffusion

would be necessary to empty the interstitial channels.

Figure 3.10 displays the temperature dependence of the thermoelectric power for the

same sample at the points A to D indicated in Figure 3.9. The series of curves S(T), also labeled

A, B, and C in Figure 3.10, are observed after successively longer periods of vacuum-degassing

that removes successively larger amounts of O2 from the ropes. We note that it is possible to tune

reliably to any intermediate metallic thermopower between the air-saturated state and the fully

degassed state, including an almost zero thermopower state (curve B).

For example, the thermopower for an initially air-saturated sample (under ambient

conditions) and the same sample O2-doped by exposure at T = 500 K to 1 atm O2, are both

positive and almost linear over the entire temperature range. A positive “knee” is observed

around ~ 100 K, and changes sign tracking the sign of the linear background on which it is

superimposed.

-61-

25

20

15

10

5

0

-5

-10

-15

S (µ

V/K

)

5004003002001000T (K)

(D) High-T O2-doped(irreversible)

(A) Ambient O2-doped(reversible)

(B) Compensated

(C) Degassed @500 Kin vacuum

Figure 3.10. Temperature dependence of the thermopower S for a SWNT thin film after successively longer periods of O2 degassing at T = 500 K in vacuum. The labels A, B, and C refer to a vacuum-degassing interval indicated in Figure 3.9. Curve D is for the same sample exposed to 1 atm O2 at T = 500 K for about 4 h after being fully degassed to point C.108

The thermoelectric power of the degassed sample, on the other hand, is negative over the

entire temperature range and also shows a linear metallic variation with temperature, with a

superimposed negative “hump” around ~ 80 K. A low-T hump is often identified with phonon

drag,70 which enhances the diffusion thermopower. Our assignment of phonon drag is consistent

with the sign change of the hump. Phonon drag thermoelectric power in SWNTs will be

discussed in the following chapter.

-62-

3.3.3. Compensating Doping and Defect Chemistry

We assume that each rope consists of a mixture of metallic and semiconducting tubes, in

the approximate ratio 1:2. There are probably defect states in the tubes, among which some are

donors and some are acceptors. Prior calculations have shown that defects in metallic nanotubes

introduce resonances in the density of states at the Fermi energy.112,113 They are discussed in a

later section.

Donors in a semiconducting tube introduce an additional electron into the system, while

acceptors contribute an additional hole. In both cases, we show that the additional electron or

hole is transferred to the metallic tubes and controls the chemical potential of the rope.108 We

assign the acceptor states to chemisorbed oxygen. Calculations by Jhi el al.92 predicted a charge

transfer of about 0.1 electrons to the O2 molecules in contact with the semiconducting tube wall.

The origin of the donor state is less clear. It may be associated with wall defects.

Consider the case of doped semiconducting carbon nanotubes with donor states, in close

contact with the metallic nanotubes in a rope. The total negative charges (electrons) must equal

the total positive charges (holes and ionized donors). Hence, the following charge neutrality

condition governs the position of the Fermi level in a rope:

(3.19) ,222 shDmese nNnn +=+ +

where nme is the electron density of the metallic tubes, nse and nsh are respectively the electron

and hole density from the semiconducting tubes, and is the density of ionized donors. +DN

Next, we use the definition of nse for a non-degenerate semiconductor,

( )( )

( ) , 2

2

F

F

20

TkEE

TkEkEse

Bg

B

en

edkn

−−

−−

= ∫ (3.20)

-63-

where Eg is the semiconducting gap, E(k) is the tight-binding energy dispersion, and n0 is the

effective density of states in the conduction band of the semiconducting tube. Similarly, the hole

density can be calculated to give,

( ) . 20

F TkEEsh

Bgenn +−= (3.21)

The concentration of ionized donors is given to a good approximation by

( ) , 21 F TkEE

DD BDe

NN −+

+= (3.22)

where ND is the density of donors, ED is the donor binding energy with respect to the conduction

band minimum, and the Fermi-Dirac distribution function is modified by the presence of the

factor 2 before the exponential term in the denominator because of the spin degeneracy of the

donor states, which can be occupied by either a spin-up or a spin-down electron.

We can rewrite the above equation to obtain

,2 γ+

γ=+

xNN DD (3.23)

where ( ) TEEgex BF k2−−= and ( ) TkEE BDge −−=γ 2 .

Next we consider the metallic carbon nanotube. We assume that it is charge neutral if the

chemical potential is at the energy zero where the bands cross. If this point is defined as EF = 0,

then the excess charge density on the metallic tube is

, lnF ⎟⎠⎞

⎜⎝⎛

α==

xTgkEgn mBmme (3.24)

where gm is the total density of states for the four metallic bands and TkE Bge 2−=α .

Using Eqs. (3.20)-(3.24), we can finally rewrite Eq. (3.19) in the form

,2

ln02

γ+γ

−⎟⎠⎞

⎜⎝⎛

α+

α−=

xrxs

xx (3.25)

-64-

with 02nTgks mB= and 0nNr D= .

At r = 0 (i.e., no donors), the solution is x = α. In this case, the chemical potential is at

midgap, and there are equal numbers of electrons and holes in the semiconductor. There is no

charge transfer to the metallic tubes.

In the limit of low doping, r is small. Assuming that α << γ, the last two terms in Eq.

(3.25) are the largest and must cancel to give srex α= . In this case, there is complete charge

transfer. All the donors ionize and transfer their electrons to the metallic tube. The chemical

potential rises in the metallic tube, but by a small amount. In equilibrium, the chemical potential

must be the same for all tubes. If the donor density becomes large, then the chemical potential

approaches the donor density in the semiconductor and the charge transfer decreases for

additional donors. A similar analysis applies for acceptor states. Because of the charge transfer,

there is negligible electrical conductivity in the semiconducting nanotubes.

3.3.4. Model Calculations of the Thermoelectric Power of SWNTs

Since S ~ T, the thermoelectric power of the SWNT film (Figure 3.10) is consistent with

a diffusion thermopower dominated by metallic tubes in a rope. The metallic character of the

thermopower can be understood from the following argument.

The thermopower for a SWNT rope can be written as the sum of the conductance-

weighted contributions from all nanotubes in a rope because they are connected in parallel (c.f. p.

45),

,1

1∑

=

=N

jjjSG

GS (3.26)

-65-

where

(3.27) ,1

∑=

=N

jjGG

and the index j runs over all N tubes in a rope, Gj and Sj are the conductance and the

thermopower, respectively, of the jth tube and G is the total conductance of the entire SWNT

rope.

If the semiconducting nanotubes are not degenerately doped, then Gj(metal) >>

Gj(semiconductor) and we find that )metal(~ jSS and 3(metal)~ NGG j , where L

indicates the average number of metallic nanotubes in a rope.

If some semiconducting nanotubes were degenerately doped, they would mimic, to some

extent, the temperature dependence of the conductivity and the thermopower of the metallic

nanotubes. Specifically, the thermopower for the degenerately doped semiconducting nanotubes

would exhibit a F~ ETS behavior. However, EF is small and hence, the thermopower for these

tubes would be high. The relative contribution of the degenerately doped semiconducting

nanotubes would be controlled by their conductance, and we anticipate that this conductance is

significantly lower than those for intrinsic metallic nanotubes. There is no clear evidence that

some of the nanotubes in our samples are degenerately doped semiconducting nanotubes.

However, if they exist they may enhance a smaller metallic thermopower.

We note that the thermopower has been shown (Figure 3.8) to depend only weakly on the

rope-rope contacts, but the film resistance is affected. The value of G one might compute from

Eq. (3.27) does not represent the mat/film conductance, i.e., the rope-rope contact resistance is in

series and must be added into the calculation.

We next show that the magnitude of the experimental thermoelectric power cannot be

-66-

explained by a simple two-band model for a metallic tube with electron-hole symmetry, except

for the fully compensated sample. To do this, we take the simplifying assumption that our

samples are mainly composed of metallic (10,10) tubes, whose band structure near EF is

characterized by two pairs of 1D tight-binding bands crossing at zero energy (cf. Eq. (1.6), n = q

= 10)

,

2cos21

, 2

cos21

00

00

⎥⎦

⎤⎢⎣

⎡⎟⎠⎞

⎜⎝⎛−γ−=

⎥⎦

⎤⎢⎣

⎡⎟⎠⎞

⎜⎝⎛−γ=

kaE

kaE

h

e

(3.28)

where e, h refer to the electron and the hole bands.

In the frequently used two-band model for the thermoelectric power, one obtains69

, hh

ee ssS

σσ

+σσ

= (3.29)

where σi and si are, respectively, the conductivity and the thermopower of the i = e, h (electron,

hole) band, and is the total conductivity. Now, using Eqs. (3.3)-(3.5), we obtain he σ+σ=σ

( )

( )( )( )

( ) ( )

( ) ( ) ( ) ( ).

114114

11414

11

, 3

02

002

0

10

202

0

00

0

22

mm

mm

B

m

K

KKe

TkS

ε−ε+−+ε+ε−−

ε±ε−+ε−

ε±ε

±=

π−=

±

−+

mm

m (3.30)

Here, 00 γ=ε E is the reduced energy.

In Figure 3.11, we plot the result for S at T = 300 K for m = 3/2 as a function of EF. We

allowed the Fermi energy to move up and down in rigid π bands in response to the balance

between donor and acceptor states. We note that on moving EF by as much as ± 1 eV relative to

the mirror symmetry plane in the band structure, we obtain a thermoelectric power value S ~ 1

-67-

µV/K, which is a factor of 40 less than the experimental data for the O2-doped material or the

fully degassed material. On the other hand, the thermoelectric power of a fully compensated

sample (S ~ 0) is consistent with this calculation (curve B, Figure 3.10), in agreement with early

theoretical calculations.74 In our model, transport in states above and below the Fermi level

cancel to give no net contribution to the diffusion thermopower when the Fermi level lies at the

band crossing. This is due to the exact electron-hole symmetry assumed in the system.

-2

-1

0

1

2

S (µ

V/K

)

-1.0 -0.5 0.0 0.5 1.0

EF/γ0

-3 -2 -1 0 1 2 3EF (eV)

Figure 3.11. Calculated thermoelectric power of a (10,10) carbon nanotube as a function of the Fermi level position.

Therefore, though the thermoelectric power of a compensated sample can be understood

on the basis of our model and using mirror symmetry bands [Eq. (3.28)], the same calculation is

unable to explain the large positive or negative values of S in the doped material.

-68-

3.3.5. Thermopower from Enhanced D(EF) due to Impurities

Previous calculations of the thermoelectric power in a doped metallic tube provide an

explanation for the large values of the thermoelectric power.112-114 These calculations have

reported that the density of states of metallic SWNTs containing nitrogen impurity donor states

exhibit a broad resonance in the density of states near the chemical potential. These broad

resonances have been identified as the donor bound states derived from the next highest-lying

electron band. Electron states in the metallic bands that cross near EF overlap in energy with

these bound states and create a broad resonance state. The conduction electrons in the metallic

bands can spend part of their time in a virtual bound state of the donor. The broad resonances in

the density of states D(E) overlap with the chemical potential and provide a larger nonzero value

for S, consistent with an enhanced D(E). An additional contribution to the thermoelectric power

may be provided by the carrier lifetime τ(E). This quantity has not yet been calculated.

According to Lammert et al.,113 boron is an acceptor impurity and its broad D(E)

resonances are from holes bound to the next lowest-lying hole-like band in the metallic tubes.

Here the resonance is below the chemical potential and the thermoelectric power will have the

opposite sign when compared to the case where donor impurities are dominant. We suspect that

these metallic DOS resonances located near EF can occur for many different impurities. An

additional contribution to the thermoelectric power from phonon drag will be discussed next.

-69-

Chapter 4.

Phonon Drag Thermoelectric Power of Single-Walled Carbon

Nanotubes

4.1. Introduction

Until now, it has been assumed implicitly that the flow of charge carriers and phonons in

response to a temperature gradient could be treated separately. However, many cases have been

studied where the interdependence of the flows must be taken into account, particularly at low

temperatures. The resultant phenomenon is known as the phonon drag effect.69,71

In the phonon drag effect, the flux of phonons proceeding from the hot end to the cold

end of the conductor sweeps or drags additional electrons along than what could occur under

normal diffusion. More specifically, phonons can impart momentum and energy to the electrons

via the phonon-electron interaction. In such scattering events, phonons are absorbed (or emitted)

and the electrons gain (or lose) the appropriate energy and momentum. This effect, often

extending up to room temperature and beyond, may bring about an increase in the thermoelectric

power, which usually takes the form of a “hump” in the temperature dependence of the

thermopower.

The phonon drag effect on the thermoelectric power of metals was first proposed by

Gurevich.115,116 The interpretation of these effects in semiconductors was given by Herring.117

For metals, a simple derivation of the phonon drag component of the thermoelectric power Sg

leads to69

-70-

, 31

α=NeC

S gg (4.1)

where Cg is the lattice specific heat per unit volume, N is the density of conduction electrons, and

the transfer factor α represents a measure of the relative probability of a phonon colliding with

the conduction electrons as compared to colliding with something else (e.g., phonons, impurity

centers, physical defects, etc.)

The probability of phonons to collide with each other increases as the temperature

increases (due to increasing anharmonic coupling), so that phonon-phonon collisions rapidly

become more frequent compared to the vital phonon-electron collisions which are responsible for

the phonon drag effect. Roughly speaking, the number of phonon-phonon collisions increases

with T. In turn, α in Eq. (4.1) should diminish as 1/T. At sufficiently low temperatures

(comparable to the Debye temperature TD) in reasonably pure metals, phonon collisions with

conduction electrons become dominant, so that α approaches unity and the phonon drag

contribution increases significantly. But, on the other hand, the lattice heat capacity for bulk

metals at low temperatures begins to fall off very rapidly as T 3 and thus Sg decays to zero as T

→ 0. A maximum in the phonon drag contribution would be expected when both the probability

of phonon collisions with other phonons and that with conduction electrons are comparable. The

qualitative variation in Sg as a function of temperature is shown in Figure 4.1.

-71-

Figure 4.1. Sketch of the thermoelectric power of a simple quasi-free electron pure metal as a function of temperature. A: Electron diffusion component of thermoelectric power approximately proportional to T. B: Phonon drag component with magnitude increasing as T 3 at very low temperatures (T << TD), and decaying as 1/T at “high” temperatures (T > TD) (Adapted from MacDonald69).

The shape of the temperature dependence of the thermoelectric power for a SWNT

network (Figure 3.10) suggests that phonon drag may be responsible for the peak observed at ~

100 K. However, the lattice specific heat of SWNTs is nearly linear in temperature118 and does

not show the strong temperature dependence required to explain the experimental data. Moreover,

in the case of a metallic SWNT, if we consider only intraband scattering, the contributions from

states filling the electron and the hole bands should cancel to give no net phonon drag

thermopower.

It is therefore clear that Eq. (4.1) for the phonon drag contribution probably cannot be

successful for metallic carbon nanotubes for at least two reasons: (1) the derivation of Eq. (4.1)

relies on a free, electronic band structure where transitions lie only within the parabolic bands,

and (2) the dominant decay mechanism for phonons is assumed to be phonon-electron scattering.

Applying Eq. (4.1) to metallic nanotubes with electron and hole bands yields a negligible

contribution to the thermopower due to phonon drag because of the electron-hole mirror band

-72-

symmetry.

4.2. Phonon Drag Model

A more general formulation of the phonon drag is obtained by considering the phonon

driven interband transitions near the Fermi level and by assuming that mechanisms other than

electron-phonon scattering limit phonon lifetimes. To do this, Scarola and Mahan119 have used a

variational solution to the coupled electron-phonon Boltzmann equations developed by

Baylin.120-122 The phonon drag thermoelectric power, Sg, can be obtained from the resulting

transport coefficients. The most general form is (see derivation in Appendix 2),122

( ) ( )[ . '',;

2

'';''''

0 ∑∑ ⋅τ−τα∂

∂σΩ

=jj

jjjjjjT

Nd

eS

kkqkkkk

q

Vvvkkqq ] (4.2)

Here σ is the electrical conductivity, d is the dimensionality of the system, Ω is the volume, N0 is

the Bose distribution function, and and are, respectively, the electron relaxation time and

the group velocity at wave vector k in the band j. Similarly V

jkτ jvk

q is the phonon group velocity at

wave vector q. The factor of 2 results from a sum over the spin degrees of freedom. For

convenience, it has been assumed that only one phonon branch contributes to Sg.

In Eq. (4.2), details of the phonon-electron interaction are included in the factor α, which

as mentioned before, is the relative probability that a phonon of wave vector q will scatter an

electron from the state kj to the state k’j’. Symbolically,

, 11

1

∑ −−

τ+τ

τ=α

ppe

pe (4.3)

where τpe is the phonon relaxation time due to the phonon-electron interaction and τp is the

-73-

phonon relaxation time due to all other interactions (e.g., phonon-phonon, phonon-boundary, and

phonon-impurity scattering).

Using first order perturbation, one finds121

( )

( )

, '',;

', ',0

',

∑+∂

∂τω

jj jjp

jj

IT

NTI

jj

kk kkq

kk

q

kkq

h

(4.4)

( ) ( )[ ] ( )[ ]( ) ( ) ( ) ( ) ( )[ ] ( , '''''12 0'', qkkqkkqkkq

qkk −−δ−ε−εδε−ε

ω= EjjNjfjf

TkMIB

jjh

) (4.5)

where is the electron-phonon coupling matrix element, f is the Fermi-Dirac distribution

function, and and are the electron and the phonon energies, respectively. In the

deformation ion model,

( )qM

( )jkε ( ) qq ω= hE

( ) ( )2qq ω= hDM where D is a constant related to the deformation

energy, and other tube parameters including the radius and lattice spacing. This constant has

been evaluated for (10,10) carbon nanotubes.123

Assuming that the phonon relaxation time is dominated by mechanisms other than

phonon-electron scattering, then τp is relatively small. If phonon-electron scattering dominates

phonon decay, one can show that at low temperatures, Sg is nearly independent of temperature.

Under this assumption,

( ) . '', '',

0 ∑>>∂∂

τω jj jjp

ITNT

kk kkq qh

(4.6)

The above approximation has been made in phonon drag studies of GaAs quantum wires.124,125

4.2.1. Phonon Lifetimes

The phonon-drag thermoelectric power depends, as we have seen, on the magnitudes of

-74-

the various relaxation times associated with phonons, which also determine the lattice thermal

conductivity. In this section, we will briefly examine the phonon interactions. Three different

mechanisms contribute to the scattering of a phonon:

1. Phonon scattering by electrons.*

2. Phonon-phonon anharmonic scattering.

3. Phonon scattering by mass defects such as impurities and/or isotopes.

Phonon-phonon scattering is usually the most important phonon scattering at higher

temperatures. It arises from the fact that the normal modes of the lattice are weakly coupled to

one another by the anharmonic part of the lattice potential. Thus, the anharmonic terms can cause

transitions between acoustic phonon modes. Under certain conditions, the relaxation time due to

phonon-phonon scattering satisfies70

(4.7) Dpp TTT >∝τ − ,1

and

. ,exp Du

pp TTTT

<⎟⎠⎞

⎜⎝⎛∝τ (4.8)

Here Tu is the temperature of the onset behavior of Umklapp processes and TD is the Debye

temperature.

