© 2014 zuanyi li

133
© 2014 Zuanyi Li

Upload: others

Post on 02-Jan-2022

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: © 2014 Zuanyi Li

© 2014 Zuanyi Li

Page 2: © 2014 Zuanyi Li

THERMAL TRANSPORT IN GRAPHENE-BASED NANOSTRUCTURES AND OTHER

TWO-DIMENSIONAL MATERIALS

BY

ZUANYI LI

DISSERTATION

Submitted in partial fulfillment of the requirements

for the degree of Doctor of Philosophy in Physics

in the Graduate College of the

University of Illinois at Urbana-Champaign, 2014

Urbana, Illinois

Doctoral Committee:

Associate Professor Nadya Mason, Chair

Adjunct Associate Professor Eric Pop, Director of Research

Professor David Cahill

Associate Professor Aleksei Aksimentiev

Page 3: © 2014 Zuanyi Li

ii

ABSTRACT

Heat conduction in nanomaterials is an important area of study, with both fundamental and

technological implications. However, little is known about heat flow in two-dimensional (2D)

materials or devices with dimensions comparable to the phonon mean free path (MFP). Here, we

investigated thermal transport in graphene-based nanostructures and several other 2D materials.

First, we measured heat conduction in nanoscale graphene by a substrate-supported

thermometry platform. Short, quarter-micron graphene samples reach ~35% of the ballistic

thermal conductance limit (Gball) up to room temperature, enabled by the relatively large phonon

MFP (~100 nm) in SiO2 substrate-supported graphene. In contrast, patterning similar samples

into nanoribbons leads to a diffusive heat flow regime that is controlled by ribbon width and

edge disorder. These results show how manipulation of device dimensions on the scale of the

phonon MFP can be used to achieve full control of their heat-carrying properties, approaching

fundamentally limited upper or lower bounds.

We also examined the possibility of using this supported platform to measure other materials

through finite element simulations and uncertainty analysis. The smallest thermal sheet

conductance that can be sensed by this method within a 50% error is ~25 nWK-1

at room

temperature, indicating this platform can be applied to most thin films like polymer and nanotube

networks, as well as nanomaterials like nanotube or nanowire arrays, even a single Si nanowire.

Moreover, the platform can be extended to plastic substrates, not limited to the SiO2/Si substrate.

Last, we calculated in-plane (for monolayer and bulk) and cross-plane (for bulk) ballistic

thermal conductances Gball of graphene/graphite, h-BN, MoS2, and WS2, based on full phonon

dispersions from first-principles approach. Then, a rigorous and proper average of phonon mean

free path, λ was simply obtained in terms of Gball and the diffusive thermal conductivity.

Moreover, length-dependent thermal conductivity (k) was estimated using a ballistic-diffusive

model, which agrees with available experimental data and shows increasing k with length until

~100λ before convergence. This indicates that, to observe theoretically predicted k divergence in

low-dimensional systems, simulations and experiments should extend beyond length ~100λ.

Our work provides a comprehensive study of thermal conduction in 2D layered materials in

micro- and nanoscale with an emphasis on ballistic conduction and size effects. The findings

extend our understanding of thermal conduction and how to tune it to reach the requirements for

potential applications like thermal management and thermoelectric conversion.

Page 4: © 2014 Zuanyi Li

iii

To my parents

Page 5: © 2014 Zuanyi Li

iv

ACKNOWLEDGEMENTS

I would like to thank my advisor, Professor Eric Pop, for his invaluable guidance and

support. Throughout my graduate study, he has been a great mentor and inspired me in various

aspects. His advice and support on research and career development has played an invaluable

role in my graduate life and will be precious to my future life. I would also like to thank the

committee members of my final and preliminary exams: Prof. Nadya Mason, Prof. David Cahill,

Prof. Aleksei Aksimentiev, and Prof. Karin Dahmen, whose valuable feedback is very helpful to

my research projects.

I want to thank all my colleagues in the Pop group, who create a friendly and cooperative

atmosphere within the group. In particular, I would like to thank Dr. Myung-Ho Bae, Dr. Vincent

Dorgan, Dr. David Estrada, and Dr. Feng Xiong, who taught me experimental skills and gave a

lot of valuable advice. I also thank Dr. Albert Liao, Dr. Ashkan Behnam, Dr. Zhun-Yong Ong,

Sharnali Islam, Kyle Grosse, Enrique Carrion, Andrey Serov, Feifei Lian, Chris English, Ning

Wang, and Michal Mleczko for all their help and support. Meanwhile, I would like to thank my

collaborators out of the Pop group: Prof. Wenhui Duan, Prof. Zlatan Aksamija, Prof. Irena

Knezevic, Dr. Lucas Lindsay, Dr. Pierre Martin, Dr. Bing Huang, Dr. Yong Xu, and Yizhou Liu.

I appreciate the resources available to me at the University of Illinois at Urbana-Champaign

(UIUC), especially those at the Micro and Nanotechnology Lab (MNTL) and Materials Research

Lab (MRL). I acknowledge funding support from National Science Foundation (NSF),

Nanoelectronics Research Initiative (NRI), and all scholarships and awards I received from

UIUC.

Finally, I would like to extend special thanks to my parents for their unconditional love and

unwavering support. They have sacrificed so much to make me have better life and education,

and they always believed in me. They are the most important part of my life. Moreover, I thank

all my friends for their help and the wonderful time that we had together, which is an

unforgettable part in my life.

Page 6: © 2014 Zuanyi Li

v

TABLE OF CONTENTS

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

1.1 Power Dissipation and Thermal Management ...................................................1

1.2 Two-Dimensional Materials ..............................................................................2

1.3 Organization of Dissertation ..............................................................................4

CHAPTER 2: REVIEW OF THERMAL AND THERMOELECTRIC PROPERTIES OF

GRAPHENE .......................................................................................................................6

2.1 Methods to Measure Thermal Conductivity ......................................................7

2.2 Intrinsic Thermal Conductivity of Graphene ...................................................12

2.3 Extrinsic Thermal Conductivity of Graphene ..................................................14

2.4 Thermoelectric Properties of Graphene ...........................................................22

CHAPTER 3: BALLISTIC TO DIFFUSIVE CROSSOVER OF HEAT FLOW IN

GRAPHENE RIBBONS ...................................................................................................26

3.1 Introduction ......................................................................................................26

3.2 Experimental Method.......................................................................................27

3.3 Model to Analyze Experimental Data ..............................................................36

3.4 Results ..............................................................................................................43

3.5 Discussion and BTE Calculations ....................................................................53

3.6 Summary ..........................................................................................................62

CHAPTER 4: SUBSTRATE-SUPPORTED THERMOMETRY PLATFORM FOR

NANOMATERIALS LIKE GRAPHENE, NANOTUBES, AND NANOWIRES ...........64

4.1 Introduction ......................................................................................................64

4.2 Thermometry Platform Demonstration ............................................................65

4.3 Uncertainty Analysis and Optimization ...........................................................70

4.4 Applicability of the Platform ...........................................................................74

4.5 Summary ..........................................................................................................78

CHAPTER 5: BALLISTIC THERMAL CONDUCTANCE AND LENGTH-DEPENDENT

SIZE EFFECTS IN LAYERED TWO-DIMENSIONAL MATERIALS .........................79

5.1 Introduction ......................................................................................................79

5.2 Phonon Dispersions by First-Principles Calculations ......................................81

5.3 Formalism of Thermal Physical Quantities .....................................................86

5.4 Results and Discussion ....................................................................................93

5.5 Summary ........................................................................................................106

CHAPTER 6: CONCLUSIONS AND OUTLOOK ........................................................108

6.1 Conclusions ....................................................................................................108

6.2 Future Work ...................................................................................................110

REFERENCES ................................................................................................................112

Page 7: © 2014 Zuanyi Li

1

CHAPTER 1

INTRODUCTION

1.1 Power Dissipation and Thermal Management

In today’s information technology (IT) era, demands for power usage in electronics are

growing exponentially, and power dissipation has been a big issue in energy consumption.

Taking the United States as an example, the total IT power use in 2007 was estimated to be

~20.6 gigawatts (GW), where the biggest usage is from data centers (~7 GW) [1], as shown in

Fig. 1.1a. The power usage in US data centers continuously increases year-by-year, and has been

tripled from 2000 to 2011 (Fig. 1.1b) [1]. Figure 1.1a also shows that the power usage by work

and home personal computers (PCs) is significant high as well. These issues of power dissipation

are due to increased power density in microprocessor (CPUs) when the transistor size is scaled

down following the Moore’s law. Figure 1.2 shows that the typical CPU power density increased

exponentially over time and flattened out after 2004 when it reached ~100 W/cm2 due to the

introduction of multi-core CPUs [1]. This power density is approximately an order of magnitude

higher than that of a hot plate, implying incredible heat generated in devices. This is clearly

FIG. 1.1: (a) Estimate of US total IT power use in 2007. (b) Estimate of US data center power

use from 2000 to 2011. Adapted from Ref. [1].

a bData centers

(7 GW)

Home PCs

(2.6 GW)Home displays (1.3 GW)

Work displays

(3.2 GW)

Work PCs

(6.5 GW)

Total power

use (2007)

Page 8: © 2014 Zuanyi Li

2

reflected in the power usage of US data centers, where about half of it is used for cooling.

To accommodate the increasing power dissipation, measures must be taken to ensure that

the generated waste heat is properly managed. One strategy of managing this waste heat is to

modify the design of the transistor and its surrounding packaging, and to use thermally conduc-

tive materials which can help to transport heat away from regions of localized Joule heating.

Another strategy is to use thermoelectric materials with low thermal conductivity to convert

extra heat to electrical power [2, 3]. Both require effective control of thermal conduction ability

in materials, which relies on comprehensive and deep understanding of thermal transport [4, 5].

1.2 Two-Dimensional Materials

As an ultimately two-dimensional (2D) crystal, graphene, an atomic single-layer of graphite

(Fig. 1.3a), has attracted huge research attention since its discovery in 2004 [6]. It has unique

band structure with a linear dispersion relation near the Brillouin corner (Dirac point), leading to

tunable electron and hole doping [7, 8]. Graphene thus exhibits extraordinary electronic and

FIG. 1.2: Power density vs. time for computer processors manufactured by AMD, Intel, and

Power PC over the past two decades. The exponential trend in power density, although flattened

by the introduction of multi-core CPUs, is a limiting factor for the future scaling of semiconduc-

tor technology. Adapted from Ref. [1].

1

10

100

1990 1994 1998 2002 2006 2010

AMD

Intel

Power PC

CP

U P

ow

er D

ensi

ty (

W/c

m2)

Year

Pentium 4(2005)

Core 2 Duo(2006) Atom

(2008)

Page 9: © 2014 Zuanyi Li

3

optical properties like high carrier mobility ~104 cm

2V

-1s

-1 and optical transparency, and presents

promising applications in electronics, optoelectronics, and photonics [9-15]. A drawback of

graphene is the absence of a band gap, which leads to low on/off ratios in transistor performance.

Stimulated by the extensive studies of graphene, other 2D materials start to be fabricated and

investigated. For example, hexagonal boron nitride (h-BN, see Fig. 1.3b) is a 2D insulator (band

gap ~5 eV) and serves as the best dielectric substrate for graphene, because it reduces surface

roughness compared to other substrates like SiO2 [16]. In addition, transition metal dichalcogen-

ides (TMDs, with the form of MX2: M = transition metal, and X = S, Se, or Te) can be exfoliated

into 2D layers (Fig. 1.3c), and they are mostly semiconductors. Among them, 2D molybdenum

disulfide MoS2 was first achieved in experiments and shown a high on/off ratio of ~108 when

used as transistors [17], due to the presence of a band gap ~1.8 eV [18, 19].

These 2D materials can be used independently or further assembled layer-by-layer into van

der Waals (vdW) heterostructures [20] (Fig. 1.3d), which widens their potential applications like

hetero-junctions and tunneling devices. However, compared to extensive studies of the electrical

FIG. 1.3: Schematic structures of (a) graphene [8], (b) single layer of hexagonal boron nitride,

and (c) single layer of transition-metal dichalcogenides MX2 [17], where M is transition metal,

and X is chalcogen atoms. (d) Schematic of vdW heterostructure assembled by graphene, h-BN,

MoS2, and WSe2 [20].

B N

X

M

a

c

bgraphene h-BN

TMD

d vdW heterostructure

graphene

h-BN

graphene

WSe2

MoS2

Page 10: © 2014 Zuanyi Li

4

and optical properties of these 2D materials, our understanding of their thermal properties is

relatively lacking, especially for micro and nanoscale in which these materials are used for

devices [21-24]. Their thermal properties inevitably affect heat dissipation in devices, and hence

should be systematically and carefully investigated to understand new physics in micro and

nanoscale as well as explore possible applications in thermal management and thermoelectric

energy conversion.

1.3 Organization of Dissertation

In this dissertation, we investigated thermal transport in graphene and several other 2D lay-

ered materials, with a focus on nanoscale ballistic conduction and size effects. This dissertation

is a combination of experiments, finite element simulations, and theoretical calculations.

In Chapter 2, we briefly reviewed experimental studies on thermal and thermoelectric prop-

erties of graphene, including up-to-date experimental methods, intrinsic and extrinsic thermal

conductivity, interface thermal conductance, and thermoelectric power.

In Chapter 3, we measured thermal conduction in SiO2-supported graphene and graphene

nanoribbons by the substrate-supported thermometry platform. Quasi-ballistic thermal conduc-

tion was observed in quarter-micron long graphene up to room temperature. Graphene nanorib-

bons show clear width-dependent thermal conductivity due to phonon scattering with edge

disorder.

In Chapter 4, we discussed the applicability of the substrate-supported thermometry plat-

form to other materials, based on finite element simulations. We estimated the lower limit of the

measurable thermal sheet conductance by this method, and demonstrated it can be used to

measure most thin films and some nanomaterials like nanotube/nanowire arrays, even a single

nanowire.

Page 11: © 2014 Zuanyi Li

5

In Chapter 5, we calculated the ballistic thermal conductances of graphene, h-BN, MoS2,

and WS2 (including their monolayers and bulk) based on full phonon dispersions. From obtained

ballistic thermal conductance, we estimated their average phonon mean free paths and length-

dependent thermal conductivity due to ballistic-diffusive transition.

In Chapter 6, we conclude the dissertation with a summary of key findings and discussions

of possible future directions.

Page 12: © 2014 Zuanyi Li

6

CHAPTER 2

REVIEW OF THERMAL AND THERMOELECTRIC PROPERTIES OF GRA-

PHENE*

Since graphene is a single atomic layer of graphite, its thermal properties can be understood

from those of graphite. Figure 2.1a shows the atomic structure of the typical AB (i.e., Bernal)

stacking of graphene layers in graphite. The adjacent carbon atoms within one layer are connect-

ed by strong covalent sp2 bonds with a bonding energy of approximately 5.9 eV [27]. In contrast,

the adjacent graphene layers are stacked via weak van der Waals interactions (~50 meV) [27].

The lattice vibrational modes (phonons) of graphene, directly related to its thermal properties,

are described by the phonon dispersion in the Brillouin zone, as shown in Fig. 2.1b. Due to N = 2

carbon atoms in graphene unit cell, the dispersion contains 3 acoustic (A) and 3N-3 = 3 optical

(O) phonon modes. Besides the usual longitudinal (L) and transverse (T) modes, the unique 2D

nature of graphene leads to flexural (Z) phonon modes, corresponding to out-of-plane atomic

* This chapter is reprinted from Y. Xu, Z. Li, and W. Duan, “Thermal and Thermoelectric Properties of Graphene,”

Small, vol. 10, pp. 2182-2199 (2014). Copyright 2014, Wiley-VCH Verlag GmbH & Co.

FIG. 2.1: (a) Schematic of atomic arrangement in graphene sheets [23]. Dashed lines in the

bottom sheet represent the outline of the unit cell. (b) Calculated phonon dispersions along high-

symmetry lines for monolayer graphene, where red circles [25] and blue triangles [26] are

experimental data plotted for comparison. Inset is the Brillouin zone of graphene.

0

50

100

150

200

Phonon E

(m

eV

)

LO

LA

TA

TO

ZA

ZO

Γ M

K

A plane

B plane

h = 3.35 Å

A planea b

Page 13: © 2014 Zuanyi Li

7

displacements. At low phonon wave vector q, the TA and LA modes have linear dispersions of

ωTA ≈ vTAq and ωLA ≈ vLAq, where vTA ≈ 14 km/s and vLA ≈ 21 km/s [28]. These group velocities

are 4-6 times higher than those in Si or Ge due to the strong in-plane sp2 bonds and small mass of

carbon atoms [23]. Whereas, ZA modes show an approximately quadratic dispersion, ωZA ≈ αq2

with α ≈ 6.2 × 10-7

m2/s [29].

2.1 Methods to Measure Thermal Conductivity

Thermal conduction through a solid material is usually modeled on the basis of Fourier's law

that relates the local heat flux J to the local temperature gradient T: J = −k T, where k is the

thermal conductivity and has the units of [Wm-1

K-1

]. The minus sign in front of k implies that

heat flows in the direction opposite to that of the temperature gradient i.e., heat flows from the

hotter to the colder regions of the solid.

Measuring nanoscale thermal transport is quite challenging due to high requirements of

sample fabrication and temperature sensing [30, 31]. So far, methods used to probe thermal

conduction in graphene include optothermal Raman thermometry [32-43], thermoreflectance

technique [44-47], 3ω method [48], micro-resistance thermometry [49-57], electrical self-heating

method [58, 59], and scanning thermal microscopy (SThM) [60-62]. Here we mainly discuss and

compare two techniques widely used to measure in-plane thermal conductivity of graphene, i.e.,

optothermal Raman thermometry and micro-resistance thermometry.

2.1.1 Optothermal Raman Thermometry

The optothermal Raman thermometry technique was developed to measure suspended, mi-

crometer scale (> 2 µm) graphene. A laser light was focused at the center of the suspended

graphene flake to generate a heating power PH and raise temperature locally (Figs. 2.2a-b).

Meanwhile, the Raman spectrum of graphene was recorded and the temperature rise ΔT could be

Page 14: © 2014 Zuanyi Li

8

monitored by calibrating it with Raman G peak position [32-36], 2D peak position [37-40] (only

for monolayer), or Stokes/anti-Stokes ratio [41, 42]. By knowing the correlation between PH and

ΔT, as well as the geometry size of suspended graphene, its in-plane thermal conductivity k can

be extracted through the solution of the heat diffusion equation. The early experiments [32-35]

were carried out on graphene strips suspended over a trench (Fig. 2.2a). This was modified in

subsequent experiments [36-41], by adopting circular holes with graphene over them (Fig. 2.2b),

which matches with the radial symmetry of the laser spot and allowed for an analytic solution of

the temperature distribution [22].

FIG. 2.2: (a) Schematic of the optothermal Raman thermometry set-up, where a graphene strip is

suspended over a trench and heated up by a focused laser light [34]. (b) Schematic of the Raman

thermometry set-up with addition of a laser power meter to measure optical transmittance. Inset

is the Raman G peak map of graphene suspended over a circular hole [36]. (c) Scanning electron

microscopy (SEM) image of micro-resistance thermometry device with SLG supported on a

suspended SiO2 membrane between thermometers [49]. Scale bar is 3 µm. (d) Optical image of

bilayer graphene (BLG) suspended over two thermometer pads [53]. Scale bar is 10 µm. (e)

SEM image of a SiO2/Si-supported micro-resistance thermometry device to measure encased

few-layer graphene (FLG) [54]. (f) False-colored SEM image of a GNR array on SiO2/Si with

micro-resistance thermometers. Inset is a zoom-in atomic force microscopy (AFM) image of

GNRs [55].

1 μm

ca

b d

e

f

suspended

BLG

suspended

SLG/SiO2

thermometers

encased FLG

thermometerheater

thermometersheater

Page 15: © 2014 Zuanyi Li

9

A major source of uncertainty in different works using the Raman thermometry technique is

determining the laser power absorbed by graphene, that is, determining optical absorbance of

graphene. Balandin et al. [33] and Ghosh et al. [32, 34, 35] evaluated this number by comparing

the integrated Raman G peak intensity of graphene with that of highly oriented pyrolytic graphite

(HOPG), leading to ~9% if converted to optical absorbance for exfoliated single-layer graphene

(SLG) [22]. The measured value of ~9% considered two passes of light (down and reflected back)

and resonance absorption effects due to close proximity of graphene to the substrate, and corre-

sponded to 488-nm wavelength where absorption is higher than the 2.3% long-wavelength limit

[63]. Faugeras et al. [41], Lee et al. [37], and Vlassiouk et al. [42] did not measure the optical

absorbance under the conditions of their experiments and assumed a value of 2.3% for exfoliated

SLG based on a separate optical transmission measurement [63]. Cai et al. [36] and Chen et al.

[38] obtained values of 3.3±1.1% and 3.4±0.7% for CVD SLG by directly measuring the optical

transmittance via addition of a power meter under the suspended portion of graphene (Fig. 2.2b).

The used optical absorbance is very important because it would proportionally change extracted

k. We note that both theory [64] and experiments [65-67] showed an increase of optical absorb-

ance in graphene with decreasing laser wavelength due to many-body effect, and the values of

Cai et al. [36] and Chen et al. [38] are consistent with those experimental results. To obtain

reliable k, it is thus necessary to measure optical absorbance under used laser wavelength and

specific experimental conditions.

Another uncertainty source is the calibration of temperature with features of Raman spec-

trum. It is known that strains and impurities in graphene can affect the Raman peak positions and

their temperature dependence [68], which greatly limits the temperature sensitivity of the Raman

thermometry technique. In addition, heat loss from graphene to the surrounding air was neglect-

Page 16: © 2014 Zuanyi Li

10

ed in most experiments, but Chen et al. [38] found that for a large diameter (9.7 µm) graphene

flake, the k obtained in air could be overestimated by 14−40% compared with the value obtained

in vacuum. This implies measurable errors in previous experiments, even though the influence

might be weaker due to smaller sizes of measured graphene. Furthermore, extra uncertainty

could come from the difference in k between suspended and supported portions of graphene, as

well as thermal boundary resistance between graphene and supporting substrates [36]. Overall,

the Raman thermometry technique provides an efficient way to measure k of suspended graphene

with benefits of relatively easy sample fabrication, reduced graphene contamination, and simple

data analysis, but it inevitably has limitations: (i) relatively large uncertainty (up to 40%) [21]; (ii)

difficulty to probe the low temperature regime due to significant heating in graphene by laser; (iii)

inability to be applied to nanometer scale or supported graphene, where edge and interface will

take effect.

2.1.2 Micro-Resistance Thermometry

The micro-resistance thermometry technique is a steady-state method to directly probe heat

flows in materials [30]. It is able to measure both suspended and supported graphene, as well as

at the nanometer scale and in low temperature range, with high resolution of temperature by

employing electrical resistance as thermometers. This technique can be further divided into two

kinds: suspended bridge platform and fully substrate-supported platform. The former was

developed by Shi et al. [69, 70] to measure thermal conductivity of 1D nanostructures, and it has

been widely used for nanotubes [71-76] and nanowires [77-81]. By using this platform, graphene

can be either supported by a suspended SiO2/SiNx membrane connecting two thermometers [49-

52] (Fig. 2.2c) or fully suspended over two thermometer pads [51, 53, 57] (Fig. 2.2d), enabling

measurements of both suspended and supported graphene. The fully substrate-supported plat-

Page 17: © 2014 Zuanyi Li

11

form was developed by Jang et al. [54] and Bae et al. [55], where at least two thermometers were

patterned on Si/SiO2-supported graphene (Fig. 2.2e) or GNRs (Fig. 2.2f). In both platforms, one

thermometer serves as the heater to generate heating power PH and a temperature gradient across

graphene by electrical heating, meanwhile all thermometers (including the heater) monitor

temperature changes ΔT in terms of their electrical resistance changes. Then, the thermal con-

ductance/conductivity of measured materials can be extracted in a simple analytic way by

solving its equivalent thermal resistance circuit for the suspended bridge platform [69, 70], while

a complicated 3D numerical (finite element) simulation has to be performed for the substrate-

supported platform due to significant heat leakage into the substrate [54, 55].

Although the data extraction of the suspended bridge platform is easier than that of the sub-

strate-supported platform, as a trade-off the sample fabrication is more complicated for the

former than the latter. Thus, for materials which are hard to be suspended, such as GNRs, the

latter is an advantageous method to be employed. However, the measurable length of interested

materials cannot be longer than a few micrometer for the substrate-supported platform, because

the temperature drops nearly exponentially away from the heater, leading to undetectable ΔT if

other thermometers are far away. It is worth noting that for the platform of graphene supported

by a suspended membrane (Fig. 2.2c) and the substrate-supported platform (Figs. 2.2e and 2.2f),

a control experiment has to be carried out by etching off graphene/GNRs and repeating meas-

urements to calibrate the background heat flow and thermal contact resistance between graphene

and thermometers [49, 50, 54, 55]. This improves the measurement accuracy. Whereas, for the

platform of graphene fully suspended (Fig. 2.2d) such a control experiment cannot be performed,

so the thermal contact resistance could be a main source of uncertainty in results. Overall, the

micro-resistance thermometry technique has very high resolution of temperature (<50 mK) [21,

Page 18: © 2014 Zuanyi Li

12

22] and can cover a wide temperature range. It can probe both suspended and supported gra-

phene, as well as in nanometer scale. However, attention should be paid to micro/nanofabrication,

which could introduce contaminations (like residues and defects) and increase the uncertainty

[53].

2.2 Intrinsic Thermal Conductivity of Graphene

In this section, we mainly discuss the “intrinsic” thermal conductivity of SLG based on ex-

perimental results and theoretical analysis. Here, by “intrinsic” we mean isolated, large scale,

pristine graphene without suffering impurity, defect, interface, and edge scatterings, so its

thermal conductivity is only limited by intrinsic phonon-phonon scattering due to crystal anhar-

monicity [21] and electron-phonon scattering. In experiments, suspended, micrometer scale

graphene samples have properties close to intrinsic ones. We thus first summarize current

experimental observations of k in suspended SLG.

Using the Raman thermometry technique described above, suspended micro-scale graphene

flakes obtained by both exfoliation from graphite [32-35, 37, 41] and CVD growth [36, 38-40, 42]

have been measured at room temperature and above. Some representative data versus tempera-

ture from these studies are shown in Fig. 2.3a. The obtained in-plane thermal conductivity values

of suspended SLG generally fall in the range of ~2000-4000 Wm-1

K-1

at room temperature, and

decrease with increasing temperature, reaching about 700-1500 Wm-1

K-1

at ~500 K. The varia-

tion of obtained values could be attributed to different choices of graphene optical absorbance

(see Section 2.1), thermal contact resistance, different sample geometries, sizes, and qualities.

For comparison, we also plot the experimental thermal conductivity of diamond [82], graphite

[82], and carbon nanotubes (CNTs) [71, 83] in Fig. 2.3a.

Page 19: © 2014 Zuanyi Li

13

It is clear that suspended graphene has thermal conductivity as high as these carbon allo-

tropes near room temperature, even higher than its 3D counterpart, graphite, whose highest

record of observed in-plane k in HOPG is ~2000 Wm-1

K-1

at 300 K. The presently available data

of graphene based on the Raman thermometry technique only cover the temperature range of

~300-600 K, except one at ~660 K reported by Faugeras et al. [41] showing k ≈ 630 Wm-1

K-1

.

For higher temperature, Dorgan et al. [58] used the electrical breakdown method for thermal

conductivity measurements and found k ≈ 310 Wm-1

K-1

at 1000 K for suspended SLG. The

FIG. 2.3: (a) Experimental thermal conductivity k as a function of temperature T: representative

data for suspended CVD SLG by Chen et al. [38] (solid red square), suspended exfoliated SLG

by Lee et al. [37] (solid purple asterisk) and Faugeras et al. [41] (solid brown pentagon), sus-

pended SLG by Dorgan et al. [58] (solid grey hexagon), suspended exfoliated BLG by Pettes et

al. [53] (solid orange diamond), supported exfoliated SLG by Seol et al. [49] (solid black circle),

supported CVD SLG by Cai et al. [36] (solid blue right-triangle), encased exfoliated 3-layer

graphene (3LG) by Jang et al. [54] (solid cyan left-triangle), supported exfoliated GNR of W ≈

65 nm by Bae et al. [55] (solid magenta square), type IIa diamond [82] (open gold diamond),

graphite in-plane [82] (open blue up-triangle), graphite cross-plane (open blue down-triangle),

suspended single-walled CNT (SWCNT) by Pop et al. [83] (open dark-green circle), and multi-

walled CNT (MWCNT) by Kim et al. [71] (solid light-green circle). (b) Thermal conductance

per unit cross-sectional area, G/A=k/L, converted from thermal conductivity data in (a), com-

pared with the theoretical ballistic limit of graphene (solid line), which can be approximated

analytically as Gball/A ≈ [1/(4.4×105T

1.68)+1/(1.2×10

10)]

-1 Wm

-2K

-1 over the temperature range

1−1000 K [55]. Data in (a) whose sample L is unknown or not applicable are not shown in (b).

100 1000

101

102

103

104

50

k (

Wm

-1K

-1)

T (K)300

encased

3LG

suspended

SLG

Graphite (||)

supported

GNR, W ≈ 65 nm

Graphite (┴)

suspended

BLG

supported

SLG

100 100010

7

108

109

1010

G/A

(W

m-2

K-1

)

T (K)40 300

a b

suspended

SLG

supported

SLG

G/A = k/L

T1.5

T 2

T -1

Page 20: © 2014 Zuanyi Li

14

overall trend of present graphene data from 300 K to 1000 K shows a steeper temperature

dependence than graphite (see Fig. 2.3a), consistent with the extrapolation of thermal conductivi-

ty by Dorgan et al. [58]. This behavior could be attributed to stronger second-order three-phonon

scattering (τ ~ T -2

) in graphene than graphite enabled by the flexural (ZA) phonons of suspended

graphene [84], similar to the observations in CNTs [83, 85]. For temperature below 300 K, the

micro-resistance thermometry technique needs to be employed for k measurements. Unfortunate-

ly, there is no reliable data for suspended single-layer graphene until now. However, data do

exist for suspended few-layer graphene (FLG) [51, 53, 56], which will be discussed in Section

2.3.4. It is instructive to compare experimental results with the ballistic limit of graphene as a

check, so we convert measured k in Fig. 2.3a to thermal conductance per unit cross-sectional area,

G/A = k/L, which are re-plotted in Fig. 2.3b with graphene Gball/A (discussed in the next section).

Above room temperature, measured G/A of suspended SLG are more than one order of magni-

tude lower than Gball/A, indicating the diffusive regime. The reason that the value of Faugeras et

al. [41] is much lower than others is because of a much larger L = 22 µm (radius) of their sus-

pended graphene.

2.3 Extrinsic Thermal Conductivity of Graphene

The long phonon MFP in pristine graphene would suggest that it is possible to tune thermal

conductivity more effectively by introducing extrinsic scattering mechanisms which dominate

over intrinsic scattering mechanisms in graphene. For example, isotope scattering, normally

unimportant with respect to other scattering processes, could become significant in graphene

thermal conduction. It could be easier to observe size effects on thermal transport in graphene

because samples do not have to be extremely shrunk. In the following, we discuss various

scattering mechanisms and their influences on thermal conduction separately, giving rise to

Page 21: © 2014 Zuanyi Li

15

tunable extrinsic thermal conductivity of graphene.

2.3.1 Isotope Effects

The knowledge of isotope effects on thermal transport properties is valuable for tuning heat

conduction in graphene. Natural abundance carbon materials are made up of two stable isotopes

of 12

C (98.9%) and 13

C (1.1%). Changing isotope composition can modify dynamic properties of

crystal lattices and affect their thermal conductivity [89, 90]. For instance, it has been found that

at room temperature isotopically purified diamond has a thermal conductivity of ~3300 Wm-1

K-1

[91, 92], about 50% higher than that of natural diamond, ~2200 Wm-1

K-1

[82]. Similar effects

have also been observed in 1D nanostructures, boron nitride nanotubes [73]. Very recently, the

first experimental work to show the isotope effect on graphene thermal conduction was reported

by Chen et al. [39]. By using the CVD technique, they synthesized isotopically modified gra-

phene containing various percentages of 13

C. The graphene flakes were subsequently suspended

over 2.8-µm-diameter holes and thermal conductivity was measured by the Raman thermometry

FIG. 2.4: (a) Thermal conductivity of suspended CVD graphene as a function of temperature for

different 13

C concentrations, showing isotope effect [39]. b) Thermal conductivity of suspended

CVD graphene with (red down-triangle) and without (blue up-triangle) wrinkles as a function of

temperature. Also shown in comparison are the literature thermal conductivity data of pyrolytic

graphite samples [86-88]. Inset shows the SEM image of CVD graphene on the Au-coated SiNx

holey membrane. The red arrow indicates a wrinkle. Scale bar is 10 µm [40].

a b

without wrinkle

with wrinkle

6000

300 350 400 450 500 550 600 650

5000

4000

3000

2000

1000

T (K)

k(W

m-1

K-1

)

0 100 200 300 400 500 600

4000

3000

2000

1000

T (K)

k(W

m-1

K-1

)

0

Page 22: © 2014 Zuanyi Li

16

technique. As shown in Fig. 2.4a, compared with natural abundance graphene (1.1% 13

C), the k

values were enhanced in isotopically purified samples (0.01% 13

C), and reduced in isotopically

mixed ones (50% 13

C).

2.3.2 Structural Defect Effects

Structural defects are common in fabricated graphene, especially in CVD grown graphene

[93, 94]. The effects of wrinkles [40] and grain size [42] on the thermal conduction of suspended

single-layer CVD graphene have been examined in experiments by using the Raman thermome-

try technique. Chen et al. [40] found that the thermal conductivity of graphene with obvious

wrinkles (indicated by arrows in the inset of Fig. 2.4b) is about 15-30% lower than that of

wrinkle-free graphene over their measured temperature range, ~330-520 K (Fig. 2.4b). Vlassiouk

et al. [42] measured suspended graphene with different grain sizes obtained by changing the

temperature of CVD growth. The grain sizes ℓG were estimated to be 150 nm, 38 nm, and 1.3 nm

in different samples in terms of the intensity ratio of the G peak to D peak in Raman spectra [95].

Since grain boundaries in graphene serve as extended defects and scatter phonons, graphene with

smaller grain sizes are expected to suffer more frequent phonon scattering. Their measured

thermal conductivity does show a decrease for smaller grain sizes, indicating the observation of

the grain boundary effect on thermal conduction. Whereas, the dependence on the grain size

shows a weak power law, k ~ ℓG1/3

, for which there is no theoretical explanation yet [42]. How-

ever, for SiO2-supported graphene, recent theoretical work based on NEGF method showed a

similar but stronger dependence of k on the grain size ℓG in the range of ℓG < 1 µm [96]. Further

experimental studies are required to reveal the grain size effects on the thermal transport of both

suspended and supported graphene.

2.3.3 Substrate Effects in Supported Graphene

Page 23: © 2014 Zuanyi Li

17

For practical applications, graphene is usually attached to a substrate in electronic and op-

toelectronic devices, so it is important to understand substrate effects on thermal properties of

supported graphene [97-100]. Seol et al. [49, 50] measured exfoliated SLG on a 300 nm thick

SiO2 membrane by using the micro-resistance thermometry technique with the suspended bridge

platform (Fig. 2.2c). The observed thermal conductivity is k ~ 600 Wm-1

K-1

near room tempera-

ture (solid black circles in Fig. 2.3a). This value is much lower than those reported for suspended

SLG via the Raman thermometry technique, but is still relatively high compared with those of

bulk silicon (~150 Wm-1

K-1

) and copper (~400 Wm-1

K-1

). Another study by Cai et al. [36]

showed CVD SLG supported on Au also has a decreased thermal conductivity, ~370 Wm-1

K-1

(this lower value compared to ~600 Wm-1

K-1

could be caused by grain boundary scattering in

CVD graphene [96]). The thermal conductivity reduction in supported graphene is attributed to

substrate scattering, which strongly affects the out-of-plane flexural (ZA) mode of graphene [49,

97, 101]. This effect becomes stronger in encased graphene, where graphene is sandwiched

between bottom and top SiO2. The thermal conductivity of SiO2-encased exfoliated SLG was

measured to be below 160 Wm-1

K-1

, reported by Jang et al. [54] using the micro-resistance

thermometry technique with the substrate-supported platform (Fig. 2.2e). For encased graphene,

besides the phonon scattering by bottom and top oxides, the evaporation of top oxide could cause

defects in graphene, which can further lower thermal conductivity. Knowing encased graphene k

is useful for analyzing heat dissipation in top-gated graphene devices.

2.3.4 Interlayer Effects in Few-Layer Graphene

Interlayer scattering as well as top and bottom boundary scattering could take place in few-

layer graphene, which could be another mechanism to modulate graphene thermal conductivity.

It is interesting to investigate the evolution of the thermal conductivity of FLG with increasing

Page 24: © 2014 Zuanyi Li

18

thickness, denoted by the number of atomic layers (N), and the critical thickness needed to

recover the thermal conductivity of graphite.

Several experimental studies on this topic have been conducted for encased [54], supported

[52], and suspended FLG [35, 56], and their results are summarized in Fig. 2.5. Jang et al. [54]

measured the thermal transport of SiO2-encased FLG by using the substrate-supported, micro-

resistance thermometry platform (Fig. 2.2e). They found that the room-temperature thermal

conductivity increases from ~50 to ~1000 Wm-1

K-1

as the FLG thickness increases from 2 to 21

layers, showing a trend to recover natural graphite k. This strong thickness dependence was

explained by the top and bottom boundary scattering and disorder penetration into FLG induced

by the evaporated top oxide [54]. Very recently, another similar yet less pronounced trend was

observed in SiO2-supported FLG by Sadeghi et al. [52] using the suspended micro-resistance

FIG. 2.5: Experimental in-plane thermal conductivity near room temperature as a function of the

number of layers N for suspended graphene by Ghosh et al. [35] (open blue diamond) and by

Jang et al. [56] (open green square), SiO2-supported graphene by Seol et al. [49] (open red circle)

and Sadeghi et al. [52] (solid red circle), and SiO2-encased graphene by Jang et al. [54] (solid

black triangle). The data show a trend to recover the value (dashed line) measured by Sadeghi et

al. [52] for natural graphite source used to exfoliate graphene. The gray shaded area shows the

highest reported k values of pyrolytic graphite [86-88]. Adapted from Ref. [52].

1 10

100

1000

k (

Wm

-1K

-1)

number of layers

Suspended

Ghosh et al.

Encased

Jang et al.

Suspended

Jang et al. Supported

Sadeghi et al.

Natural graphite

T ~ 300 K

Page 25: © 2014 Zuanyi Li

19

thermometry platform (similar to Fig. 2.2c). As shown by red dots in Fig. 2.5, the measured

room-temperature k increases slowly as increasing thickness, and the recovery to natural graphite

would occur even more than 34 layers. The difference between the results by Jang et al. [54] and

Sadeghi et al. [52] is not unexpected, because encased FLG k could be suppressed much more in

thin layers than thick layers due to the effect of top oxide, and hence shows a stronger thickness

dependence.

For suspended FLG, there are two contradictory observations in the thickness dependence.

At first, based on the Raman thermometry technique (Fig. 2.2a) Ghosh et al. [35] showed a

decrease of suspended FLG k from the SLG high value to regular graphite value as thickness

increases from 2 to 8 layers (open diamonds in Fig. 2.5). The k reduction was explained by the

interlayer coupling and increased phase-space states available for the phonon Umklapp scattering

in thicker FLG [35]. However, a very recent study by Jang et al. [56] seems to show a different

thickness trend for suspended FLG. They measured thermal conductivity of suspended graphene

of 2-4 and 8 layers by using a modified T-bridge micro-resistance thermometry technique. The

obtained room-temperature k for 2-4 layers is about 300-400 Wm-1

K-1

with no apparent thickness

dependence, while k for 8-layer shows an increase to ~600 Wm-1

K-1

(open squares in Fig. 2.5).

Surprisingly, this trend is qualitatively in agreement with that of Sadeghi et al. [52] for supported

FLG; both show similar increasing amounts of k from 2 to 8 layers (Fig. 2.5), despite a small

decrease from 2 to 4 layers in the former, which could arise from different sample qualities and

measurement uncertainty. Given opposite thickness trends of Ghosh et al. [35] and Jang et al.

[56], further experimental works are required to clarify the real thickness-dependent k in sus-

pended FLG. Moreover, we want to point out that suspended FLG k values of Jang et al. [56] are

close to those reported by Pettes et al. [53] for suspended bilayer graphene (BLG), ~600 Wm-1

K-

Page 26: © 2014 Zuanyi Li

20

1 at room temperature (see Fig. 2.3a). Both are much lower than suspended SLG k. The latter

attributed this to phonon scattering by a residual polymeric layer on graphene [53], even though

the former claimed that electrical current annealing was used to remove polymer residues [56].

2.3.5 Cross-Plane Thermal Conduction

A remarkable feature of graphite and graphene is that their thermal properties are highly ani-

sotropic. Despite high thermal conductivity along the in-plane direction, heat flow along the

cross-plane direction (c-axis) is hundreds of times weaker, limited by weak van der Waals

interactions between layers (for graphite) or with adjacent materials (for graphene). For example,

the thermal conductivity along the c-axis of pyrolytic graphite is only ~6 Wm-1

K-1

at room

temperature [82] (Fig. 2.3a). For graphene, it is often attached to a substrate or embedded in a

FIG. 2.6: Experimental thermal interface conductance G vs. temperature for SLG/SiO2 by Chen

et al. [48] (open purple diamond), FLG/SiO2 by Mak et al. [44] (open purple square), CNT/SiO2

by Pop et al. [102] (solid purple right-triangle), Au/SLG by Cai et al. [36] (solid gold diamond),

Au/Ti/SLG/SiO2 (solid blue circle) and Au/Ti/graphite (solid orange circle) by Koh et al. [45],

interfaces of graphite with Au (solid magenta square), Al (solid gray up-triangle), Ti (solid green

asterisk) by Schmidt et al. [103], interfaces of Al/SLG/SiO2 without treatment (open black up-

triangle), with oxygen treatment (Al/O-SLG/SiO2, open green up-triangle), and with hydrogen

treatment (Al/H-SLG/SiO2, open red up-triangle) by Hopkins et al. [46].

100 100010

100

50

20

40

G (

MW

m-2

K -

1)

T (K)

300

Al/O-SLG/SiO2

Au/SLG

Au/Graphite

Au/Ti/SLG/SiO2

Al/SLG/SiO2

Al/H-SLG/SiO2

SLG/SiO2

CNT/SiO2

Ti/Graphite

FLG/SiO2

Au/Ti/Graphite

Page 27: © 2014 Zuanyi Li

21

medium for potential applications. Heat conduction along the cross-plane direction is character-

ized by the thermal interface/boundary conductance between graphene and adjacent materials,

which could become a limiting dissipation bottleneck in highly scaled graphene devices and

interconnects [23, 104-107].

The thermal interface conductance across graphene/graphite and other materials has been

measured by using 3ω method [48], time-domain thermoreflectance (TDTR) technique [44-47,

103, 108], and Raman-based method [36, 43]. Most experimental data available to date are

shown in Fig. 2.6, and they are consistent with each other in general, given the variations of

sample qualities and measurement techniques. Chen et al. [48] and Mak et al. [44] showed the

thermal interface conductance per unit area of graphene/SiO2 is G ~ 50-100 MWm-2

K-1

at room

temperature, with no strong dependence on the FLG thickness. Their values are close to that of

CNT/SiO2 [102], reflecting the similarity between graphene and CNT. Schmidt et al. [103]

measured G of the graphite/metal interfaces, including Au, Cr, Al, and Ti. Among them, the

graphite/Ti has the highest G ~120 MWm-2

K-1

, and the graphite/Au interface has the lowest G

~30 MWm-2

K-1

near room temperature. Their G of graphite/Au is consistent with the value by

Norris et al. [108] and values of SLG/Au by Cai et al. [36] and FLG/Au by Ermakov et al. [43].

Koh et al. [45] later measured heat flow across the Au/Ti/N-LG/SiO2 interfaces with the layer

number N=1−10. Their observed room-temperature G is ~25 MWm-2

K-1

, which shows a very

weak dependence on the layer number N and is equivalent to the total thermal conductance of

Au/Ti/graphite and graphene/SiO2 interfaces acting in series. This indicates that the thermal

resistance of two interfaces between graphene and its environment dominates over that between

graphene layers. Interestingly, Hopkins et al. [46] showed the thermal conduction across the

Al/SLG/SiO2 interface could be manipulated by introducing chemical adsorbates between the Al

Page 28: © 2014 Zuanyi Li

22

and SLG. As shown in Fig. 2.6, their measured G of untreated Al/SLG/SiO2 is ~30 MWm-2

K-1

at room temperature, in agreement with Zhang et al. [47]. The G increases to ~42 MWm-2

K-1

for oxygen-functionalized graphene (O-SLG), while decreases to ~23 MWm-2

K-1

for hydrogen-

functionalized graphene (H-SLG). These effects were attributed to changes in chemical bonding

between the metal and graphene, and are consistent with the observed enhancement in G from

the Al/diamond [109] to Al/O-diamond interfaces [110].

2.4 Thermoelectric Properties of Graphene

Thermoelectric materials can convert waste heat into electricity by the Seebeck effect and

use electricity to drive electronic cooling or heating by the Peltier effect. Thermoelectric devices

are all-solid-state devices with no moving parts, thus are silent, reliable and scalable. However,

they only find limited applications due to their low efficiency. The efficiency of a thermoelectric

material is determined by the thermoelectric figure of merit (ZT), which typically is defined as

2S

ZT Tk

(2.1)

where σ is the electrical conductivity, S is the Seebeck coefficient [also called thermoelectric

power (TEP) or thermopower], T is the absolute temperature, and the thermal conductivity k =

ke+kl have contributions from electrons (ke) and lattice vibrations (kl). ke is usually extracted

based on the Wiedemann-Franz Law ke/σ = L0T, where the Lorenz number L0 is equal to

2.44×10-8

WΩK-2

for free electrons. This law does not always hold. For example, ke becomes

zero for a delta-shaped transport distribution [111]. However, in graphene ke is negligible with

respect to kl [112-116], similar to CNTs [71, 117]. Currently, the state-of-art commercial

thermoelectric materials, like Bi2Ti3, have room-temperature ZT around 1 [3].

Page 29: © 2014 Zuanyi Li

23

Thermoelectric transport in graphene has been experimentally investigated in the past five

years [118-129]. The Seebeck coefficient S and electrical conductance Ge of graphene can be

measured against the gate voltage Vg (thus, carrier density nc) simultaneously via a widely-used

microfabricated structure (inset of Fig. 2.7a), which was developed by Small et al. [130] to

measure thermoelectric transport in CNTs. Figure 2.7 shows typical results of measured Ge and S

as a function of Vg in graphene at different temperatures [118]. The Seebeck coefficient S shows

two peaks near the Dirac point (charge neutrality point) and changes its sign across the Dirac

point as the majority carrier switches from electron to hole. The room-temperature peak values

of S for SLG and BLG are observed to be ~50-100 µVK-1

in different experiments [118-122].

For high carrier density nc (i.e., high |Vg|), the measured Seebeck coefficient scales as S~|nc|-1/2

for SLG due to its linear dispersion [119], while S~1/|nc| for BLG due to its hyperbolic disper-

FIG. 2.7: (a) Electrical Conductance Ge and (b) thermopower TEP of a graphene sample as a

function of back gate voltage Vg for T = 300 K (square), 150 K (circle), 80 K (up triangle), 40 K

(down triangle), and 10 K (diamond). Upper inset: SEM image of a typical device for thermoe-

lectric measurements, scale bar is 2 µm. Lower inset: TEP values taken at Vg = −30 V (square)

and −5 V (circle). Dashed lines are linear fits to the data. Adapted from Ref. [118].

Conduct

ance

a

b

Page 30: © 2014 Zuanyi Li

24

sion [121], consistent with theories. Importantly, the simultaneous measurements of Ge and S

enable testing the validation of the semiclassical Mott relation [131]:

2 2 1

3 | | F

geBE E

e g

dVdGk TS

e G dV dE

(2.2)

where kB is the Boltzmann constant, e is the electron charge, and EF is the Fermi energy. For

SLG the measured S shows a linear T dependence (inset of Fig. 2.7b) and matches calculated S

from measured Ge by Eq. 2.2 [118-120], indicating an agreement with the Mott relation. For

BLG, however, the agreement only holds for high carrier density; for low carrier density there is

an obvious difference between measured and calculated S as well as a deviation from the linear T

dependence at high temperature [121, 122]. This failure of the Mott relation was attributed to the

low Fermi temperature in BLG.

The thermoelectric properties of materials can be also probed by using a conducting tip to

measure the thermoelectric voltage between the sample and tip, induced by a given temperature

difference between them. By employing atomic force microscopy (AFM) and scanning tunneling

microscopy (STM) techniques, Cho et al. [132] and Park et al. [133] measured the thermopower

of epitaxial graphene on SiC, respectively. The advantage of this method is the simultaneous

imaging of the sample structure and thermoelectric signals with a spatial resolution of atomic-

scale. Since the Seebeck coefficient relies on the sample local density of states (LDOS) near the

Fermi energy, and LDOS can be quite different in the presence of boundaries and disorders [134-

138], thermoelectric imaging allows us to probe grain boundaries, wrinkles, defects, and impuri-

ties in graphene, which may not be reflected in topography images [132, 133].

For practical applications, the Seebeck coefficient and power factor σS 2

of graphene should

be improved. Some experimental efforts have been made in this direction. Wang et al. [121]

observed enhanced S below room temperature in a dual-gated BLG device, resulting from the

Page 31: © 2014 Zuanyi Li

25

opening of a band gap by applying a perpendicular electric field on BLG. Additionally, the

Seebeck coefficient and power factor of FLG could be enhanced at high temperature (>500 K)

by molecular attachments [139] and oxygen plasma treatment [140], attributed to the band gap

opening. By constructing the c-axis preferentially oriented nanoscale Sb2Te3 film on monolayer

graphene, both S and σ were increased, benefiting from a highway for carriers provided by

graphene [141]. From a practical point of view, Hewitt et al. [142] focused on maximizing the

power output of FLG/polyvinylidene fluoride composite thin films by considering the absolute

temperature, temperature gradient, load resistance, and physical dimensions of films.

Page 32: © 2014 Zuanyi Li

26

CHAPTER 3

BALLISTIC TO DIFFUSIVE CROSSOVER OF HEAT FLOW IN GRA-

PHENE RIBBONS*

3.1 Introduction

The thermal properties of graphene are derived from those of graphite, and are similarly ani-

sotropic. The in-plane thermal conductivity of isolated graphene is high, ~2000 Wm-1

K-1

at room

temperature, due to the strong sp2 bonding and relatively small mass of carbon atoms [21, 23, 24,

58]. Heat flow in the cross-plane direction is nearly a thousand times weaker, limited by van der

Waals interactions with the environment (for graphene) [45] or between graphene sheets (for

graphite) [21, 23]. Recent studies have suggested that the thermal conductivity of graphene is

altered when in contact with a substrate through the interaction between vibrational modes

(phonons) of graphene and those of the substrate [49, 54, 97, 101]. However an understanding of

heat flow properties in nanometer scale samples of graphene [or any other two-dimensional (2D)

materials] is currently lacking.

By comparison, most graphene studies have focused on its electrical properties when con-

fined to scales on the order of the carrier mean free path (MFP) [105, 143-149]. For example,

these have found that “short” devices exhibit near-ballistic behavior [145], Fabry-Perot wave

interference [146], and “narrow” nanoribbons display a steep reduction of charge carrier mobility

[144, 147]. Previous studies do exist for heat flow in three-dimensional (3D) structures like

nanowires and nanoscale films. For instance, ballistic heat flow was observed in suspended

GaAs bridges [150] and silicon nitride membranes [151] at low temperatures, of the order 1 K.

* This chapter is reprinted from M.-H Bae

†/Z. Li

†, Z. Aksamija, P. N. Martin, F. Xiong, Z.-Y. One, I.

Knezevic, and E. Pop, “Ballistic to Diffusive Crossover of Heat Flow in Graphene Ribbons,” Nature Communica-

tion, vol. 4, p. 1734 (2013). Copyright 2014, Macmillan Publishers Limited. †Denotes equal contribution.