At low temperatures, the Umklapp processes freeze out and the phonon relaxation time is

dominated by the impurity scattering. The transition between the Umklapp region and the

impurity scattering region manifests as a peak in the temperature dependence of the thermal

conductivity. For carbon nanotubes, this peak is observed at T ~ 320 K.8 Thus, impurity

* We must be careful not to confuse the phonon-electron relaxation time, τp, with the electron-phonon

relaxation time which we will continue to describe by the generic symbol τ.

-75-

scattering becomes the dominant scattering mechanism at low temperatures, i.e., T < 300 K for

carbon nanotubes.

In one dimension, Mahan126 have found that the lifetime due to impurity scattering is

( ) , 1 2

0qr ω=

τh (4.9)

where the constant r depends on the density of defects.

4.3. Baylin Formalism Applied to Metallic Carbon Nanotubes

The above formalism was applied to (10,10) carbon nanotubes where the dispersion of

electron and phonon modes is linear near EF, with group velocities υ and c, respectively. Only

the two pairs of nearly linear electronic bands crossing at the zero energy point are considered. In

the linear electronic band approximation, Eq. (4.2) vanishes when only intraband scattering

between the electronic bands is allowed. Therefore, only the interband scattering is considered.

For cTkTk DBB 2υ<<µ<< , one has119

( )( ) ,

21

1esgn

Bk ⎥⎦

⎤⎢⎣

⎡+

−−

=TkT

BSB

Tgµ

ε

εµµ

(4.10)

where

Dr

cLeB

υσπ

τ≡ 22 h

(4.11)

is independent of temperature and may be taken as a fitting parameter. Here υµ=εµ c2 , τ is

the electron relaxation time, TD is the Debye temperature, µ = EF is the chemical potential, and

( ) ( )cc −υ+υ≡γ . The above expressions were obtained taking into account the fact that the

-76-

scattering is strongest for phonons which have cq hµε= .

The thermoelectric power data for a SWNT sample can be fitted with an equation of the

form

( ) , gd SSTS += (4.12)

where Sd is given by Eq. (3.3). To determine the parameters A, B, and µ in Eqs. (3.3) and (4.10),

we used c = 20.35×103 m/s (longitudinal acoustic mode),53 υ = 8.1×105 m/s.37

20

15

10

5

0

-5

-10

S (µ

V/K

)

300250200150100500T (K)

(1) High-T O2-doped(irreversible)

(2) Ambient O2-doped(reversible)

(3) Compensated

(4) Degassed @500 Kin vacuum

Figure 4.2. Temperature dependence of the thermoelectric power for a purified SWNT thin film after successively longer periods of O2 degassing at 500 K in vacuum. Curve 1 corresponds to the same sample exposed to 1 atm O2 at 500 K for about 4 h, after being fully degassed (curve 4). The solid lines in the figure represent the fits to the data using Eq. (4.12).

In Figure 4.2, we show the experimental temperature dependence (below 300 K) for the

-77-

same sample dealt with in Figure 3.10. The solid lines in Figure 4.2 are fits to the data using Eq.

(4.12). The best fits to the data are achieved for values of the parameters (A, B, and µ) given in

Table 4–1. Unfortunately, the magnitude of the parameter B is difficult to estimate from first

principles. The sign of the parameter A (temperature coefficient of the diffusive thermopower)

indicates a positive diffusion thermoelectric power for curves 1 and 2, and a negative

thermoelectric power for curve 4. The diffusion thermoelectric power for the sample represented

by curve 3 is negligibly small, in agreement with previous observations that this curve

corresponds to a sample almost fully compensated by the balance between an unidentified

positively charged donor state, tentatively assigned to wall defects, and charged species (δ =

fractional charge), which can be removed by vacuum-degassing.

δ-2O

108 Note from Table 4–1 that

both the diffusion and the phonon drag contributions to the thermoelectric power are of the same

sign for a particular curve as indicated by the parameters A and B·sgn(–µ). The third column in

Table 4–1 represents the Fermi energy, measured with respect to the band crossing point for the

armchair nanotube. These values for µ are computed from the phonon drag term [Eq. (4.10)]. In

principle, µ could also be calculated from the parameter A in Eq. (3.3). However, µ should be

located near impurity states, and this complicates the problem considerably because impurity

state resonances are possible.112,114

The temperature dependence of Sg induces a smooth change of slope or a small “knee” on

the temperature dependence of the thermoelectric power at low T. Sg, with a temperature-

independent scattering mechanism for phonons, cannot introduce a more pronounced “hump” or

a broad peak superimposed on the linear Sd. The thermoelectric power data in Figure 4.3,

reported by Grigorian et al.,76 show this kind of behavior. These data were obtained from mats of

“as-prepared” SWNT ropes using the specific catalyst indicated in the figure. Best fit curves

-78-

derived from our model [Eq. (4.12)] are given by the solid lines, except for the sample grown

using Fe-Y catalyst.

Table 4–1. Best fit parameter values achieved with Eq. (4.12)

Curve A (µV/K2) B (µV) µ (eV)

(1) 3.9×10-2 1400 – 0.44

(2) 6.3×10-3 1125 – 0.37

(3) – 1.4×10-3 — ~ 10-4

(4) – 6.4×10-3 425 0.17

The broad peaks and the small knees in Figure 4.3 for the samples grown from Fe-Y, Co-

Y and Ni-Y catalysts were previously assigned to the Kondo effect involving residual magnetic

catalyst (e.g., Fe, Ni, Co) residing as small magnetic particles on the bundles or rope surfaces, or

trapped as atoms or small clusters within the bundles.76 The Kondo mechanism was also

suggested because the particular catalyst caused a simultaneous change in the magnitude of the

upturn in the electrical resistivity with decreasing temperature below ~ 100 K.76 The drawback of

the Kondo proposition of Grigorian et al. is that the thermoelectric power peak should occur at T

= TK ~ ∆E/kB, where ∆E is the width of the Kondo resonance near EF. One might expect a

stronger variation in TK with various magnetic impurities (Ni, Co, Fe), unless the effect is

associated with Y, which is a common element in all three samples in Figure 4.3. That a

relatively small knee was observed in the thermoelectric power of iodine-treated material

(bottom trace, Figure 4.3) was tentatively explained by Grigorian et al. as to be due to the fact

that iodine either complexed with the residual metal catalyst or vapor-transported the metal away

as, e.g., FeI. As our fits to the data of the unpurified SWNT material of Grigorian et al.76 indicate,

-79-

a phonon drag contribution superimposed on a linear metallic diffusion thermoelectric power

background fits the data very well, except for the more pronounced peaks in samples grown with

Co-Y and Fe-Y catalysts.

80

60

40

20

0

S (µ

V/K

)

300250200150100500T (K)

Fe-Y

Co-Y

Ni-Y

Iodine-treated(Fe-,Co-,Ni-Y)

Figure 4.3. Temperature dependence of the thermoelectric power for SWNT mats prepared using different catalysts. The samples were not purified and contained ~ 5 at% residual catalyst. The data were measured by Grigorian et al.76 The solid lines represent the best fits to the data using Eq. (4.12).

The solid line in Figure 4.4 represents the fit of Eq. (4.12) to typical thermoelectric power

data (curve 1 in Figure 4.2) of a purified SWNT material. The dashed lines in this figure

represent the fits of Eqs. (3.3) and (4.10) to the data. Note that the diffusion contribution to the

thermoelectric power fits well to the observed linear metallic thermoelectric power below 50 K.

-80-

In this temperature range, the phonon drag contribution is nearly zero and increases smoothly

with T before decreasing slowly above ~ 150 K. The contribution from phonon drag flattens out

for large temperatures. The lack of suppression of the phonon drag at high temperatures clearly

results in only a smooth change of slope or small knee, but not a pronounced peak characteristic

of SWNT samples grown with Co-Y and Fe-Y catalysts. It is likely that these impurities impose

a different scattering mechanism (for either electrons or phonons) in nanotubes which can

significantly alter the temperature dependence of the phonon drag thermopower.

18

15

12

9

6

3

0

S (µ

V/K

)

300250200150100500T (K)

Sd

Sg

Figure 4.4. Fits to the measured thermoelectric power data (curve 1 in Figure 4.2) using a model involving diffusion and phonon drag contributions to the thermoelectric power. The solid curve represents a fit to the data using Eq. (4.12). The dashed lines represent the contributions from Sd [Eq. (3.3)] and Sg [Eq. (4.10)].

-81-

Chapter 5.

Carbon Nanotubes: A Thermoelectric Nano-Nose

5.1. Introduction

Chemical doping effects on the electrical properties of SWNTs have been investigated by

several groups. SWNT doping experiments with electron withdrawing (Br2, I2) and donating

species (K, Cs) were first carried out on bundled SWNT mats by Lee et al.127 and Grigorian et

al.128 Individual bundles of SWNTs have also been studied after doping in situ with

potassium.129,130 The early studies have demonstrated the amphoteric character of carbon

nanotubes.† In particular, Rao et al.131 first demonstrated the amphoteric character of SWNTs by

observing the sign of frequency change in the tangential Raman modes.

In general, chemical doping can change the electronic behavior of SWNTs from p-type to

n-type or vice versa, accompanied by orders of magnitude changes in the resistance of the

material. The largest changes are expected for semiconducting nanotubes. The doping species

can also absorb and charge transfer with the nanotube surfaces and/or intercalate into the

interstitial sites of bundles of SWNTs.

We will show that carbon nanotubes are also sensitive to gas molecule physisorption,

exhibiting significant changes in their electrical transport properties. In previous chapters, we

already studied how molecular oxygen adsorption (which probably is weakly chemisorbed due to

charge transfer) affects the thermoelectric power and electrical resistance of carbon nanotubes.

† Amphoteric means that it can be doped to produce additional electrons and holes.

-82-

Elegant work by Kong et al.18 have found that individual semiconducting SWNT

transistors (CHEM-FET) can be used in miniature chemical sensors to detect small

concentrations (2-200 ppm) of gas molecules (NO2 and NH3) with high sensitivity at room

temperature. Exposure to 200 ppm of NO2 can increase the electrical conductance by up to three

orders of magnitude in a few seconds. On the other hand, exposure to 2% NH3 caused the

conductance to decrease by up to two orders of magnitude. Thus, CHEM-FET sensors made

from SWNTs have high sensitivity and fast response time at room temperature, which are

important advantages for sensing applications. NO2 and NH3 are known to be an electron

acceptor and an electron donor, respectively. Therefore, Kong et al. have proposed that the

charge transfer between the tube wall and the adsorbed molecules was driving the observed

changes in the electrical conductance of semiconducting nanotubes. Interestingly, changes in the

electrical resistance and the thermoelectric power of SWNTs were observed in cases where gas

adsorption (i.e., N2, He, H2) should not induce any charge transfer.89,132 In such cases, the

changes in the electrical properties upon gas adsorption were tentatively assigned to changes in

the electron and the hole free carrier lifetime (or equivalently, to the carrier mobility). We have

assigned these changes in the carrier lifetime to increased carrier scattering from dynamic defect

states associated with either physisorbed gas molecules or collisions of the gas molecules with

the tube walls.89

Despite all these considerations, no microscopic or “atomistic” explanation of the

transport changes induced by molecular adsorption on SWNTs has been given yet. It is

reasonable to expect the effects of molecules on the transport properties of SWNTs to be an

outcome of a delicate interplay among various factors including the charge transfer, possible

pinning of the Fermi energy, the creation of impurity band and its location relative to EF.

-83-

However, the contribution from each one of these factors to the transport properties of SWNTs

has not been established yet. It is apparent that a quantitative understanding of their contributions

to the transport properties of SWNTs is essential and timely in its own right, as well as for

understanding the true intrinsic properties of SWNTs.

Figure 5.1. Schematic structure of a SWNT bundle showing the sites available for gas adsorption. The dashed line indicates the nuclear skeleton of the nanotubes. Binding energies EB and specific surface area contributions σ for hydrogen adsorption on these sites are indicated.133

The bundle structure of SWNTs produces at least four distinct sites in which gas

molecules can adsorb, as shown in Figure 5.1: on the external bundle surface, in a groove formed

at the contact between adjacent tubes on the outside of the bundle, within an interior pore of an

individual tube and inside an interstitial channel formed at the contact of three tubes in the

bundle interior. For a particular gas molecule, some sites can be excluded on size considerations

surface

pore

groove

channel

EB = 0.119 eVσ = 45 m2/g

EB = 0.089 eVσ = 22 m2/g

EB = 0.062 eVσ = 783 m2/g

EB = 0.049 eVσ = 483 m2/g

surface

pore

groove

channel

EB = 0.119 eVσ = 45 m2/g

EB = 0.089 eVσ = 22 m2/g

EB = 0.062 eVσ = 783 m2/g

EB = 0.049 eVσ = 483 m2/g

-84-

alone (assuming the bundle or tube does not swell to accommodate the adsorbed molecule). For

molecular hydrogen, calculations ignoring swelling have ordered the binding energy EB in these

various sites as EB (channels) > EB (grooves) > EB (pores) > EB (surface).56,134 Access of

molecules to the internal tube pores is either through open SWNT ends or defects (holes) in the

tube walls. It is commonly believed that these gateways must be produced by post-synthesis

chemical treatment. Small molecules have access to the interstitial channels between nanotubes,

and their adsorption there could conceivably lead to a swelling of the bundle diameter.

5.2. Effects of Gas Adsorption on the Electrical Transport Properties of

SWNTs

Figure 5.2 shows the thermoelectric power response over time of a degassed “as-

prepared” SWNT mat to 1 atm overpressure of He gas at T = 500 K (filled symbols). The initial

thermopower (or S0) is due, in part, to defects in the structure, as discussed above. S0 is therefore

expected to be sample dependent. The thermopower is seen to rise exponentially with time,

saturating at ∼ 12 µV/K above the initial thermopower S0. Removing the He overpressure above

the SWNT mat induces an exponential decay of S with time (open symbols). The dashed lines in

the figure represent exponential fits to S(t). The four-probe resistance R (not shown) was found

to exhibit a similar exponential rise and fall. A concomitant increase of ∼ 10% in R was also

observed.

Note that the response times to these treatments are long (several hours). We interpret

these long time constants as due to the slow diffusion of the gas into and out of the internal pores

and the channels of SWNT ropes. Simulations by Tuzun et al.135 of the dynamic flow of helium

-85-

and argon atoms through nanotubes have predicted that the flow slows down rapidly when both

the nanotube and the fluid are kept at high temperatures. When the tube moves (due to thermal

vibrations), it perturbs the motion of the nearest fluid atoms. This causes the fluid motion to

randomize faster, leading to hard collisions with the tube. Thus, fluid-nanotube collisions tend to

slow down the fluid flow. In general, fluid-fluid interactions are stronger for heavier atoms,

leading to the excitation of larger amplitude vibrations in the tube. This, in turn, randomizes the

fluid motion faster.

-45

-40

-35

-30

-25

S (µ

V/K

)

543210Time (h)

He (1atm) T = 500 K

τads = 0.28 h τdes = 1.09 h

Figure 5.2. The time dependence of the thermoelectric power response of a SWNT mat to 1 atm overpressure of He gas (filled circles), and to the subsequent application of a vacuum over the sample (open circles). The dashed lines are exponential fits of the data (see text).136

The dashed lines in Figure 5.2 are the fits to the S(t) data using exponential functions of

-86-

the form,

( ), 1max0τ−−∆+= teSSS (5.1)

and

, max0τ−∆+= teSSS (5.2)

for adsorption and desorption, respectively. Here, S0 is the initial or degassed thermopower,

∆Smax is the maximum response to gas exposure (t → ¶), and τ is the time constant for the

response. The dashed curves are seen to fit the data for adsorption and desorption rather well. It

should be noted that the desorption time is ~ 3 times larger than the adsorption time constant, in

agreement with the results of Tuzun et al.135 A collective effort is needed for the atoms to find

their way out of the internal pore and channels. We therefore have indirect evidence for single-

file diffusion during the bundle desorption process.

Figure 5.3 shows the time response of the thermopower of a vacuum-degassed SWNT

mat to a sudden 1 atm overpressure of H2 gas at T = 500 K (solid symbols). The response of the

H2-loaded mat to a vacuum (open symbols) is also shown. With increasing exposure time to H2,

the thermopower is driven to more negative values, eventually saturating after ~ 6 h. Note that

in contrast to He exposure (Figure 5.2). The negative thermoelectric response of

SWNTs to H

desads 3τ≈τ

2 is truly special. Exposure of carbon nanotubes to inert gases will be discussed later

in this thesis.

We found that the initial thermopower of the degassed sample and its maximum change

upon H2 exposure depends somewhat on the post-synthesis processing (Figure 5.4).132 We

believe that this difference is most likely due to different concentrations of wall defects, perhaps

introduced during post-synthesis (acid) purification. It is interesting to note from Figure 5.4 that

the equilibration time for S(t) in 1 atm H2 is reduced with increasing reflux time in HCl. This

-87-

would be consistent with the introduction of physical holes in the tube wall with exposure to HCl,

possibly at defect sites associated with carbidic (Ni-C) bonds to residual growth catalyst. It may

also have something to do with “spillover” involving residual catalyst in the sample.137 Spillover

describes the catalytic process of the dissociation of molecular into atomic hydrogen. It could be

that atomic hydrogen is changing the moieties attached to the tube wall.

-58

-56

-54

-52

-50

-48

S (µ

V/K

)

876543210Time (h)

H2 (1atm) T = 500 K

τads = 0.81 h τdes = 0.28 h

Figure 5.3. In situ thermoelectric power versus time after exposure of a vacuum-degassed SWNT mat to 1 atm overpressure of H2 at T = 500 K (solid symbols). The response of the H2-loaded SWNT sample to a vacuum is also represented (open symbols). The dashed lines are fits to the data using exponential functions (see text).

These two observations document a rather remarkable sensitivity of the electrical

transport parameters to adsorbed gases, even an inert gas such as He. Both the thermoelectric

power and the electrical resistance were completely reversible in all these experiments.

-88-

-60

-50

-40

-30

-20

-10S

(µV

/K)

1614121086420Time (h)

(0.2 at%)

(2 at%)

(5 at%)

∆Smax = 4 µV/K

∆Smax = 6 µV/K

∆Smax = 7 µV/K

Figure 5.4. In situ thermoelectric power as a function of time after exposure of degassed SWNT mats to a 1 atm overpressure of H2 at T = 500 K (solid symbols). The open symbols are the response of the H2 loaded SWNT system to a vacuum. Data are shown for three samples: not purified (bottom), HCl reflux for 4 h (middle), HCl reflux for 24 h (top). The dashed lines are guides to the eye. The catalyst residue in at% is indicated.