Page 33: © 2014 Zuanyi Li

27

Conversely, suppression of thermal conductivity due to strong edge scattering effects was noted

in narrow and rough silicon nanowires [80, 81], up to room temperature. Yet such effects have

not been studied in 2D materials like graphene, and ballistic heat conduction has not been

previously observed near room temperature in any material.

In this work we find that the thermal properties of graphene can be tuned in nanoscale de-

vices comparable in size to the intrinsic phonon MFP. (By “intrinsic” thermal conductivity or

phonon MFP we refer to that in large samples without edge effects, typically limited by phonon-

phonon scattering in suspended graphene, and by substrate scattering in supported graphene, here

λ ≈ 100 nm at room temperature). The thermal conductance of “short” quarter-micron graphene

reaches up to 35% of theoretical ballistic upper limits [96]. However, the thermal conductivity of

“narrow” graphene nanoribbons (GNRs) is greatly reduced compared to that of “large” graphene

samples. Importantly, we uncover that nanoengineering the GNR dimensions and edges is

responsible for altering the effective phonon MFP, shifting heat flow from quasi-ballistic to

diffusive regimes. These findings are highly relevant for all nanoscale graphene devices and

interconnects, also suggesting new avenues to manipulate thermal transport in 2D and quasi-one-

dimensional (1D) systems.

3.2 Experimental Method

3.2.1 Fabrication and Characterization of Graphene Nanoribbons

Graphene monolayers were deposited on SiO2/Si (~290 nm/0.5 mm) substrates by mechani-

cal exfoliation from natural graphite. Graphene thickness and GNR edge disorder were evaluated

with Raman spectroscopy [45, 152, 153]. Samples were annealed in Ar/H2 at 400 °C for 40

minutes. We used two approaches to define and fabricate graphene nanoribbon (GNR) arrays

with pitch ~150 nm and varying widths: one with a PMMA mask (Fig. 3.1a), the other with an

Page 34: © 2014 Zuanyi Li

28

Al mask (Fig. 3.1b) [154]. Double poly(methyl methacrylate) (PMMA) layers (PMMA 495K

A2/PMMA 950K A4) were coated on the Si/SiO2 substrate. For the electron (e)-beam lithogra-

phy, we used 30 keV e-beam accelerating voltage. After opening 40 nm wide PMMA windows,

we etched the graphene exposed through the windows with an oxygen plasma, creating GNRs of

width W (Fig. 3.1a). This PMMA mask method was used for the W ≈ 130 nm, ~85 nm, and ~65

nm wide GNRs. For the narrower ~45 nm GNRs we used Al masks (Fig. 3.1b). In this case, after

opening the PMMA windows, instead of plasma etching, we deposited 30 nm thick Al and

obtained ~45 nm wide Al strips on graphene. After plasma etching of exposed graphene and Al

etching (type A, Transene Company) we obtained ~45 nm wide GNRs. Figure 3.2 shows the

atomic force microscopy (AFM) images of fabricated GNR arrays with W ≈ 130 nm, 85 nm, 65

nm, and 45 nm, respectively. The bottom and top regions of Figs. 3.2a and 3.2c correspond to the

un-etched pristine graphene.

FIG. 3.1: Process to define the GNR widths. (a) PMMA mask method. (b) Al mask method.

Plasma etching

W

Al deposition and lift-off Plasma etching

Al etching W W

a

b

PMMA

grapheneSiO2

Al

Page 35: © 2014 Zuanyi Li

29

To characterize the prepared GNRs, we performed Raman spectroscopy with a 633 nm

wavelength laser (~1 µm spot size) as shown in Fig. 3.3. Even before patterning into GNRs, we

selected only monolayer graphene flakes, identifiable through their 2D (G’) to G Raman peak

ratio, and through a single fitted Lorentzian to their 2D (G’) peak. The unpatterned graphene

samples had no identifiable D peak, indicative of little or no disorder [155]. On the other hand,

the GNR arrays showed a pronounced D band consistent with the presence of edge disorder

[152]. The peak intensity of the D band with respect to that of the G band increases with narrow-

er GNR width. Because the edges of graphene serve as defects by breaking the translational

symmetry of the lattice, the larger fraction of the edge in narrow GNRs will enhance the D peak

[156]. The inset of Fig. 3.3a quantitatively shows the behavior by calculating the ratio of inte-

grated D band (ID) to G band (IG), ID/IG, as a function of GNR width (symbols). The width

FIG. 3.2: Atomic force microscopy (AFM) images of GNR arrays. (a) W ~ 130 nm GNRs. (b) W

~ 85 nm GNRs. (c) W ~ 65 nm GNRs. Inset: AFM image near metal electrodes. (d) W ~ 45 nm

GNRs. The axis units are given in microns on each panel.

graphene

a

d

0.0 0.2 0.4 0.6 0.8 1.0 1.2

c

0.0 0.2 0.4 0.6 0.8 1.0 1.2

[µm]b

Page 36: © 2014 Zuanyi Li

30

dependence of the peak ratio follows a relation of ID/IG = cW-1

with c = 210 nm (dashed line),

which is consistent with previous reports of GNR characterization [95, 152]. Figure 3.3b shows

the D, G, and D’ peaks in detail of fabricated GNR arrays and un-patterned graphene with 633

nm wavelength laser. Figures 3.3c-g show the Raman 2D band spectra (scattered points) for the

samples. All 2D bands are fit by a single Lorentzian peak (solid red curves) with ~2650 cm-1

peak position, which is consistent with previous reports of monolayer graphene and GNRs [152].

FIG. 3.3: Raman spectra of GNR arrays and un-patterned graphene. (a) Raman signal for W ~

130, 85, 65, 45 nm GNRs and un-patterned graphene. Each spectrum is vertically offset for

clarity. Inset is the ID/IG ratio as a function of GNR width, consistent with the enhanced role of

edge disorder in narrower GNRs [105, 152, 153]. (b) Zoomed-in D, G, and D’ bands of all

samples. (c-g) 2D bands with a single Lorentzian fit for all samples, consistent with the existence

of monolayer graphene.

a

2400 2600 2800

Inte

nsity (

a.

u.)

Raman shift (cm-1)

2400 2600 2800

Raman shift (cm-1)

2400 2600 2800

Raman shift (cm-1)

2400 2600 2800

Raman shift (cm-1)

2400 2600 2800

Raman shift (cm-1)

W=45 nm GNRW=65 nm GNRW=85 nm GNRW=130 nm GNR

wide graphene

1300 1400 1500 1600 1700

Inte

nsity (

a.u

.)

Raman shift (cm-1)

45 nm GNR

65 nm GNR

85 nm GNR

130 nm GNR

Graphene

D

G

D’

b c

d e f g

0 100 2000

2

4

6

8

W (nm)

I D/

I G

DG

2D (G’)W=45 nm

W=65 nm

W=85 nm

W=130 nm

graphene

1500 2000 2500

Raman Shift (cm-1)

Inte

nsity (a

.u.)

~ 1/W

Page 37: © 2014 Zuanyi Li

31

3.2.2 Thermometry Platform and Measurements

To fabricate thermometry platform for thermal measurements, electron (e)-beam lithography

was used to pattern the heater and sensor thermometers [54, 157] as long, parallel, ~200-nm-

wide electrodes with current and voltage probes, with a separation of L ≈ 260 nm. Electrodes

were deposited by successive evaporation of SiO2 (20 nm) for electrical insulation and Ti/Au

(30/20 nm) for temperature sensing. Figure 3.4 illustrates scanning electron microscopy (SEM)

images of our several experimental test structures, showing graphene and GNR arrays supported

FIG. 3.4: Measurement of heat flow in graphene ribbons. (a) Scanning electron microscopy

(SEM) image of parallel heater and sensor metal lines with ~260 nm separation, on top of

graphene sample (colorized for emphasis). A thin SiO2 layer under the metal lines provides

electrical insulation and thermal contact with the graphene beneath. (b) Similar sample after

graphene etch, serving as control measurement for heat flow through contacts and SiO2/Si

underlayers. (c) Heater and sensor lines across array of graphene nanoribbons (GNRs). (d)

Magnified portion of array with GNR widths ~65 nm; inset shows atomic force microscopy

(AFM) image of GNRs. Scale bars of (a-d) are 2 μm, 1 μm, 2 μm and 1 μm, respectively. (e)

Three-dimensional (3D) simulation of experimental structure, showing temperature distribution

with current applied through heater line.

a

b

∆T (K)

c d

e

1010.1

290 nm SiO2 Silicon

graphene

SiO2

Page 38: © 2014 Zuanyi Li

32

on a SiO2/Si substrate, as well as bare SiO2 with graphene etched off by an oxygen plasma

(however, graphene still exists under the metal electrodes, consistent with the other samples).

Figure 3.5 shows the measurement set-up for the SiO2 sample (Fig. 3.4b) as an example. To

block environmental noise including electrostatic discharge, π-filters with a cut-off frequency of

2 MHz were inserted across all measurement lines. To control the temperature, Physical Property

Measurement System (PPMS, Quantum Design) was used with a temperature range of 10 K –

363 K. Inside the PPMS, the vacuum environment is always a few ~10-3

Torr, rendering convec-

tive heat losses negligible.

The measurement proceeds as follows. In the heater, we apply a sinusoidal voltage with fre-

quency lower than 2 mHz through a standard resistor of 1 kΩ to flow current with a range of

±1.5 mA, generating a temperature gradient across the sample. To obtain the response of the

FIG. 3.5: Scanning electron micrograph (SEM) of thermometry platform and measurement

configuration. Scale bar is 4 µm. Image taken of sample after graphene was etched off, and after

all electrical and thermal measurements were completed. Dark region around “part 1” is substrate

charging due to previous SEM imaging performed to obtain Fig. 3.4b in the main text.

VH

Lock-in amplifier

2.147 kHz

1 k

Ω

< 2

mH

z

π-filter

sensor

heater

1 MΩ

Page 39: © 2014 Zuanyi Li

33

sensor (Fig. 3.6a), we measured its resistance change by a standard lock-in method with excita-

tion frequency 2.147 kHz and root-mean-square (rms) current 1 μA (carefully chosen to avoid

additional heating). All electrical measurements were performed in a four-probe configuration.

Both electrode resistances are calibrated over the full temperature range for each sample (Fig.

3.6b), allowing us to convert measured changes of resistance into changes of sensor temperature

ΔTS as a function of heater power PH (Fig. 3.6c). We sometimes found that the electrical re-

sistance of the sensor slowly drifted (increased) with time at room temperature. However, this

effect was stabilized after annealing the sample at 363 K for 5 min, eliminating resistance drift at

room temperature. Therefore, this behavior could be related to the absorbed water on the metal

electrodes.

3.2.3 Experimental Data and Error

Figure 3.6a shows the measured sensor resistance change, ∆RS, as a function of the power

applied to the heater, PH, at T = 100 K for the SiO2 sample (Fig. 3.4b). The black (for negative

heater current, IH) and red (for positive IH) lines overlap with each other, indicating the meas-

urement is symmetric and reliable. The calibration for sensor and heater resistance vs. tempera-

ture is shown in Fig. 3.6b; thus, sensor heating due to heater power ∆RS can be converted to a

temperature rise, ∆TS, as shown in Fig. 3.6c by using the resistance-temperature calibration

curve. The fitted slope of the ∆TS vs. PH curve in Fig. 3.6c is 0.01797 K/µW, which is then used

for the extraction of thermal properties through simulations (see Section 3.3). Figure 3.6d shows

the measured ratio of PH to ∆TS for all representative samples as a function of ambient tempera-

ture from 20 K to 300 K. The uncertainty of the electrical thermometry measurement is ~2%

(Tables 3.1 and 3.2), which is comparable to the symbol size on this plot. Thus, although data are

available down to 20 K, the values are distinguishable without ambiguity only when T ≥ ~70 K,

Page 40: © 2014 Zuanyi Li

34

which is the main temperature range shown in Section 3.5.

We note that PH/ΔTS shown in Fig. 3.6d is not the thermal conductance through graphene,

because ΔTS is the temperature rise in the sensor, not the temperature drop from the heater to

sensor, and PH is the heat generated in the heater, not the one flowing in graphene. The thermal

FIG. 3.6: Measurement process. (a) Sensor resistance change, ∆RS as a function of heater power,

PH at T =100 K for the SiO2 sample (Fig. 3.5). Red and black lines are taken with current flow in

opposing direction. (b) Calibration of sensor and heater resistances as a function of temperature.

The inset shows the R-T curve and slope of the sensor near T = 100 K. (c) Converted sensor

temperature rise, ∆TS as a function of heater power, PH at T = 100 K from (a) and (b). The slope

of the fitted red line is ∆TS/PH = 0.01797 K/µW, which is later used to extract the thermal

properties of the SiO2 layer (see Fig. 3.9). (d) Measured ratio of heater power to sensor tempera-

ture rise for all representative samples. The uncertainty of these data is ~2% (Tables 3.1 and 3.2),

comparable to the symbol size. Although this plot shows all raw data taken, the values can be

distinguished without ambiguity only at T ≥ 70 K, which is the temperature range displayed in

the Figs. 3.12 and 3.16.

0 50 100 150 200 250 300 350

45

50

55

60

65

70

R (

Oh

m)

T (K)

Sensor

Heater

0 20 40 60 80 100 120 1400

50

100

150

200

R

S (

m

)

PH (W)

0 20 40 60 80 100 120 1400.0

0.5

1.0

1.5

2.0

2.5

T

S (

K)

PH (W)

90 95 100 105 11051.5

52.0

52.5

53.0

53.50.0866 Ω/K

ΔTS/PH = 0.01797 K/µW

T=100 K

T=100 K

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90

100

SiO2 (part 2)

GNR, W = 45 nm

GNR, W = 65 nm

GNR, W = 85 nm

GNR, W = 130 nm

Graphene (L=0.26 um)

PH/

TS (W

/K)

T (K)

a b

c ddistinguishable

Page 41: © 2014 Zuanyi Li

35

conductance of the graphene cannot be immediately extracted from our raw data, due to heat

leakage into the substrate (a drawback of the substrate-supported thermometry method). Instead,

we employ 3D simulations to carefully account for all heat flow paths and, by comparison with

the experiments, to obtain the thermal conductance of the graphene samples (see Section 3.4).

Figure 3.7a shows the sensor resistance as a function of count number (time) without apply-

ing current to the heater at T = 102 K. The standard deviation of the scattered data points is δR =

3.1 mΩ, which corresponds to δT ~ 36 mK by using calibration coefficient 0.0866 Ω/K obtained

in Fig. 3.6b. Thus, the error of the temperature reading is ±36 mK, primarily due to slight ambi-

ent temperature fluctuations in the PPMS (consistent with a fluctuation of ±30 mK of the dis-

played temperature on the PPMS monitor). Zooming into the circled region of Fig. 3.7a, we note

a resistance fluctuation δR = 0.17 mΩ, corresponding to a temperature uncertainty ±2 mK due to

electrical measurement instruments. Therefore, during the time scales of most of our measure-

ments our temperature accuracy is limited by the ambient temperature control of the PPMS

rather than by the electrical measurements themselves.

FIG. 3.7: Measurement error and thermal steady-state. (a) Sensor resistance as a function of

count number (time) at background T = 102 K. (b) Heater power, PH and corresponding re-

sistance change in sensor, ΔRS as a function of time.

0 500 1000 1500 2000 2500

53.385

53.390

53.395

53.400

RS (

)

count number

0 2 4 6 8 10 12 14 160

20

40

60

80

100

120

Time (min.)

PH (W

)

0

50

100

150

200

R

S (

m

)

a b

Page 42: © 2014 Zuanyi Li

36

The sweep speed of the heater power is chosen to be sufficiently slow to reach thermal

steady-state between the heater and sensor. Figure 3.7b shows the heater power (PH) sweep with

time and corresponding resistance change in the sensor, ∆RS. Data shown here correspond to the

linear ramp in Fig. 3.6a. After ~15 minutes, the heater power reaches its maximum, and the

change of sensor resistance follows the same trend without delay, indicating that the thermal

steady-state between the heater and the sensor is established during the entire sweep process. If

the sweep speed of the heater power is too fast to reach the steady-state, the data point at PH ~

110 µW in Fig. 3.6a will deviate from the linear trend. We also verified this by a comparison

between the corresponding constant DC power and the above methods.

3.3 Model to Analyze Experimental Data

3.3.1 3D Simulation to Extract Thermal Properties

To extract the thermal properties of graphene, GNRs, or the SiO2 substrate from the meas-

ured ∆TS vs. PH, we use a commercial software package (COMSOL) to set up a three-

dimensional (3D) finite element method (FEM) model of the entire structure. A typical setup is

shown in Fig. 3.8, where only a half of the sample is included due to the symmetry plane which

bisects the region of interest. The size of the Si substrate is 100×50×50 μm3, covered by a 290

nm thick SiO2 layer. While the simulated Si substrate is slightly smaller than the actual Si chip

employed in practice (to manage computational complexity and meshing), the size of the simu-

lated structure has been carefully chosen and verified to reproduce all heat flow through the

substrate itself. Figure 3.8b shows the zoomed-in structure containing the core area of the

thermometry, with GNRs, heater, and sensor highlighted by different colors. A more zoomed-in

structure is shown in Fig. 3.8c, where from top to bottom different layers are 40 nm metal, 25 nm

top oxide insulator, GNRs, 290 nm bottom oxide, and silicon, respectively.

Page 43: © 2014 Zuanyi Li

37

To perform the simulation, the bottom surface and three side surfaces (except symmetry

plane) of the Si substrate are held at the ambient temperature, i.e. isothermal boundary condition.

Other outer surfaces of the whole structure are treated as insulated, i.e. adiabatic boundary

condition. The Joule heating in the heater is simulated by applying a power density within the

heater metal, and the stationary calculation is performed to obtain the temperature distribution in

the steady state, as shown in Figs. 3.4e and 3.8d typically. After calculating the average tempera-

ture rise in the measured segment of the sensor, we obtain the simulated value of (∆TS/PH). Thus,

the simulation effectively fits the thermal conductance (G) of the test sample between heater and

thermometer. The thermal conductivity (k = GL/A) of the test sample is thus an effective fitting

parameter within the FEM simulator, ultimately adjusted to yield the best agreement between the

FIG. 3.8: 3D Finite element method (FEM) model. (a) Whole structure of 3D FEM model. (b)

Zoomed-in structure to show the core area of the thermometry. (c) More zoomed-in structure to

show different layers. (d) Typical distribution of temperature rise due to heating in simulations

which matches with measurements.

Si

50

μm

GNRs

a b

c∆T (K)

101

100

10-1

10-2

Silicon290 nm SiO2

GNR

SensorHeaterd

Page 44: © 2014 Zuanyi Li

38

simulated and measured ∆TS vs. PH. This fitting process is implemented by using MATLAB to

interface directly with the COMSOL software, taking ~0.5 hour on a single desktop computer to

converge to a best-fit value at a single temperature point for a typical calculation.

Before performing a substantial amount of calculations, a series of optimizations were car-

ried out. First, the mesh was optimized. Due to the extreme ratio of the graphene/GNRs thickness

(~0.34 nm) to their typical in-plane dimensions (~10 μm), this subdomain was optimized using a

swept mesh strategy rather than the typical free mesh. Other subdomains were optimized careful-

ly using the free mesh strategy, and in the bottom oxide and Si substrate the mesh size grows

gradually from the heating region to the boundaries. Second, the real substrate size is about

8×8×0.5 mm3, which can be regarded as a semi-infinite substrate relative to the small heating

region (~10 μm). In FEM modeling, however, we have to select a finite size for the substrate due

to the computational limitation. By choosing the distance from the center of the heater to the side

and bottom surfaces of the substrate as a testing variable, we found 50 μm is large enough to

model this 3D heat spreading, consistent with the recent work by Jang et al. [54]. Third, the

length of the six probe arms attached to the heater and sensor (see Fig. 3.8b) was chosen as 2 μm

(shorter than their real counterparts), which was found to be sufficiently long to mimic any

peripheral heat loss. Fourth, it was confirmed that the simulated ∆TS/PH is independent of the

power PH applied in the heater, which means the final results do not rely on the choice of the

power PH.

3.3.2 Uncertainty Analysis

We can estimate the uncertainty of our analysis with the classical partial derivative method:

2

ixki

i i

uus

k x

(3.1)

Page 45: © 2014 Zuanyi Li

39

where uk is the total uncertainty in the extracted thermal conductivity k, uxi is the uncertainty in

the i-th input parameter xi, and the dimensionless sensitivities si are defined by

(ln )

(ln )

ii

i i

x k ks

k x x

. (3.2)

The partial derivatives are evaluated numerically by giving small perturbation of each parameter

around its typical value and redoing the extraction simulation to obtain the change of k. To

highlight the relative importance of each input parameter, we define their absolute contributions

Table 3.1: Uncertainty analysis for heat flow in the unpatterned “short” graphene. Example of

calculated sensitivities and uncertainty analysis for the thermal conductivity of the graphene

sample GS1 at 300 K. The extracted k is 319 Wm-1

K-1

and its overall uncertainty is 19%.

Units Values x i

Uncertainty

u xi

u xi /x i

Sensitivity

s i

Contribution

c i =|s i |×u xi /x i

c i2/Σc i

2

Expt. Sensor response ΔT S/P H K/μW 0.01635 0.0002 1.2% 4.49 5.5% 8.6%

k ox 1.267 0.04 3.2% 4.25 13.4% 51.6%

k Si 115 10 8.7% 0.55 4.8% 6.6%

k met 55 4 7.3% 0.34 2.5% 1.8%

R gox 1.15E-08 2.0E-09 17.4% -0.12 2.1% 1.3%

R oxs 9.92E-09 3.0E-09 30.2% -0.25 7.6% 16.5%

R mox 1.02E-08 3.0E-09 29.5% 0.05 1.5% 0.7%

t box 288 1 0.3% -5.65 2.0% 1.1%

t tox 20 1 5.0% -0.01 0.1% 0.0%

t met 50 2 4.0% 0.35 1.4% 0.5%

D met 494 5 1.0% 5.38 5.4% 8.5%

D tox 486 5 1.0% 2.39 2.5% 1.7%

Wmet 186 4 2.2% -0.16 0.3% 0.0%

W tox 224 4 1.8% -0.17 0.3% 0.0%

Half length of H/S

electrodesL HS/2 5.86 0.03 0.5% 0.04 0.0% 0.0%

Distance of 2

Voltage probesD pVV 4.23 0.02 0.5% 3.75 1.8% 0.9%

Distance of

Current and

Voltage probes

D pIV 1.04 0.02 1.9% 0.003 0.0% 0.0%

Geo

metr

ica

l

Thickness of

bottom and top

SiO2, and metal

nmDistance of H/S

metal lines and

H/S SiO2 lines

Width of metal and

SiO2 lines

μm

Input parameters (T = 300 K)

Th

erm

al

Thermal

conductivity of

SiO2, Si, metal

W/m/K

TBR of

graphene/SiO2,

SiO2/Si, metal/SiO2

interfaces

m2K/W

Page 46: © 2014 Zuanyi Li

40

as ci = |si|×(uxi/xi), and relative contributions as ci2/Σci

2. The uncertainty analysis for extracting

kox and Roxs from the SiO2 control experiment is performed in the same way.

Table 3.1 summarizes the sensitivities and uncertainty analysis for the unpatterned graphene

(GS1) at 300 K. All input parameters can be separated into 3 classes: experimental data, thermal

parameters, and geometric parameters. Their uncertainties are our estimates considering both

random and systematic errors, and those of experimental and thermal parameters are updated

appropriately as the temperature changes. The calculated sensitivities show that the graphene

thermal conductivity is the most sensitive (|si| > 2) to the measured sensor response (ΔTS/PH),

thermal conductivity (kox) and thickness (tbox) of bottom SiO2, center-to-center distances between

metal lines of heater and sensor (Dmet), between top SiO2 lines of heater and sensor (Dtox), and

between two voltage probes (DpVV). These findings are consistent with previous work by W.

Jang et al. [54] using similar substrate-supported thermometry structures. The input parameters

with the greatest relative uncertainty (uxi/xi) are all three thermal boundary resistances (TBRs),

thicknesses of top SiO2 (ttox) and metal (tmet), and the thermal conductivity of the Si substrate

(kSi) and metal (kmet). The combined effects of both sensitivities and relative uncertainties show

that five largest contributions (ci > ≈ 5%) to the overall uncertainty of our thermal measurement

are from ΔTS/PH, kox, kSi, Roxs, and Dmet. In slight contrast to Jang et al. [54], we find that uncer-

tainties introduced by Roxs, Rmox and tmet are non-negligible for our structure and should be

considered. On the other hand, geometric parameters related to the shape and size of the gra-

phene sheet have very small sensitivities (|si| < 0.001), so their contributions are negligible in

uncertainty analysis and not listed in Table 3.1.

Page 47: © 2014 Zuanyi Li

41

For the extraction of GNR thermal conductivity, an example of calculated sensitivities and

uncertainty analysis at 300 K is summarized in Table 3.2 (here for the sample with W ~ 65 nm).

Compared with Table 3.1, we have two more parameters: the thermal conductivity of outer

graphene connected to GNRs (kg) and the width of GNRs (WGNR). The parameters with the

largest sensitivities (|si| > 5) are the same as those in Table 3.1, but their values increase because

Table 3.2: Uncertainty analysis for GNR thermal conductivity. Example of calculated sensitivi-

ties and uncertainty analysis for the thermal conductivity of GNRs with W = 65 nm at 300 K.

The extracted k is 101 Wm-1

K-1

and its overall uncertainty is 60%.

Units Values x i

Uncertainty

u xi

u xi /x i

Sensitivity

s i

Contribution

c i =|s i |×u xi /x i

c i2/Σc i

2

Expt. Sensor response ΔT S/P H K/μW 0.0124 0.0002 1.6% 13.22 21.3% 12.9%

k ox 1.267 0.04 3.2% 10.38 32.8% 30.5%

k Si 115 10 8.7% 2.61 22.7% 14.6%

k met 22 2 9.1% 0.68 6.2% 1.1%

k g 320 60 18.8% 0.001 0.0% 0.0%

R gox 1.15E-08 2.0E-09 17.4% -0.04 0.6% 0.0%

R oxs 9.92E-09 3.0E-09 30.2% -1.00 30.3% 26.1%

R mox 1.02E-08 3.0E-09 29.5% 0.11 3.4% 0.3%

t box 291 1 0.3% -22.51 7.7% 1.7%

t tox 25 1 4.0% -0.17 0.7% 0.0%

t met 40 2 5.0% 0.70 3.5% 0.3%

D met 517 5 1.0% 14.73 14.2% 5.8%

D tox 509 5 1.0% 14.25 14.0% 5.6%

Wmet 218 4 1.8% -0.64 1.2% 0.0%

W tox 266 4 1.5% -2.27 3.4% 0.3%

WGNR 65 3 4.6% -0.80 3.7% 0.4%

Half length of H/S

electrodesL HS/2 5.77 0.03 0.5% 0.20 0.1% 0.0%

Distance of 2

Voltage probesD pVV 4.17 0.02 0.5% 7.80 3.7% 0.4%

Distance of

Current and

Voltage probes

D pIV 1.04 0.02 1.9% -0.286 0.5% 0.0%

Geo

metr

ica

l

Thickness of

bottom and top

SiO2, and metal

Distance of H/S

metal lines and

H/S SiO2 lines

μm

Width of metal

lines, top SiO2

lines, and GNRs

nm

Input parameters (T = 300 K)

Th

erm

al

TBR of

graphene/SiO2,

SiO2/Si, metal/SiO2

interfaces

m2K/W

Thermal

conductivity of

SiO2, Si, metal, and

outter graphene

W/m/K

Page 48: © 2014 Zuanyi Li

42

the total width of the GNR array is smaller than that of the unpatterned graphene. Due to the

significant increase of sensitivities, the total uncertainty increases from 19% for unpatterned

graphene to 60% for GNRs, while the input parameters with the greatest contributions (ci > 10%)

to the total uncertainty are ΔTS/PH, kox, kSi, Roxs, Dmet, and Dtox, the same as those for graphene

along with Dtox. From the 60% total uncertainty, approximately 21% is due to measurement

uncertainty and the remainder from geometric and temperature-dependent variables as listed in

Table 3.2. [we note that the geometric parameters related to the shape and size of GNRs and

graphene outside the heater and sensor region have very small sensitivities (|si| < 0.01) and are

not listed in Table 3.2, which is also consistent with the GNR results being insensitive to kg.]

For other GNR samples with different widths, the total uncertainties gradually increase from

22% to 83% at 300 K as W decreases from ~130 to ~45 nm, as less heat flows in the GNR array

rather than the substrate. As the temperature decreases, the uncertainties of all graphene and

GNR thermal conductivities also increase due to either increased sensitivities or increased

relative uncertainties of input parameters.

As mentioned earlier and shown in Tables 3.1 and 3.2, all input parameters are classified in-

to three groups, not all of which need to be included when comparing relative GNR thermal

conductivities (e.g. with width or temperature). For instance, Fig. 3.12b compares the thermal

conductivities of the samples at different temperatures considering the contributions of experi-

mental and thermal parameters to the error bars. Similarly, Fig. 3.12d compares thermal conduc-

tivities of different samples at the same temperature, considering the contributions of experi-

mental and geometric parameters to the error bars. Different samples share the same thermal

parameters and these uncertainties would only shift all k values in the same direction without

affecting their relative values. In the end, the differences are relatively subtle, and within the

Page 49: © 2014 Zuanyi Li

43

fitting capabilities of our models, all based on the initial data above 70 K which are clearly

distinguishable from one another as seen from the raw thermometry in Fig. 3.6d.

3.4 Results

Based on experimental method described above, we performed heat flow measurements

from 20 to 300 K on unpatterned graphene (Fig. 3.4a), control samples with the graphene etched

off (Fig. 3.4b), and arrays of GNR widths W ≈ 130, 85, 65 and 45 nm (Figs. 3.2 and 3.4c-d). To

obtain the thermal properties of our samples we used the above mentioned 3D simulations of the

structures with dimensions obtained from measurements by SEM and AFM, as shown in Figs.

3.4e and 3.8. The results for the SiO2 control sample, unpatterned graphene, and GNR arrays are

presented in the following.

3.4.1 Thermal Properties of the SiO2 Underlayer

To validate our thermometry approach, we first carefully focused on a control experiment to

measure the thermal properties of the SiO2 underlayer supporting our samples. The sample was

first prepared as described before, including the graphene under the heater and thermometer

electrodes, to reproduce the thermal contacts encountered in all samples. However, the exposed

graphene was then etched by oxygen plasma, leaving the bare SiO2 as shown in Figs. 3.4b and

3.5. This allows us to obtain the thermal properties of the parallel heat flow path through the

contacts, supporting SiO2 and substrate. Measurements were performed on part 1 and part 2 of

the metal electrodes (see Fig. 3.5), and the analyzed sensor and heater temperature rises normal-

ized by heater power as a function of the ambient temperature are shown in Fig. 3.9a.

Page 50: © 2014 Zuanyi Li

44

To compare our 3D simulations to this experimental data set, we needed to fit the thermal

properties of the SiO2 layer (which dominate), and to a lesser degree those of the SiO2-Si TBR at

the bottom of the SiO2 layer (which also plays a role). While the thermal conductivity of SiO2 is

well-known and easy to calibrate against [30, 158], to the best of our knowledge no consistent

data for the TBR of the SiO2-Si interface (Roxs) exist as a function of temperature. Two recent

studies [159, 160] suggested Roxs ~ 5−7×10-8

m2KW

-1 at room temperature, but some earlier

efforts [161-165] found the total TBR of metal-SiO2-Si interfaces as low as ~1−3×10-8

m2KW

-1,

putting an upper bound on Roxs without being able to separate it from the total TBR. Due to this

contradiction, we set out to obtain the temperature-dependent Roxs, treating it as another fitting

parameter of our simulations in addition to the thermal conductivity of SiO2 (kox). Other thermal

parameters well characterized in the literature are the thermal conductivity of highly doped Si

[166-168], thermal boundary resistances of the graphene-SiO2 interface [44, 48] and the Au-Ti-

SiO2 interfaces [45]. In addition, the effective thermal conductivity of the metal electrodes

(Au/Ti) was calculated from the measured electrical resistance according to the Wiedemann-

FIG. 3.9: Control experiment to extract SiO2 thermal properties. (a) Sensor and heater tempera-

ture rise normalized by heater power from measurements taken at part 1 and part 2 of the SiO2

sample (see Fig. 3.5). (b) Extracted thermal conductivity of SiO2 from two measurements

compared with well-known data from Cahill [158]. The green solid line is the polynomial fit up

to the 7th order to our data. (c) Extracted thermal boundary resistance of the SiO2-Si interface,

Roxs. The green solid line is the fit to our data by Eq. 3.3.

0 50 100 150 200 250 300 3500.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4 Cahill (1990)

Polynomial Fit

Part 1

Part 2

ko

x (

Wm

-1K

-1)

T (K)10 100

10-8

10-7

Part 1

Part 2

Fit

Roxs (

m2K

/W)

T (K)

a b c

0 50 100 150 200 250 300

0.01

0.1

1

Part 1

Part 2

T

/PH (

K/

W)

T (K)

Sensor

Heater

Page 51: © 2014 Zuanyi Li

45

Franz Law, where an average Lorentz number Lz = 2.7×10-8

WΩK-2

is used for Au/Ti electrodes

[169]. (all parameters were allowed to vary within known experimental bounds, leading to the

uncertainty analysis in Section 3.3.2.)

Our extracted kox and Roxs of two data sets (part 1 and 2) are shown in Figs. 3.9b and 3.9c,

respectively. Our kox data are in a good agreement with well-established values reported by

Cahill [158], and the typical uncertainty is ~5% at most temperatures. By fitting our kox data with

a polynomial up to the 7th

-order, we obtained a smooth dependence of kox on T (green solid line),

and this was used to extract the thermal properties of our graphene and GNRs. Our extractions

suggest Roxs ~ 10-8

m2KW

-1 at room temperature, in agreement with the reported upper bound in

Refs. [161-165]. For the subsequent thermal analysis, our Roxs data are best fit by a simple

expression,

4

9

2.25

1.046 10( ) 9.67 10

( 13.4)oxsR T

T

m2K/W (3.3)

as shown by the green solid line in Fig. 3.9c. Thus, this control experiment demonstrates the

feasibility and reliability of our thermometry platform, also giving the first report of the tempera-

ture-dependent TBR of SiO2-Si interfaces.

3.4.2 Quasi-Ballistic Thermal Transport in Short Graphene

After demonstrating the feasibility of our thermometry method by the control sample, we

next examined three “short” graphene samples (GS1, 2, 3) which were not patterned into GNRs;

all had length L ~ 260 nm and width W ~ 12 μm between the heater and sensor thermometers

(corresponding to schematic in Fig. 3.10a). Besides GS1 which is shown in Fig. 3.4a of the main

paper, the SEM image of another sample (GS2) is shown in Fig. 3.11a. The third one (GS3)

broke after measurements at two temperature points, which are nevertheless listed among the

Page 52: © 2014 Zuanyi Li

46

data in Fig. 3.11. Figure 3.11b displays raw data taken as ratio of heater power to sensor temper-

ature (PH/ΔTS) for all three samples; the corresponding data of the SiO2-only sample (part 2 of

Fig. 3.9a) is also plotted here for comparison. The presence of graphene notably “heats” the

sensor (higher ΔTS) and is distinguishable from the SiO2-only sample all the way down to ~20 K

(although the GNRs become harder to distinguish below ~70 K as mentioned earlier). The

extracted graphene thermal conductivities are shown in Fig. 3.11c, and the three samples show

very similar values. They all decrease from ~300 Wm-1

K-1

at 300 K to ~10 Wm-1

K-1

at 30 K and

show similar temperature dependence. The data of the “long” graphene with L ~10 μm from Seol

et al. [49] are also plotted in Fig. 3.11c (black dots).

FIG. 3.10: Schematic of size effects and different heat flow regimes. (a) Diffusive heat transport

in “large” samples with dimensions much greater than the intrinsic phonon MFP (L, W ≫ λ).

This regime corresponds to the samples measured in both substrate-supported [49] and suspend-

ed graphene [58] studies to date. (b) Quasi-ballistic heat flow in “short but wide” samples (L ~ λ

and W ≫ λ). These correspond to our geometry shown in Fig. 3.4a, with L ≈ 260 nm and W ≈ 12

μm. (c) Return to a diffusive heat transport regime as the sample width is narrowed down, and

phonon scattering with edge roughness (of rms Δ) begins to dominate. These correspond to our

arrays of GNRs from Figs. 3.4c-e (L ≈ 260 nm and W varying from 45 nm to 130 nm). A fourth

regime (long L, narrow W) is not shown here, but it can be easily understood from the above.

‘Large’ L, W » λ

(diffusive)

‘Short’ L ~ λ, W » λ

(quasi-ballistic)

‘Short and narrow’

L, W ~ λ

(diffusive)

λ W

Δ

λ

a b

c

λL

HOT

COLD

Page 53: © 2014 Zuanyi Li

47

Next we compare the sample thermal conductance to the theoretical ballistic limits in Fig.

3.11d, recalling the relationship between conductance and conductivity, G = kA/L, where A is the

cross-sectional area of heat flow, A = WH, with W and H = 0.335 nm as the width and thickness

of the graphene samples, respectively. The theoretical ballistic conductance per unit cross-

sectional area Gball/A of graphene (solid line in Fig. 3.11d) is calculated by using the full phonon

dispersion of graphene and integrating over the entire 2D Brillouin zone (see Sections 5.4.1 and

5.3.2). For convenience, we note that the Gball/A of graphene can be approximated analytically as

FIG. 3.11: Data of three unpatterned graphene samples (all L ~ 260 nm). (a) SEM image of the

unpatterned graphene sample GS2 (colorized for emphasis). (b) Measured ratio of heater power

to sensor temperature rise of all three graphene samples compared with that of SiO2 sample from

the control experiment (Fig. 3.9). (c) Extracted graphene thermal conductivity of all our “short”

samples (L ~ 260 nm) compared with the “long” graphene (L ~ 10 μm) reported by Seol et al.

[49]. (d) Thermal conductance per unit area compared to theoretical ballistic limit of graphene.

10 1001

10

100

1000 Graphene (L~10 m)

Graphene (GS1)

Graphene (GS2)

Graphene (GS3)

k (

Wm

-1K

-1)

T (K)

~T2

~T1.5

~T

Seol (2010)

10 100

10

100 SiO

2 (part 2)

Graphene (GS1)

Graphene (GS2)

Graphene (GS3)

PH/

TS (W

/K)

T (K)

a b

c d

10 100

107

108

109

1010

Graphene (L~10 m)

Graphene (GS1)

Graphene (GS2)

Graphene (GS3)

G/A

(W

m-2K

-1)

T (K)

GS2

Seol (2010)

Page 54: © 2014 Zuanyi Li

48

Gball/A ≈ [1/(4.4 × 105 T

1.68) + 1/(1.2 × 10

10)]

-1 WK

-1m

-2 over the temperature range 1-1000 K, as

a fit to full numerical calculations. As shown in Fig. 3.11d, our three “short” graphene samples

all display on average ~35% of the ballistic conductance limit, indicating they reach a quasi-

ballistic conduction regime (schematic in Fig. 3.10b). The data for the 10-μm “long” sample

from Seol et al. [49] show <2% of ballistic limit on average, indicating a diffusive transport

regime (schematic in Fig. 3.10a) as would be expected for a sample much longer than the phonon

MFP (W, L ≫ λ). Both percentages are consistent with a simple estimation of transmission

probability ~λbs/(λbs + L) using their own lengths and back scattering mean free path λbs = (π/2)λ ~

160 nm (see Section 3.4.3) where the intrinsic phonon MFP λ ~ 100 nm for most temperatures.

3.4.3 Length dependence of thermal conductivity

To understand different transport behaviors in short and long graphene, we recall that in the

ballistic limit (L ≪ λ) the conductance rather than the conductivity approaches a constant at a

given temperature [29, 112, 170], Gball(T). Nevertheless, the thermal conductivity is the parame-

ter typically used for calculating heat transport in practice, and for comparing different materials

and systems. Thus, the well-known relationship k = (G/A)L imposes the conductivity k to be-

come a function of length in the ballistic regime and to decrease as L is reduced. This situation

becomes evident when we plot the thermal conductivity in Figs. 3.11c and 3.12b, finding k ≈ 320

Wm-1

K-1

for our “short” and wide samples at room temperature (schematic Fig. 3.10b), almost a

factor of two lower than the large graphene [49] (schematic Fig. 3.10a). We note that both

unpatterned samples here and in Ref. [49] were supported by SiO2, showed no discernible

defects in the Raman spectra, and the measurements were repeated over three samples (Fig. 3.11)

with similar results obtained each time.

Page 55: © 2014 Zuanyi Li

49

The transition of thermal conductivity from diffusive to ballistic can be captured through

FIG. 3.12: Thermal conduction scaling in GNRs. (a) Thermal conductance per cross-sectional

area (G/A) vs. temperature for our GNRs (L ≈ 260 nm, W as listed, see Fig. 3.10c), a “short”

unpatterned sample (L ≈ 260 nm, W ≈ 12 μm, see Fig. 3.10b), and a “large” sample from Seol et

al. [49] (L ≈ 10 μm, W ≈ 2.4 μm, see Fig. 3.10a). The short but wide graphene sample attains up

to ~35% of the theoretical ballistic heat flow limit [29, 112, 170]. (b) Thermal conductivity for

the same samples as in (a) (also see Fig. 3.14a). (c) Thermal conductivity reduction with length

for “wide” samples (W ≫ λ), compared to the ballistic limit (kball = GballL/A) at several tempera-

tures. Symbols are data for our “short” unpatterned graphene samples (Figs. 3.4a and 3.10b), and

“large” samples from Seol et al. [49] (Fig. 3.10a). Solid lines are model from Eq. 3.4. (d) Ther-

mal conductivity reduction with width for GNRs, all with L ≈ 260 nm (Figs. 3.4c-d and 3.10c).

Solid symbols are experimental data from (b), open symbols are interpolations for the listed

temperature; lines are fitted model from Eq. 3.5. The thermal conductivity of plasma-etched

GNRs in this work appears lower than that estimated for GNRs from unzipped nanotubes [144]

at a given width, consistent with a stronger effect of edge disorder [153]. Also see Fig. 3.13a.

100 200 300 40010

7

108

109

G/A

(W

m-2K

-1)

T (K)50

10 100 1000 100000

50

100

150

200

250

300

350

k (

Wm

-1K

-1)

W (nm)10 100 1000 10000

0

100

200

300

400

500

600

700

k (

Wm

-1K

-1)

L (nm)

ballistic

limit kballT = 300 K

T = 190 K

T = 150 K

T = 70 K

L ~ 260 nm

100 K

c d

large graphene

(L ~ 10 μm)

Seol (2010)

W~130 nm

W~85 nm

W~65 nm

W~45 nm

~T1.5

b

100 200 300 40010

100

1000

k (

Wm

-1K

-1)

T (K)50

~T1.5

~T

W~130 nm

W~85 nm

W~65 nm

W~45 nm

large graphene (L ~ 10 μm)

Seol (2010)

short graphene

(L ~ 260 nm)

a

Seol (2010)

Page 56: © 2014 Zuanyi Li

50

simple models [171], similar to the apparent mobility reduction during quasi-ballistic charge

transport observed in short-channel transistors [172, 173]:

1 1

ball

ball diff

1 1 1.

( / 2)b bb

GAk L

LG k A L

(3.4)

The first equality is a “three-color” model with b the phonon branch (longitudinal acoustic,

LA; transverse, TA; flexural, ZA), ball

bG calculated using the appropriate dispersion [29], and ∑

diff

bk = kdiff the diffusive thermal conductivity. At T = 300 K, kdiff ≈ 560 ± 50 Wm-1

K-1

, and diff

bk =

148, 214, 198 Wm-1

K-1

for b = ZA, TA, LA modes, respectively [49]. A simpler “gray” approx-

imation can also be obtained by dropping the b index, k(L) ≈ [A/(LGball) + 1/kdiff]-1

, where Gball/A

≈ 4.2 × 109 WK

-1m

-2 at room temperature [96] (see Section 3.4.2). The second expression in Eq.

3.4 is a Landauer-like model [171, 174], with π/2 accounting for angle averaging [175] in 2D to

obtain the phonon backscattering MFP.

We compare the simple models in Eq. 3.4 with the experiments in Fig. 3.12c and find good

agreement over a wide temperature range. The Eq. 3.4 also yields our first estimate of the

intrinsic phonon MFP in SiO2-supported graphene, λ ≈ (2/π)kdiff/(Gball/A) ≈ 90 nm at 300 K and

115 nm at 150 K. (The same argument estimates an intrinsic phonon MFP λ ≈ 300-600 nm in

freely suspended graphene at 300 K, if a thermal conductivity 2000-4000 Wm-1

K-1

is used [21,

23, 24, 58].) This phonon MFP is the key length scale which determines when the thermal

conductivity of a sample becomes a function of its dimensions, in other words when L and W are

comparable to λ. Based on Fig. 3.12c, we note that ballistic heat flow effects should become non-

negligible in all SiO2-supported graphene devices shorter than approximately 1 μm.

3.4.4 Width dependence of thermal conductivity

Figure 3.12a also displays in-plane thermal conductance per area (G/A) for our measured

Page 57: © 2014 Zuanyi Li

51

GNRs (L ≈ 260 nm, W ≈ 130, 85, 65 and 45 nm). As the width decreases, thermal conduction

returns from quasi-ballistic to diffusive regimes gradually (corresponding to schematic Fig.

3.10c). In parallel, Fig. 3.10 summarizes schematics of the size effects and all three transport

regimes discussed, corresponding to the samples measured in Fig. 3.12.

We now turn to the width-dependence of heat flow in narrow GNRs. Our experimental data

in Figs. 3.12b and 3.12d show a clear decrease of thermal conductivity as the width W is reduced

to a regime comparable to the intrinsic phonon MFP. For instance, at room temperature k ≈ 230,

170, 100, and 80 Wm-1

K-1

for GNRs of W ≈ 130, 85, 65 and 45 nm, respectively, and same L ≈

260 nm. To understand this trend, we consider k limited by phonon scattering with edge disorder

[176, 177] through a simple empirical model with a functional form suggested by previous work

on rough nanowires [178, 179] and GNR mobility [180]:

1

eff

1 1,

( )

n

k W Lc W k L

. (3.5)

Here Δ is the rms edge roughness and k(L) is given by Eq. 3.4 [k(L) = 346, 222, 158, 37 Wm-1

K-1

for T = 300, 190, 150, 70 K, respectively]. The solid lines in Fig. 3.12d show good agreement

with our GNR data (L ≈ 260 nm) using Δ = 0.6 nm and a best-fit exponent n = 1.8 ± 0.3, as well

as c fitted to be 0.04019, 0.02263, 0.01689, and 0.00947 Wm-1

K-1

for T = 300, 190, 150, and 70

K, respectively. Note that we cannot assign overly great physical meaning to the parameter c

because the empirical model can only fit Δn/c, not Δ or c independently. The simple model

appears to be a good approximation in a regime with Δ ≪ W, where the data presented here were

fitted. However it is likely that this simple functional dependence would change in a situation

with extreme edge roughness [81], where the roughness correlation length (which cannot be

directly quantified here) could also play an important role.

Page 58: © 2014 Zuanyi Li

52

Nevertheless, the nearly W-squared dependence of thermal conductivity in narrow GNRs

with edge roughness is consistent with previous findings for rough nanowires [178, 179], and

also similar to that suggested by theoretical studies of GNR electron mobility [180]. The precise

scaling with Δ is ostensibly more complex [176, 177] than can be captured in a simple model, as

it depends on details of the phonon dispersion, the phonon wave vector, and indirectly on tem-

perature. However the Δ estimated from the simple model presented above is similar to that from

extensive numerical simulations below, and to that measured by transmission electron microsco-

py (TEM) on GNRs prepared under similar conditions [153]. Thus, the simple expressions given

above can be taken as a practical model for heat flow in substrate-supported GNRs with edge

roughness (Δ ≪ W) over a wide range of dimensions, corresponding to all size regimes in Fig.

3.10.

It is instructive to examine some similarities and differences between our findings here vs.

previous results regarding size effects on charge carrier mobility in GNRs with dimensions

comparable to the phonon or electron MFP. Figure 3.13 displays the width W dependence of our

GNR thermal conductivity side-by-side with the electrical mobility measured by Yang and

Murali [147] on similar samples (Fig. 3.13a is re-plot of Fig. 3.12d, here using log axis). Both

the GNR thermal conductivity and electrical mobility show a similar trend with W, starting to

decrease significantly when scattering becomes edge limited. However it is apparent that their

fall-off occurs at different critical widths: W ~ 200 nm for thermal conductivity and W ~ 40 nm

for electrical mobility. Above these widths, the thermal conductivity is limited by phonon-

substrate scattering, while the electrical mobility is limited by electron scattering with substrate

impurities. The difference between their critical W is consistent with the intrinsic phonon MFP

λph ~ 100 nm being approximately five times larger than the intrinsic electron MFP [104] λel ~ 20

Page 59: © 2014 Zuanyi Li

53

nm in SiO2-supported graphene. (We note phonons are entirely responsible for the thermal

conductivity of these GNRs, with a negligible electronic contribution [144]). Thus, the fall-off of

thermal conductivity and electrical mobility corresponds approximately to GNR widths approx-

imately twice the phonon and electron MFPs. Interestingly, this also suggests a GNR width

regime (~40 < W < ~200 nm) where the thermal conductivity is reduced from intrinsic values but

the electrical mobility is not yet affected by edge disorder. This suggests the possibility of

manipulating heat and charge flow independently in such narrow edge-limited structures. Further

control of such behavior could also be achieved if substrates with different roughness, impurity

density, and vibrational (phonon) properties are used.

3.5 Discussion and BTE Calculations

3.5.1 Low-T Scaling and Comparison with CNT

FIG. 3.13: Width dependence of GNR thermal conductivity and electrical mobility. (a) Thermal

conductivity k vs. W at several temperatures from this study (L ~ 260 nm); same plot as Fig.

3.12d but on a log-log scale. (b) Electrical mobility μ vs. W at room temperature for several layer

(L) GNRs, adapted from Yang et al. [147]. Solid line is a fit to data points with μ = (1/0.0163W3

+ 1/3320)-1

. Both data sets show a decreasing trend as W is narrowed, but with different fall-offs.

10 100 1000 1000010

100

1000

k (

Wm

-1K

-1)

W (nm)

T = 300 K

T = 190 K

T = 150 K

T = 70 K

a

10 100 100010

2

103

104

ML

BL

3-5L

6-8L

mob

ility

(cm

2V

-1s

-1)

W (nm)

bedge limited

impurity limited

edge limited

substrate limited

Page 60: © 2014 Zuanyi Li

54

Figure 3.14a shows the extracted thermal conductivity of our “short” graphene (GS1) and

GNRs for the full temperature range measured, down to ~20 K (however, we recall that GNR

measurements are challenging to distinguish below ~70 K, as previously mentioned). We can fit

the thermal conductivity as k = αTβ below ~200 K, and the obtained power β is shown as an inset

of Fig. 3.14a. We find that β decreases from ~1.6 for the unpatterned graphene to ~1 for narrow

GNRs. However, we note that this does not necessarily mark a transition to one-dimensional (1-

D) phonon flow, as the GNRs here are much wider than the phonon wavelengths (few nm).

Thus, the simple model is given as a convenient analytic estimate, and the exponent β represents

the complex physical behavior of GNR heat flow due to the increasing heat capacity (which

scales as ~T1.5

) and the slightly decreasing phonon MFP in this T range.

We also compare our extracted graphene and GNR thermal conductance with carbon nano-

tubes (CNTs) in Fig. 3.14b. We perform this comparison with the calculated ballistic upper limit,

with a single-wall carbon nanotube (SWCNT) by M. Pettes et al. [75], and with multi-wall

FIG. 3.14: Complete data sets including low-T range and comparison with CNTs. (a) Extracted

thermal conductivity of graphene (GS1) and GNRs, same as Fig. 3.12b but with low-T (20~60

K) data included for completeness. Long graphene data are from Seol et al. [49]. Inset: power

exponent β vs. GNR width fit from k = αTβ (see text). (b) Thermal conductance (G/A) of our data

compared with those of long graphene [49], ballistic limit, SWCNT of M. Pettes et al. [75], and

MWCNTs from studies of P. Kim et al. [71], M. Pettes et al. [75], and M. Fujii et al. [181].