5.3. Thermoelectric Power from Multiple Scattering Processes

As mentioned before, the electrical transport response of a bundle of SWNTs to a variety

of gases can be understood in terms of the change in the thermoelectric power of the metallic

tubes due to either a charge-transfer-induced change in the Fermi energy (i.e., molecule donates

an electron to the conduction band) or the creation of an additional scattering channel for

conduction electrons in the metallic nanotube wall. This scattering channel might be identified

-89-

with impurity sites associated with the adsorbed gas molecules or be attributed to gas collisions.

We briefly develop the equations necessary to understand this point of view.

We have shown earlier that the metallic behavior of the SWNT mat thermopower is a

consequence of the percolating pathways through the metallic tube components in the mats.

According to the Mott relation (derived in Appendix A), the thermoelectric power associated

with the diffusion of free carriers in a metal can be written compactly as a logarithmic energy

derivative of the electrical resistivity ρ,

( ) , ln

FEEdEEdCTS

=

⎟⎠⎞

⎜⎝⎛ ρ

= (5.3)

where ekC B 322π= .

For our purposes, it is convenient to explicitly separate the contributions to the resistivity

from (a) scattering intrinsic to the degassed tube, ρ0 (identified with phonons and permanent tube

wall defects), and (b) additional carrier scattering processes associated with perturbations in the

local tube wall potential due to adsorbed gas molecules or collisions of gas molecules with the

tube wall, ρI. We assume that these scattering contributions follow Mattheissen’s rule, which is

equivalent to the additive nature of independent scattering rates, i.e.

. 0 Iρ+ρ=ρ (5.4)

If these contributions are incorporated into Eq. (5.3), it follows that

( , 000

00 SSSSS

S II

I

II −ρ

ρ+=

ρ+ρ)ρ+ρ

= (5.5)

where

, 13

F

0

0

22

0EE

B

dEd

eTkS

=

⎥⎦

⎤⎢⎣

⎡ ρρ

π= (5.6)

-90-

and

. 13

F

22

EE

I

I

BI dE

de

TkS=

⎥⎦

⎤⎢⎣

⎡ ρρ

π= (5.7)

The variables S0 and SI are, respectively, the thermopower of the degassed tube and the

additional impurity contribution from adsorbed gas molecules. It is usually understood

that , i.e., the intrinsic resistivity is much greater than the additional resistivity due to

impurities. This is certainly the case here, as verified by experiment.

Iρ>>ρ0

Equation (5.5) has the same form as the well-known Nordheim-Gorter (N-G) expression,

developed to explain the thermoelectric power of binary alloys.70 In this case ρ and ρI refer,

respectively, to resistivity contributions from the host and a dopant. The significance of Eq. (5.5)

for our work is that, for fixed T, the thermopower is linear in ρI, if ( )0SSI − is constant and not

affected by the contact with the gas and if ρI << ρ. This should occur if the gas contact leaves the

SWNT band structure intact and EF unchanged, i.e., charge transfer between the adsorbed gas

and the host lattice does not occur. This situation is consistent with physisorption, NOT a

chemisorption process.

If the particular molecules under study are physisorbed, i.e., van der Waals bonded to the

tube walls, they will induce only a small perturbation on the SWNT band structure and an almost

linear N-G plot should be obtained. If, on the other hand, the N-G plot for a particular adsorbed

gas on SWNTs were strongly curved, this nonlinearity would indicate that the molecules are

chemisorbed onto the tube walls. Chemisorption, of course, has a much more pronounced effect

on the host band structure and/or the value of EF, and thus ( )0SSI − must then depend on gas

coverage or storage and the linearity of a N-G plot is lost. N-G plots, therefore, should be very

valuable in identifying the nature of the gas adsorption process in SWNTs. Below we further

-91-

develop Eq. (5.5) and the validity of these remarks will be more apparent.

-70

-60

-50

-40

-30

-20

-10

0

10S

(µV

/K)

50x10-3403020100ρI / ρo

N2He

H2

T = 500 K

-60

-40

-20

0

20

S (µ

V/K

)80x10-36040200

ρI / ρo

O2

NH3

Figure 5.5. Nordheim-Gorter plots showing the effect of gas adsorption on the electrical transport properties of a SWNT mat. The amount of gas stored in the bundles increases to the right, tracking the increase in ρ. For the H2 data, the open circles are from the time dependent response to 1 atm of H2 at T = 500 K and the closed circles are from a pressure study at the same temperature. The inset shows the Nordheim-Gorter plots for O2 (electron acceptor) and NH3 (electron donor). Note that the data in the inset, as opposed to that in the main plot, is non-linear. The non-linearity is consistent with charge transfer and Fermi energy shifts.

-92-

In Figure 5.5, we display the N-G plots (S vs. ρI) for isothermal adsorption of He, N2, and

H2 in SWNTs at 500 K. As can be seen in the figure, the data are linear for these three gases,

consistent with molecular physisorption and Eq. (5.5). In the inset to Figure 5.5, we display N-G

plots for NH3 and O2; these are strongly curved, indicating, as discussed above, that these

molecules must chemisorb on the tube walls. These results confirm the point of view, previously

discussed, that the large changes in the thermoelectric power of SWNT exposed to O2 can be

identified with chemisorption.

Returning to the discussion of the linear N-G plots in Figure 5.5 (for He, N2, and H2),

some interesting points remain. First, the N-G slope for He and N2 are positive, while that for H2

is negative. According to Eq. (5.5), the sign of the slope is determined by . This

conclusion is best seen by writing down the form of the thermopower explicitly. We use the Mott

relation [Eq. (5.3)] and the well known expression

( 0SSI − )

( ) ( )[ ] ( ) ( ) ( ). 221 EEDEveEE τ=ρ=σ − (5.8)

Then, the Nordheim-Gorter equation [Eq. (5.5)] becomes

. 113

F

0

0

22

0EE

I

I

IB

dEd

dEd

eTkSS

=

⎥⎦

⎤⎢⎣

⎡ ττ

−τ

τ⎟⎟⎠

⎞⎜⎜⎝

⎛ρρπ

+= (5.9)

We can further understand the equation above by allowing ( ) ( )EgfE jjj =τ1 , where f

and g are functions and f is not a function of E. For impurity scattering we might expect that

( )EgNI ατ ~1 , where N is the number of molecules adsorbed per unit length of tube and ( )Egα

is the scattering cross-section. With this factorization in mind we notice that

. 11

FF EEEE

I

I dEdg

gdEd

==⎥⎦

⎤⎢⎣

⎡=⎥

⎤⎢⎣

⎡ ττ

(5.10)

Therefore, the first term between square brackets in Eq. (5.9) is independent of the constant

-93-

prefactor α of the scattering cross-section and the molecular coverage N, hence independent of ρI,

as long as EF is constant (no charge transfer). We can now anticipate either a positive or a

negative slope to the data S vs. ρI collected at fixed temperature, depending on the sign and

magnitude of the derivatives in Eq. (5.9). Fundamentally different energy dependence for the

electron scattering rates associated with He and H2 adsorption sites must exist. In Figure 5.5 we

also note a significant difference in slope (but not sign) for He and N2 sites. Indeed, it is the

sensitivity of the N-G plots at fixed temperature to different molecules that can be the basis for

the utility of a SWNT thermoelectric “nano-nose”.

-94-

Chapter 6.

Effects of Molecular Physisorption on the Transport Properties of

Carbon Nanotubes

6.1. Introduction

Molecular adsorbates that engage in charge transfer with the nanotube wall might be

expected to have a significant impact on the transport of charge and heat down the nanotube wall.

Although weaker than common dopants for carbon materials, the chemical doping effects of

small organic molecules are far from negligible. Recent work138 has showed that adsorption of

several small amine-containing organic molecules on nanotubes can cause significant changes in

the electrical conductance of the nanotube samples. Modulated chemical gating of individual

semiconducting SWNTs by these molecules has also been demonstrated.138

In this chapter we discuss the results of a systematic study of the changes in the

thermoelectric power and electrical resistance of vacuum-degassed films of nanotube bundles

induced by adsorption of six-membered ring molecules (C6H2n; n = 3-6), alcohols (CnH2n+1OH; n

= 1-4) and water molecules. For six-membered ring molecules, as n increases from 3 (benzene)

to 6 (cyclohexane), π electrons are removed from the molecule. For n = 6, cyclohexane, only σ

bonds remain. Thus the C6H2n/SWNT is an interesting model system to study the coupling

between a molecule and a carbon nanotube, as regulated by the π-electron character of the

adsorbed molecule. For polar molecules, we show that the measured perturbation on the

electronic properties of the nanotubes is sensitive to these molecules. Interestingly, exposure to

-95-

water, which is also strongly polar in nature, produces virtually no change in the thermoelectric

power, though the electrical resistance shows a change of ~ 4%, typical for the alcohols. We also

present the results here for exposure of SWNTs to water vapor.

6.2. Effects of Adsorption of Six-Membered Ring Molecules

The experiments began with an in situ vacuum-degassing of the SWNT film in the

measurement apparatus at 500 K. After the thermopower remained constant and negative for ~ 8

h, the sample was cooled to 40 ºC and the vapors of the particular six-membered ring compounds

(C6H2n; n = 3-6) were admitted. A sample temperature of 40 ºC was chosen to avoid

condensation of liquid on the nanotube bundles. The molecular vapor pressure of the C6H2n is

essentially independent of n and equal to that of the vapor in equilibrium with the liquid (vapor

pressure ∼ 12 kPa at 24 ºC).

Figure 6.1(a) shows the in situ thermopower response with time to the vapors of benzene

(C6H6), 1,3-cyclohexadiene (C6H8), cyclohexene (C6H10), and cyclohexane (C6H12). The sample

temperature was maintained at 40 ºC. After each curve in Figure 6.1 was collected, the sample

was heated again under vacuum at 500 K for a few hours in order to fully recover the original

degassed values S0 and R0. For benzene, with increasing exposure time, the thermopower

increases with time from its initial degassed value at 40 ºC (S0 = – 6.4 µV/K), eventually

saturating after ~ 6 h at a positive value Smax = +1.3 µV/K. Subsequent exposure to 1,3-

cyclohexadiene leads to a similar time dependence of the thermopower and a saturation at Smax ~

– 3.6 µV/K. Cyclohexene was found to induce a smaller change in the thermopower, saturating

at Smax ~ – 4.6 µV/K. Cyclohexane, which has no π electrons, was found to produce no

-96-

detectable change in the thermopower.

-8

-6

-4

-2

0

2S

(µV

/K)

6543210Time (h)

C6H2n n = 3

n = 4

n = 5

n = 6

(a)

2.9

2.8

2.7

2.6

2.5

2.4

R (Ω

)

543210Time (h)

n = 3

n = 4

n = 5

n = 6

C6H2n(b)

Figure 6.1. In situ (a) thermoelectric power and (b) resistance responses at 40 ºC as a function of time during successive exposure of a degassed SWNT thin film to vapors of six-membered ring molecules C6H2n; n = 3-6. The dashed lines are guides to the eye. The vapor pressure was ~ 12 kPa.

-97-

Figure 6.1(b) shows the concurrent time evolution of the four-probe resistance for

exposure to each molecular vapor at 40 ºC. Exposure to benzene produces the largest change in

resistance with an increase of ∆R/R0 ~ 13% at saturation. 1,3-cyclohexadiene and cyclohexene

induce increases in resistance saturating at ∆R/R0 ~ 10% and 7%, respectively. Exposure to

cyclohexane induces essentially no change in the four-probe resistance, consistent with the

thermopower results. It is clear that both the thermoelectric power and the electrical resistance

follow similar trends with varying n in C6H2n. The results are tabulated in Table 6–1.

Table 6–1. Comparison of the T = 40 ºC thermoelectric power and resistive responses of a SWNT thin film to adsorbed C6H2n molecules. The vapor pressure at 24 ºC and the adsorption energy Ea of the corresponding molecule (measured on graphitic surfaces) are also listed. S0 and R0 refer to the degassed film before exposure to C6H2n molecules.

Molecule

C6H2n

n

Smax

(µV/K)a

∆Smax

(µV/K)

(∆R/R0)max

(%)b

p

(kPa)139

Ea

(kJ/mol)140,141

Benzene 3 + 1.3 7.7 13 12.7 9.42 ± 0.06

1,3-Cyclohexadiene 4 – 3.6 2.8 10 13.0 8.87 ± 0.08

Cyclohexene 5 – 4.6 1.8 7 11.8 8.71 ± 0.08

Cyclohexane 6 – 6.4 0 0 13.0 7.59 ± 0.07 a . 0maxmax SSS −=∆

b . 0maxmax RRR −=∆

All of these C6H2n molecules are almost of the same size and have approximately the

same molecular weight. We therefore suggest that the observed differences in the thermopower

and the resistive responses should be attributed to the number of π electrons in the molecule.

Adsorption of benzene (n = 3) induces the largest increase in the thermopower and the resistance.

As the number of π electrons per molecule is reduced (increasing n), the impact of the molecular

adsorption on the transport properties coefficients of the SWNT disappears. The data are

-98-

therefore consistent with the idea of a new scattering channel created by the molecular adsorbate,

and the size of the effect is apparently driven by the coupling of π electrons in the molecule to π

electrons in the metallic nanotube wall. We have discussed the connection of S and R in metallic

tubes in section 5.3 above.

8

7

6

5

4

3

2

1

0

∆Sm

ax (µ

V/K

)

9.59.08.58.07.5Ea (kJ/mol)

C6H12

C6H10

C6H8

C6H6

Figure 6.2. Maximum change of the thermoelectric power of a SWNT film as a function of the adsorption energy of the adsorbed molecule. The dashed line is a guide to the eye.

It is interesting to compare the values of the thermoelectric power and the resistance at

saturated coverage to the heat of adsorption Ea of C6H2n on graphitic carbons (Table 6–1).140,141

The tabulated data show that the maximum responses ∆Smax and (∆R/R0)max at 40 ºC correlate

with the adsorption energy of the C6H2n molecules to a sp2 carbon surface (see also Figure 6.2).

Zhao et al.142 have used first principles calculations to study the interaction between carbon

-99-

nanotubes and organic molecules including benzene and cyclohexane. They have found that

benzene and cyclohexane are very weak charge donors (0.01-0.04e) to carbon nanotubes, similar

to most inorganic gas molecules. However, adsorption of benzene molecules on the carbon

nanotube surface brings, as a consequence, a hybridization between molecular levels and

nanotube valence bands. In contrast, the electron density in the top valence band of a carbon

nanotube is localized on the nanotube and has no density on an adsorbed cyclohexane molecule.

This result clearly shows that the difference between the electronic configurations of benzene

and cyclohexane molecules plays a role in the perturbation of the transport properties.

-6

-4

-2

0

2

S (µ

V/K

)

0.150.100.050.00∆R/R0

C6H6

C6H8

C6H10

T = 40 ºC

Figure 6.3. S vs. ∆R/R0 plots during exposure to C6H2n (n = 3-6). The dashed curve is a fit to the data using a quadratic function.

In Figure 6.3, we plot the evolution of the thermopower as a function of the fractional

-100-

change in the four-probe resistance ∆R/R0 at 40 ºC and at fixed molecular vapor pressure p ~ 12

kPa. As the coverage of the molecules on the SWNTs increases with increasing exposure time to

the respective molecular vapor, both the thermoelectric power and the relative change of the

resistance ∆R/R0 increase, as shown in Figure 6.1. Data for the adsorption of three molecules

C6H2n (n = 3-5) are shown. Cyclohexane C6H12 did not produce a detectable thermoelectric

response (i.e., a change in S and R). Although the maximum variations in the thermoelectric

power and the resistance observed in the same SWNT film sample after long term exposures to

various molecules are very different, a universal behavior (i.e., independent of n) is observed for

the dependence of the thermoelectric power on the change in resistance ∆R/R0 (Figure 6.3).

We appeal to the following expression we derived earlier,

. 113

F

0

00

22

0EE

I

I

IB

dEd

dEd

eTkSS

=

⎥⎦

⎤⎢⎣

⎡ ττ

−τ

τ⎟⎟⎠

⎞⎜⎜⎝

⎛ρρπ

+= (6.1)

First, we associate 0ρρI in the above equation with the experimental quantity 0RR∆ , assuming

that the C6H2n adsorption does not change the contact resistance between bundles in the SWNT

film. Therefore, if the lifetime factor in brackets is independent of the molecular coverage and

the constant prefactor of the scattering cross-section, Eq. (6.1) predicts a universal relationship

between S and 0RR∆ .

For small values of 00 ρρ=∆ IRR , Eq. (6.1) predicts a linear relationship between S and

0RR∆ . At low coverage, i.e., for 06.00 <∆ RR we do observe an approximately linear behavior.

However, at higher values of 0RR∆ , S increases faster than a linear variation. In fact, we can fit

the data reasonably well with the quadratic relation ( ) ( 200 RRcRRbaS ∆+∆+= ) over most of

the data range. This quadratic function is plotted as the dashed line in Figure 6.3. The nonlinear

-101-

behavior is tentatively assigned to multiple scattering processes associated with larger molecular

coverage. Further work is necessary to understand the observed non-linearity in the S vs. ∆R/R0

plots of Figure 6.3.

-10

-8

-6

-4

-2

0

2

4

S (µ

V/K

)

300250200150100500T (K)

S0 (T)

C6H2n

n = 3

n = 4

n = 5

Figure 6.4. Temperature dependence of the thermoelectric power of the degassed SWNT after saturation coverage of the various C6H2n molecules. The dashed lines are guides to the eye.

In Figure 6.4, we display the temperature dependence of the thermoelectric power after

saturation coverage at 40 ºC. Before cooling the sample to collect the data, the valve to the

hydrocarbon bulb was closed. As the temperature of the sample was reduced, residual vapor in

the thermopower probe should first condense at the bottom of the sample compartment. Some of

the molecules may condense on the saturated surface of the bundles forming a second monolayer

-102-

over the initial primary layer. However, the effect of this second monolayer or overlayer on the

transport properties should be small. As shown in Figure 6.4, the thermoelectric power S0(T) for

the degassed film is nearly constant down to 100 K and approaches zero quasi-linearly at lower

temperatures.

The temperature dependence of the thermoelectric power for the SWNT film after

exposure to C6H2n, can also be understood from Eq. (6.1). We assume that metallic tubes exhibit

an intrinsic resistivity, i.e.

(6.2) ⎩⎨⎧

∝ρTT

Thigh ,

low const,0

and the impurity resistivity ρI is independent of temperature. Then, we have

(6.3) ( )⎩⎨⎧

∝−T

TTSTS

high const, low ,

0

Thus, the thermoelectric power should vanish at low temperatures and should be a constant at

high temperatures, as observed in Figure 6.4.

6.3. Effects of Adsorption of Polar Molecules

As in the case of six-membered ring molecules, the experiments started with an in situ

vacuum-degassing (for ~ 15 h) of a purified SWNT thin film in the measurement apparatus at

500 K, before water or the various CnH2n+1OH molecular vapors were introduced. A glass bulb

containing the water or alcohol was connected via a valve to the measurement apparatus [see

inset to Figure 6.5(b)]. All the alcohols were spectra grade (Sigma-Aldrich, Co) and had been

previously vacuum-degassed. The water was de-ionized and had a resistivity of ~ 18 MΩ-cm.