10 1001

10

100

1000 Long Graphene (L~10 m)

Short Graphene (L~0.26 m)

GNR, W = 130 nm

GNR, W = 85 nm

GNR, W = 65 nm

GNR, W = 45 nm

k (

Wm

-1K

-1)

T (K)10 100

106

107

108

109

1010

MWCNT

MWCNT

SWCNT

MWCNT

G/A

(W

m-2K

-1)

Long Graphene (L~10 m)

Short Graphene (L~0.26 m)

GNR, W = 130 nm

GNR, W = 85 nm

GNR, W = 65 nm

GNR, W = 45 nm

T (K)

a b

~T2

~T1.5

~T

, Seol (2010)

Kim (2001)

Fujii (2005)

Pettes (2009)10 100 1000 10000

1.0

1.2

1.4

1.6

W (nm)

, Seol (2010)

Page 61: © 2014 Zuanyi Li

55

carbon nanotubes (MWCNTs) by P. Kim et al. [71], M. Pettes et al. [75], and M. Fujii et al.

[181]. We find that our short graphene data (red filled dots) and P. Kim’s MWCNT data [71]

(open black squares) have the highest values, both reaching up to ~35% of the ballistic limit of

graphene.

3.5.2 Effects of Contacts

We note that the 3D simulations automatically include all known contact resistance effects,

including those of the graphene-SiO2 and SiO2-metal interfaces, matched against data from the

literature and our control experiments (see Section 3.4.1). To provide some simple estimates, the

contact thermal resistance (per electrode width) is RC ≈ 0.7 m∙KW-1

, the “wide” unpatterned

graphene thermal resistance is RG ≈ 2.5 m∙KW-1

, and that of the GNR arrays is in the range RGNR

≈ 4 to 32 m∙KW-1

(from widest to narrowest). The graphene is not patterned under the electrodes,

thus the contact resistance remains the same for all samples. The 3D simulations also account for

heat spreading through the underlying SiO2, and our error bars include various uncertainties in

all parameters (see Section 3.3.2).

We also consider the effects of measurement contacts and how they relate to the interpreta-

tion of sample length in the quasi-ballistic heat flow regime. As in studies of quasi-ballistic

electrical transport [172, 173], we defined the “channel length” L as the inside edge-to-edge

distance between the heater and thermometer electrodes (Fig. 3.10c). Simple ballistic theory

assumes contacts with an infinite number of modes, and instant thermalization of phonons at the

edges of the contacts. The former is well approximated here by electrodes two hundred times

thicker than the graphene sheet, however phonons may travel some distance below the contacts

before equilibrating. The classical, continuum analog of this aspect is represented by the thermal

transfer length (LT) of heat flow from the graphene into the contacts [129], which is automatical-

Page 62: © 2014 Zuanyi Li

56

ly accounted in our 3D simulations (Fig. 3.4e). However, a sub-continuum perspective [182]

reveals that phonons only thermally equilibrate after traveling one MFP into the graphene under

the contacts. Previous measurements of oxide-encased graphene [54] had estimated a thermal

conductivity kenc = 50-100 Wm-1

K-1

, which suggests a phonon MFP λenc = (2kenc/π)/(Gball/A) ≈ 8-

15 nm under the contacts. This adds at most 12% to our assumption of edge-to-edge sample

length (here L ≈ 260 nm), a small uncertainty which is comparable to the sample-to-sample

variation from fabrication, and to the size of the symbols in Fig. 3.12c. (The relatively low

thermal conductivity of encased monolayer graphene [54] is due to scattering with the SiO2

sandwich, although some graphene damage from the SiO2 evaporation [183] on top is also

possible.)

3.5.3 BTE Calculations

To gain deeper insight into our experimental results, we employ a numerical solution of the

Boltzmann transport equation (BTE) with a complete phonon dispersion [177, 184]. Our ap-

proach is similar to previous work [28, 49, 184], but accounting for quasi-ballistic phonon

propagation and edge disorder scattering in short and narrow GNRs, respectively. We obtain the

thermal conductivity by solving the Boltzmann transport equation in the relaxation time approx-

imation including scattering at the rough GNR edges [177]. The simulation uses the phonon

dispersion of an isolated graphene sheet, which is a good approximation for SiO2-supported

graphene within the phonon frequencies that contribute most to transport [101], and at typical

graphene-SiO2 interaction strengths [97]. (However, we note that artificially increasing the

graphene-SiO2 coupling, e.g. by applying pressure [185], could lead to modifications of the

phonon dispersion and hybridized graphene-SiO2 modes [97].) We assume a graphene monolay-

er thickness H = 0.335 nm, and a concentration of 1% 13

C isotope point defects [21, 49]. The

Page 63: © 2014 Zuanyi Li

57

FIG. 3.15: Computed phonon scattering rates at T = 300 K for a GNR with W = 65 nm, L = 260

nm, and Δ = 0.65 nm. Substrate contact patch radius is a = 8.75 nm. Rates are plotted as a

function of energy for each scattering mechanism and their total; however, each dot represents

one phonon mode. Note that angle-dependent mechanisms like contact and edge roughness

scattering have additional dependence on the angle of the phonon velocity vector, which can lead

to different rates for the same value of phonon energy.

0 0.05 0.1 0.15 0.210

8

109

1010

1011

1012

Energy (eV)

Scattering r

ate

(s-1

)

0 0.1 0.20.05 0.1510

8

109

1010

1011

1012

Energy (eV)

Scattering r

ate

(s-1

)

0 0.05 0.1 0.15 0.210

5

107

109

1011

1013

1015

Energy (eV)

Scattering r

ate

(s-1

)

0 0.1 0.20.05 0.1510

8

109

1010

1011

1012

Energy (eV)

Scattering r

ate

(s-1

)

0 0.05 0.1 0.15 0.210

10

1011

1012

1013

1014

Energy (eV)

Scattering r

ate

(s-1

)

0 0.1 0.20.05 0.1510

10

1011

1012

1013

1014

Energy (eV)

Scattering r

ate

(s-1

)

Edge Roughness

Substrate Isotopes

Total

LA

a b

c d

e f

Umklapp

Contacts

ZO LO TOZA

TA

Page 64: © 2014 Zuanyi Li

58

interaction with the substrate is modeled through perturbations to the scattering Hamiltonian [49]

at small patches where the graphene is in contact with the SiO2, with nominal patch radius a =

8.75 nm. Anharmonic three-phonon interactions of both normal and umklapp type are included

in the relaxation time. An equivalent 2D ballistic scattering rate [171, 175] ~2vx/L is used in the

numerical solution (x is direction along graphene) to account for transport in short GNRs.

Figure 3.16a finds good agreement of thermal conductivity between our measurements and

the BTE model across all samples and temperatures. We obtained the best fit for GNRs of width

130 and 85 nm with rms edge roughness Δ = 0.25 and 0.3 nm, where the gray bands in Fig. 3.16a

correspond to ±5% variation around these values. For GNRs of widths 65 and 45 nm the gray

bands correspond to edge roughness ranges Δ = 0.35-0.5 and 0.5-1 nm, respectively. We note

that, unlike the empirical model of Eq. 3.5, the best-fit BTE simulations do not use a unique

value of edge roughness Δ. This could indicate some natural sample-to-sample variation in edge

roughness from the fabrication conditions, but it could also be due to certain edge scattering

physics (such as edge roughness correlation [81] and phonon localization [186]) which are not

yet captured by the BTE model.

Figure 3.16b examines the scaling of MFPs by phonon mode, finding they are strongly re-

duced as the GNR width decreases below ~200 nm, similar to the thermal conductivity in Fig.

3.12d. The frequency-dependent phonon MFP is directly obtained from the calculated total

scattering rate (e.g. Fig. 3.15f). The MFP for each phonon mode is calculated as an average over

the entire frequency spectrum, weighted by the frequency-dependent heat capacity cb(ω) and

group velocity vb(ω),

( ) ( ) ( )

( ) ( )

b b b

b

b b

c v d

c v d

. (3.6)

Page 65: © 2014 Zuanyi Li

59

We note that LA and TA modes, which have larger intrinsic MFPs, are more strongly affected by

GNR edge disorder. On the other hand ZA modes are predominantly limited by substrate scatter-

ing and consequently suffer less from edge disorder, consistently with recent findings from

FIG. 3.16: Insights from numerical simulations. (a) Comparison of Boltzmann transport model

(lines) with experimental data (symbols, same as Fig. 3.12b, but here on linear scale). (b) Com-

puted scaling of phonon mean free path (MFP) vs. width, for two sample lengths as listed. LA

and TA phonons with long intrinsic MFP are subject to stronger size effects from edge scattering

than ZA modes, which are primarily limited by substrate scattering. (c) Estimated contribution of

modes to thermal conductivity vs. edge roughness Δ, for a wide sample and a narrow GNR. Like

W, changes in Δ also more strongly affect LA and TA modes, until substrate scattering begins to

dominate. (d) Phonon MFP vs. energy (frequency) for a large sample and a narrow GNR. Low-

frequency modes are strongly affected by substrate scattering, such that effects of edge rough-

ness are most evident for ℏω > 15 meV. Panels (b-d) are all at room temperature. Also see Figs.

3.15 and 3.17.

0 50 100 1501

10

100

1000

Energy (meV)

mfp

(nm

)

101

102

103

104

0

50

100

150

200

W (nm)

mfp

(nm

)

0 100 200 3000

100

200

300

400

500

600

700

T (K)

k (

Wm

-1K

-1)

130 nm

85 nm

65 nm

large graphene (L ~ 10 μm)

Seol (2010)

short graphene

(L ~ 260 nm)

GNRs

GNR W ~

45 nm

a b

TA long

LA long

ZA long

ZA short

long: L=10 μm

short: L=260 nm (Δ~0.65 nm)

0.1 10.510

1

102

103

(nm)

k (

Wm

-1K

-1)

GNR

W ~ 65 nm

TA wide

LA wide

LA GNR

TA GNR

ZA wide

ZA GNR

total wide

total GNR

Seol (2010)

wide: W=2 μm, L=10 μm

GNR: W=65 nm, L=260 nm

c d

TA large

LA large

LA GNR

TA GNR

ZA large

Energy ħω (meV)

ZA GNR

large: W=2 μm, L=10 μm

GNR: W=65 nm, L=260 nm (Δ=0.65 nm)

Page 66: © 2014 Zuanyi Li

60

molecular dynamics (MD) simulations [97, 101].

Increasing edge disorder reduces phonon MFPs (Fig. 3.17d), and the thermal conductivity is

expected to scale as shown in Fig. 3.16c. In the BTE model the edge roughness scattering is

captured using a momentum-dependent specularity parameter,

2 2 2( ) exp 4 sin Ep q q , (3.7)

where q is the magnitude of the phonon momentum and θE is the angle between the direction of

phonon momentum and the edge. This specular parameter indicates that small wavelength (large

q) phonons are more strongly affected by line edge roughness. However, as Δ increases the

specularity parameter saturates, marking a transition to fully diffuse edge scattering, and also to a

regime where substrate scattering begins to dominate long wavelength phonons in substrate-

supported samples. This transition cannot be captured by the simplified Δn dependence in the

empirical model of Eq. 3.5.

To further illustrate such distinctions, Fig. 3.16d displays the energy (frequency ω) depend-

ence of phonon MFPs for a “small” GNR and a “large” SiO2-supported graphene sample (corre-

sponding to Figs. 3.10c and 3.10a, respectively). Low-frequency substrate scattering (propor-

tional to ~1/ω2) dominates the large sample [49, 97], while scattering with edge disorder affects

phonons with wavelengths comparable to, or smaller than, the roughness Δ. Therefore, larger Δ

can affect more long-wavelength (low energy) phonons, but only up to Δ ~ 1 nm, where the

effect of the substrate begins to dominate in the long wavelength region. (Also seen in Fig.

3.16c.) Such a separation of frequency ranges affected by substrate and edge scattering could

provide an interesting opportunity to tune both the total value and the spectral components of

thermal transport in GNRs, by controlling the substrate and edge roughness independently.

Page 67: © 2014 Zuanyi Li

61

We also show that the lattice thermal conductivity (Fig. 3.17a) and the phonon MFP (Fig.

3.17b) both scale with the length of the GNR in “short” ribbons due to the MFP being limited to

about half the length (τball = ½L/|vb(q)|). A difference exists between wide (W = 2 μm) and narrow

FIG. 3.17: Dimension scaling behaviors. (a) Variation of lattice thermal conductivity and (b)

mean free path with the length of graphene samples for a wide (W=2 μm) and narrow (W=65 nm)

ribbon, showing that the length effect is more pronounced in wide ribbons. Phonons in narrow

ribbons suffer more scattering at the rough edges; hence, the effect of length is weaker in narrow

GNRs. (c) Dependence of thermal conductivity on ribbon width shows good agreement with

experimental data (symbols). Our model shows that thermal conductivity in short (L=260 nm)

ribbons is independent of width when ribbons are wide (W>L), but strongly dependent on width

when they are narrow (W<L), consistent with strong diffuse scattering at the rough edges. (d)

Effect of rms edge roughness Δ is confirmed in the dependence of the MFP, where we can see

that the phonon MFP in wide ribbons is largely independent of Δ, while the MFP in narrow

ribbons decreases with increasing Δ, indicating the strong role of edge roughness in thermal

transport in narrow GNRs.

10 100 1000 10,0000

100

200

300

400

500

600

700

W (nm)

k (

Wm

-1K

-1)

TA long

LA long

LA short

TA short

ZA longZA short

Total long

Total short

0.1 10.510

1

102

LER (nm)

mfp

(nm

)

TA wide

LA wide

LA narrow

TA narrow

ZA wide

ZA narrow

Total wide

Total narrow

101

102

103

104

101

102

103

L (nm)

k (

Wm

-1K

-1)

TA wide

LA wide

LA narrowTA narrow

ZA wide

ZA narrow

Total wide

Total narrow

GNR W=65 nmSeol (2010)

graphene

101

102

103

104

101

102

L (nm)

mfp

(nm

)

TA wide

LA wide

LA narrow

TA narrow

ZA wide

ZA narrow

Total wide

Total narrow

a b

c d

Δ

Seol (2010)

Page 68: © 2014 Zuanyi Li

62

(W = 65 nm) ribbons due to the presence of edge roughness scattering in the narrow ribbons. In

the wide ribbons, the most significant scattering mechanism is substrate scattering, which limits

the phonon MFP to around 100 nm (the value to which the total MFP converges for “long”

ribbons in Fig. 3.17b). We note that this value is very nicely consistent with the average sub-

strate-limited (or intrinsic) MFP deduced through a simple comparison with the ballistic con-

ductance limits in the main text (Fig. 3.12 and surrounding discussion).

For a sample length L ~ 200 nm, or approximately twice the intrinsic MFP, the length-

dependent ballistic scattering rate is comparable to the substrate scattering rate and the effective

thermal conductivity and phonon MFP become one-half of the substrate-limited values (Fig.

3.17a-b for “wide” sample). However, in narrow GNRs, the diffuse scattering at the edges limits

the phonon MFP to approximately one-half of the width W (the edge-roughness-limited MFP

also depends on the rms edge roughness, as shown in Fig. 3.17d). Consequently, thermal

transport in narrow GNRs is mostly diffusive until the length is reduced to values comparable to

the width, and the transport again becomes quasi-ballistic (shown by solid lines in Fig. 3.17a-b).

3.6 Summary

In conclusion, we investigated heat flow in SiO2-supported graphene samples of dimensions

comparable to the phonon MFP. Short devices (L ~ λ, corresponding to Fig. 3.10b schematic)

have thermal conductance much higher than previously found in micron-sized samples, reaching

35% of the ballistic limit at 200 K and 30% (~1.2 GWK-1

m-2

) at room temperature. However,

narrow ribbons (W ~ λ, corresponding to Fig. 3.10c schematic) show decreased thermal

conductivity due to phonon scattering with edge disorder. Thus, the usual meaning of thermal

conductivity must be carefully interpreted when it becomes a function of sample dimensions.

The results also suggest powerful means to tune heat flow in 2D nanostructures through the

Page 69: © 2014 Zuanyi Li

63

effects of sample width, length, substrate interaction and edge disorder.

Page 70: © 2014 Zuanyi Li

64

CHAPTER 4

SUBSTRATE-SUPPORTED THERMOMETRY PLATFORM FOR NANO-

MATERIALS LIKE GRAPHENE, NANOTUBES, AND NANOWIRES*

4.1 Introduction

To probe nanoscale thermal conduction, different measurement techniques have been devel-

oped, such as the 3ω method, scanning thermal microscopy, time-domain thermoreflectance, and

various bridge platforms [30]. Among them, using a microfabricated suspended bridge can

directly measure in-plane heat flow through nanostructures by electrical resistance thermometry

[69, 187-189]. This suspended thermometry has good measurement accuracy and has been

widely used for nanofilms [189-191], one-dimensional (1D) materials like nanowires (NWs)

[187, 188] and nanotubes [73, 76], as well as two-dimensional (2D) materials like hexagonal

boron nitride [192] and graphene [51, 52]. A major drawback of this platform is that the fabrica-

tion is complicated and the test sample has to be either fully suspended [51, 69, 73, 76, 187-192]

or supported by a suspended dielectric (e.g. SiNx) membrane [52, 78]. This limits the diversity of

measureable materials and makes the suspended platform fragile.

To overcome these limitations, substrate-supported platforms could be preferred and have

been recently employed in thermal studies of Al nanowires [157], encased graphene [54] and

graphene nanoribbons in our previous work [55]. Nanomaterials are almost always substrate-

supported in nanoscale electronics [193], thus a substrate-supported platform has the advantage

of testing devices including extrinsic substrate effects [1, 106], which could be different from

their intrinsic thermal properties (probed by suspended platforms) [21, 23, 24]. Substrate-

* This Chapter is reprinted from Z. Li, M.-H. Bae, and E. Pop, “Substrate-Supported Thermometry Platform for

Nanomaterials Like Graphene, Nanotubes, and Nanowires,” Applied Physics Letters, vol. 105, p. 023107 (2014).

Copyright 2014, American Institute of Physics.

Page 71: © 2014 Zuanyi Li

65

supported thermometry platforms could also readily be incorporated in industrial mask designs

and fabrication processes as thermal test structures in addition to existing electrical test struc-

tures. Thus, measuring the thermal conduction of nanomaterials on a substrate is crucial from a

practical viewpoint.

In this work, we critically examine the applicability and limitations of nanoscale thermal

measurements based on a substrate-supported platform utilizing electrical resistance thermome-

try. As a prototype, the thermal conductivity of graphene on a SiO2/Si substrate was experimen-

tally tested. The fabrication of this supported platform is much easier than that of suspended

platforms, but as a trade-off the thermal conductivity extraction is slightly more challenging and

must employ a three-dimensional (3D) heat conduction simulation of the test structure. Through

careful uncertainty analysis, we find that the supported platform can be optimized to improve the

measurement accuracy. The smallest thermal sheet conductance that can be measured by this

method within a 50% error is ~25 nWK-1

at room temperature, which means the supported

platform can be applied to nanomaterials like carbon nanotube (CNT) networks, nanotube or

nanowire arrays, and even a single Si nanowire. Additionally, it is suitable for materials which

cannot be easily suspended, like many polymers, and the substrate is not limited to SiO2/Si but

can be extended to other substrates such as flexible plastics.

4.2 Thermometry Platform Demonstration

4.2.1 Platform Structure and Measurement

Figure 4.1a shows a scanning electron microscopy (SEM) image of a typical supported

thermometry platform, here applied to a monolayer graphene sample. (In general, the sample to

be measured is prepared on a SiO2/Si substrate, though this is not always necessary, as we will

show below.) Then, two parallel, long metal lines with at least four probe arms are patterned by

Page 72: © 2014 Zuanyi Li

66

electron-beam (e-beam) lithography as heater and sensor thermometers. If the sample is conduc-

tive (here, graphene), then the heater and sensor must be electrically insulated by a thin SiO2

layer, as seen in the Fig. 4.1b cross-section. To perform measurements, a DC current is passed

through one metal line (heater) to set up a temperature gradient across the sample, and the

electrical resistance changes of both metal lines (heater and sensor) due to the heating are

monitored. After temperature calibration of both metal line resistances the measured changes in

resistance (ΔR) can be converted into changes in temperature of the heater and sensor, ΔTH and

ΔTS, as a function of heater power PH.

FIG. 4.1: (a) Scanning electron microscopy image of supported thermometry platform designed

to measure thermal conductivity of a graphene sample (purple) on a SiO2/Si substrate. (b, c) 2D

and 3D finite element models used to simulate heat conduction in the supported thermometry

platform, respectively. In the 2D model, only the cross-section is included and the zoom-in

shows the typical temperature distribution with heating current applied through the heater. (d)

Zoomed-in temperature distribution around heater and sensor obtained from 3D simulation,

which matches with measured temperature. The white dashed lines indicate the outline of the

sample, which is highlighted by the red line and rectangle in (b) and (c), respectively. The

detailed shape and size of the sample will not affect the simulation.

a

c

b

d

sample

2 μm

2D Model

2LS = 6 μm

LS

= 3

μm

tbox

Si

sensorheater

sample

SiO2LHS

SiO2

Si

∆T (K)10-2 10-1 100 101

3D Model

LS

= 2

0 μ

m

3D Model

thin SiO2

zx

y

LDmet

sample

Page 73: © 2014 Zuanyi Li

67

As the sample is not suspended, a control experiment should be performed after removing

the exposed parts of the sample and repeating the above measurements to independently find the

thermal properties of the heat flow path through the contacts and substrate. Etching away the

sample is important, rather than simply performing the measurement without the original sample,

because it preserves the sample portion beneath the heater and sensor electrodes, i.e. the same

contact resistance in both configurations. Using this approach, in previous work [55] we obtained

the thermal conductivity of the underlying SiO2 within 5% error of widely known values, which

also helps support the validity of this approach.

4.2.2 Model to Extract Thermal Conductivity

We note that the fabrication of the supported platform can be performed with greater yield

than that of suspended platforms, but the thermal conductance between the heater and sensor

cannot be obtained analytically as in the suspended case, due to non-negligible heat leakage into

the substrate. Therefore, numerical modeling of such heat conduction must be employed to

extract the thermal conductivity of the sample. For comparison, we considered both 2D and 3D

finite element models of the sample, which are implemented by a commercial software package

(COMSOL). In the 2D model, only the cross-section of the platform is simulated, and the Si

substrate size is chosen as 2LS×LS (Fig. 4.1b). In the 3D model, half of the platform needs to be

simulated due to the symmetry plane which bisects the region of interest, and the Si substrate

size is chosen as 2LS×LS×LS (Fig. 4.1c). To perform the simulation, in both 2D and 3D models

the bottom and side boundaries (except symmetry plane in 3D) of the Si substrate are held at the

ambient temperature, i.e. isothermal boundary condition. Other outer boundaries of the whole

structure are treated as insulated, i.e. adiabatic boundary condition. Joule heating is simulated by

applying a power density within the heater metal, and the calculation is performed to obtain the

Page 74: © 2014 Zuanyi Li

68

temperature distribution in steady state, as shown in Figs. 4.1b and 4.1d for the 2D and 3D

models, respectively. After calculating the average temperature rises in the measured segments

of the heater and sensor, we obtain the simulated values of ΔTH and ΔTS vs. PH. Then, we can

match these with the measured values by fitting the thermal conductance G of the test sample

between the heater and sensor. The thermal conductivity of the sample, k = GL/A, can then be

extracted. Here L is the sample length, i.e. the heater-sensor separation and A is the cross-

sectional area, i.e. the sample width W times thickness h (see Figs. 4.1b and 4.1d).

To correctly obtain the sample thermal conductivity, the simulated size LS of the Si substrate

must be carefully chosen because the real Si chip is ~0.5 mm thick and several mm wide. The

chip dimensions are semi-infinite compared to the small heating region (~10 μm) and cannot be

fully included due to computational grid limits. Thus, the simulated LS should be large enough to

model the heat spreading and give a converged value of the extracted k. For the 2D and 3D

models, the extracted thermal conductivities as a function of simulated LS from our graphene

measurement [55] at 270 K are shown in Fig. 4.2a. Here, for the 3D model, we further consid-

ered two cases: heater and sensor with probe arms (as shown in Fig. 4.1d), and without probe

FIG. 4.2: (a) Extracted graphene thermal conductivity as a function of the Si substrate size LS for

different models. (b) Simulated temperature profiles along the sensor from 3D models with and

without probes. Two solid lines are obtained by using the same graphene k. Changing from solid

to dash red lines corresponds to the red arrow in (a). (c) Extracted graphene k converges as the

simulated probe length Lprobe increases [corresponding to the red arrow in (a)].

0 1 2 3 4 5 60.0

0.2

0.4

0.6

0.8

1.0

T

s (

K)

y (m)

a cb

0 20 40 60 80 10050

100

150

200

250

300

350

Simulated LS (m)

Extr

acte

d k

(W

m-1K

-1)

3D with probe

3D without probe

2D

3D with probe

3D without probe

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6150

200

250

300

Extr

acte

d k

(W

m-1K

-1)

Simulated Lprobe (m)

T=270 K LS=50 μmSimulated Lprobe

W/2

Page 75: © 2014 Zuanyi Li

69

arms (as shown in Fig. 4.3a), because the latter has a more direct correspondence to the 2D

model (reflected by the similar extracted k for small LS). Comparing these two cases also allows

us to test how many details of the electrode geometry should be included.

It is clear that the extracted k by the 2D model continues decreasing as the simulated LS in-

creases, whereas the two 3D models give converged k when the simulated LS is sufficiently large

(≥ 50 μm). The 2D model is insufficient because it neglects the heat spreading along the y-

direction perpendicular to the 2D plane. Although the heater and sensor length (~10 μm or

similar to the sample width) are long compared to their separation LHS (~0.5 μm), we find that

3D heat spreading about 10 μm away from the heating center cannot be neglected (see Fig. 4.3b).