-103-

The vapor pressure p above the SWNT sample for each liquid was that known to be in

equilibrium with the liquid in the bulb at 24 ºC (Table 6–2). After the thermoelectric and

resistive responses to a particular molecular vapor were recorded, the sample was then degassed

in situ at 500 K again until the thermoelectric power and four-probe resistance of the sample

returned to the original “degassed” values (S0, R0). Then the same film was exposed to the next

molecular vapor and so on. Data are presented here from one such SWNT thin film; other

samples, prepared in the same way, showed similar behavior.

Figure 6.5 shows the in situ thermoelectric power and the normalized four-probe

resistance responses with time t to the vapors of methanol (CH3OH), ethanol (C2H5OH),

isopropanol (C3H7OH), butanol (C4H9OH), and H2O. Dashed lines in Figure 6.5 are fits to the

data using a simple exponential function

( ), 1max0SteSSS τ−−∆+= (6.4)

where S0 is the initial or degassed thermopower, ∆Smax is the maximum response to physisorption

(t → ¶), and τS is the time constant for the response. The same function is used for the resistive

response, but R0, ∆Rmax, and τR replace their counterparts in Eq. (6.4).

After each set of curves in Figure 6.5 was collected for a specific adsorbate, the sample

was then heated again in situ under vacuum (∼ 10-7 Torr) at 500 K to remove the molecules.

After a few hours at 500 K, the sample was found to fully recover the original degassed values S0

and R0. In Figure 6.5(a), it is seen that, for methanol, ethanol, isopropanol, and butanol, the

thermoelectric power also rises exponentially with time from the degassed value S0 ~ − 2.7 µV/K

to a higher plateau after ~ 1 h. For methanol and ethanol, S is even driven positive, saturating at

Smax ~ 1.1 and 0.1 µV/K, respectively. Exposure to larger alcohol molecules, i.e., isopropanol

and butanol, is found to lead to smaller changes in S and a saturation at Smax ~ – 0.5 and – 1.0

-104-

µV/K, respectively.

-4

-3

-2

-1

0

1

2

S (µ

V/K

)

1.00.80.60.40.20.0Time (h)

CnH2n+1OH

H2O

n = 1

n = 2

n = 4

n = 3

S0

(a)

1.10

1.08

1.06

1.04

1.02

1.00

0.98

R/R 0

1.00.80.60.40.20.0Time (h)

T1

V2

V1

T2

vacuum

T1

V2

V1

T2

vacuum

CnH2n+1OH

H2O

n = 1

n = 4

(b)

Figure 6.5. Time dependence of the (a) thermoelectric power and (b) normalized four-probe resistance responses to vapors of water and alcohol molecules (CnH2n+1OH; n = 1-4) at 40 ºC. The dashed lines are fit to S(t) and R(t) data using an exponential function. The inset shows a simple schematic of the measurement apparatus. The liquid temperature T2 establishes the vapor pressure in the sample chamber which is at a temperature T1 > T2. The system is evacuated through V2. After degassing, V2 is closed and V1 is opened. The responses of S and R are then measured simultaneously.

-105-

Interestingly, exposure to water vapor (another small, but very polar molecule) induces

virtually no change in the thermoelectric power. Bradley et al.88 have also found very weak or no

response of the thermoelectric power of mats of bundled SWNTs to water vapor. This lack of

sensitivity of the thermoelectric power to water is very interesting and will be discussed later.

Table 6–2 shows some relevant parameters of the molecules including the molecular projection

area and the dipole moment.

Table 6–2. Comparison of the T = 40 ºC thermoelectric power and resistive responses of a SWNT thin film to adsorbed water and CnH2n+1OH; n = 1-4. The vapor pressure p at 24 ºC, the molecular area A, the static dipole moment µ, and the adsorption energy Ea of the corresponding molecule (measured on graphitc surfaces) are also listed. S0 and R0 refer to the degassed film before exposure to water and alcohols. An increase in vapor pressure did not change the values of ∆Smax or ∆Rmax; see text.

Molecule ∆Smax

(µV/K)a

(∆R/R0)max

(%)b

p

(kPa)139

A

(Å2)143,144

Ea

(eV) 143-145

µ

(Debye)139

Methanol 3.7 8.2 16.9 16.0 0.43 1.70

Ethanol 2.8 7.5 7.9 22.0 0.49 1.69

2-propanol 2.2 6.8 5.8 29.0 0.56 1.68

Butanol 1.7 5.4 2.3 36.0 0.62 1.66

Water ~ 0 4.4 3.2 10.8 0.04 1.85 a . 0maxmax SSS −=∆

b . 0maxmax RRR −=∆

In a separate study, we have investigated the effect of an increase in molecular vapor

pressure on Smax and Rmax for each alcohol. After the values Smax and Rmax were observed from

exposure to vapor pressure in equilibrium with the liquid at 24 ºC, and before any vacuum-

degassing, the bulb containing the alcohol [inset to Figure 6.5(b)] was heated from 24 ºC to a

higher temperature (~ 60 ºC) to increase the vapor pressure. After ~ 30 min of exposure to the

higher vapor pressure, no further changes in S and R were observed. This suggests that the values

-106-

of ∆Smax and ∆Rmax observed in earlier experiments correspond to the response of a maximum

molecular coverage attainable for our bundled SWNT sample at 40 ºC. In effect, our experiments

suggest that the surface saturates at 40 ºC. Also, we should mention that we have no direct

evidence as to what extent the nanotubes are “open” or “closed” at their ends, although step (1)

in the nanotube purification process is expected to open the tubes. Furthermore, all the molecules

investigated in this work satisfy the inequality DK > dI, where DK is the kinetic diameter of the

molecule and dI is the diameter of a typical interstitial channel (dI ~ 0.21 nm for (10,10) tube

bundles). This suggests that the molecule cannot easily enter the channel, unless the bundles

swell to accommodate these molecules. However, they are all small enough to enter an internal

pore of a (10,10) or larger tube, if the tube end is open, or if a large hole is present in the tube

wall.

Figure 6.5(b) shows the time-evolution of the normalized four-probe resistance. The data

for each molecule type were taken concurrently with the thermoelectric power data in Figure

6.5(a). The trends for ∆Rmax vs. n for the alcohols (CnH2n+1OH; n = 1-4) match those observed

for ∆Smax [Figure 6.5(a)], i.e., exposure to methanol shows the largest change in R, with an

increase of ~ 8.2 %. Ethanol, isopropanol, butanol and water induce an increase in R, with ∆R/R0

saturating at 7.5 %, 6.8 %, 5.4 %, and 4.4 %, respectively. As can be seen from the fits in Figure

6.5, both R(t) and S(t) exhibit a simple exponential behavior, as described by Eq. (6.4). The time

constants obtained from the fits to R(t) are all in excellent agreement with those obtained from

the fits to S(t) (Table 6–3). According to simple molecular kinetic theory, the diffusion time

should be proportional to the square root of the molecular mass, i.e., M~τ . However, the

time constants obtained in our study do not exhibit any systematic dependence. This result

indicates that the rate limiting step may not be ordinary diffusion, but perhaps the success rate to

-107-

enter the tube pore through an open end. In a computational study of molecular diffusion through

carbon nanotubes, Mao and Sinnott146 have shown that the intermolecular and molecule-

nanotube interactions strongly affect the molecular diffusion ranging from normal mode

(individual molecules can pass each other within the pore) to single-file diffusion (individual

molecules cannot pass each other in the pore due to their large size relative to the pore diameter).

Table 6–3. Adsorption time constants for thermoelectric ( Sτ ) and resistive ( ) response of a SWNT thin film to adsorbed water and alcohol molecules.

Molecule Rτ (min) Sτ (min) 2

SR τ+τ=τ (min)

OH

alcohol

2M

M

Methanol 11.7 10.1 10.9 ± 0.8 1.33

Ethanol 15.1 16.0 15.6 ± 0.5 1.60

2-propanol 11.3 12.6 12.0 ± 0.6 1.83

Butanol 9.1 10.5 9.8 ± 0.7 2.03

Water 9.8 – 9.8 1

The increase in ∆Rmax is identified with an additional impurity scattering of conduction

electrons in metallic tubes within the bundles due to physisorbed molecules. This will be

discussed in detail later. Interestingly, when exposed to water vapor, the resistance of the SWNT

films increased by ~ 4.4%, even though the thermoelectric power was constant and equal to its

initial degassed value. Although we see no change in the thermoelectric power (∆Smax = 0) for

H2O, in agreement with Bradley et al.,88 we do see a strong response and saturation in R for the

same exposure to H2O. This result is in contrast to the results of Zahab et al.,147 who have

reported an initial increase of resistance of the SWNTs when exposed to water vapor, with an

-108-

eventual crossover to a decrease of resistance for increasing exposure, reaching a resistance

value lower than the starting value. We have not observed this crossover in three separate studies

of H2O/SWNT systems. Furthermore, Zahab et al.147 have interpreted their results on the basis

that the outgassed SWNTs are p-type semiconductors and water molecules act as compensating

donors. It is difficult to speculate about the different behavior on R(t) observed in our samples

with respect to Zahab et al.’s samples. We do note, however, that they have initially degassed

their sample at 220 ºC in a vacuum of 3×10-6 mbar for only 5 h. According to our experiments,

this may not be sufficient time to remove all the weakly chemisorbed oxygen. We also do not

know if they have annealed their samples at 1000 ºC as we have done. In our work, we have

monitored S and R vs. t during vacuum-degassing and have waited for an exponential approach

to a lower plateau in S(t) and R(t) before exposing the sample to a particular vapor for study.

Previous studies on the thermoelectric power behavior of SWNT films have been found

to be consistent with a diffusion thermoelectric power dominated by metallic tubes in a rope.108

Recently, a broad peak in S(T), observed below 100 K and superimposed on a linear T

background, has been attributed to an additional contribution from phonon drag.108,148 The reader

is referred to section 3.3 for further details. As our measurements in this study were made at T =

40 ºC, we ignore a phonon drag contribution which is a low-temperature effect. We have

discussed the connection of S and R in metallic tubes in section 6.2 above.

-109-

-4

-3

-2

-1

0

1

2S

(µV

/K)

8x10-26420∆R/R0

n = 1

n = 2

n = 3

n = 4

CnH2n+1OH

H2O

Figure 6.6. S vs. ∆R/R0 plots during exposure of degassed SWNT bundles to water and CnH2n+1OH (n = 1-4). The solid lines are linear fits to the data until saturation is established.

Figure 6.6 shows the evolution of the thermoelectric power versus the fractional change

in the four-probe resistance (∆R/R0) at fixed temperature (40 ºC). As the coverage of the

molecules on the SWNTs increases with increasing exposure time to the respective molecular

vapor, both S and 0RR∆ increase. It is very important to note that the data for all the alcohols

show linear behavior for S vs. ∆R/R0, consistent with Eq. (6.1) (i.e., S ~ ρI for ρI << ρ0).136 It

should be noted that the Fermi level EF is kept constant in the derivation of Eq. (6.1). Therefore,

this result [Eq. (6.1)] is appropriate for physisorption and NOT for a chemisorption process

involving significant charge transfer. Thus, the linearity of S vs. ∆R/R0 implies that little or no

-110-

charge transfer is taking place between the adsorbed molecules and the SWNTs, i.e., H2O and

the alcohols that are physisorbed onto high-T annealed films do not chemically dope the SNWTs.

In our previous study on the effects of physisorption of six-membered ring molecules (C6H2n; n =

3-6) on SWNTs, we have found a slightly non-linear behavior of S vs. ∆R/R0 data (Figure 6.3).

This non-linear character in the C6H2n/SWNT system is not well understood and we have

tentatively identified it with a multiple electron scattering process.149

From Figure 6.5, it seems that the physisorbed behavior of water on the surface of carbon

nanotubes is markedly different from that of the alcohols. Adsorption of strongly polar molecules,

such as water vapor, is thought to occur by hydrogen bonding on graphitic surfaces and on

carbon nanotubes,145 but it appears that the predominant interaction for all alcohols is the relative

contribution (i.e., van der Waals) contribution from the alkyl chains, which increases with alkyl

chain length.150 In fact, H2O has a behavior different from all the molecules we have studied, i.e.,

a zero response of the thermoelectric power and yet a normal resistive response. At this time, all

we can conclude is that Eq. (6.1) may hold the answer for our observations (i.e., ∆Smax ~ 0),

although we do not have a microscopic model for the scattering mechanism required to apply Eq.

(6.1).

From a similar study on the effects of physisorption of C6H2n family of molecules on S

and R for bundled SWNTs, we were able to correlate the strengths of the thermoelectric power

and resistive responses to the molecular adsorption energy on graphitic surfaces.149 In the former

case,149 all the C6H2n molecules have adsorption energy Ea that is related to the number of π

electrons on the molecule and is therefore a measure of the coupling of the molecule to the

nanotube surface. Ea was then presumed to be a measure of the perturbative interaction of the gas

molecules on the nanotube wall potential, responsible for the enhanced electron scattering rate.149

-111-

In the present study, all the molecules share a dipolar character, but have different projection

area A (see Table 6–2). Furthermore, we have found that the surface appears saturated. We

presume that the scattering rate w, and thus ρI in Eq. (6.1), is related to the product of the

molecular coverage ξ and the adsorption energy, i.e.,

, ~ a ξEw (6.5)

where ξ is the areal density of physisorbed molecules on the nanotube surfaces. We furthermore

expect that jA1~ξ , where Aj is the projection area of the particular molecule j. Thus, from Eqs.

(6.1) and (6.5) we expect that AES I amax ~βρ~ β∆ (β is the slope of the S vs. ∆R/R0 straight

lines in Figure 6.6), i.e., the maximum change in S is proportional to the adsorption energy and

inversely proportional to the projection area of the molecule. Yang et al.151 have recently studied

the adsorption of butanol and methanol on HiPCO (High-Pressure Carbon Monoxide Synthesis)

SWNTs at 30 ºC, and have found that the number of adsorbed moles of molecules of butanol per

unit weight is smaller than that of methanol. This result has been identified with the difference in

molecular volumes.151 The explanation should be equivalent to one involving molecular

projection areas.

Thus, in an attempt to explain the systematics of ∆Smax against the molecular properties,

we have plotted ∆Smax versus the quantity AEaβ in Figure 6.7. Interestingly, all the data fall on

a quasi-linear curve, motivating the concept that the extra electron scattering in the nanotube

wall due to physisorption is proportional to the product of the adsorption energy and the

molecular coverage. The curvature of this quasi-linear curve at high AEaβ could be an

indication of the saturation of the thermoelectric and resistive responses. On the other hand,

methanol might be too small to be expected to follow the linear trend established for butanol,

-112-

isopropanol, and ethanol in Figure 6.7 (dashed line).

4

3

2

1

0

∆Sm

ax (µ

V/K

)

1.61.20.80.40.0βEa/A (eV/Å2·µV/K)

CnH2n+1OH

H2O

n = 4n = 3

n = 2

n = 1

Figure 6.7. Maximum thermoelectric power change ∆Smax of a SWNT thin film successively exposed to vapors of water and alcohol molecules (CnH2n+1OH; n = 1-4) as a function of the quantity AEaβ , where Ea and A are, respectively, the molecular adsorption energy and the projection area. The solid and dashed lines are guides to the eye.

In conclusion, we have utilized in situ measurements of the thermoelectric power and

electrical resistance to investigate the adsorption of various polar molecules (alcohol and water)

in bundled SWNTs. We observe a strong effect on both the thermoelectric power and electrical

resistance for methanol, ethanol, isopropanol, and butanol. Surprisingly, water vapor does not

have any effect on the thermoelectric power, i.e., ∆Smax ~ 0, but has a significant impact on the

resistance, i.e., (∆R/R0)max ~ 4.4%. The fact that ∆Smax ~ 0 may be due to a fortuitous cancellation

-113-

of scattering terms in Eq. (6.1). We have also observed that S exhibits a linear relationship with

∆R/R0, consistent with creation of a new impurity scattering channel via physisorption, and that

the slopes of the S vs. ∆R/R0 data are specific to the particular molecules. In an effort to correlate

what we have observed with molecular properties, we have found that, for water and the C1−C4

alcohols, the maximum change in the thermoelectric power is proportional to the product of the

molecular adsorption energy (measured on graphitic surfaces) and the molecular coverage A1~ ,

where A is the molecular projection area on the host surface.

-114-

Chapter 7.

Effects of Gas Collisions on the Transport Properties of Carbon

Nanotubes

7.1. Introduction

Recently, much attention has been focussed on the problem of gas adsorption within

bundles of carbon nanotubes, as evidenced by the wealth of theoretical and computer

calculations studies on the physisorption of rare gases152-159 and methane160 in SWNTs. Phase

transitions, capillary condensation, adsorption capacities, and effects of dimensionality have

been investigated over a range of tube radii and temperatures. Classical and path integral

molecular simulations have also been used to study physisorption and fluid dynamics of

helium,135,161,162 neon,163,164 argon,135,164-166 xenon,163,164,167 krypton,166 methane,163,164,168 and

nitrogen165,169 in SWNTs and SWNT bundles for a range of pressures, temperatures, tube radii,

and bundle structures. Adsorption capacities and molecular density distribution in different

adsorption sites have been reported in these studies. Experimental studies have dealt with

adsorption of helium,170-173 neon,171,174,175 argon,171,176 xenon,171,174,175,177-179 kypton,171,180

methane,174,175,180-184 and nitrogen.111,185 Adsorption and storage of hydrogen on carbon

nanotubes have also been studied extensively.133,186,187

There are only a few reports in the literature on collisions of atoms or molecules with

carbon nanotubes.188,189 The importance of these studies stems from the fact that production and

growth of carbon nanotubes often take place in gas environments at elevated temperatures.190 For

-115-

instance, inert atmospheres (helium and argon in most cases) have been used for preparation of

nanotubes by the arc-discharge method.191,192 Methane, hydrogen and nitrogen atmospheres have

also been used to grow carbon nanotubes.193-196 In all cases, the quality, yield, and growth rate of

nanotubes depend sensitively on the gas environment and the pressures.

In this chapter, the results of a systematic study of the effects of collisions of inert gas

atoms, CH4, and N2 on the electrical transport properties of SWNTs are presented. We have

observed unusually strong and systematic changes in the electrical transport properties in

metallic SWNTs that are undergoing collisions with inert gas atoms. At a fixed gas temperature

(~ 500 K) and pressure (1 atm), the changes in the resistance and the thermoelectric power are

observed to scale as ~ M1/3, where M is the mass of the colliding gas atom (He, Ar, Ne, Kr, Xe).

The results of molecular dynamics simulations carried out in collaboration with Göteborg

University and Chalmers University of Technology are also presented here. They show that the

radial energy transfer between the colliding atom and the nanotube also exhibits ~ M1/3

dependence. A significant transient population of low-frequency optical phonons is observed to

stem from a single collision of an atom with the nanotube wall. These long-lived vibrations may

provide a new scattering mechanism needed to explain the collision-induced changes in the

electrical transport.

7.2. Collision-Induced Electrical Transport of Carbon Nanotubes

Thermoelectric power and four-probe resistance measurements were carried out on

samples in the form of thin films of bundled nanotubes (CarboLex, Inc.; arc-discharge method

(ARC)) and purified “buckypaper” (Rice University; pulsed laser vaporization (PLV)).66 The

-116-

arc-material was also purified,44 and thin films were prepared by deposition of a sonicated

ethanol solution containing purified SWNT bundles onto a warm (~ 50 ºC) quartz substrate. The

films and buckypaper samples (Rice University) were vacuum annealed at ~ 1000 °C for 12 h

before attaching thermocouples (chromel-Au/7 at% Fe) and electrical (copper) leads with silver

epoxy to four corners of the sample for the TEP and resistance measurements. The 2 mm × 2 mm

specimens that contained ropes of 10’s to 100’s SWNTs of 1.0-1.6 nm in diameter and several

microns long were placed in a turbo-pumped vacuum chamber (~ 10-7 Torr) where transport

measurements were made in situ in the presence of various gases. The gases (e.g., Ar, N2, etc.)

were first passed through a purification cartridge (OT-4-SS, R&D Separations, Inc.) to remove

residual O2 and H2O. Details of the electrical measurements are available in Chapter 2.

Before collecting data, the samples were first vacuum-degassed in situ at 500 K to

remove adsorbed oxygen and water. During the vacuum-degassing process, the thermoelectric

power S was observed to decrease slowly over several days from a positive initial value, change

sign, and then asymptotically approach a negative value representative of the “degassed state” S0.

This behavior is in agreement with previous results on similar mats or film samples.89,90,108 For

the interpretation of the “degassing” effects on S and R see Chapter 3.

-117-

-12

-8

-4

0

4

8

12

S (µ

V/K

)

302520151050Time (h)

He

Ar

Xe

CH4

N2

S0 Arc SWNTs

(a)

(b)

-50

-45

-40

-35

-30

-25

S (µ

V/K

)

24201612840Time (h)

He

Ne

Ar

Kr

Xe T = 500 K

S0 PLV SWNTs

Figure 7.1. Time dependence of the thermoelectric power response of (a) PLV buckypaper and (b) arc-derived thin film exposed to 1 atm of inert gas (closed symbols), and to subsequent application of vacuum over the sample (open symbols) at T = 500 K. The different values of S0 in (a) and (b) reflect differences in defect densities in the PLV and the arc-derived material (see Chapter 5).

-118-

In Figure 7.1 we show the reversible thermoelectric response of a vacuum-degassed PLV

buckypaper (a) and an arc-derived thin film (b) to sequential exposure of a sudden pressure (p =

1 atm) of various gases at T = 500 K. The samples were degassed in vacuum at 500 K between

successive exposures to the various gases. The data from exposure to the series of gases have

been superimposed in Figure 7.1. The thermoelectric power of the degassed state is indicated as

S0. The difference between the S0 values (S0 ~ – 11 µV/K: ARC; S0 ~ – 45 µV/K: PLV) for the

two samples depends on the different concentrations of tube wall defects and the

functionalization introduced during growth or post-synthesis (acid) purification.108,132 The

reversible response on exposure to the various inert gases and the molecular gases N2 and CH4 is

also shown in Figure 7.1(b) for the arc-derived film sample. Under vacuum, S can be seen to

return slowly to the degassed value S0. The long time constants for ∆S > 0 and ∆S < 0 of the

system, which also depend on the mass of the gas atoms/molecules, are identified with the slow

diffusion of gas into, and out of, the pore structure of the SWNT bundles (i.e., into and out of

interstitial channels between tubes and also the internal pores of the tubes). We emphasize that

the same sample was sequentially exposed to a series of gases, so that the relative response of a

single sample to each gas could be observed. As can be seen in Figure 7.1, the effect of exposure

to each gas (1 atm, 500 K) on S is fully reversible. That is, the system can be returned to the

degassed value S0. Furthermore, it is clear that the maximum change in thermopower ∆Smax

increases with the mass M of the colliding gas species.

-119-

-50

-45

-40

-35

-30

-25

-20

-15S

(µV

/K)

0.200.150.100.050.00∆R/R0

He

Ne

Ar

Kr

Xe

20

15

10

5

0

∆Sm

ax (µ

V/K

)

2.01.61.20.80.40.0p (atm)

He

Xe

Ar

Figure 7.2. S vs. ∆R/R0 plots showing the effect of inert gases on the transport properties of a SWNT buckypaper prepared from PLV material. The closed symbols are from the time evolution of S and R to 1 atm of gas at T = 500 K and the open symbols are from a pressure study at the same temperature, where the maximum response of S and R to a given pressure was measured. The inset shows the pressure dependence of the maximum change of thermopower for the same sample.

Figure 7.2 shows the thermopower vs. the fractional change in the four-probe resistance

-120-

∆R/R0 for the PLV buckypaper sample at 500 K. 0RRR −=∆ represents the “extra resistance”

due to the colliding atoms and R0 is the initial sample resistance in the degassed state. In the

experiments, both S and R were measured simultaneously as they evolve with time. R (not

shown) was found to exhibit the same time behavior as S, and therefore a linear relationship

between them was observed, i.e., S ~ ∆R/R0. The data plotted as S vs. ∆R/R0 in Figure 7.2 show

that the slope is related to the mass M of the particular gas. The data represented by open circles

were taken with a gas pressure p = 1 atm in the chamber and the closed circles refer to data taken

at various pressures in the range 0 < p < 2 atm (p is a measure of the collision frequency of the

atoms with the nanotube walls). Since both sets of data (open and closed symbols) fall on the

same line for a particular gas, it is clear that the slope of the lines S vs. ∆R/R0 in Figure 7.2

depends on the mass of the gas atom/molecule and not on the chamber pressure. As discussed

previously, the thermoelectric power of bundles of SWNTs should be dominated by the metallic

tubes.108 Furthermore, we have shown on the basis of Boltzmann transport theory that the linear

relationship between S and ∆R/R0 is consistent with the creation of a new scattering channel for

the conduction electrons in the metallic tubes, provided that the nanotube Fermi energy EF

remains constant.136 In the data presented here, the scattering channel is identified with gas

collisions with the tube walls. The variation in slope with mass M observed in Figure 7.2

suggests that the impulse delivered to the tube wall per collision may be an important variable.

The inset to Figure 7.2 shows ∆Smax vs. p for selected inert gases (He, Ar, Xe), where

∆Smax is the maximum change of the thermopower measured relative to the degassed state (S0).

We see that ∆Smax saturates with pressure at relatively low pressure (~ 1 atm) and that the

saturation value depends on the gas species (e.g., He, Ar, Xe). A related saturation of the

electrical resistance (not shown) was also observed. Resistivity saturation phenomena are well

-121-

known in solid state physics, and several reviews197,198 have been published on this topic. The

basic idea usually invoked is that resistivity saturation is associated with a minimum mean free

path for the conduction electrons. This saturation can be associated with many scattering

mechanisms and it has been concluded that resistivity saturation usually occurs when the

electron mean free path approaches the interatomic spacing in the material.197,198 In the inset to

Figure 7.2, the saturation with pressure may be due to a limiting mean free path resulting from

increased collisions of the gas atoms with the nanotube wall and might represent the pressure at

which neighboring transient deformations in the same tube begin to overlap.

1.1. Molecular Dynamics Simulations

Because of the extreme flexibility of the nanotube wall, it is interesting to consider what

wall deformations might be generated by the collisions of gas atoms or molecules with the tube

wall. We have used molecular dynamics simulations to study this question.189 Here we have

considered the effect of approaching Xe, Ar, Ne, and He atoms on a finite length (10,0) carbon

nanotube at 0 K. The (10,0) nanotube is semiconducting and has a diameter of 7.83 Å. It is

sufficiently small to facilitate the simulation of a large number of scattering events required for

statistical analysis. Calculations were made at 0 K so that complications from the thermal

phonon background can be eliminated. Phonons generated in the “dent” created by a gas

collision will propagate slowly away from the collision site and are absorbed in the “thermal

reservoirs” at the tube ends. This was done by scaling the velocity of the carbon atoms at the

tube ends to zero at each trajectory time step. In this way, the energy flowing along the tube axis

direction was adsorbed, but not energy that flows along the tube radial or circumferential

-122-

directions. The phonons are primarily low frequency (q = 0) optical modes that have almost zero

group velocity. For the dent to “diffuse”, it may be necessary for these phonons to decay to two

oppositely directed acoustic phonons. Further details of the molecular dynamics methods are

described elsewhere.189 To study the transient character of the tube deformations (or phonons)

induced by a collision, the displacements of the carbon atoms in a short (400 atoms) nanotube

were followed over time.

Figure 7.3 shows the power spectrum of the radial C-atom motion generated by collisions

of Xe, Ne, and He during (a) the first 5 ps of the collision–which includes the gas-nanotube

impact–and (b) during the second 5 ps. The C-atom monitored was the one closest to the impact

site, but the features of the power spectra are insensitive to the C-atom chosen. The colliding

atoms were incident at θi = 45º to the tube surface normal and with an initial energy of Ei = 13

kcal/mol (θi = 0º is an exactly radial trajectory). The power spectra, shown in Figure 7.3, are

weighted so that the area under the spectrum equals the average total energy of the carbon

nanotube. The average vibrational energies of the carbon nanotube during the first 5 ps of the

collision (see Figure 7.3(a)) are 4.2, 3.7, and 2.7 kcal/mol for Xe, Ne, and He, respectively.

During the next 5 ps interval (see Figure 7.3), we find that the nanotube still retains 2.0, 2.2, and

1.8 kcal/mol for Xe, Ne, and He, respectively.

-123-

Inte

nsity

Xe Ne He

(a)In

tens

ity

5004003002001000Frequency (cm-1

)

Xe Ne He

(b)

Figure 7.3. Computed power spectra of the radial motion of a C-atom nearest the point of contact in a (10,0) carbon nanotube at 0 K. The figure shows the phonons induced during (a) the first 5 ps of the collision (and includes the gas-tube impact) and (b) the second 5 ps after the collision. The inset to (a) shows the side view of a collision between a Xe atom (θi = 0º, Ei = 13 kcal/mol) and a nanotube. The inset to (b) shows the schematic representation of the tube wall deformation in response to an atom collision.

-124-

Clearly the most prominent peaks in the power spectra increase in strength with the mass

M of the colliding atoms. Furthermore, snapshots of the tube wall motion in time show that the

gas atom impact locally flattens one side of the tube, and then the tube wall near the impact

begins to “ring” or oscillate in an elliptically-shaped deformation identifiable with the Raman-

active “squash” mode (E2g), and the frequency we observe (~ 43 cm-1) is in reasonable agreement

with the calculated squash-mode frequency for a (10,0) nanotube.50 Saito et al.50 have reported

on the strong diameter dependence of the squash mode. Using their calculated squash-mode

frequency for a (10,10) nanotube, we estimate that the squash-mode frequency for a (10,0) tube

is 29 cm-1. The short length of the nanotube, clamped at the ends, may upshift the squash-mode

frequency relative to that obtained for an infinite tube.

Schematics of the tube wall deformation prior to, during, and after collision are shown in

Figure 7.3(b). The side view of a collision is shown in Figure 7.3(a). We find that the energy

dissipation from these squash modes is slow, i.e., the ringing takes place for a long time. These

optical modes have zero group velocity, and for the energy to propagate away from the collision

site, the optical squash mode must first decay into two acoustic modes. In all cases, however, our

simulations show that a significant amount of energy remains near the collision site for 10 ps

after the collision.

Three features from the power spectra in Figure 7.3 are worth mentioning. First, Xe

imparts more energy to the carbon nanotube at the time of impact than does Ne, which, in turn,

imparts more energy than He. However, the strong mass dependence we observe for the

transferred collision energy is lost soon after impact. Second, the low frequency vibrations at ~

43 cm-1 (squash mode) dominate the vibrational power spectrum, and the energy (amplitude) of

this mode is significantly larger after collision with Xe than after collisions with lighter atoms,

-125-

e.g., Ne and He. Third, there are fewer peaks in the spectra obtained 5-10 ps after collision, than

in the spectra obtained over the collider-nanotube impact. This shows that (i) either the energy in

the short-lived transient phonons created upon impact flows into the squash-mode phonons, or

(ii) the energy in these transient phonons flows axially along the tube and, in the simulations, is

lost preferentially to the tube ends.

1

2

3

456

10

2

3

∆Sm

ax (µ

V/K

)

12 4 6 8

102 4 6 8

1002

M (g/mol)

1

2

4

6

810

2

4

∆E (kcal/mol)

3

456

0.1

2

3

456

1

Dm

ax (Å)

He NeAr

Kr

Xe

Ne

He

Xe

CH4

N2

Ar

∆Smax

∆E

Dmax

Figure 7.4. Maximum thermoelectric power change ∆Smax of two SWNT samples exposed to gases indicated (ARC: open circles and PLV: closed circles; data from Figure 7.1), calculated total energy gained by a (10,0) nanotube upon collision with a gas atom (θi = 0º, Ei = 3.97 kcal/mol, squares), and maximum radial displacement ∆Dmax of the tube C-atom immediately after impact with a gas atom (θi = 45º, Ei = 1.99 kcal/mol, triangles) as a function of the mass of the colliding inert gas. The lines are power law fits to the data of the forms , , and . 35.0

max 08.3 MS =∆ 39.091.0 ME =∆ 35.0max 04.0 MD =∆

Our simulation studies on collisions between the carbon nanotube and Xe atoms (θi = 0º,

Ei = 13 kcal/mol) suggest that all power spectra (e.g., Figure 7.3) should be relatively insensitive

-126-

to the incident angle θi of the colliding species. The sole difference we observed is that more

total energy is transferred at smaller incident angles. This is expected, since a smaller incident

angle corresponds to a larger radial component of the gas kinetic energy. Similarly, simulations

of collisions between a Xe atom and a carbon nanotube, where the incident angle was 45º and the

initial energies were 0.40 kcal/mol (200 K), 0.99 kcal/mol (500 K), and 1.99 kcal/mol (1000 K),

showed that the features of the power spectra are insensitive to the initial kinetic energy. The

only difference we observed is that more total energy is transferred to vibrational motion from

higher kinetic energy colliders.

In Figure 7.4, we consider the mass (M) dependence of the colliding atoms or molecules

on ∆Smax (1 atm, 500 K, c.f. Figure 7.1). The straight solid lines in Figure 7.4 represent least-

squares fits to the experimental ∆Smax vs. M (circles) data. Both solid lines exhibit a ~ M1/3

dependence. The closed and the open circles represent the data for the buckypaper and arc-

derived sample, respectively. For comparison, as observed in our molecular dynamics

simulations, we plot the mass dependences of the total energy transfer ∆E (squares) to a short

(10,0) nanotube upon collision (i.e., the total energy lost by the colliding atom), as well as the

maximum radial displacement ∆Dmax (triangles) of the tube C-atom (with which the gas atom

collides) on impact. We find that the slopes of these theoretical lines are insensitive to the

simulation conditions (i.e., incident angle, initial energy, and point of impact). As can be seen in

the figure, the perturbation of the transport properties, the total energy gained by the tube, and

the maximum amplitude of the dent obtained from our simulations all share the same

approximate power law dependence on the collider mass (i.e., M1/3). We also found that the slope

of the theoretical lines of the average tube energy and the maximum radial displacement as a

function of the collider’s mass decreases with time. The polarizability α of inert gases also

-127-

follows a power law dependence with the mass (see Figure 7.5).‡ However, data from small

molecules (N2, CH4) deviate from the power law trend.

0.1

2

4

68

1

2

4

68

10

α (Å

3 )

12 3 4 5 6

102 3 4 5 6

1002

M (g/mol)

α = 0.083×M0.79

HeNe

Ar

Xe

KrN2CH4

Figure 7.5. Dipole polarizability α as a function of the mass of the inert atom or small molecule.

In conclusion, we have observed remarkable experimental effects of collisions of inert

gas atoms (He, Ne, Ar, Kr, and Xe) and small molecules (N2 and CH4) on the thermopower S and

the resistance R of samples of tangled carbon nanotube bundles. At 500 K and 1 atm, the

maximum changes in S and R are proportional to ~ M1/3, where M is the mass of the colliding

species. Our simulations also revealed that the energy exchanged between the tube wall and the

‡ The polarizability of an atom or molecule describes the response of the electron cloud to an external

electric field, E. The induced electric dipole moment is Eα .

-128-

colliding atoms, as well as the maximum deformation of the tube wall as a result of this collision,

exhibits an approximate M1/3 behavior. We propose that the effects we observe in S and R are

due to a new scattering channel for conduction electrons created by collision-induced transient

dents in the tube wall. The pressure saturation of the changes in the transport parameters (R and

S) may be due to the eventual overlap of adjacent dents.

-129-

Chapter 8.

Conclusions and Future Work

Single-walled carbon nanotubes (SWNTs) are now widely accepted as promising

candidates for nanoscale electronic devices on account of their unprecedented electronic

properties. SWNTs provide a unique opportunity to study the interaction of molecules with

conducting surfaces. This stems from the unique structure of the nanotube, i.e., a monolayer

sheet of sp2-bonded carbon (graphene) that is rolled into a small diameter (0.4-2 nm) seamless

cylinder. Since all the carbon atoms reside at the tube surface, the chemical environment in

contact with the nanotube should be expected to influence the transport of electrons in the tube

wall. In addition, although quite strong in tension, the cross-section of a SWNT is easily

deformed, and gas collisions should also be expected to affect the transport of electrons in the

tube wall. This thesis has examined the effects of gas interactions (adsorption and collisions) on

the electrical transport properties (electrical resistance and thermoelectric power) of random

networks of bundled SWNTs.

It has been shown that the measured electronic properties of nanotubes are very sensitive

to their chemical environments. In fact, exposure to air or oxygen (at room temperature)

dramatically influences the electrical resistance and the thermoelectric power of nanotubes.

SWNT samples consisting of randomly oriented bundles of hundreds of nanotubes exhibit a

gradual crossover of thermopower from a large positive value to a large negative value, as

adsorbed oxygen is removed from the sample in high vacuum at elevated temperatures (~ 500 K),

in agreement with previous results.89,90 In addition, we have observed that the thermoelectric

power can be reliably and reversibly “tuned” simply by controlling the amount of oxygen

-130-

adsorbed on the tube walls at low temperatures. Interestingly, independent of the amount of

oxygen adsorbed, the thermoelectric power retains the typical behavior of purified, macroscopic

samples: a dominant quasi-linear component with a superimposed knee or smooth change of

slope at ~ 100 K. The quasi-linear behavior suggests that metallic tubes in the sample dominate

the thermoelectric power. To explain these results, a weak charge transfer between the O2

molecules and the SWNTs to form a complex where δ ≤ 0.1, has been proposed.