As the simulated LS increases, the effect of neglecting the heat spreading along the y-direction

becomes stronger in the 2D model. Thus, we find that the 3D simulation is preferable in order to

fully capture all heat spreading effects due to the finite size of the sample.

FIG. 4.3: (a) Structure and temperature distribution for the 3D model without probe arms, which

can be regarded as the one extruded from the 2D model, and hence has a better correspondence

to the 2D model than the 3D with probe model. (b) Evolvement of temperature isosurface shape

in 3D simulation. When heat spreads out from the heater, the isosurface is close to a cylinder at

the beginning (inset), behaving as 2D heat conduction, while it changes to a sphere after ~10 μm

from the center, indicating heat spreading along the third direction cannot be ignored, which is

the reason why the 2D model fails.

b

LS

= 2

0 μ

m

∆T (K)

101

100

SiO2Silicon

∆T (K)10-2 10-1 100 101

a

Page 76: © 2014 Zuanyi Li

70

Figure 4.2a also shows that simulating the effect of heat loss through the voltage probe arms

is necessary. As shown in Fig. 4.2b, if the same k of the sample is used, the simulated tempera-

ture rise along the sensor for the “3D with probe” case (solid red line) is lower than that for the

“3D without probe” case (solid black line), and in the former case there are temperature dips at

the points where the probe arms are connected due to heat leakage through them. Figure 4.2c

shows that the extracted k increases and saturates gradually as the simulated length of the probe

arms Lprobe increases, indicating Lprobe ≥ 1.5 μm is sufficiently long to catch its effect and this

converged value finally provides a correct k of the sample.

4.3 Uncertainty Analysis and Optimization

4.3.1 Uncertainty Analysis of Measurements

Next, we turn to the estimation of uncertainty in the extracted thermal conductivity k, which

can be accomplished by the classical partial derivative method: uk/k = [Σi (si×uxi/xi)2]1/2

, where uk

is the total uncertainty of extracted thermal conductivity k, uxi is the estimated uncertainty for

each input parameter xi of the simulation, and the sensitivity si is defined by si = (xi/k)∂k/∂xi =

∂(lnk)/∂(lnxi). The sensitivity is evaluated numerically by giving a small perturbation for each

input parameter around its typical value and redoing the extraction simulation to obtain the

change of k [54]. To highlight the relative importance of each input parameter, we define its

absolute contribution as ci = |si|×(uxi/xi), and a relative contribution as ci2/Σci

2. As an example, the

calculated sensitivities and uncertainty analysis for the extracted k in Fig. 4.2c are shown in

Table 4.1. The total uncertainty in this case is ~20.8%, and it mainly arises from the contribu-

tions (ci > 5%) of the thermal conductivity of bottom SiO2 (kox), thermal boundary resistance

(TBR) of the SiO2/Si interface (Roxs), measured sensor response (ΔTS/PH), thermal conductivity

of Si substrate (kSi), and heater-sensor midpoint separation (LHS). The values and their uncertain-

Page 77: © 2014 Zuanyi Li

71

ties of kox and Roxs are from our previous measurements [55]. The thermal conductivity of metal

lines (kmet) is calculated from the measured electrical resistance according to the Wiedemann-

Franz Law. kSi is well-known data from Ref. [168]. The TBR values and their uncertainties for

the graphene/SiO2 interface (Rgox) and metal/SiO2 interface (Rmox) are based on measurements in

Refs. [48] and [45], respectively. These two interface resistances have small contributions to the

total uncertainty (ci ≤ 2%) due to their small sensitivities (si). Even when their TBR values

Table 4.1: Calculated sensitivities and uncertainty analysis for the extracted graphene k from our

measurement at 270 K [correspond to the converged value in Fig. 4.2c with LS = 50 µm and

Lprobe = 1.6 µm]. The total uncertainty is ~20.8%.

Units Values x i

Uncertainty

u xi

u xi /x i

Sensitivity

s i

Contribution

c i =|s i |×u xi /x i

c i2/Σc i

2

Sensor response ΔT S/P H 0.01623 0.0003 1.8% 4.41 8.2% 15.4%

Heater response ΔT H/P H 0.14080 0.001 0.7% 2.42 1.7% 0.7%

k ox 1.213 0.04 3.3% 4.32 14.2% 46.9%

k Si 127 15 11.8% 0.46 5.5% 6.9%

k met 49 4 8.2% 0.40 3.3% 2.5%

R gox 1.15E-08 2.0E-09 17.4% -0.11 1.8% 0.8%

R oxs 9.99E-09 3.5E-09 35.0% -0.25 8.8% 17.7%

R mox 1.02E-08 3.5E-09 34.3% 0.06 2.0% 0.9%

t box 294 1 0.3% -5.90 2.0% 0.9%

t tox 25 1 4.0% -0.02 0.1% 0.0%

t met 50 2 4.0% 0.40 1.6% 0.6%

Distance of H/S L HS 505 5 1.0% 5.09 5.0% 5.9%

D met 199 4 2.0% -0.15 0.3% 0.0%

D tox 243 4 1.6% -0.31 0.5% 0.1%

Half length of H/S

metal linesL met/2 5.9 0.04 0.7% 0.19 0.1% 0.0%

Distance of 2

Voltage probesD pV 4.21 0.02 0.5% 3.58 1.7% 0.7%

Distance of

Current and

Voltage probes

D pIV 1.04 0.02 1.9% -0.12 0.2% 0.0%

Geo

metr

ica

l

Thickness of

bottom and top

SiO2, and metal

nm

Width of metal and

top SiO2 lines

μm

Input parameters (T = 270 K)

Th

erm

al

Thermal

conductivity of

SiO2, Si, metal

W/m/K

TBR of

graphene/SiO2,

SiO2/Si, metal/SiO2

interfaces

m2K/W

Expt. K/μW

Page 78: © 2014 Zuanyi Li

72

change by a factor of 2 (i.e., uxi =100%), their contributions become ci = 10% and 6%, which

only changes the total uncertainty from ~20.8% to ~23% and ~21.5%, respectively. Thus, their

contributions are not shown in the following results.

4.3.2 Optimization of the Platform

The accuracy of our supported thermometry platform can be optimized through two im-

portant geometric parameters: i) the center-to-center distance between the heater and sensor

(LHS) and ii) the bottom insulator (here oxide) thickness (tbox) (see Fig. 4.1b). If LHS is too large

then too much of the heater power is dissipated into the substrate; if it is too short, then the

temperature drop between heater and sensor is not large compared with the temperature variation

under the heater/sensor. If tbox is too thin, significant heat leakage will occur into the substrate; if

it is too thick, then its lateral thermal conductance will dominate the heat flow between heater

and sensor, overwhelming that of the supported sample.

The optimized values of LHS and tbox can be found by monitoring the uncertainty change

FIG. 4.4: (a, b) Estimated uncertainty of extracted sample thermal conductivity as a function of

the heater-sensor midpoint separation LHS and the bottom oxide thickness tbox, respectively. This

gives the optimized design with LHS = 600 nm and tbox = 300 nm for the partly-etched sample

case (inset of b, red part indicates the sample). (c) Estimated uncertainty of extracted sample k

increases as its thermal sheet conductance G□ decreases for LHS = 600 nm and tbox = 300 nm,

showing the measureable G□ (blue region) by this SiO2/Si supported thermometry platform. Inset

is a schematic for CNT/NW array measurements, where green lines are CNTs/NWs, and red and

blue lines are heater and sensor, respectively.

200 400 600 800 10000

5

10

15

20

25

30

35

Un

ce

rta

inty

(%

)

LHS

(nm)0 200 400 600

0

10

20

30

40

50

Un

ce

rta

inty

(%

)

tbox (nm)10 100

0

20

40

60

80

100

Uncert

ain

ty (

%)

G (nWK

-1)

total

ΔTS

Roxs

kSi

kox

LHS

DpV

tbox

a b c

tbox=300 nm

LHS=600 nm

Page 79: © 2014 Zuanyi Li

73

(due to sensitivity change) of extracted sample k, and the results are shown in Figs. 4.4a and

4.4b. Here we consider T = 300 K and thin film sample of thermal sheet conductance G□ = kh =

100 nWK-1

, and assume the sample is partly etched off so that only the part between the heater

and sensor is preserved (see the inset of Fig. 4.4b). The optimization is calculated at heater and

sensor linewidth Dmet = 200 nm, and only the input parameters whose contributions ci are larger

than 2% are included. From the estimated total uncertainty, we find that the optimized values of

LHS and tbox are ~ 600 nm and ~300 nm, respectively, leading to the minimized uncertainty

~18%. The non-etched case has the similar results but with a slightly higher uncertainty ~25%

(see Figs. 4.5a and 4.5b). By looking at the uncertainty contributed by each input parameter, it is

clear that the optimization is achieved mainly due to the competition between the measured

sensor response (ΔTS) and the role of the bottom oxide (kox), as we explained above. Additional

calculations (not shown) indicate that narrower heater and sensor linewidths (~100 nm) give

almost the same optimized values of LHS and tbox, but slightly lower total uncertainty.

FIG. 4.5: Results for the sample non-etched case. (a, b) Optimize the heater-sensor distance LHS

and the bottom oxide thickness tbox, respectively. The optimized LHS and tbox are almost the same

as those for the sample non-etched case (Fig. 4.4), but with a slightly higher uncertainty ~25%.

(c) Estimated uncertainty of extracted sample k increases as its thermal sheet conductance G□

decreases for LHS = 600 nm and tbox = 300 nm, showing the measureable G□ within a 50% error

(blue region) is G□ ≥ 32 nWK-1

. Inset is a schematic for CNT/NW array measurements, where

green lines are CNTs/NWs, and red and blue lines are heater and senor, respectively.

200 400 600 800 10000

10

20

30

40

50

60

70

Uncert

ain

ty (

%)

LHS

(nm)

0 100 200 300 400 500 6000

10

20

30

40

50

60

Un

ce

rta

inty

(%

)

tbox

(nm)10 100

0

20

40

60

80

100

Un

ce

rta

inty

(%

)

G (nWK

-1)

a b ctotal

ΔTS

Roxs

kSi

kox

LHS

tbox

DpV

tbox=300 nm LHS=600 nm

Page 80: © 2014 Zuanyi Li

74

4.4 Applicability of the Platform

4.4.1 Measureable Thermal Sheet Conductance

After optimizing our supported thermometry platform, the next question concerns the small-

est in-plane thermal conductance that can be sensed by this method. To address this, we calculate

the uncertainty change of extracted thermal conductivity as a function of the sample thermal

sheet conductance G□ = kh. When the sample is partly etched to match the dimensions of the

heater-sensor width and spacing (inset of Fig. 4.4b) the results at T = 300 K are shown in Fig.

4.4c. As expected, the estimated measurement uncertainty increases as G□ decreases. If we set

the maximum uncertainty to ~50%, the sensible range of G□ enabled by this platform is G□ > 25

nWK-1

(the blue region in Fig. 4.4c). For samples that are not etched to conform to the heater-

sensor width and separation (e.g., Fig. 4.1a), the uncertainty is slightly higher, and the smallest

sensible G□ within a 50% error is ~32 nWK-1

(see Fig. 4.5c). This requirement could be satisfied

in most thin film materials, such as polymer films (k > 0.2 Wm-1

K-1

and h > 200 nm) [194] and

CNT networks (k > 20 Wm-1

K-1

and h > 2 nm) [195, 196], etc.

4.4.2 Nanotube or Nanowire Array

We emphasize that this supported thermometry platform can be applied not only to thin

films but also to arrays of quasi-one-dimensional materials (inset of Fig. 4.4c). In our previous

work [55], we had shown its application to graphene nanoribbon arrays. Here we give estima-

tions of minimum array density required to apply the platform to carbon nanotube and Si nan-

owire arrays. For single-wall CNT arrays, assuming array density p (the number of CNTs per

unit width), the equivalent thermal sheet conductance is G□ = k(πdδ)p, where k, d, and δ are the

thermal conductivity, diameter, and wall thickness of single-wall CNTs, respectively. Then, the

array density is given by p = G□/(kπdδ). Considering single-wall CNTs with k = 1000 Wm-1

K-1

, d

Page 81: © 2014 Zuanyi Li

75

= 2 nm, and δ = 0.34 nm, as well as G□ = 25 nWK-1

(the best case), we obtain the density re-

quired for CNT array measurements is p ≥ 12 μm-1

. This CNT array density is achievable exper-

imentally today, as some studies [197, 198] have demonstrated CNT densities up to ~50 μm-1

.

For NW arrays and for thicker multi-wall CNTs the array density is given by p =

G□/(kπd2/4), where k and d are the thermal conductivity and diameter of the NWs, respectively.

For 20 nm diameter Si NWs [77] with k ≈ 7 Wm-1

K-1

, the required array density is p ≥ 11 μm-1

;

for 50 nm diameter Si NWs, smooth and rough edges lead to k ≈ 25 and 2 Wm-1

K-1

, respectively

[77, 79], and the required array density is p ≥ 0.5 μm-1

and 6 μm-1

. Nanowire arrays can be

fabricated much denser than these required densities [78, 199]. The above estimations indicate

that the supported thermometry platform can be easily applied to CNT arrays and Si NW arrays

with both smooth and rough edges. In addition, such arrays do not require uniform spacing.

4.4.3 Single Si Nanowire

We note that the minimum array density for 50 nm diameter smooth Si NWs is very low

(~0.5 μm-1

), which implies that it is possible to measure a single Si NW by using this platform.

To confirm this idea, we performed the simulation with just one NW between the heater and

sensor (inset of Fig. 4.6a). To achieve the best measurement accuracy, we first optimize the

dimensions of the heater and sensor, that is, the midpoint distance between them (LHS) and the

distance between two voltage probe arms (DpV) (inset of Fig. 4.6a). The calculated uncertainty

contributed from the measured sensor temperature rise (ΔTS) as a function of LHS and DpV is

shown in Fig. 4.6a. In the calculation, the bottom oxide thickness (tbox) and electrode linewidth

(Dmet, see Fig. 4.1b) are chosen as 300 nm and 200 nm, respectively, and a Si NW with d = 50

nm and k = 25 Wm-1

K-1

is used. The minimum of the uncertainty indicates the optimized struc-

ture is LHS = 600 nm and DpV = 1000 nm. By using these values, the total uncertainty as a

Page 82: © 2014 Zuanyi Li

76

function of kA (A is the NW cross-sectional area) is calculated and shown in Fig. 4.6b. For highly

conductive NWs (kA > 5×10-14

WmK-1

) the measurement uncertainty is around 60%, indicating

that obtaining an estimate of the thermal properties of a single NW is possible. However, this

also indicates that it is not possible to measure an individual single-wall CNT with the supported

platform because its kA is low (< 2×10-14

WmK-1

), although it may be possible to measure one

multi-wall CNT as long as the kA condition above is satisfied.

4.4.4 Plastic Substrate

Before concluding, we note that the supported thermometry platform is not limited to

SiO2/Si substrates, but could be extended to thermally insulating plastic substrates like Kapton,

polyethylene terephthalate (PET) and so on. Here, we consider measurements with a 25 μm thick

Kapton substrate on a heat sink (inset of Fig. 4.7a). We note that results for other Kapton thick-

ness or PET are similar. In simulations, the ambient (isothermal) boundary condition should be

applied to at least one surface of the substrate, but due to its low thermal conductivity the outer

FIG. 4.6: (a) Optimizing the design for measuring a single nanowire (see inset) by estimating the

uncertainty of extracted nanowire thermal conductivity as a function of LHS and DpV. Here only

the uncertainty contributed from the measured sensor temperature rise (ΔTS) is calculated. The

optimized design is LHS = 600 nm and DpV = 1000 nm. (b) Estimated uncertainty of extracted

nanowire k increases as its cross-sectional thermal conductance kA decreases.

400 600 800 1000 0

2000

4000

30

40

50

60

70U

ncert

ain

ty (

%)

10-14

10-13

0

20

40

60

80

100

120

140 total

TS

Wmet

kSi

kox

tbox

Dmet

kmet

DpVV

dwire

Un

ce

rta

inty

(%

)

kA (WmK-1)

a b

DpV

NW

LHS DpV=1000 nm

LHS=600 nm

total

ΔTS

Dmet

kSi

kox

tbox

LHS

kmet

DpV

dwire

Page 83: © 2014 Zuanyi Li

77

surfaces of the plastic substrate cannot reach the ambient temperature. Thus, a heat sink with

high thermal conductivity is added for this purpose. On the other hand, from a practical

viewpoint, measurements are generally performed in vacuum and on a chuck for ambient

temperature control, that is, on a heat sink. Since the plastic substrate is generally tens of microns

thick, its background thermal conductance will be typically larger than that of the sample; thus

the sample should be trimmed (etched) leaving just the portion between the heater and sensor

(inset of Fig. 4.7b), otherwise it will be difficult to sense the difference between the sample

measurement and the control experiment without the sample, resulting in a large uncertainty. By

using kps = 0.37 Wm-1

K-1

for Kapton (DuPontTM

Kapton® MT) and G□ = 100 nWK-1

for the

sample, we calculate the measurement uncertainty as a function of the heater-sensor distance

LHS, as shown in Fig. 4.7a. The minimized uncertainty is ~ 38% at LHS = 1.2 μm. With this

optimized structure, we further calculate the uncertainty change as a function of the sample

thermal sheet conductance (Fig. 4.7b), and find the smallest sensible G□ within a 50% error for a

FIG. 4.7: (a) Optimizing the heater-sensor midpoint separation LHS with the platform supported

by a plastic substrate on a heat sink (see inset). The estimated uncertainty of extracted sample k

is minimized at LHS = 1.2 µm for a 25 μm thick Kapton substrate. (b) Estimated uncertainty of

extracted sample k increases as its thermal sheet conductance G□ decreases, giving the sensible

range (blue) of G□ by this platform. Inset shows the optimized structure of the platform applied

to the plastic substrate, with LHS = 1.2 µm and the sample (red) only between heater and sensor.

10 1000

20

40

60

80

100

Uncert

ain

ty (

%)

G (nWK

-1)

0 1 2 30

10

20

30

40

50

60

70

80

Un

ce

rta

inty

(%

)

LHS (m)

a btotal

ΔTS

kps

kmet

tmet

LHS

DpV

plastic sub

heat sink

Page 84: © 2014 Zuanyi Li

78

Kapton substrate is ~60 nWK-1

, which is ~2.5 times higher than for the optimized SiO2/Si

substrate. Correspondingly, the required density for CNT and NW arrays will be also 2.5 times

higher, which remains achievable in experiments.

4.5 Summary

In conclusion, we demonstrated that a relatively simple, substrate-supported platform can be

used to measure heat flow in nanoscale samples like graphene, CNT or NW arrays. This platform

requires fewer fabrication efforts and is useful for materials which are difficult to suspend, but

the sample thermal conductivity must be extracted by 3D finite element analysis. Based on

careful uncertainty analysis we find the platform design can be optimized and the smallest

thermal sheet conductance measurable by this method within 50% error is estimated to be ~25

nWK-1

at room temperature. This thermometry platform can also be applied to individual nan-

owires, and can be implemented both on SiO2/Si (or similar) and flexible plastic substrates.

Page 85: © 2014 Zuanyi Li

79

CHAPTER 5

BALLISTIC THERMAL CONDUCTANCE AND LENGTH-DEPENDENT SIZE

EFFECTS IN LAYERED TWO-DIMENSIONAL MATERIALS

5.1 Introduction

Stimulated by extensive studies of graphene, enormous interest is being generated in the

properties of other two-dimensional (2D) layered materials, such as hexagonal boron nitride (h-

BN) and transition-metal dichalcogenides (TMDs, e.g., MoS2, and WS2). They exhibit remarka-

ble physical and chemical properties, and are promising candidates for applications in electron-

ics, optoelectronics, and photonics [9, 18, 19, 143]. Understanding thermal properties of materi-

als is important for improving thermally limited performance in devices and efficiency of energy

conversion [1, 5, 106]. The thermal properties of these layered materials are unique and highly

anisotropic, including high in-plane but very low cross-plane thermal conductivities, due to

strong in-plane chemical bonding and weak interlayer van der Waals interactions [23, 24]. Like

other materials, as sample dimensions are comparable to the average phonon mean free path

(MFP) λ, size effects of thermal transport become important [5]. However, compared to bulk

properties our understanding in this scale is relatively lacking in general (not limited to layered

materials).

First, if the sample length L is shorter than average phonon MFP λ, heat conduction will be

ballistic (i.e., no scattering) and governed by an upper limit, ballistic thermal conductance (Gball),

but the corresponding values for these layered materials (even for most materials) are still

unknown except for graphene [29, 55]. Second, in the range of L > λ, how thermal conductivity k

evolves from the ballistic to diffusive regime as L increases is not well established, and when it

will eventually enter the diffusive regime is not clear. Particularly, some theoretical models

Page 86: © 2014 Zuanyi Li

80

predicted that in low-dimensional momentum-conserving systems k will diverge with L, i.e., ~Lα

for one-dimension (1D) [200, 201] and ~logL for 2D [201-203]. Whereas, in what length scale

the divergence might occur and should be experimentally tested has not been carefully discussed.

Third, as an important characteristic length, MFP itself has not been estimated consistently. For

example, the “textbook” phonon MFP value for Si is ~40 nm at room temperature [204], ob-

tained from the kinetic theory by using the sound velocity as phonon group velocity [205].

However, by analyzing accumulative thermal conductivity as a function of MFP, phonons with

MFP longer than 1 µm contribute ~40% to the bulk k in Si [206-209], so a “median” MFP of 0.5-

1 µm is suggested [204]. Thus, it is crucial to systematically study these L-dependent size effects

as well as the proper estimation of phonon MFP.

In this work, we study ballistic-diffusive thermal conduction in monolayer graphene, h-BN,

MoS2, and WS2, as well as their three-dimensional (3D) bulk counterparts. Based on full phonon

dispersions obtained from ab initio simulations, we calculate the in-plane (for monolayer and

bulk) and cross-plane (for bulk) ballistic thermal conductance Gball of these materials for the first

time. Due to their stronger chemical bonding, graphene/graphite and h-BN show higher Gball than

MoS2 and WS2 above ~100 K in general. From our calculated Gball, we can obtain L-dependent

thermal conductivity by using a phenomenological ballistic-diffusive model, which shows good

agreement with very recent measurements [57] of k vs. L up to 9 µm in suspended graphene. We

also show that a proper estimation of the overall MFP λ should be a rigorous average of all

phonon modes, which can be condensed to an expression just including Gball and diffusive

thermal conductivity kdiff. We point out that the kinetic theory can reach the same result as long

as the correctly averaged phonon group velocity is used. Importantly, for anisotropic layered

materials, 2D and 1D forms of the kinetic theory should be used for in-plane and cross-plane,

Page 87: © 2014 Zuanyi Li

81

respectively. Based on the ballistic-diffusive model, k eventually converges to its diffusive value,

kdiff, around L ~ 100λ, much longer than the commonly assumed length. Thus, whether k is

divergent in low-dimensional materials should be examined beyond ~100λ to distinguish from

the intrinsic increase of k due to the ballistic-to-diffusive transition. All these results broaden our

understanding of thermal transport in low-dimensions and short length scales, and will help

guide the use of these layered materials for tailored applications.

5.2 Phonon Dispersions by First-Principles Calculations

The equilibrium atomic structures are calculated within density functional theory (DFT) us-

ing projector-augmented wave (PAW) potentials [210], as implemented in the VASP code [211].

The local density approximation (LDA) is used for the exchange-correlation functional [212].

The plane-wave basis set with kinetic energy cutoff of 400 eV is used except for h-BN (450 eV

Fig. 5.1: Atomic structures for graphite with AB stacking (a), h-BN with AA' stacking (b), and

TMD with 2H-phase (c), where three atomic layers are plotted for each of them. Gray lines

indicate primitive cells for their bulk, and lattice constants a and c are labeled. Unrepeatable

atoms within the primitive cells are highlighted by bigger and darker colored balls: 4 atoms for

graphite and h-BN, 6 atoms for TMDs. (d) Schematic of their Brillouin zone with high-symmetry

lines highlighted. For their 2D monolayers, the material thickness is generally defined as c/2, and

the 2D BZ is simply a regular hexagon (see inset of Fig. 5.2a).

A

MK

LHΓ

ca

d

bgraphite h-BN TMD

A

B

A

A

A’

A

N

B

C

S, Se

Mo, W

A

A’

A

c

c

a

a a

c

Page 88: © 2014 Zuanyi Li

82

is used). The convergence for energy is chosen as 10-8

eV between two consecutive steps, and

Table 5.1: Optimized lattice constants from our DFT calculations compared to experimental

values from literature [26, 213-220].

Experiments DFT simulations

a (Å) c (Å) a (Å) c (Å)

graphite [26, 213-216] 2.46 6.71 2.446 6.414

h-BN [217, 218] 2.50 6.66 2.490 6.488

MoS2 [219] 3.15 12.29 3.121 12.087

WS2 [220] 3.153 12.323 3.123 12.115

Fig. 5.2: Calculated phonon dispersions along high-symmetry lines for monolayer graphene (a),

h-BN (b), MoS2 (c), and WS2 (d). For graphene, red circles [25] and blue triangles [26] are

experimental data plotted for comparison. Other experimental data are also shown for h-BN

[221] and MoS2 [219]. The 2D Brillouin zone for monolayers is shown as the inset in (a).

0

50

100

150

200

Phonon E

(m

eV

)

0

10

20

30

40

50

60

Phonon E

(m

eV

)

0

10

20

30

40

50

60 Phonon E

Phonon E

(m

eV

)

0

50

100

150

200

Phonon E

(m

eV

)

LO

LA

TA

TO

ZA

ZO

a b

c d

Γ M

K

graphene h-BN

MoS2 WS2

Page 89: © 2014 Zuanyi Li

83

full geometry relaxation is carried out until remaining forces on atoms smaller than 0.01 eV/Å.

For monolayers, a vacuum spacing of 20 Å is used to prevent interlayer interactions.