Model calculations of the thermoelectric power has been presented which show that, in the

general case, the thermoelectric power is determined by coordinated effects in both the

semiconducting and the metallic nanotubes in the SWNT rope. That is, the sign and magnitude

of the thermoelectric power can be determined by the near mirror symmetry bands in metallic

nanotubes and density of states resonances near the Fermi energy which, in turn, are determined

by the balance between acceptor (charged species that can be removed by vacuum-

degassing) and donor states on the semiconducting nanotubes, possibly due to wall defects.

δδ+ − -2p OC

δ-2O

Before one can hope to fully rationalize the aforementioned effects of exposure to

molecular oxygen on the electronic transport properties of SWNTs, the phenomenon of O2

adsorption needs to be better understood. Specifically, the issue of whether O2 molecules are

physisorbed or chemisorbed on the SWNT walls needs to be resolved. In addition to resolving

this controversy, one should be able to model the effects of O2 adsorption on the electrical

conductance of SWNTs and to include potential effects of the contacts between SWNTs and

leads attached to their ends during experimental measurements. The importance of the sensitivity

of carbon nanotubes to O2 need not be over-emphasized, as it raises questions about the stability

of carbon nanotube devices to air exposure. On the other hand, the observed effects of O2

exposure on the carbon nanotubes properties raises the possibility that unintentional oxygen

-131-

contamination during sample preparation may lead to incorrect analysis of the experimental data.

It has been suggested that O2 in proximity to a wall defect may lead to the formation of the

charge transfer complex. Further work is needed to prove this hypothesis.

The second contribution to the “intrinsic” thermoelectric power we have observed is in

the form of a knee or a broad peak at ~ 100 K (superimposed on a linear, metallic background).

We propose that this peak should not be identified with a Kondo anomaly, as proposed

previously,76 but rather should be attributed to the phonon drag effect. To calculate the phonon

drag contribution, a variational solution to the coupled electron-phonon Boltzmann equations

was used.119 The calculation also involved the use of a simple tight-binding model for the

electronic structure of metallic armchair tubes, a linear acoustic phonon dispersion, and a

temperature-independent phonon relaxation time. The relevant phonon wave vectors contributing

to the phonon drag effect are determined by the Fermi level position which, in turn, is

determined by the oxygen content in the SWNTs. This model describes very well the unusual

behavior of the experimental thermopower in purified and as-grown SWNT materials, as long as

the broad peak is not so pronounced, as in the case of SWNT material grown and containing Co-

Y and Fe-Y catalysts. For these Co-Y and Fe-Y samples, quantitative agreement with our simple

theory is lacking, but the discrepancy may be reconciled if an appropriate temperature-

dependence of the phonon lifetime is incorporated into the model.

The transport properties of SWNTs have also been shown to be most sensitive to gases

that chemisorb and undergo weak charge transfer reactions (e.g., O2 and NH3) with the

nanotubes. Surprisingly, the transport parameters (thermoelectric power and electrical resistance)

haver even been found to be sensitive to physisorption of gases and chemical vapors (e.g., H2,

alcohols, and cyclic hydrocarbons) and gas collisions with the carbon nanotube walls (e.g., inert

-132-

gases, N2, CH4). For polar molecules (alcohols and water), the trends in the measured

perturbations of the transport properties could be explained on the basis of the interplay between

the adsorption energy and the molecular coverage on the nanotube surfaces. Interestingly, the

thermoelectric power response to water vapor was found to be very weak, although the resistive

response to water vapor exposure was similar to that to simple alcohols exposure. For six-

membered-ring molecules, the magnitudes of the changes in the transport properties were found

to be related to the π electron population of the molecule, suggesting that the coupling between

these π electrons in the adsorbed molecules and those in the nanotube wall may be responsible

for the observed effects. In our studies on gas collisions (He, Ne, Ar, Kr, Xe, N2, CH4) on the

SWNT surface, the unusually strong and systematic changes in the transport properties have

been found to scale as ~ M1/3 (M is the mass of the colliding gas atom or molecule)–the same

dependence exhibited by the total energy transfer to the nanotube and the maximum deformation

of the tube upon collision. Molecular simulations by Kim Bolton and Arne Rosén (Göteborg

University and Chalmers University of Technology) have supported the idea that a significant

transient population of low-frequency phonons (squash-mode phonons) is created upon collisions,

which could then provide the new scattering mechanism for the conduction electrons in the tube

walls. This could be direct atom-electron scattering, but the observed M1/3 dependence of the

wall deformation is compelling evidence for deformation scattering.

The mechanism we have proposed to explain the thermoelectric changes in SWNTs upon

interaction with gases and chemicals is the creation of an additional scattering channel for

conduction electrons in the nanotube wall created by physisorption or gas collisions with the tube

wall. The expression developed for this model follows from the Mott relation for the

thermoelectric power of metals and the Matheissen’s rule, which is equivalent to the additive

-133-

nature of independent scattering mechanisms. The same expression was developed previously to

discuss the thermoelectric behavior of Au and Ag alloys, and is known as the Nordheim-Gorter

relation. We believe that the sensitivity of Nordheim-Gorter plots at a fixed temperature to

different molecules can be the basis for the utility of a SWNT thermoelectric “nano-nose”. In

fact, the sensitivities of S and R to coverage must be related to the quasi-one-dimensional nature

of the transport in SWNTs, and to the fact that almost all carbon atoms are associated with one

adsorption site or another. Detailed theoretical calculations to investigate the scattering

mechanisms for various adsorption sites and molecules are now needed. These calculations

would allow understanding why, unlike other gases (e.g., N2, He, etc.), exposure of the

nanotubes to H2 give rise to negative slopes in the S vs. ∆R/R0 plots. It would be interesting to

study the effects of H2, D2, and HD exposures on the same SWNT sample to determine the

relative contributions from collisions and chemical interactions to the electrical transport

properties.

The purity of SWNT samples is limited by the extensive use of metal catalysts in the

growth process and chemicals during purification and dispersion procedures. Unfortunately, in

spite of accurate protocols for elimination of chemicals, traces may still remain in the sample as

undesired and often unrevealed contaminants, which might interference with SWNT interactions

with adsorbates.109 This indicates the extreme importance of careful determination of the

chemical purity of samples subjected to purification processes and thus unintentionally

contaminated, as the presence of extrinsic contaminant species on the tube walls can mimic

electronic properties different from the intrinsic electronic properties of nanotubes. It is thus

important to investigate the thermoelectric power of functionalized nanotubes. Moreover, the

structure of SWNTs is increasingly altered by the introduction of defects due to the effect of

-134-

chemical treatments with increasing strength.199 Although, structural defects and side openings

can be cured to some extent by subsequent annealing, the effects of different types of defects on

the electronic properties of carbon nanotubes in the presence of gases and chemicals need to be

further investigated.

There are significant disadvantages to our measurements on “bulk” samples for

understanding the intrinsic properties of a single nanotube. One problem is that these

measurements yield an ensemble average over the different tubes in a sample. More importantly,

the numerous tube-tube junctions present in these macroscopic samples make it difficult to

extract absolute values from the transport properties. Mesoscopic scale measurements able to

probe individual nanotubes are necessary in order to elucidate their intrinsic transport properties.

In conclusion, the work presented here emphasizes the effects of gaseous molecules on

the electrical transport of carbon nanotubes. The knowledge acquired may have important

implications on the production and growth, as well as the applications of carbon nanotubes.

Theoretical calculations to investigate the scattering mechanisms for various adsorption sites and

molecules are now needed. As these calculations will have to deal with the details of the

molecule-SWNT interactions, it is hoped that these calculations and the data presented here will

provide quantitative insight into the details of gas-SWNT interactions.

-135-

Appendix A:

Derivation of the Mott Relation

An alternative and general expression for the thermoelectric power S is to write:

( ) , )(1F ε

εεσ−ε

σ= ∫ d

ddfE

eTS (A.1)

where

. )()()(

, )(

εµε≡εσ

εε

εσ−=σ ∫eN

dddf

(A.2)

Here σ is the electrical conductivity, N(ε) is the density of states, µ(ε) is the mobility at a given

energy level ε, and f(ε) is the equilibrium electron distribution function. If the scattering is weak

and the mean free path is long, Eqs. (A.1) and (3.2) give essentially identical results.

Let us consider a general equation of the form,

. )(0∫∞

εε

ε−= dddfhI (A.3)

When only states near EF contribute to the current, εddf will be appreciable only within

a few kBT of EF. In this case, it is very reasonable to evaluate Eq. (A.3) by expanding the

integrand about EF which leads to a rapidly converging series:

, 1

1!0 0

∫ ∑∞

= =⎟⎟⎠

⎞⎜⎜⎝

⎛+⎥

⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛−=

TkE

yn y

n

nn

BF

dyedy

ddy

hdnyI (A.4)

where ( ) TkEy BF−ε= . Since the expansion converges rapidly only if 1F >>TkE B , the lower

-136-

limit of the integral on this integral can be replaced by −∞. The result is

( )

( ) , )2(2

12

, )(

12

12

22

F

F

nC

dhdTkCEhI

nn

n En

nn

Bn

ζ⎟⎠⎞

⎜⎝⎛ −=

⎟⎟⎠

⎞⎜⎜⎝

⎛ε

+=

= =ε

∑ (A.5)

where ξ(n) is the Riemann zeta function.

Keeping only the first terms in the summation,

( ) . 6

)(F

2

22

2

FE

B dhdTkEhI

=ε⎟⎟⎠

⎞⎜⎜⎝

⎛ε

π+=

Let ( )[ ]TkEh BF)()( −εεσ=ε , hence

( ) . )(2)()(2)(

FFF

2

2F

2

2

EBEBBE dd

Tkdd

TkE

dd

Tkdhd

=ε=ε=ε

⎟⎠⎞

⎜⎝⎛

εεσ

=⎥⎦

⎤⎢⎣

⎡ε

εσ−ε+

εεσ

ε

Substituting this result into Eq. (A.1),

( ) ( )

, )(13

)(26

1

F

F

2

22

F

E

B

EBB

dd

eTk

dd

TkTk

EeS

⎟⎠⎞

⎜⎝⎛

εεσ

σπ

=

⎥⎥⎦

⎢⎢⎣

⎡⎟⎠⎞

⎜⎝⎛

εεσπ

σ=

or

. ln3

F

22

E

B

dd

eTkS

⎟⎠⎞

⎜⎝⎛

εσπ

= (A.6)

This is the familiar Mott formula for metallic conduction.

Although Eq. (A.6) has been derived here with the aid of the relaxation time

approximation, the result is generally valid provided the conduction electrons constitute a

degenerate Fermi gas, so that higher-order terms in the expansion (A.5) are negligible.

-137-

Appendix B:

Derivation of the Phonon Drag Thermopower

Simple derivations of the expression for phonon drag thermopower are usually based on

the ideas of “balance forces” and “momentum transfer” between the phonon and electron

systems. Here, we provide an alternative picture based on “current balance” model for the

Seebeck effect, i.e.

, 0total =++= ges JJJJ (B.1)

where Je is the electric current density arising from the maintained temperature gradient

produced by the electron diffusion mechanism, Jg is produced by the phonon drag mechanism

and Js is the current density produced in the steady state to counteract Je + Jg, driven by the

Seebeck voltage.

The Seebeck current can be written as:

, 'EJs σ= (B.2)

where σ is the electrical conductivity and ( ) µ∇−= eEE 1' ; E is the electric field, and µ is the

chemical potential. Hence,

. '

0=

⎟⎠⎞

⎜⎝⎛

∇=

JTES (B.3)

The current Jg is defined in the form

( ) [ , 2

'',''

coll ∑∑ −⎟

⎠⎞

⎜⎝⎛

∂∂

−Ω

=− jj

jjep

g llet

NJkk

kkq

q ] (B.4)

-138-

where is the phonon distribution function, is the mean free path of an electron with

wave vector k in the band j, and Ω is the volume. The term

( )qN jlk

( )coll ep

tN−

∂∂− q is the number of

phonons leaving state q per unit time due to the phonon-electron collision mechanism; [ ]'' jj ll kk −

is the average change in electron mean free path in the direction of the temperature gradient

arising from absorption of a phonon in state q.

We now make the following assumptions:

(1) ( ) ( ) ( ) ; '',;coll allcoll t

Njjt

N

p-e ∂∂

α=∂

∂ qkkqq (B.5)

in other words we introduce the factor ( )'',; jj kkqα which is equal to the fraction of phonon-

electron collisions to all phonon collisions.

(2) A Boltzmann equation for the phonon distribution ( )qN can be used:

( ) ( ) . 0driftcoll all

=∂

∂+

∂∂

tN

tN qq (B.6)

(3) An electron relaxation time treatment is valid:

(B.7) . jjj vl kkk τ=

The conventional treatment of ( )drift

tN ∂∂ q is:

( ) ( ) . 0

drift

TVT

Nt

N∇

∂∂

−=∂

∂q

qq (B.8)

Hence using the three assumptions listed above, Eq. (B.4) becomes

( ) ( ) [ . '',;2

'',''''

0 ∑∑ τ−τα∇∂

∂Ω

−=jj

jjjjg vvVjjTT

NeJ

kkkkκkq

q

kkqq ] (B.9)

Returning to the current balance picture [Eq. (B.1)] and neglecting Je, we finally obtain

-139-

( ) ( ) [ ], '',;2

''

'',''''

0

0

∑∑ τ−τα∂

∂Ωσ

=

∇−=⎟

⎠⎞

⎜⎝⎛

∇=

=

jjjjjj

B

e

g

Jg

vvVjjT

NkeEJTJ

TES

kkkkkkq

q

kkqq (B.10)

and assuming an isotropic material,

( ) ( )[ , '',;2

'',''''

0q

kkkkkk

q

Vvvkkqq⋅τ−τα

∂∂

σΩ= ∑∑

jjjjjjg jj

TN

de

S ] (B.11)

where d is the dimensionality of the system.

-140-

Bibliography

[1] S. Iijima, "Helical microtubules of graphitic carbon." Nature 354, 56 (1991).

[2] M. Endo, "The growth mechanism of vapor-grown carbon fibers." Ph.D. thesis,

University of Orleans, Orleans, France (1975).

[3] M. S. Dresselhaus, G. Dresselhaus, and P. C. Eklund, Science of fullerenes and carbon

nanotubes. (Academic Press, San Diego, 1996).

[4] S. Frank, P. Poncharal, Z. L. Wang, and W. A. de Heer, "Carbon nanotube quantum

resistors." Science 280, 1744 (1998).

[5] W. Liang, M. Bockrath, D. Bozovic, J. H. Hafner, M. Tinkham, and H. Park, "Fabry-

Perot interference in a nanotube electron waveguide." Nature 411, 665 (2001).

[6] P. Poncharal, C. Berger, Y. Yi, Z. L. Wang, and W. A. de Heer, "Room temperature

ballistic conduction in carbon nanotubes." Journal of Physical Chemistry B 106 (47),

12104 (2002).

[7] D. Mann, A. Javey, J. Kong, Q. Wang, and H. Dai, "Ballistic transport in metallic

nanotubes with reliable Pd ohmic contacts." Nano Letters, ACS ASAP.

[8] P. Kim, L. Shi, A. Majumdar, and P. L. McEuen, "Thermal transport measurements of

individual multiwalled nanotubes." Physical Review Letters 87 (21), 215502 (2001).

[9] E. W. Wong, P. E. Sheehan, and C. M. Lieber, "Nanobeam mechanics: elasticity, strength,

and toughness of nanorods and nanotubes." Science 277, 1971 (1997).

[10] B. I. Yakobson and P. Avouris, "Mechanical properties of carbon nanotubes." Topics in

Applied Physics 80 (Carbon Nanotubes), 287 (2001).

[11] M. S. Dresselhaus, "Future directions in carbon science." Annual Review of Materials

Science 27, 1 (1997).

-141-

[12] R. H. Baughman, A. A. Zakhidov, and W. A. de Heer, "Carbon nanotubes-the route

toward applications." Science 297, 787 (2002).

[13] P. M. Ajayan and O. Z. Zhou, "Applications of carbon nanotubes." Topics in Applied

Physics 80 (Carbon Nanotubes), 391 (2001).

[14] D. Qian, E. C. Dickey, R. Andrews, and T. Rantell, "Load transfer and deformation

mechanisms in carbon nanotube-polystyrene composites." Applied Physics Letters 76

(20), 2868 (2000).

[15] M. J. Biercuk, M. C. Llaguno, M. Radosavljevic, J. K. Hyun, A. T. Johnson, and J. E.

Fischer, "Carbon nanotube composites for thermal management." Applied Physics

Letters 80 (15), 2767 (2002).

[16] C. Niu, E. K. Sichel, R. Hoch, D. Moy, and H. Tennent, "High power electrochemical

capacitors based on carbon nanotube electrodes." Applied Physics Letters 70 (11), 1480

(1997).

[17] R. H. Baughman, C. Cui, A. A. Zakhidov, Z. Iqbal, J. N. Barisci, G. M. Spinks, G. G.

Wallace, A. Mazzoldi, D. De Rossi, A. G. Rinzler, O. Jaschinski, S. Roth, and M. Kertesz,

"Carbon nanotube actuators." Science 284, 1340 (1999).

[18] J. Kong, N. R. Franklin, C. Zhou, M. G. Chapline, S. Peng, K. Cho, and H. Dai,

"Nanotube molecular wires as chemical sensors." Science 287, 622 (2000).

[19] A. Modi, N. Koratkar, E. Lass, B. Wei, and P. M. Ajayan, "Miniaturized gas ionization

sensors using carbon nanotubes." Nature 424, 171 (2003).

[20] W. A. de Heer, A. Chatelain, and D. Ugarte, "A carbon nanotube field-emission electron

source." Science 270, 1179 (1995).

[21] A. G. Rinzler, J. H. Hafner, P. Nikolaev, L. Lou, S. G. Kim, D. Tomanek, P. Nordlander,

D. T. Colbert, and R. E. Smalley, "Unraveling nanotubes: field emission from an atomic

wire." Science 269, 1550 (1995).

-142-

[22] Q. H. Wang, A. A. Setlur, J. M. Lauerhaas, J. Y. Dai, E. W. Seelig, and R. P. H. Chang,

"A nanotube-based field-emission flat panel display." Applied Physics Letters 72 (22),

2912 (1998).

[23] Y. Saito, S. Uemura, and K. Hamaguchi, "Cathode ray tube lighting elements with carbon

nanotube field emitters." Japanese Journal of Applied Physics, Part 2: Letters 37 (3B),

L346 (1998).

[24] R. Rosen, W. Simendinger, C. Debbault, H. Shimoda, L. Fleming, B. Stoner, and O.

Zhou, "Application of carbon nanotubes as electrodes in gas discharge tubes." Applied

Physics Letters 76 (13), 1668 (2000).

[25] N. S. Lee, D. S. Chung, I. T. Han, J. H. Kang, Y. S. Choi, H. Y. Kim, S. H. Park, Y. W.

Jin, W. K. Yi, M. J. Yun, J. E. Jung, C. J. Lee, J. H. You, S. H. Jo, C. G. Lee, and J. M.