After obtaining relaxed lattice constants and atomic structures, the interatomic force con-

stants are calculated within a supercell using the frozen-phonon approach [223]. For gra-

phene/graphite and h-BN, the plane-wave kinetic energy cutoff is increased to 1000 eV, and the

supercell size is 5×5×1 and 5×5×2 for their monolayers and bulk, respectively. For MoS2 and

WS2, the energy cutoff is still 400 eV, and the supercell size is 7×7×1 and 4×4×3 for their

monolayers and bulk, respectively. Last, phonon dispersions are calculated by the PHONOPY

Fig. 5.3: Calculated phonon dispersions along high-symmetry lines for bulk graphite (a), h-BN

(b), MoS2 (c), and WS2 (d). Insets in (a) and (b) are zoom-in dispersions along the cross-plane

direction. For graphite, green diamonds [222], red circles [25], and blue triangles [26] are

experimental data plotted for comparison. Other experimental data are also shown for h-BN

[221] and MoS2 [219].

0

50

100

150

200

Ph

on

on

E (

me

V)

0

10

20

30

40

50

60

Phonon E

(m

eV

)

0

10

20

30

40

50

60

Phonon E

(m

eV

)

0

50

100

150

200

Phonon E

(m

eV

)

M KA

graphite h-BN

MoS2 WS2

a b

c d

0

10

20

A 0

5

10

15

A

LO1,LO2

TO1,TO2

ZO1,ZO2

Page 90: © 2014 Zuanyi Li

84

package [224] interfaced with VASP.

The atomic lattices of bulk graphite, h-BN, and TMDs are shown in Figs. 5.1a-c. Graphite is

well known to crystalize in the AB stacking, while h-BN favors the AA' stacking [217, 225]. For

TMDs, the most common 2H phase [18, 19] (similar to the AA' stacking in h-BN) is considered

Fig. 5.4: (a) Brillouin zone of calculated bulk layered materials. High-symmetry lines and points

are highlighted. M1 and L1 are mid-points of lines Γ–M and A–L, respectively. (b-d) Zoomed-in

cross-plane phonon dispersions along Γ–A for graphite, h-BN, and MoS2. (e-h) Calculated in-

plane phonon dispersions on the ALH plane and cross-plane dispersions along Γ–A and M1–L1

directions for graphite, h-BN, MoS2, and WS2. For h-BN cross-plane (M1–L1), optical branches

with high phonon group velocity are highlighted in red and labeled. In (b-h) symbols are experi-

mental data for graphite (blue triangles [26] and green diamonds [222]), h-BN [221] and MoS2

[219].

0

10

20

30

40

50

60

L

Phonon E

(m

eV

)

H M1

L1

0

10

20

30

40

50

60

H

Phonon E

(m

eV

)

M1

L1

L

0

50

100

150

200

M1

Phonon E

(m

eV

)

L1

L H

0

50

100

150

200

M1

L1

Phonon E

(m

eV

) L H

a

0

5

10

15

20

Ph

ono

n E

(m

eV

)

A

graphite h-BN

MoS2 WS2

e f

g h

LO2

A

M1K

L1H

Γ

M

L

LO1

ZO2

ZO1

b d

0

5

10

15

20

A

cgraphite h-BN

0

2

4

6

8

A

MoS2

LA

TA(2)

TO’(2)

LO’

LA

TA(2)

TO’(2)

LO’

LO1,LO2

TO1,TO2

ZO1,ZO2

Page 91: © 2014 Zuanyi Li

85

here. The Brillouin zone (BZ) of their bulk is shown in Fig. 5.1d. Table 5.1 summarizes our

relaxed lattice constants, and experimental results [26, 213-220] are listed as well for compari-

son. Our calculations show that using LDA gives underestimated lattice constants than experi-

mental values, but it gives better agreement with experimental phonon dispersions than using

generalized gradient approximation (GGA).

Figure 5.2 shows the calculated phonon dispersions along high-symmetry lines in BZ (see

inset) for monolayer graphene, h-BN, MoS2 and WS2. The phonon dispersions of their bulk are

shown in Fig. 5.3. Our calculations are in excellent agreement with the experimental data [25,

26, 219, 221, 222] plotted as symbols in Figs. 5.2 and 5.3. For monolayer graphene and h-BN,

there are six phonon branches due to two atoms in their primitive cells (Figs. 5.1a-b); whereas,

there are nine phonon branches for monolayer MoS2 and WS2 because of three atoms in their

primitive cells (Fig. 5.1c). For the bulk counterparts, each branch in their monolayer phonon

dispersions splits into two branches because two layers form a primitive cell in bulk (Fig. 5.1).

The two branches are distinguishable on the ΓMK plane, but they become degenerate on the

ALH plane in BZ (see Figs. 5.4e-h). Moreover, the dispersion along the cross-plane direction

appears in bulk materials (see Fig. 5.4). For cross-plane dispersions along the Γ–A direction,

transverse acoustic (TA) and transverse optical (TO’) branches are double degenerate, as shown

in Figs. 5.4b-d. Interestingly, some high-frequency optical branches (ZO1, ZO2, LO1, LO2) of h-

BN have significantly nonzero cross-plane group velocity, different from other layered materials,

which can be noticed along the M1–L1 direction (Fig. 5.4a) for example (see Figs. 5.4e-h). Thus,

when these modes in h-BN are excited as temperature T increases, they will provide extra

contributions to ballistic thermal conductance Gball, resulting in a “bump” in Gball versus T of h-

BN (see Fig. 5.6b).

Page 92: © 2014 Zuanyi Li

86

5.3 Formalism of Thermal Physical Quantities

5.3.1 Heat Capacity

After obtaining full phonon dispersion in the Brillouin zone (BZ), heat capacity can be cal-

culated by

V ,

1( )b

b q

fC q

V T

, (5.1)

where the material volume V=WLH, a product of width W, length L, and hight H; ħ is the re-

duced Planck constant; q and ( )b q are phonon wavevector and frequency (b denotes different

branches), respectively; f=1/[exp(ħωb/kBT)-1] is the Bose-Einstein distribution; T is temperature

and kB is the Boltzmann constant. The sum Σ is over different branches (b) and all wave vectors (

q ) in the entire Brillouin zone. In practice, the sum over q is converted to an integral, and the

general expressions for one-dimension (1D), two-dimension (2D), and three-dimension (3D) are

given by

V

BZ

1

2

b

b

fC dq

A T

, (1D) (5.2a)

2

V 2 BZ

1

(2 )

b

b

fC dq

H T

, (2D) (5.2b)

3

V 3 BZ

1

(2 )

b

b

fC dq

T

, (3D) (5.2c)

where the integral is calculated in the entire BZ. For 2D and 3D, considering isotropic and

anisotropic cases, the expressions change to

V

1

2

b

b

fC qdq

H T

, (isotropic 2D) (5.3a)

2

V 2

1

2

b

b

fC q dq

T

, (isotropic 3D) (5.3b)

Page 93: © 2014 Zuanyi Li

87

V 2

1

4

b

zb

fC dq qdq

T

, (anisotropic layered 3D) (5.3c)

These equations will be used in the determination of pre-factors for averaged phonon group

velocity v in Section 5.3.5. Figure 5.5 shows our calculated heat capacity from Eq. 5.2b (mono-

layer) and Eq. 5.2c (bulk) for graphite, h-BN, MoS2, and WS2.

5.3.2 Ballistic Thermal Conductance

Ballistic thermal conductance Gball can be calculated from full phonon dispersion without

any approximation by

ball ,

1( ) | ( ) |

2

b b

nb q

fG q v q

L T

, (5.4)

Fig. 5.5: Calculated heat capacity as a function of temperature for graphite (a), h-BN (b), MoS2

(c), and WS2 (d) from Eq. 5.2b (for monolayer) and Eq. 5.2c (for bulk).

1 10 100 100010

-3

10-2

10-1

100

101

102

103

Cv (

JK

-1kg

-1)

T (K)

101

102

103

104

105

106

Cv (

JK

-1m

-3)

1 10 100 100010

-3

10-2

10-1

100

101

102

103

Cv (

JK

-1kg

-1)

T (K)

101

102

103

104

105

106

107

Cv (

JK

-1m

-3)

1 10 100 100010

-3

10-2

10-1

100

101

102

103

Cv (

JK

-1kg

-1)

T (K)

101

102

103

104

105

106

107

Cv (

JK

-1m

-3)

1 10 100 100010

-3

10-2

10-1

100

101

102

103

Cv (

JK

-1kg

-1)

T (K)

101

102

103

104

105

106

Cv (

JK

-1m

-3)

~T3

~T2

~T1

~T3

~T2

~T1

h-BN

MoS2 WS2

graphene/graphitea b

c d

~T3

~T2

~T1

~T3

~T2

~T1

Page 94: © 2014 Zuanyi Li

88

where ( )b

nv q = ˆ ( )b

qn q is phonon group velocity along the direction n (a normalized vector).

After converting the sum to an integral and dividing it by the cross-sectional area A=WH to get a

value independent of material sizes, the general expressions of Gball/A for 1D-3D are given by

ball

BZ

1( ) | ( ) |

4

b b

b

G fdq q v q

A A T

, (1D) (5.5a)

2ball

2 BZ

1( ) | ( ) |

8

b b

nb

G fdq q v q

A H T

, (2D) (5.5b)

3ball

3 BZ

1( ) | ( ) |

16

b b

nb

G fdq q v q

A T

, (3D) (5.5c)

In our calculations, ( )b

zv q is taken for cross-plane (3D); ( )b

xv q and ( )b

yv q are taken for in-plane

(2D and 3D) and they show almost the same results (within 2%), demonstrating the isotropy of

in-plane thermal conduction. The average of two results using ( )b

xv q and ( )b

yv q is adopted for in-

plane Gball/A.

For isotropic 2D and 3D, the velocity along a certain direction, ( )b

nv q can be expressed in

terms of the radial velocity scalar vb(q) as ( )b

nv q = vb(q)cosφ and ( )b

nv q = vb(q)sinθcosφ, respec-

tively, i.e., ( )b

nv q is the projector of vb(q) along the direction of temperature gradient. For aniso-

tropic 3D layered materials, we have 2( )b

xv q +2( )b

yv q =2

|| ( , )b

zv q q with || ( , )b

zv q q as the in-plane

velocity scalar, and | ( )b

zv q |= ( , )b

zv q q with ( , )b

zv q q as the cross-plane velocity scalar. Thus, Eq.

5.5b-c in different cases change to

2ball

2 0

2

1| ( )cos |

8

1 ( )

2

b b

b

b b

b

G fqdq d v q

A H T

fqdq v q

H T

, (isotropic 2D) (5.6a)

Page 95: © 2014 Zuanyi Li

89

22ball

3 0 0

2

2

1sin | ( )sin cos |

16

1 ( )

8

b b

b

b b

b

G fq dq d d v q

A T

fq dq v q

T

, (isotropic 3D) (5.6b)

2ball, ||

||3 0

||3

1| ( , ) cos |

16

1 ( , )

4

b b

z zb

b b

z zb

G fdq qdq d v q q

A T

fdq qdq v q q

T

, (layered 3D in-plane) (5.6c)

2ball,

3 0

2

1| ( , ) |

16

1 ( , )

8

b b

z z zb

b b

z zb

G fdq qdq d v q q

A T

fdq qdq v q q

T

, (layered 3D cross-plane) (5.6d)

These equations will be used in the determination of pre-factors for averaged phonon mean free

path (MFP) λ and group velocity v in Sections 5.3.4 and 5.3.5.

5.3.3 Diffusive Thermal Conductivity

In the Boltzmann transport equation (BTE) approach with the relaxation time approxima-

tion, the diffusive thermal conductivity along a transport direction n is generally given by

2

diff tot,

1( )[ ( )] ( )b b b

nb q

fk q v q q

V T

, (5.7)

where ( )b

tot q is phonon relaxation time. After converting the sum of q to an integral, the general

expressions of kdiff for 1D-3D are then given by

2

diff tot

1( )[ ( )] ( )

2

b b b

b BZ

fk dq q v q q

A T

, (1D) (5.8a)

2 2

diff tot2

1( )[ ( )] ( )

(2 )

b b b

nb BZ

fk dq q v q q

H T

, (2D) (5.8b)

3 2

diff tot3

1( )[ ( )] ( )

(2 )

b b b

nb BZ

fk dq q v q q

T

. (3D) (5.8c)

For isotropic 1D-3D, phonon MFP is λb(q)=v

b(q)

b

tot . For anisotropic 3D layered materials, in-

Page 96: © 2014 Zuanyi Li

90

plane and cross-plane phonon MFPs are || ||( , ) ( , )b b b

z z totq q v q q and ( , ) ( , )b b b

z z totq q v q q ,

respectively. Similar to the treatment of ( )b

nv q in Gball/A, kdiff in different cases can be further

expressed as

22

diff tot2 0

1[ ( )cos ]

4

1 ( ) ( )

4

b b b

b

b b b

b

fk qdq d v q

H T

fqdq v q q

H T

, (isotropic 2D) (5.9a)

22 2

diff tot3 0 0

2

2

1sin [ ( )sin cos ]

(2 )

1 ( ) ( )

6

b b b

b

b b b

b

fk q dq d d v q

T

fq dq v q q

T

, (isotropic 3D) (5.9b)

22

diff, || || tot3 0

|| ||2

1[ ( , )cos ]

(2 )

1 ( , ) ( , )

8

b b b

z zb

b b b

z z zb

fk dq qdq d v q q

T

fdq qdq v q q q q

T

, (layered 3D in-plane) (5.9c)

22

diff, tot3 0

2

1[ ( , )]

(2 )

1 ( , ) ( , )

4

b b b

z z zb

b b b

z z zb

fk dq qdq d v q q

T

fdq qdq v q q q q

T

, (layered 3D cross-plane) (5.9c)

These equations will be used in the determination of pre-factors for averaged phonon MFP λ in

Section 5.3.4.

5.3.4 Averaged Phonon Mean Free Path

If we look at the ratio of kdiff to Gball/A, based on Eqs. 5.5a, 5.6, 5.8a, and 5.9, we can find

that besides a pre-factor the ratio is just the defined average of phonon MFPs with all modes

weighted properly, which is shown below for all cases:

diff

ball

| ( ) |

2 2 ( ) 2/

| ( ) |

( )b b

bb

b b

b

bs

b

fdq v q

k T qfG A

dq v qT

q

, (1D) (5.10a)

Page 97: © 2014 Zuanyi Li

91

diff

ball

( )( )

( )/ 2 2 2

( )

b b

bb

bs

b

b b

b

fqdq v q

k T qfG A

qdq v qT

q

, (isotropic 2D) (5.10b)

2

diff

2ball

( ()4 4 4

( )/ 3 3 3

)

( )

b b

bb

b

b

b

bsb

qf

q dq v qk T q

fG Aq dq v q

T

, (isotropic 3D) (5.10c)

||diff, ||

|| ||

ball, ||||

|| ( , )( , )

( , )/ 2 2 2

( , )

b b

z zbb

z bsb b

z

z

b

b

z

fdq qdq v q qk T q q

fG Adq qdq

q q

v q qT

,

(layered 3D in-plane) (5.10d)

diff,

ball,

( , )

2 2 ( , ) 2/

(

, )

)

(

,

b b

z zbb

z bsb b

b

z

z

zb

fdq qdq v q qk T q q

fG Adq qdq v q q

T

q q

,

(layered 3D cross-plane) (5.10e)

Thus, in general we have

diffbs

ball /

k

G A , (5.11)

which is the rigorous and correct average of phonon MFPs of all modes. The pre-factor βλ is 1/2,

2/π, and 3/4 for isotropic 1D, 2D, and 3D materials, consistent with Ref. [175]. Interestingly, for

layered 3D materials the in-plane pre-factor is 2/π, same as isotropic 2D, and the cross-plane pre-

factor is 1/2, same as 1D.

5.3.5 Averaged Phonon Group Velocity

When we look at the ratio of Gball/A to Cv, based on Eqs. 5.2a, 5.3, 5.5a, and 5.6, we can find

that besides a pre-factor the ratio is just the defined average of phonon group velocity with all

modes weighted properly, which is shown below for all cases:

Page 98: © 2014 Zuanyi Li

92

ball

V

/ 1 1 1( )

2

( )

2 2

b

bb

b

b

b fdq

G A T v q vf

v

Cd

q

qT

, (1D) (5.12a)

ball

V

/ 1 1 1(

( )

)

b

bb

b

b

b fqdq

G A T v q vfC

q

v q

dqT

, (isotropic 2D) (5.12b)

2

ball

2V

/ 1 1 1( )

4 4 4

( )b

bb

b

b

b fq dq

G A T v q vfC

dq

v

qT

q

, (isotropic 3D) (5.12c)

ball, ||

|| ||

V

||/ 1 1)

1( ,

( , )

bb

zbb

zb

z

z

b

fdq qdqG A T v q q

v q

vfC

dq q

q

dqT

,

(layered 3D in-plane) (5.12d)

ball,

V

/ 1 1 1( , )

2 2 2

( , )b

zbb

zb

zb

b

z

fdq qdqG A T v q q v

f

v q q

Cdq qdq

T

,

(layered 3D cross-plane) (5.12e)

Thus, in general we have

ball

V

/v

G Av

C , (5.13)

which is the rigorous and correct average of phonon group velocity of all modes. The pre-factor

βv is 2, π, and 4 for isotropic 1D, 2D, and 3D materials. Particularly, for layered 3D materials the

in-plane and cross-plane pre-factors are π and 2, respectively, similar to the choice of βλ. We note

that to use Eq. 5.13 the units of Cv should be JK-1

m-3

.

After having averaged phonon MFP λ and group velocity v, we can examine if they can lead

Page 99: © 2014 Zuanyi Li

93

to a consistent result with the kinetic theory. From Eqs. 5.11 and 5.13, we have

ball diffdiff V V diff

V ball

/1 1

/

vv

G A kk C v C k

d d C G A d

, (5.14)

where d=1−3 is the material dimension. Based on our derived pre-factors βv and βλ, we find that

βvβλ /d=1 holds for isotropic 1D, 2D, and 3D, meaning the reproduced kinetic theory. Important-

ly, for anisotropic layered 3D materials d has to be 2 and 1 for in-plane and cross-plane, respec-

tively, to hold the kinetic theory. This is a very important conclusion which has not been pointed

out before.

5.4 Results and Discussion

5.4.1 Ballistic Thermal Conductance

Based on the obtained phonon dispersions, the monolayer and bulk Gball/A of layered mate-

rials are calculated by Eqs. 5.5b and 5.5c, respectively. The results of Gball/A [including mono-

layer, bulk in-plane (||) and cross-plane ()] as a function of T for graphene/graphite, h-BN,

MoS2, and WS2 are shown in Fig. 5.6. For each material, its monolayer Gball/A is higher than that

of bulk (||) at low T, but they will become almost overlapped at high T, because the in-plane

phonon dispersions of monolayer and bulk are nearly degenerate except at low frequencies. For

bulk (), its Gball/A is similar to that of bulk (||) at low T, but the former is about ten times lower

than the latter at high T, reflecting the significant anisotropy in heat conduction. For better

comparison among four kinds of materials, Gball/A of the most widely-used semiconductor

material, silicon [174], is also plotted (dashed line) in each panel of Fig. 5.6 as a reference.

Above ~100 K, graphene, graphite (||), h-BN monolayer and bulk (||) show much larger Gball/A

than Si (Figs. 5.6a-b), due to their very strong in-plane covalent bonds and more dispersive

phonon dispersions, whereas, the monolayer and bulk (||) of MoS2 and WS2 show smaller Gball/A

Page 100: © 2014 Zuanyi Li

94

than Si at T > 100 K (Figs. 5.6c-d) due to relatively weak metal-sulfur bonds. All cross-plane

Gball/A of four materials are lower than that of Si because of their weak van der Waals interac-

tions between layers. The room-temperature Gball/A values and Debye temperature ΘD [219, 226,

227] of these materials are listed in Table 5.2.

In the high T limit, all Gball/A vs. T curves flatten out (saturate) when all phonon modes are

fully excited. The temperatures at which Gball/A reaches 90% of its saturated (maximum) value

are nearly the same for the monolayer and bulk (||) of each material. They are ~1150 K, ~1060 K,

~320 K and ~310 K for graphite, h-BN, MoS2 and WS2, respectively, which are roughly 60% of

Fig. 5.6: Calculated ballistic thermal conductance per unit cross-sectional area as a function of

temperature for graphene/graphite (a), h-BN (b), MoS2 (c), and WS2 (d). Each panel includes

Gball/A of their monolayer, bulk (||), and bulk (), as well as that of silicon [174] for comparison.

1 10 100 100010

3

104

105

106

107

108

109

1010

Gball/

A (

WK

-1m

-2)

T (K)

1 10 100 100010

3

104

105

106

107

108

109

Gball/

A (

WK

-1m

-2)

T (K)1 10 100 1000

103

104

105

106

107

108

109

Gball/

A (

WK

-1m

-2)

T (K)

1 10 100 100010

3

104

105

106

107

108

109

1010

Gball/

A (

WK

-1m

-2)

T (K)

graphene/graphite h-BN

MoS2 WS2

~T 3

~T 2.5

~T 2

~T 1.5

graphite (||)graphite ()

bulk (||)bulk ()

bulk ()bulk (||)

bulk ()

Si Si

Si Si

a b

c d

bulk (||)

~T 3

~T 2.5

~T 2

~T 1.5

~T 3

~T 2.5

~T 2

~T 1.5

~T 3

~T 2.5

~T 2

~T 1.5

Page 101: © 2014 Zuanyi Li

95

their Debye temperature ΘD (Table 5.2). The corresponding temperatures for bulk () are ~780

K, ~1270 K, ~345 K and ~360 K for graphite, h-BN, MoS2 and WS2, respectively. Interestingly,

there is a “bump” in Gball/A of h-BN bulk () before it flattens out, different from other bulk ()

curves. The reason is that the high frequency optical branches (ZO1, ZO2, LO1, LO2) of bulk h-

BN have noticeable non-flatten dispersions along the cross-plane direction (see Fig. 5.4f),

leading to non-zero phonon velocity vz, unlike the nearly zero vz in other bulk materials. Thus,

when temperature increases and these phonon modes start to contribute to cross-plane conduc-

tion, a significant increase (bump) appears in the Gball/A vs. T curve.

In the low T limit, all Gball/A vs. T curves show power law scaling, ~T n

(Fig. 5.6). For the

monolayers, the power exponent is n ≈ 1.59−1.68, which is a combined effect of n = 1.5 from the

quadratic ZA branch and n = 2 from the linear TA and LA branches [29, 55]. The power law

only applies to monolayers below ~90 K for graphene and h-BN, and below ~40 K for MoS2 and

WS2. For bulk (||), the power exponent is n ≈ 2.91−2.96, which is a combined effect of n = 2.5

Table 5.2: Our calculated room-temperature Gball/A of interested layered materials, as well as

their Debye temperature ΘD from literature [219, 226, 227]. Since Debye temperature ΘD is

actually a function of temperature T [219, 226], here we mainly list their values of high T limit

for our discussion, but values of low T limit for MoS2 and WS2 (in parentheses) are listed due to

lack of high T limit value for WS2.

Debye

temperature

ΘD (K)

Gball/A (GWK-1

m-2

)

monolayer bulk (||) bulk ()

graphite [226] 1930 4.37 4.34 0.294

h-BN [226] 1740 3.85 3.65 0.291

MoS2 [219, 227] 570 (253) 0.88 0.83 0.135

WS2 [227] (210) 0.72 0.67 0.114

Page 102: © 2014 Zuanyi Li

96

from the ZA branch (if purely quadratic) [29] and n = 3 from linear TA and LA branches (see

Supporting Information Section 7). The reason that linear TA and LA branches lead to n = 3 for

bulk layered materials can be understood by rewriting Eq. 5.5c as

ball

3

2 2

3 2

1( ) | ( ) |

16

( ) | ( ) |16 ( 1)

b b b

nb

xb bB

n xb

G fd D v

A T

k T x edxD v

e

, (5.15)

where x=ħωb/kBT is a dimensionless number. For TA and LA branches at low-frequency, ( )b

nv

is constant, and Db(ω) ω

2 [228], which will give additional ~T

2 when the integral is converted

to a dimensionless one. Thus, TA and LA branches will have a ~ T 3

scaling in Gball/A at low T.

The reason that the overall n is closer to 3 than 2.5 is that the real ZA branch for bulk has a

linear component (not purely quadratic) at very low frequency [171], resulting in n larger than

2.5. The power law for bulk (||) is valid only below ~35 K for graphite and h-BN, and below ~20

K for MoS2 and WS2. For bulk (), the power exponent is n ≈ 2.73−2.89 with the same physical

reason as bulk (||), and it applies only below ~10 K for graphite and h-BN, and below ~7 K for

MoS2 and WS2.

Calculated ballistic thermal conductance is a useful quantity in the investigation of heat con-

duction in materials. It reflects the intrinsic anisotropic conduction and gives the upper limit of

heat flow at a certain temperature. Any measured and calculated thermal conductivity k can be

formally converted to G/A=k/L, and compared to Gball/A, from which we can know if obtained

thermal conductivity satisfies the law of quantum mechanics (i.e., below the ballistic limit) [22]

and what percentage of ballistic conduction it reaches [55, 57]. More importantly, it can be used

to predict thermal conductivity as a function of transport length L and also to estimate the

average phonon MFP λ.

Page 103: © 2014 Zuanyi Li

97

5.4.2 Length-Dependent Thermal Conductivity

We first discuss the L-dependent thermal conductivity. In the ballistic regime (L ≪ λ), since

the conductance rather than the conductivity approaches a constant, the ballistic thermal conduc-

tivity kball = (Gball/A)L becomes linearly dependent on length L. In the diffusive regime (L ≫ λ),

the conductivity is generally independent of length L (the case of k divergent with L will be

discussed later), becoming a constant kdiff (diffusive thermal conductivity). Therefore, in the

intermediate ballistic-diffusive regime (i.e., quasi-ballistic regime), the thermal conductivity

should increase with L and gradually converge to kdiff, similar to the mobility change during

quasi-ballistic charge transport observed in short-channel transistors [172, 173]. This transition

can be captured through a phenomenological model, called ballistic-diffusive (BD) or quasi-

ballistic (QB) model [55, 171]:

1 1

ball diff ball diff

1 1 1 1

( / ) ( / )b bb

k LG A L k G A L k

, (5.16)

where the first expression is a “1-color” model, and the second one is a branch-resolved “multi-

color” model taking into account different phonon branch contributions separately. Apparently,

such L-dependent k expressions can correctly reproduce both ballistic and diffusive regimes

stated above. As we will show later, this model is essentially a Landauer-like model and it can

yield an expression of diffusive thermal conductivity consistent with the kinetic theory.