Kim, "Application of carbon nanotubes to field emission displays." Diamond and Related

Materials 10 (2), 265 (2001).

[26] A. C. Dillon, K. M. Jones, T. A. Bekkedahl, C. H. Kiang, D. S. Bethune, and M. J. Heben,

"Storage of hydrogen in single-walled carbon nanotubes." Nature 386, 377 (1997).

[27] P. Chen, X. Wu, J. Lin, and K. L. Tan, "High H2 uptake by alkali-doped carbon

nanotubes under ambient pressure and moderate temperatures." Science 285, 91 (1999).

[28] C. Liu, Y. Y. Fan, M. Liu, H. T. Cong, H. M. Cheng, and M. S. Dresselhaus, "Hydrogen

storage in single-walled carbon nanotubes at room temperature." Science 286, 1127

(1999).

[29] Y. Ye, C. C. Ahn, C. Witham, B. Fultz, J. Liu, A. G. Rinzler, D. Colbert, K. A. Smith,

and R. E. Smalley, "Hydrogen adsorption and cohesive energy of single-walled carbon

nanotubes." Applied Physics Letters 74 (16), 2307 (1999).

[30] Y. Chen, D. T. Shaw, X. D. Bai, E. G. Wang, C. Lund, W. M. Lu, and D. D. L. Chung,

"Hydrogen storage in aligned carbon nanotubes." Applied Physics Letters 78 (15), 2128

(2001).

-143-

[31] G. G. Tibbetts, G. P. Meisner, and C. H. Olk, "Hydrogen storage capacity of carbon

nanotubes, filaments, and vapor-grown fibers." Carbon 39 (15), 2291 (2001).

[32] J. H. Hafner, C. Li Cheung, and C. M. Lieber, "Growth of nanotubes for probe

microscopy tips." Nature 398, 761 (1999).

[33] S. S. Wong, E. Joselevich, A. T. Woolley, C. Li Cheung, and C. M. Lieber, "Covalently

functionalized nanotubes as nanometer-sized probes in chemistry and biology." Nature

394, 52 (1998).

[34] P. Kim and C. M. Lieber, "Nanotube nanotweezers." Science 286, 2148 (1999).

[35] M. S. Fuhrer, J. Nygard, L. Shih, M. Forero, Y.-G. Yoon, M. S. C. Mazzoni, H. J. Choi, J.

Ihm, S. G. Louie, A. Zettl, and P. L. McEuen, "Crossed nanotube junctions." Science 288,

494 (2000).

[36] T. Rueckes, K. Kim, E. Joselevich, G. Y. Tseng, C.-L. Cheung, and C. M. Lieber,

"Carbon nanotube-based nonvolatile random access memory for molecular computing."

Science 289, 94 (2000).

[37] S. J. Tans, A. R. M. Verschueren, and C. Dekker, "Room-temperature transistor based on

a single carbon nanotube." Nature 393, 49 (1998).

[38] R. Martel, T. Schmidt, H. R. Shea, T. Hertel, and P. Avouris, "Single- and multi-wall

carbon nanotube field-effect transistors." Applied Physics Letters 73 (17), 2447 (1998).

[39] A. Bachtold, P. Hadley, T. Nakanishi, and C. Dekker, "Logic circuits with carbon

nanotube transistors." Science 294, 1317 (2001).

[40] S. Iijima and T. Ichihashi, "Single-shell carbon nanotubes of 1-nm diameter." Nature 363,

603 (1993).

[41] D. S. Bethune, C. H. Kiang, M. S. de Vries, G. Gorman, R. Savoy, J. Vazquez, and R.

Beyers, "Cobalt-catalyzed growth of carbon nanotubes with single-atomic-layer walls."

Nature 363, 605 (1993).

-144-

[42] A. Thess, R. Lee, P. Nikolaev, H. Dai, P. Petit, J. Robert, C. Xu, Y. H. Lee, S. G. Kim, D.

T. Colbert, G. Scuseria, D. Tománek, J. E. Fischer, and R. E. Smalley, "Crystalline ropes

of metallic carbon nanotubes." Science 273, 483 (1996).

[43] C. Journet, W. K. Maser, P. Bernier, A. Loiseau, M. Lamy de la Chapells, S. Lefrant, P.

Deniard, R. Lee, and J. E. Fischer, "Large-scale production of single-walled carbon

nanotubes by the electric-arc technique." Nature 388, 756 (1997).

[44] C. Dekker, "Carbon nanotubes as molecular quantum wires." Physics Today 52 (5), 22

(1999).

[45] J. Nygard, D. H. Cobden, M. Bockrath, P. L. McEuen, and P. E. Lindelof, "Electrical

transport measurements on single-walled carbon nanotubes." Applied Physics A 69 (3),

297 (1999).

[46] Z. Yao, C. Dekker, and P. Avouris, "Electrical transport through single-wall carbon

nanotubes." Topics in Applied Physics 80 (Carbon Nanotubes), 147 (2001).

[47] L.-C. Qin, X. Zhao, K. Hirahara, Y. Miyamoto, Y. Andos, and S. Iijima, "Materials

science: The smallest carbon nanotube." Nature 408, 50 (2000).

[48] N. Wang, Z. K. Tang, G. D. Li, and J. S. Chen, "Materials science: Single-walled 4 Å

carbon nanotube arrays." Nature 408, 50 (2000).

[49] B. Wei, R. Vajtai, Y. Y. Choi, P. M. Ajayan, H. Zhu, C. Xu, and D. Wu, "Structural

characterizations of long single-walled carbon nanotube strands." Nano Letters 2 (10),

1105 (2002).

[50] R. Saito, G. Dresselhaus, and M. S. Dresselhaus, Physical properties of carbon

nanotubes. (Imperial College Press, London, 1998).

[51] T. W. Odom, J.-L. Huang, P. Kim, and C. M. Lieber, "Atomic structure and electronic

properties of single-walled carbon nanotubes." Nature 391, 62 (1998).

[52] P. R. Wallace, "The band theory of graphite." Physical Review 71, 622 (1947).

-145-

[53] M. S. Dresselhaus and P. C. Eklund, "Phonons in carbon nanotubes." Advances in

Physics 49 (6), 705 (2000).

[54] A. M. Rao, E. Richter, S. Bandow, B. Chase, P. C. Eklund, K. A. Williams, S. Fang, K. R.

Subbaswamy, M. Menon, A. Thess, R. E. Smalley, G. Dresselhaus, and M. S.

Dresselhaus, "Diameter-selective Raman scattering from vibrational modes in carbon

nanotubes." Science 275, 187 (1997).

[55] M. S. Dresselhaus, G. Dresselhaus, P. Eklund, and R. Saito, "Carbon nanotubes." Physics

World 11 (1), 33 (1998).

[56] K. A. Williams and P. C. Eklund, "Monte Carlo simulations of H2 physisorption in finite-

diameter carbon nanotube ropes." Chemical Physics Letters 320 (3/4), 352 (2000).

[57] A. C. Dillon and M. J. Heben, "Hydrogen storage using carbon adsorbents: past, present

and future." Applied Physics A 72 (2), 133 (2001).

[58] H. Dai, "Carbon nanotubes: opportunities and challenges." Surface Science 500 (1-3),

218 (2002).

[59] C. W. Zhou, J. Kong, E. Yenilmez, and H. J. Dai, "Modulated chemical doping of

individual carbon nanotubes." Science 290, 1552 (2000).

[60] P. C. Eklund, J. M. Holden, and R. A. Jishi, "Vibrational modes of carbon nanotubes;

spectroscopy and theory." Carbon 33 (7), 959 (1995).

[61] S. Bandow, S. Asaka, Y. Saito, A. M. Rao, L. Grigorian, E. Richter, and P. C. Eklund,

"Effect of the growth temperature on the diameter distribution and chirality of single-wall

carbon nanotubes." Physical Review Letters 80 (17), 3779 (1998).

[62] U. D. Venkateswaran, A. M. Rao, E. Richter, M. Menon, A. Rinzler, R. E. Smalley, and

P. C. Eklund, "Probing the single-wall carbon nanotube bundle: Raman scattering under

high pressure." Physical Review B 59 (16), 10928 (1999).

[63] L. Henrard, E. Hernandez, P. Bernier, and A. Rubio, "Van der Waals interaction in

-146-

nanotube bundles. Consequences on vibrational modes." Physical Review B 60 (12),

R8521 (1999).

[64] D. Kahn and J. P. Lu, "Vibrational modes of carbon nanotubes and nanoropes." Physical

Review B 60 (9), 6535 (1999).

[65] L. Alvarez, A. Righi, T. Guillard, S. Rols, E. Anglaret, D. Laplaze, and J. L. Sauvajol,

"Resonant Raman study of the structure and electronic properties of single-wall carbon

nanotubes." Chemical Physics Letters 316 (3/4), 186 (2000).

[66] A. G. Rinzler, J. Liu, H. Dai, P. Nikolaev, C. B. Huffman, F. J. Rodriguez-Macias, P. J.

Boul, A. H. Lu, D. Heymann, D. T. Colbert, R. S. Lee, J. E. Fischer, A. M. Rao, P. C.

Eklund, and R. E. Smalley, "Large-scale purification of single-wall carbon nanotubes:

Process, product, and characterization." Applied Physics A 67 (1), 29 (1998).

[67] P. C. Eklund and A. K. Mabatah, "Thermoelectric power measurements using analog

subtraction." Review of Scientific Instruments 48 (7), 775 (1977).

[68] G. U. Sumanasekera, L. Grigorian, and P. C. Eklund, "Low-temperature thermoelectrical

power measurements using analogue subtraction." Measurement Science & Technology

11 (3), 273 (2000).

[69] D. K. C. MacDonald, Thermoelectricity: an introduction to the principles. (Wiley, New

York, 1962).

[70] R. D. Barnard, Thermoelectricity in metals and alloys. (Taylor & Francis, London, 1972).

[71] F. J. Blatt, P. A. Schroeder, C. L. Foiles, and D. Greig, Thermoelectric power of metals.

(Plenum Press, New York, 1976).

[72] H. Fritzsche, "General expression for the thermoelectric power." Solid State

Communications 9 (21), 1813 (1971).

[73] F. J. Blatt, Physics of electronic conduction in solids. (McGraw-Hill, New York, 1968).

[74] J. Hone, I. Ellwood, M. Muno, A. Mizel, M. L. Cohen, A. Zettl, A. G. Rinzler, and R. E.

-147-

Smalley, "Thermoelectric power of single-walled carbon nanotubes." Physical Review

Letters 80 (5), 1042 (1998).

[75] L. Grigorian, K. A. Williams, S. Fang, G. U. Sumanasekera, A. L. Loper, E. C. Dickey, S.

J. Pennycook, and P. C. Eklund, "Reversible intercalation of charged iodine chains into

carbon nanotube ropes." Physical Review Letters 80 (25), 5560 (1998).

[76] L. Grigorian, G. U. Sumanasekera, A. L. Loper, S. L. Fang, J. L. Allen, and P. C. Eklund,

"Giant thermopower in carbon nanotubes: A one-dimensional Kondo system." Physical

Review B 60 (16), R11309 (1999).

[77] B. T. Kelly, Physics of graphite. (Applied Science Publishers, London, 1981).

[78] A. B. Kaiser, Y. W. Park, G. T. Kim, E. S. Choi, G. Dusberg, and S. Roth, "Electronic

transport in carbon nanotube ropes and mats." Synthetic Metals 103 (1-3), 2547 (1999).

[79] G. C. McIntosh and A. B. Kaiser, "Calculating thermopower due to fluctuation-assisted

tunneling with application to carbon nanotube ropes." Current Applied Physics 1, 145

(2001).

[80] M. Baxendale, K. G. Lim, and G. A. J. Amaratunga, "Thermoelectric power of aligned

and randomly oriented carbon nanotubes." Physical Review B 61 (19), 12705 (2000).

[81] J. W. G. Wildoer, L. C. Venema, A. G. Rinzier, R. E. Smalley, and C. Dekker,

"Electronic structure of atomically resolved carbon nanotubes." Nature 391, 59 (1998).

[82] P. L. McEuen, M. Bockrath, D. H. Cobden, Y.-G. Yoon, and S. G. Louie, "Disorder,

pseudospins, and backscattering in carbon nanotubes." Physical Review Letters 83 (24),

5098 (1999).

[83] P. G. Collins, M. S. Arnold, and P. Avouris, "Engineering carbon nanotubes and

nanotube circuits using electrical breakdown." Science 292, 706 (2001).

[84] Y. Zvyagin, in Hopping transport in solids, edited by B. I. Shklovskii (North Holland,

New York, 1991).

-148-

[85] E. S. Choi, D. S. Suh, G. T. Kim, D. C. Kim, and Y. W. Park, "Magnetothermopower of

single wall carbon nanotube newtwork." Synthetic Metals 103 (1-3), 2504 (1999).

[86] P. Sheng, "Fluctuation-induced tunneling conduction in disordered materials." Physical

Review B 21 (6), 2180 (1980).

[87] G. D. Mahan, Many-particle physics, 3rd ed. (Kluwer Academic/Plenum Publishers, New

York, 2000).

[88] K. Bradley, S. -H. Jhi, P. G. Collins, J. Hone, M. L. Cohen, S. G. Louie, and A. Zettl, "Is

the intrinsic thermoelectric power of carbon nanotubes positive?" Physical Review

Letters 85 (20), 4361 (2000).

[89] G. U. Sumanasekera, C. K. W. Adu, S. Fang, and P. C. Eklund, "Effects of gas

adsorption and collisions on electrical transport in single-walled carbon nanotubes."

Physical Review Letters 85 (5), 1096 (2000).

[90] P. G. Collins, K. Bradley, M. Ishigami, and A. Zettl, "Extreme oxygen sensitivity of

electronic properties of carbon nanotubes." Science 287, 1801 (2000).

[91] S. Peng and K. Cho, "Chemical control of nanotube electronics." Nanotechnology 11 (2),

57 (2000).

[92] S.-H. Jhi, S. G. Louie, and M. L. Cohen, "Electronic properties of oxidized carbon

nanotubes." Physical Review Letters 85 (8), 1710 (2000).

[93] X. Y. Zhu, S. M. Lee, Y. H. Lee, and T. Frauenheim, "Adsorption and desorption of an

O2 molecule on carbon nanotubes." Physical Review Letters 85 (13), 2757 (2000).

[94] D. C. Sorescu, K. D. Jordan, and P. Avouris, "Theoretical study of oxygen adsorption on

graphite and the (8,0) single-walled carbon nanotube." Journal of Physical Chemistry B

105 (45), 11227 (2001).

[95] J. Zhao, A. Buldum, J. Han, and J. P. Lu, "Gas molecule adsorption in carbon nanotubes

and nanotube bundles." Nanotechnology 13 (2), 195 (2002).

-149-

[96] A. Ricca and J. A. Drocco, "Interaction of O2 with a (9,0) carbon nanotube." Chemical

Physics Letters 362 (3,4), 217 (2002).

[97] S.-P. Chan, G. Chen, X. G. Gong, and Z.-F. Liu, "Oxidation of carbon nanotubes by

singlet O2." Physical Review Letters 90 (8), 086403 (2003).

[98] S. Dag, O. Gulseren, T. Yildirim, and S. Ciraci, "Oxygenation of carbon nanotubes:

atomic structure, energetics, and electronic structure." Physical Review B 67 (16),

165424 (2003).

[99] G. E. Froudakis, M. Schnell, M. Muhlhauser, S. D. Peyerimhoff, A. N. Andriotis, M.

Menon, and R. M. Sheetz, "Pathways for oxygen adsorption on single-wall carbon

nanotubes." Physical Review B 68 (11), 115435 (2003).

[100] M. Grujicic, G. Cao, A. M. Rao, T. M. Tritt, and S. Nayak, "UV-light enhanced oxidation

of carbon nanotubes." Applied Surface Science 214 (1-4), 289 (2003).

[101] M. Grujicic, G. Cao, and R. Singh, "The effect of topological defects and oxygen

adsorption on the electronic transport properties of single-walled carbon-nanotubes."

Applied Surface Science 211 (1-4), 166 (2003).

[102] P. Giannozzi, R. Car, and G. Scoles, "Oxygen adsorption on graphite and nanotubes."

Journal of Chemical Physics 118 (3), 1003 (2003).

[103] A. Ricca, C. W. Bauschlicher, Jr., and A. Maiti, "Comparison of the reactivity of O2 with

a (10,0) and a (9,0) carbon nanotube." Physical Review B 68 (3), 035433 (2003).

[104] V. Derycke, R. Martel, J. Appenzeller, and P. Avouris, "Carbon nanotube inter- and

intramolecular logic gates." Nano Letters 1 (9), 453 (2001).

[105] R. Martel, V. Derycke, C. Lavoie, J. Appenzeller, K. K. Chan, J. Tersoff, and P. Avouris,

"Ambipolar electrical transport in semiconducting single-wall carbon nanotubes."

Physical Review Letters 87 (25), 256805 (2001).

[106] V. Derycke, R. Martel, J. Appenzeller, and P. Avouris, "Controlling doping and carrier

-150-

injection in carbon nanotube transistors." Applied Physics Letters 80 (15), 2773 (2002).

[107] A. B. Kaiser, G. U. Flanagan, D. M. Stewart, and D. Beaglehole, "Heterogeneous model

for conduction in conducting polymers and carbon nanotubes." Synthetic Metals 117 (1-

3), 67 (2001).

[108] H. E. Romero, G. U. Sumanasekera, G. D. Mahan, and P. C. Eklund, "Thermoelectric

power of single-walled carbon nanotube films." Physical Review B 65 (20), 205410

(2002).

[109] A. Goldoni, R. Larciprete, L. Petaccia, and S. Lizzit, "Single-wall carbon nanotube

interaction with gases: sample contaminants and environmental monitoring." Journal of

the American Chemical Society 125 (37), 11329 (2003).

[110] H. Ulbricht, G. Moos, and T. Hertel, "Physisorption of molecular oxygen on single-wall

carbon nanotube bundles and graphite." Physical Review B 66 (7), 075404 (2002).

[111] A. Fujiwara, K. Ishii, H. Suematsu, H. Kataura, Y. Maniwa, S. Suzuki, and Y. Achiba,

"Gas adsorption in the inside and outside of single-walled carbon nanotubes." Chemical

Physics Letters 336 (3/4), 205 (2001).

[112] T. Kostyrko, M. Bartkowiak, and G. D. Mahan, "Reflection by defects in a tight-binding

model of nanotubes." Physical Review B 59 (4), 3241 (1999).

[113] P. E. Lammert, V. H. Crespi, and A. Rubio, "Stochastic heterostructures and diodium in

B/N-doped carbon nanotubes." Physical Review Letters 87 (13), 136402 (2001).

[114] T. Kostyrko, M. Bartkowiak, and G. D. Mahan, "Localization in carbon nanotubes within

a tight-binding model." Physical Review B 60 (15), 10735 (1999).

[115] L. Gurevich, "Thermoelectric propoerties of conductors. I." Journal of Physics 9, 477

(1945).

[116] L. Gurevich, "Thermoelectric properties of conductros. II." Journal of Physics 10, 67

(1946).

-151-

[117] C. Herring, "Theory of the thermoelectric power of semiconductors." Physical Review 96

(5), 1163 (1954).

[118] J. Hone, B. Batlogg, Z. Benes, A. T. Johnson, and J. E. Fischer, "Quantized phonon

spectrum of single-wall carbon nanotubes." Science 289, 1730 (2000).

[119] V. W. Scarola and G. D. Mahan, "Phonon drag effect in single-walled carbon nanotubes."

Physical Review B 66 (20), 205405 (2002).

[120] M. Baylin, "Transport in metals: Effect of the nonequilibrium phonons." Physical Review

112 (5), 1587 (1958).

[121] M. Baylin, "Tranport in metals. II. Effect of the phonon spectrum and Umklapp processes

at high and low temperature." Physical Review 120 (2), 381 (1960).

[122] M. Baylin, "Phonon-drag part of the thermoelectric power in metals." Physical Review

157 (3), 480 (1967).

[123] L. M. Woods and G. D. Mahan, "Electron-phonon effects in graphene and armchair

(10,10) single-wall carbon nanotubes." Physical Review B 61 (16), 10651 (2000).

[124] S. S. Kubakaddi and P. N. Butcher, "A calculation of the phonon-drag thermopower of a

1D electron gas." Journal of Physics 1 (25), 3939 (1989).

[125] M. Tsaousidou and P. N. Butcher, "Phonon-drag thermopower of a ballistic quantum

wire." Physical Review B 56 (16), R10044 (1997).

[126] in G. D. Mahan (2003).

[127] R. S. Lee, H. J. Kim, J. E. Fischer, A. Thess, and R. E. Smalley, "Conductivity

enhancement in single-walled carbon nanotube bundles doped with K and Br." Nature

388, 255 (1997).

[128] L. Grigorian, G. U. Sumanasekera, A. L. Loper, S. Fang, J. L. Allen, and P. C. Eklund,

"Transport properties of alkali-metal-doped single-wall carbon nanotubes." Physical

Review B 58 (8), R4195 (1998).

-152-

[129] R. S. Lee, H. J. Kim, J. E. Fischer, J. Lefebvre, M. Radosavljevic, J. Hone, and A. T.

Johnson, "Transport properties of a potassium-doped single-wall carbon nanotube rope."

Physical Review B 61 (7), 4526 (2000).

[130] M. Bockrath, J. Hone, A. Zettl, P. L. McEuen, A. G. Rinzler, and R. E. Smalley,

"Chemical doping of individual semiconducting carbon-nanotube ropes." Physical

Review B 61 (16), R10606 (2000).

[131] A. M. Rao, P. C. Eklund, S. Bandow, A. Thess, and R. E. Smalley, "Evidence for charge

transfer in doped carbon nanotube bundles from Raman scattering." Nature 388, 257

(1997).

[132] G. U. Sumanasekera, C. K. W. Adu, B. K. Pradhan, G. Chen, H. E. Romero, and P. C.

Eklund, "Thermoelectric study of hydrogen storage in carbon nanotubes." Physical

Review B 65 (3), 035408 (2001).

[133] M. S. Dresselhaus, K. A. Williams, and P. C. Eklund, "Hydrogen adsorption in carbon

materials." MRS Bulletin 24 (11), 45 (1999).

[134] G. Stan and M. W. Cole, "Hydrogen adsorption in nanotubes." Journal of Low

Temperature Physics 110 (1/2), 539 (1998).

[135] R. E. Tuzun, D. W. Noid, B. G. Sumpter, and R. C. Merkle, "Dynamics of fluid flow

inside carbon nanotubes." Nanotechnology 7 (3), 241 (1996).

[136] C. K. W. Adu, G. U. Sumanasekera, B. K. Pradhan, H. E. Romero, and P. C. Eklund,

"Carbon nanotubes: A thermoelectric nano-nose." Chemical Physics Letters 337 (1-3), 31

(2001).

[137] A. Lueking and R. T. Yang, "Hydrogen spillover from a metal oxide catalyst onto carbon

nanotubes - Implications for hydrogen storage." Journal of Catalysis 206 (1), 165 (2002).

[138] J. Kong and H. Dai, "Full and modulated chemical gating of individual carbon nanotubes

by organic amine compounds." Journal of Physical Chemistry B 105 (15), 2890 (2001).

-153-

[139] D. R. Lide, CRC handbook of chemistry and physics, 83rd. ed. (CRC Press, Boca Raton,

2002).

[140] P. A. Elkington and G. Curthoys, "Heats of adsorption on carbon black surfaces." Journal

of Physical Chemistry 73 (7), 2321 (1969).

[141] E. V. Kalashnikova, A. V. Kiselev, R. S. Petrova, and K. D. Shcherbakova, "Gas-

chromatographic investigation of the adsorption equilibrium on graphitized thermal

carbon black. I. Henry constants and heats of adsorption of C1-C6 hydrocarbons at zero

coverage." Chromatographia 4 (11), 495 (1971).

[142] Jijun Zhao, Jian Ping Lu, Jie Han, and Chih-Kai Yang, "Noncovalent functionalization of

carbon nanotubes by aromatic organic molecules." Applied Physics Letters 82 (21), 3746

(2003).

[143] N. N. Avgul, G. I. Berezin, A. V. Kiselev, and I. A. Lygina, "Adsorption and heat of

adsorption of normal alcohols on graphitized charcoal." Bulletin of the Academy of

Sciences of the USSR, Division of Chemical Science, 186 (1961).

[144] N. N. Avgul, A. V. Kiselev, and I. A. Lygina, "Energy of adsorption for water, alcohols,

ammonia, and methylamine on graphite." Bulletin of the Academy of Sciences of the

USSR, Division of Chemical Science, 1308 (1961).

[145] R. Pati, Y. Zhang, S. K. Nayak, and P. M. Ajayan, "Effect of H2O adsorption on electron

transport in a carbon nanotube." Applied Physics Letters 81 (14), 2638 (2002).

[146] Z. Mao and S. B. Sinnott, "A computational study of molecular diffusion and dynamic

flow through carbon nanotubes." Journal of Physical Chemistry B 104 (19), 4618 (2000).

[147] A. Zahab, L. Spina, P. Poncharal, and C. Marliere, "Water-vapor effect on the electrical

conductivity of a single-walled carbon nanotube mat." Physical Review B 62 (15), 10000

(2000).

[148] J. Vavro, M. C. Llaguno, J. E. Fischer, S. Ramesh, R. K. Saini, L. M. Ericson, V. A.

Davis, R. H. Hauge, M. Pasquali, and R. E. Smalley, "Thermoelectric power of p-doped

-154-

single-wall carbon nanotubes and the role of phonon drag." Physical Review Letters 90

(6), 065503 (2003).

[149] G. U. Sumanasekera, B. K. Pradhan, H. E. Romero, K. W. Adu, and P. C. Eklund, "Giant

thermopower effects from molecular physisorption on carbon nanotubes." Physical

Review Letters 89 (16), 166801 (2002).

[150] R. H. Bradley and B. Rand, "A comparison of the adsorption behavior of nitrogen,

alcohols, and water towards active carbons." Carbon 29 (8), 1165 (1991).

[151] C.-M. Yang, H. Kanoh, K. Kaneko, M. Yudasaka, and S. Iijima, "Adsorption behaviors

of HiPCo single-walled carbon nanotube aggregates for alcohol vapors." Journal of

Physical Chemistry B 106 (35), 8994 (2002).

[152] Q. Wang, J. K. Johnson, and J. Q. Broughton, "Path integral grand canonical Monte

Carlo." Journal of Chemical Physics 107 (13), 5108 (1997).

[153] G. Stan and M. W. Cole, "Low coverage adsorption in cylindrical pores." Surface

Science 395 (2/3), 280 (1998).

[154] G. Stan, V. H. Crespi, M.W. Cole, and M. Boninsegni, "Interstitial He and Ne in

nanotube bundles." Journal of Low Temperature Physics 113 (3/4), 447 (1998).

[155] A. M. Vidales, V. H. Crespi, and M. W. Cole, "Heat capacity and vibrational spectra of

monolayer films adsorbed in nanotubes." Physical Review B 58 (20), R13426 (1998).

[156] G. Stan, J. M. Hartman, V. H. Crespi, S. M. Gatica, and M. W. Cole, "Helium mixtures in

nanotube bundles." Physical Review B 61 (11), 7288 (2000).

[157] M. W. Cole, Vincent H. Crespi, G. Stan, C. Ebner, Jacob M. Hartman, S. Moroni, and M.

Boninsegni, "Condensation of helium in nanotube bundles." Physical Review Letters 84

(17), 3883 (2000).

[158] M. K. Kostov, M. W. Cole, J. C. Lewis, P. Diep, and J. K. Johnson, "Many-body

interactions among adsorbed atoms and molecules within carbon nanotubes and in free

-155-

space." Chemical Physics Letters 332 (1/2), 26 (2000).

[159] M. Boninsegni, S. Y. Lee, and V. H. Crespi, "Helium in one-dimensional nanopores: free

dispersion, localization, and commensurate/incommensurate transitions with nonrigid

orbitals." Physical Review Letters 86 (15), 3360 (2001).

[160] H. Tanaka, M. El-Merraoui, W. A. Steele, and K. Kaneko, "Methane adsorption on

single-walled carbon nanotube. A density functional theory model." Chemical Physics

Letters 352 (5/6), 334 (2002).

[161] M. C. Gordillo, J. Boronat, and J. Casulleras, "Quasi-one-dimensional 4He inside carbon

nanotubes." Physical Review B 61 (2), R878 (2000).

[162] L. Vranjes, Z. Antunovic, and S. Kilic, "Helium molecules within carbon nanotubes."

Physica B. 329-333 (Part1), 276 (2003).

[163] M. M. Calbi, S.M. Gatica, M. J. Bojan, and M. W. Cole, "Phases of neon, xenon, and

methane adsorbed on nanotube bundles." Journal of Chemical Physics 115 (21), 9975

(2001).

[164] M. M. Calbi and M. W. Cole, "Dimensional crossover and quantum effects of gases

adsorbed on nanotube bundles." Physical Review B 66 (11), 115413 (2002).

[165] M. W. Maddox and K. E. Gubbins, "Molecular simulation of fluid adsorption in

buckytubes." Langmuir 11 (10), 3988 (1995).

[166] S. M. Gatica, M. J. Bojan, G. Stan, and M. W. Cole, "Quasi-one- and two-dimensional

transitions of gases adsorbed on nanotube bundles." Journal of Chemical Physics 114 (8),

3765 (2001).

[167] V. V. Simonyan, J. K. Johnson, A. Kuznetsova, and J. T. Yates, Jr., "Molecular

simulation of xenon adsorption on single-walled carbon nanotubes." Journal of Chemical

Physics 114 (9), 4180 (2001).

[168] X. Zhang and W. Wang, "Methane adsorption in single-walled carbon nanotubes arrays

-156-

by molecular simulation and density functional theory." Fluid Phase Equilibria 194-197,

289 (2002).

[169] I. A. Khan and K. G. Ayappa, "Density distributions of diatoms in carbon nanotubes: A

grand canonical Monte Carlo study." Journal of Chemical Physics 109 (11), 4576 (1998).

[170] W. Teizer, R. B. Hallock, E. Dujardin, and T. W. Ebbesen, "4He desorption from single

wall carbon nanotube bundles: A one-dimensional adsorbate." Physical Review Letters

82 (26), 5305 (1999).

[171] K. Ichimura, K. Imaeda, and H. Inokuchi, "Characteristic bonding of rare gases in solid

carbon nanotubes." Synthetic Metals 121 (1-3), 1191 (2001).

[172] Y. H. Kahng, R. B. Hallock, E. Dujardin, and T. W. Ebbesen, "4He binding energies on

single-wall carbon nanotube bundles." Journal of Low Temperature Physics 126 (1/2),

223 (2002).

[173] T. Wilson and O. E. Vilches, "Adsorption of 4He on carbon nanotube bundles." Physica

B 329-333 (1), 278 (2003).

[174] S. Talapatra, A. Z. Zambano, S. E. Weber, and A. D. Migone, "Gases do not adsorb on

the interstitial channels of closed-ended single-walled carbon nanotube bundles."

Physical Review Letters 85 (1), 138 (2000).

[175] S. Talapatra, V. Krungleviciute, and A. D. Migone, "Higher coverage gas adsorption on

the surface of carbon nanotubes: Evidence for a possible new phase in the second layer."

Physical Review Letters 89 (24), 246106 (2002).

[176] S. Talapatra, D. S. Rawat, and A. D. Migone, "Possible existence of a higher coverage

quasi-one-dimensional phase of argon adsorbed on bundles of single-walled carbon

nanotubes." Journal of Nanoscience and Nanotechnology 2 (5), 467 (2002).

[177] A. Kuznetsova, J. T. Yates, Jr., J. Liu, and R. E. Smalley, "Physical adsorption of xenon

in open single walled carbon nanotubes. Observation of a quasi-one-dimensional

confined Xe phase." Journal of Chemical Physics 112 (21), 9590 (2000).

-157-

[178] A. Kuznetsova, J. T. Yates, Jr., V. V. Simonyan, J. K. Johnson, C. B. Huffman, and R. E.

Smalley, "Optimization of Xe adsorption kinetics in single walled carbon nanotubes."

Journal of Chemical Physics 115 (14), 6691 (2001).

[179] H. Ulbricht, J. Kriebel, G. Moos, and T. Hertel, "Desorption kinetics and interaction of

Xe with single-wall carbon nanotube bundles." Chemical Physics Letters 363 (3/4), 252

(2002).

[180] M. Muris, N. Dufau, M. Bienfait, N. Dupont-Pavlovsky, Y. Grillet, and J. P. Palmari,

"Methane and krypton adsorption on single-walled carbon nanotubes." Langmuir 16 (17),

7019 (2000).

[181] E. B. Mackie, R. A. Wolfson, L. M. Arnold, K. Lafdi, and A. D. Migone, "Adsorption

studies of methane films on catalytic carbon nanotubes and on carbon filaments."

Langmuir 13 (26), 7197 (1997).

[182] S. E. Weber, S. Talapatra, C. Journet, A. Zambano, and A. D. Migone, "Determination of

the binding energy of methane on single-walled carbon nanotube bundles." Physical

Review B 61 (19), 13150 (2000).

[183] M. Bienfait, B. Asmussen, M. Johnson, and P. Zeppenfeld, "Methane mobility in carbon

nanotubes." Surface Science 460 (1-3), 243 (2000).

[184] S. Talapatra and A. D. Migone, "Adsorption of methane on bundles of closed-ended

single-wall carbon nanotubes." Physical Review B 65 (4), 045416 (2002).

[185] M. Eswaramoorthy, R. Sen, and C. N. R. Rao, "A study of micropores in single-walled

carbon nanotubes by the adsorption of gases and vapors." Chemical Physics Letters 304

(3/4), 207 (1999).

[186] H. M. Cheng, Q. H. Yang, and C. Liu, "Hydrogen storage in carbon nanotubes." Carbon

39 (10), 1447 (2001).

[187] F. L. Darkrim, P. Malbrunot, and G. P. Tartaglia, "Review of hydrogen storage by

adsorption in carbon nanotubes." International Journal of Hydrogen Energy 27 (2), 193

-158-

(2001).

[188] Y. Ma, Y. Xia, M. Zhao, M. Ying, X. Liu, and P. Liu, "Collision of hydrogen atom with

single-walled carbon nanotube: Adsorption, insertion, and healing." Journal of Chemical

Physics 115 (17), 8152 (2001).

[189] K. Bolton and A. Rosen, "Computational studies of gas-carbon nanotube collision

dynamics." Physical Chemistry Chemical Physics 4 (18), 4481 (2002).

[190] H. Dai, "Nanotube growth and characterization." Topics in Applied Physics 80 (Carbon

Nanotubes), 29 (2001).

[191] T. W. Ebbesen and P. M. Ajayan, "Large-scale synthesis of carbon nanotubes." Nature

358, 220 (1992).

[192] Y. Ando and S. Iijima, "Preparation of carbon nanotubes by arc-discharge evaporation."

Japanese Journal of Applied Physics, Part 2 32 (1A-B), L107 (1993).

[193] X. K. Wang, X. W. Lin, M. Mesleh, M. F. Jarrold, V. P. Dravid, J. B. Ketterson, and R. P.

H. Chang, "The effect of hydrogen on the formation of carbon nanotubes and fullerenes."

Journal of Materials Research 10 (8), 1977 (1995).

[194] X. K. Wang, X. W. Lin, V. P. Dravid, J. B. Ketterson, and R. P. H. Chang, "Carbon

nanotubes synthesized in a hydrogen arc discharge." Applied Physics Letters 66 (18),

2430 (1995).

[195] R. Yu, M. Zhan, D. Cheng, S. Yang, Z. Liu, and L. Zheng, "Simultaneous synthesis of

carbon nanotubes and nitrogen-doped fullerenes in nitrogen atmosphere." Journal of

Physical Chemistry 99 (7), 1818 (1995).

[196] J. M. Lauerhaas, J. Y. Dai, A. A. Setlur, and R. P. H. Chang, "The effect of arc

parameters on the growth of carbon nanotubes." Journal of Materials Research 12 (6),

1536 (1997).

[197] P. B. Allen, in Superconductivity in d- and f-band metals, edited by M. B. Maple

-159-

(Academic Press, New York, 1980), Vol., 525.

[198] P. L. Rossiter, The electrical resistivity of metals and alloys. (Cambridge University Press,

Cambridge, 1987).

[199] M. Monthioux, B. W. Smith, B. Burteaux, A. Claye, J. E. Fischer, and D. E. Luzzi,

"Sensitivity of single-wall carbon nanotubes to chemical processing: an electron

microscopy investigation." Carbon 39 (8), 1251 (2001).

-160-

VITA

Hugo E. Romero

Education:

Ph.D. in Physics, The Pennsylvania State University, University Park, Pennsylvania,

2004

M.S. in Physics, University of Cincinnati, Cincinnati, Ohio, 1996

B.S. in Physics, Universidad Central de Venezuela, Caracas, Venezuela, 1993

Professional Positions Held:

Research and Teaching Assistant, The Pennsylvania State University (2000–2003)

Teaching Assistant, University of Cincinnati (1995–2000)

Research Scientist, Universidad Simón Bolívar (1993–1994)

Undergraduate Teaching Assistant, Universidad Central de Venezuela (1990–1993)

Honors and Awards

David C. Duncan Graduate Fellowship, The Pennsylvania State University (2002–2003)

Charles B. Braddock Graduate Fellowship, The Pennsylvania State University (2000–

2001)

Graduate Scholarship, CONICIT, Venezuela (1995–2000)

AVVA Merit Scholarship, American-Venezuelan Association of Friendship, Venezuela

(1990–1993)

Award “Academic Achievements”, Universidad Central de Venezuela, Venezuela (1990)