We first apply this model to the thermal conductivity of graphene supported on SiO2. Figure

5.7a shows its k (T=300 K) as a function of L from both experimental measurements [49, 55,

229] (solid symbols) and Boltzmann transport equation (BTE) calculations (open symbols,

calculations follow the method described in Ref. [49], including the use of an empirical potential

to describe the atomic interactions). By using our calculated Gball/A = Σb ball

bG /A = 4.37 GWK-1

m-2

Page 104: © 2014 Zuanyi Li

98

for graphene as well as kdiff = Σb diff

bk = 578 Wm-1

K-1

from BTE simulations, the 1-color model

(solid blue line) and 6-color model (solid orange line) as well as its components of each branch

(dash-dot lines) are shown in Fig. 5.7a. They are in good agreement with experimental data and

BTE simulations, indicating that the BD model can provide reasonable L-dependent k in the

intermediate regime. We find that the simple 1-color model is sufficiently good to give almost

the same result as the 6-color model, although the latter provides more information about the

contribution of each branch. For supported graphene, besides three acoustic branches, the

flexural optical (ZO) branch also has a notable contribution, but other optical (TO and LO)

branches have negligible contributions (<1%). This is because the ZO branch has relatively low

phonon frequency, and hence larger thermal population.

We have demonstrated the BD model can successfully describe the k change with L in sup-

ported graphene, next we focus on the L-dependent k in suspended graphene. Since graphene is a

2D material, and some theoretical studies predicted in isolated low-dimensional (1D and 2D)

momentum-conserving systems k will diverge with L, i.e., ~Lα for 1D [200, 201] and ~logL for

2D [201-203]. However, other works [85, 230, 231] argued that disorder and higher-order three-

phonon scattering may eliminate the divergence, and no experiments have confidently observed

divergent k yet. For supported graphene discussed above, the ~logL divergence does not appear

due to phonon scattering with substrate vibrational modes [97, 98], but for suspended graphene it

is still an open question.

Very recently, Xu et al. [57] systematically measured k of suspended graphene as a function

of L, and their data at T=300 K (green squares) are shown in Fig. 5.7b. We find that our 1-color

BD model (Eq. 5.16) can fit their data well with the fitting parameter kdiff = 1790 Wm-1

K-1

(see

solid line in Fig. 5.7b). If assuming the divergence of k ~ logL for large L, the BD model can be

Page 105: © 2014 Zuanyi Li

99

changed by adding a logarithmic term to capture this effect:

1

ball diff 0 0

1 1

( / ) ln( / 1)k L

G A L k k L L

, (5.17)

Fig. 5.7: Thermal conductivity k as a function of length L at room temperature. (a) L-dependent k

for graphene supported on SiO2. The 1-color (solid blue line) and 6-color (solid orange line) BD

models from Eq. 5.16 show good agreement with BTE calculations (open circles) and experi-

mental data of Bae et al. [55] (solid triangle) and Seol et al. [49] (solid diamonds). The dash-dot

lines are components in the 6-color model. (b) L-dependent k for suspended graphene. The 1-

color BD model (blue solid line) from Eq. 5.16, log model A (grey dash-dot line) from Eq. 5.17,

and log model B (brown dot line) are used to fit experimental data of Xu et al. [57]. (c) L-

dependent k for SWCNTs. Experimental data of Yu et al. [72], Pop et al. [83], Wang et al. [232],

and Li et al. [233] are plotted. Dash lines represent the 1-color BD model fitted to individual data

points, and the yellow band shows the overall range covering plotted data. In (a-c), the short-

dash line represents the ballistic upper limit kball = (Gball/A)L. (d) L-dependent k for different

materials obtained from the 1-color BD model using calculated Gball/A and experimental kdiff in

literature (see text).

101

102

103

104

0

500

1000

1500

2000

2500

k (

Wm

-1K

-1)

L (nm)

10-1

100

101

102

103

104

10-2

10-1

100

101

102

103

104

k (

Wm

-1K

-1)

L (nm)

101

102

103

104

0

100

200

300

400

500

600

700

k (

Wm

-1K

-1)

L (nm)

Yu (2005)

Li (2009)

Pop (2006)

SWCNT

a b

c d

graphite (||)

graphene (suspended)

MoS2 (1L)

h-BN (||)

graphite ()

h-BN ()

Si

Wang (2007)

kball

graphene(on SiO2)

Seol (2010)

Bae (2013)

BTE

101

102

103

104

105

102

103

104

k (

Wm

-1K

-1)

L (nm)

TALAZA

ZO

TO+LO

Xu (2014)

kball

graphene(suspended)

Page 106: © 2014 Zuanyi Li

100

where k0 and L0 are parameters with units of thermal conductivity and length, respectively. We

call this model as the “log model A” and it can reproduce the ballistic limit correctly as well, i.e.,

k → (Gball/A)L when L→0. By choosing kdiff = 1590 Wm-1

K-1

, k0 = 130 Wm-1

K-1

, and L0 = 1.1

µm, the log model A can also fit the data of Xu et al. [57] well (see dash-dot line in Fig. 5.7b).

Here we also propose another log model B:

ball ln 1B

B

G Lk L L

A L

, (5.18)

where LB is a length parameter. Similar to the log model A, the log model B can also reproduce

the correct ballistic behavior at short L, that is, k → (Gball/A)L when L→0. The best fit of this

model to experimental data of Xu et al. [57] by using LB = 100 nm is shown as the dot line in

Fig. 5.7b. The fit of the log model B is not as good as that of the log model A, but both are worth

to be employed and tested. It is clear that the convergent BD model and divergent log model A

are distinguishable only at L > 10 µm, and below 10 µm the increase of k with L mainly results

from the ballistic-to-diffusive transition (or quasi-ballistic effect). However, currently available

data of k are only up to 9 µm, so further measurements beyond 10 µm are required to eventually

show whether thermal conductivity is divergent in suspended graphene. For 1D it is predicted

that k diverges with L as ~Lα, so similar to Eq. 5.17 we propose divergent k(L) with a power law

term for 1D:

1

ball diff

1 1

( / )k L

G A L k L

, (5.19)

where kdiff, γ, and α can be fitting parameters to fit experimental data or theoretical calculations.

In Fig. 5.7c we also plot experimental thermal conductivity data [72, 83, 232, 233] for sus-

pended single-wall carbon nanotubes (SWCNTs) with similar diameters as a function of length

at room temperature. There is no systematic measurement of k versus L for SWCNTs, and the

Page 107: © 2014 Zuanyi Li

101

few available data do not fall in a single trend, so here we only use the 1-color BD model (Eq.

5.16) to fit to individual data points (dash-dot lines) and give a “band” (yellow area) to show the

k range as L increases (Fig. 5.7c). As shown by Mingo and Broido [29], SWCNTs have the same

Gball/A as graphene above ~200 K, so the used room temperature Gball/A for SWCNTs is 4.37

GWK-1

m-2

(same as graphene) in the BD model. The range of obtained kdiff is ~2500−8500 Wm-

1K

-1, and thermal conductivity keeps increasing with length up to tens of microns. A more

general expression of k including both length and temperature dependence for SWCNTs is

provided in our previous study of short channel (L < 100 nm) CNT devices [193], where heat

conduction is nearly ballistic.

For graphite, h-BN, MoS2, and Si, their L-dependent k at 300 K obtained from the 1-color

BD model are summarized in Fig. 5.7d, including monolayer (1L), bulk in-plane (||) and bulk

cross-plane () k of each material where the corresponding measured kdiff are known. For sus-

pended graphene, most Raman measurements report k in the range of ~2000−4000 Wm-1

K-1

for

L ~ 1−10 µm [23, 24], so here we use kdiff = 4000 Wm-1

K-1

in the BD model to plot its L-

dependent k if assuming convergent k(L). In Fig. 5.7d, other kdiff used in the BD model are all

from experimental measurements in the literature: kdiff = 2000, 6, 400, 2, 35, 2.5, and 150 Wm-

1K

-1 for graphite (||) [82], graphite () [82], h-BN (||) [234], h-BN () [235], MoS2 (1L) [236],

MoS2 () [237], and Si [168], respectively. These layered materials show a strong anisotropy in

thermal conductivity, and their values span a wide range, more than three orders of magnitude.

The estimated k-L dependence given here will be helpful for understanding heat conduction at

scales where size effects take place.

5.4.3 Estimation of Average Phonon Mean Free Path

Next we turn to the estimation of phonon MFP λ in terms of the calculated Gball/A. We note

Page 108: © 2014 Zuanyi Li

102

that the underlying physics of the BD model is Landauer transport theory, in which the conduct-

ance (or conductivity) is given by the ballistic conductance (or conductivity) multiplied by a

transmission coefficient [24, 174, 238]:

1

ball bs

bs ball diff

1 1

( / )

Gk L L

A L G A L k

, (5.20)

where the transmission coefficient is given in terms of the transport length L and averaged

phonon back-scattering MFP λbs, that is, λbs/(L+λbs). A simple rearrangement of the first expres-

sion will reproduce the BD model, meanwhile yielding a relation λbs = kdiff/(Gball/A). This relation

can be directly derived from the definition of averaged phonon MFP with all modes weighted

properly (see Section 5.3.4). The common phonon MFP λ is shorter than the back-scattering

MFP λbs, and they are related with a factor βλ, as shown by Eq. 5.11.

As long as we know reliable kdiff from calculations or experimental measurements, the over-

all phonon MFP can be estimated based on Eq. 5.11, because Gball/A can be calculated without

approximations. Take graphite as an example, using its widely-accepted values of measured kdiff

[23, 82] and our calculated Gball/A, the obtained phonon in-plane and cross-plane MFPs (λ|| and

λ) as a function of temperature are shown as solid lines in Fig. 5.8a. They decrease rapidly as T

increases and differ by two and one orders of magnitude at low and high T, respectively. At room

temperature, λ|| = 290 nm and λ = 10 nm. The latter corresponds to ~30-layers thickness in

graphite, which means the cross-plane heat conduction in multi-layer graphene is already in the

ballistic regime. Thus, the advantage of the single average MFP (Eq. 5.11) is giving information

in a concise way, compared to mode-dependent MFPs, and it helps us understand when to

consider size effects and make corrections to thermal conductivity in a simple way, which is

quite useful especially from a device point of view.

Page 109: © 2014 Zuanyi Li

103

We point out that the traditional way to estimate average phonon MFP is through the kinetic

theory, kdiff = (1/d)Cvvλ., where d=1−3 is the material dimension, Cv is the heat capacity, and v is

the average phonon group velocity. Different from Eq. 5.11, using this method requires not only

kdiff and Cv but also careful calculations for averaged values of phonon group velocity v. Howev-

er, most studies simply used the sound velocity vs (low-frequency group velocity) averaged

among acoustic branches as v. This will underestimate phonon MFP for two reasons. First, the

sound velocity is only valid for low-frequency phonons and hence for low T. For high T, even

room temperature, the effective group velocity is much smaller. Second, for the flexural ZA

mode in layered materials, its low-frequency group velocity approaches zero, and cannot be

simply included in the usual way [239]. Thus, only including TA and LA modes will overesti-

mate v.

Take graphite as an example, through the sound velocity average [239], we have v|| = [2/(

TA

s,||v -2+

LA

s,||v -2)]

1/2 = 15.8 km/s and v = [3/(2

TA

s,v

-2+

LA

s,v

-2)]

1/2 = 2.0 km/s, where

TA

s,||v , LA

s,||v , TA

s,v ,

Fig. 5.8: (a) Phonon MFP λ as a function of temperature for graphite (||) and (). The kinetic

theory with sound velocity gives underestimated λ (dash lines) compared to the correct results

(solid lines) from Eq. 5.11. (b) Room temperature phonon MFP λ for different materials obtained

from Eq. 5.11 using calculated Gball/A and experimental kdiff in literature (see text).

10 100 100010

0

101

102

103

104

Ph

on

on

MF

P

(n

m)

T (K)

graphite ()

graphite (||)

100

101

102

103

104

Ph

on

on

MF

P

(n

m)

gra

ph

ite

(||)

gra

ph

en

e

(su

sp

en

de

d)

Mo

S2

(1L

)

h-B

N (

||)

gra

ph

ite

(

)

h-B

N (

)

Si

Mo

S2

()

gra

ph

en

e

(on S

iO2)

T = 300 Ka b

Page 110: © 2014 Zuanyi Li

104

and LA

s,v are from our calculated phonon dispersion. Using these values with Cv from our calcula-

tions (see Section 5.3.1) and kdiff given previously, the obtained λ|| and λ as a function of tem-

perature from the kinetic theory are shown as dash lines in Fig. 5.8a. We note that since graphite

is anisotropic, the equations for in-plane and cross-plane should be taken as 2D (d=2) and 1D

(d=1) forms of the kinetic theory, respectively, similar to the choice of βλ (shown in Section

5.3.5). As expected, the obtained λ|| (=145 nm at 300 K) from the kinetic theory is lower than that

from our method (Eq. 5.11) over the whole T range, mainly due to neglecting the small group

velocity of ZA mode in the average. The λ (=1.7 nm at 300 K) from the kinetic theory is also

underestimated, except for low T (<30 K), the range in which the sound velocity is valid. The

consistence between two methods for λ at low T also indicates that the 1D form of the kinetic

theory for cross-plane is correct.

To obtain correct average phonon MFP λ by the kinetic theory, the group velocity v should

be an average weighted by the heat capacity, as shown by Eq. 5.13 in Section 5.3.5. Substituting

Eq. 5.13 into the kinetic theory, the phonon MFP is then given by

diff diff

V ball /v

k d kd

C v G A

. (5.21)

We can find d/βv ≡ βλ for all cases, so when the correct v is used, the kinetic theory can give the

same estimation of MFP as our model (Eq. 5.11). Figure 5.9a shows the calculated v in terms of

Eq. 5.13 as a function of T for graphene, graphite (||), and graphite (). In the whole temperature

range, the maximum of v for graphite (||) and graphene is ~8 km/s, about two times smaller than

the averaged sound velocity shown above (15.8 km/s). For graphite (), the averaged v at 300 K

is only ~345 m/s, almost six times smaller than 2 km/s from the sound velocity average, and it

reaches 2 km/s only below ~30 K, i.e., the T range where the sound velocity is valid. The calcu-

Page 111: © 2014 Zuanyi Li

105

lated v [including monolayer, bulk (||), and bulk ()] based on Eq. 5.13 for h-BN, MoS2, and

WS2 are also shown in Fig. 5.9.

After demonstrating the BD model and kinetic theory are consistent in the MFP estimation,

we apply Eq. 5.11 to other interesting materials whose measured kdiff are available. As shown in

Fig. 5.8b, the estimated λ at 300 K are ~3, 9, 10, 25, 70, 90, 100, 290, 580 nm for h-BN (),

MoS2 (), graphite (), MoS2 (1L), h-BN (||), graphene (on SiO2), Si, graphite (||), and graphene

(suspended), respectively. The used kdiff are listed previously. We note that Eq. 5.11 gives the

“averaged” MFP including contributions from all phonon modes, but some modes (e.g., optical)

have very small contributions to thermal conductivity (heat conduction) and very short MFPs

Fig. 5.9: Calculated phonon group velocity as a function of temperature for graphite (a), h-BN

(b), MoS2 (c), and WS2 (d) from Eq. 5.13. Each material includes its monolayer, bulk (||), and

bulk ().

1 10 100 1000100

1000

10000

v (

m/s

)

T (K)1 10 100 1000

100

1000

10000

v (

m/s

)

T (K)

graphene/graphite h-BN

MoS2 WS2

graphite (||)

graphite ()

bulk (||)

bulk (||)monolayer

bulk ()

bulk (||)monolayer

a b

c d

bulk ()

bulk ()

1 10 100 1000100

1000

10000

v (

m/s

)

T (K)1 10 100 1000

100

1000

10000

v (

m/s

)

T (K)

Page 112: © 2014 Zuanyi Li

106

(<10 nm for in-plane and <1 nm for cross-plane). This means that those modes contributing to

heat conduction significantly should have MFPs longer than the value given by Eq. 5.11. Indeed,

by analyzing accumulative thermal conductivity as a function of MFP, it is found that phonons

with MFP longer than 1 µm and 100 nm contribute ~40% to the total thermal conductivity for Si

[206-209] and cross-plane graphite [240] at 300 K, respectively. Thus, a “median” MFP of 0.5-1

µm and ~100 nm (much larger than our “averaged” numbers, 100 nm and 10 nm) is suggested to

explain the strong size effects of thermal transport in Si membrane [204] and graphite (cross-

plane) [240], respectively. However, we emphasize that the value given by Eq. 5.11 is a rigorous

average with all phonon modes weighted properly (see Section 5.3.4) to represent the overall

MFP, which should be used in the kinetic theory and to understand quasi-ballistic thermal

transport. By comparing our estimated MFPs in Fig. 5.8b and k(L) in Fig. 5.7d, we can find that

when L < λ, heat conduction enters the ballistic regime (i.e., the linear region in Fig. 5.7d) as

expected. Whereas, when L is shorter than the suggested “median” MFP (~10λ), k just starts to

decrease rapidly, but is not in the ballistic regime, meaning the suggested “median” MFP is more

like a characteristic length below which the size effects (not ballistic transport) take place.

Interestingly, beyond 10λ thermal conductivity still increases slowly and will eventually reach

kdiff (i.e., become fully diffusive) at L ~ 100λ (Fig. 5.7d), which is much longer than the com-

monly assumed length. The whole ballistic-to-diffusive transition takes more than two orders of

magnitude in length (λ−100λ) to complete. This also indicates that any obtained k increase in this

range might arise from the intrinsic ballistic-to-diffusive transition, and whether k is divergent in

low-dimensional materials should be examined beyond 100λ to draw a realistic conclusion.

5.5 Summary

In conclusion, we calculated the ballistic thermal conductances Gball of graphene/graphite, h-

Page 113: © 2014 Zuanyi Li

107

BN, MoS2, and WS2 based on full phonon dispersions obtained from ab initio simulations. From

the calculated Gball, we obtained L-dependent k by using a phenomenological ballistic-diffusive

model. In particular, for suspended graphene recent measurements [57] of k vs. L can be fitted by

both the convergent BD model and divergent log model, and the two models only differ beyond

L~10 µm. We also showed that the rigorously averaged phonon MFP λ is simply determined by

Gball and kdiff, and that the kinetic theory can reach the same result as long as a proper phonon

group velocity is used. We emphasize that for anisotropic layered materials, 2D and 1D forms of

the kinetic theory should be used for in-plane and cross-plane, respectively. Based on the calcu-

lated λ and the BD model, we find that k will not fully converge to kdiff until after L~100λ, much

longer than the commonly assumed length. Thus, to verify predictions of divergent k in low-

dimensional systems, simulations and measurements should be performed beyond L~100λ.

Page 114: © 2014 Zuanyi Li

108

CHAPTER 6

CONCLUSIONS AND OUTLOOK

6.1 Conclusions

In this dissertation, we investigated thermal transport in graphene and several other 2D lay-

ered materials, with a focus on nanoscale ballistic conduction and size effects. We combined

experimental measurements, finite element simulations, and theoretical calculations to give a

comprehensive study.

First, we measured heat conduction in nanoscale graphene by a substrate-supported ther-

mometry platform. Short, quarter-micron graphene samples reach ~35% of the ballistic thermal

conductance limit up to room temperature. In contrast, patterning similar samples into nanorib-

bons (GNRs) leads to a diffusive heat flow regime that is controlled by ribbon width and edge

disorder. In the edge-controlled regime, the GNR thermal conductivity scales with width approx-

imately as ~W1.8±0.3

, being about 100 Wm-1

K-1

in 65-nm-wide GNRs, at room temperature. Thus,

the usual meaning of thermal conductivity must be carefully interpreted when it becomes a

function of sample dimensions. These findings are highly relevant for all nanoscale graphene

devices and interconnects, also suggesting new avenues to manipulate thermal transport in 2D

and quasi-one-dimensional systems.

We also examined the possibility of using this supported platform to measure other materials

through finite element simulations. The platform geometry is optimized based on uncertainty

analysis. The smallest thermal sheet conductance that can be sensed by this method within a 50%

error is ~25 nWK-1

at room temperature, indicating this platform can be applied to most thin

films like polymer and nanotube networks, as well as nanomaterials like nanotube/nanowire

arrays, even a single Si nanowire. Moreover, the platform can also be extended to plastic sub-

Page 115: © 2014 Zuanyi Li

109

strates (not limited to the SiO2/Si substrate), and could find wide applicability in circumstances

where fabrication challenges and low yield associated with suspended platforms must be avoid-

ed.

Last, we calculated in-plane (for monolayer and bulk) and cross-plane (for bulk) ballistic

thermal conductances Gball of graphene/graphite, h-BN, MoS2, and WS2, based on full phonon

dispersions from first-principles approach. We find that the anisotropy of thermal transport in

these materials comes from both anisotropic Gball and phonon MFPs, not just the effect of one of

them. A proper estimation of the overall MFP λ should be a rigorous average of all phonon

modes, which can be condensed to an expression just including Gball and diffusive thermal

conductivity kdiff. We point out that the kinetic theory can reach the same result as long as the

correctly averaged phonon group velocity is used. We emphasize that for anisotropic layered

materials, 2D and 1D forms of the kinetic theory should be used for in-plane and cross-plane,

respectively. Based on the calculated λ and the BD model, we find that k will not fully converge

to kdiff until after L~100λ, much longer than the commonly assumed length. Thus, to verify

predictions of divergent k in low-dimensional systems, simulations and measurements should be

performed beyond L~100λ to distinguish from the intrinsic increase of k due to the ballistic-to-

diffusive transition.

These results represented a comprehensive study of thermal conduction in 2D layered mate-

rials in micro and nanoscale where ballistic conduction and size effects take place. We extended

experimental methods to measure nanoscale thermal conduction and showed that thermal con-

ductivity should be redefined in nanoscale and could be tuned effectively. This dissertation

broadens our understanding of thermal conduction and how to manipulate it to reach the re-

quirements for potential applications like thermal management and thermoelectric conversion.

Page 116: © 2014 Zuanyi Li

110

6.2 Future Work

There has been an increasing interest of thermoelectric properties in low-dimensional

nanostructures. Thermoelectric transport of graphene has been experimentally studied, showing a

maximum Seebeck coefficient of S ~ 100 μV/K at room temperature (Section 2.4). However, the

thermoelectric properties of GNRs are still unknown in experiments. As pointed out in Section

3.4.4, in the intermediate width range of GNRs (40 nm < W < 200 nm), compared with graphene,

GNR thermal conductivity k is suppressed significantly while electrical conductivity σ is not

affected. Thus, GNRs may have higher ZT than graphene. The Seebeck coefficient of GNRs is

suggested to be measured to investigate the effects of carrier density and edge scattering on their

thermoelectric properties, which will be useful for potential applications in thermoelectric

devices and energy conversion.

Besides graphene, other 2D materials like h-BN and TMDs also show outstanding electrical

properties and promising applications in electronics and optoelectronics. They attracted a lot of

studies about its electronic transport properties, but few focused on their thermal properties.

Although our current study has extended to the thermal transport of these 2D materials in a

theoretical frame (Chapter 5), there are very few experimental works available in the community.

Therefore, it is crucial to measure thermal conduction of these materials in the future, which will

help to better understand the performance of devices made of them. In addition, TMDs are

predicted to have higher thermopower than graphene due to the presence of band gaps, so their

thermoelectric properties are worth to investigate experimentally.

As mentioned in Chapter 5, whether thermal conductivity diverges with transport length in

low-dimensional materials is still an open question. To clarify this issue, experimental measure-

ments should be carried out on 1D and 2D materials with length beyond ~100 times of their

Page 117: © 2014 Zuanyi Li

111

phonon MFP. Furthermore, studying the thermal rectification effect of nanostructures composed

of 2D materials is important and helpful to develop thermal diodes and phononics in the future.

Page 118: © 2014 Zuanyi Li

112

REFERENCES

[1] E. Pop, "Energy Dissipation and Transport in Nanoscale Devices," Nano Res., vol. 3, pp.

147-169 (2010).

[2] M. S. Dresselhaus, G. Chen, M. Y. Tang, R. G. Yang, H. Lee, D. Z. Wang, Z. F. Ren, J. P.

Fleurial, and P. Gogna, "New directions for low-dimensional thermoelectric materials,"

Adv. Mater., vol. 19, pp. 1043-1053 (2007).

[3] G. J. Snyder and E. S. Toberer, "Complex thermoelectric materials," Nat. Mater., vol. 7,

pp. 105-114 (2008).

[4] D. G. Cahill, W. K. Ford, K. E. Goodson, G. D. Mahan, A. Majumdar, H. J. Maris, R.

Merlin, and S. R. Phillpot, "Nanoscale thermal transport," J. Appl. Phys., vol. 93, pp. 793-

818 (2003).

[5] D. G. Cahill, P. V. Braun, G. Chen, D. R. Clarke, S. Fan, K. E. Goodson, P. Keblinski, W.

P. King, G. D. Mahan, A. Majumdar, H. J. Maris, S. R. Phillpot, E. Pop, and L. Shi,

"Nanoscale thermal transport. II. 2003–2012," Appl. Phys. Rev., vol. 1, p. 011305 (2014).

[6] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V.

Grigorieva, and A. A. Firsov, "Electric Field Effect in Atomically Thin Carbon Films,"

Science, vol. 306, pp. 666-669 (2004).

[7] A. K. Geim and K. S. Novoselov, "The rise of graphene," Nat. Mater., vol. 6, pp. 183-191

(2007).

[8] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim, "The

electronic properties of graphene," Rev. Mod. Phys., vol. 81, pp. 109-162 (2009).

[9] F. Bonaccorso, Z. Sun, T. Hasan, and A. C. Ferrari, "Graphene photonics and

optoelectronics," Nat. Photonics, vol. 4, pp. 611-622 (2010).

[10] K. S. Novoselov, V. I. Fal'ko, L. Colombo, P. R. Gellert, M. G. Schwab, and K. Kim, "A

roadmap for graphene," Nature, vol. 490, pp. 192-200 (2012).

[11] Q. L. Bao and K. P. Loh, "Graphene Photonics, Plasmonics, and Broadband

Optoelectronic Devices," ACS Nano, vol. 6, pp. 3677-3694 (2012).

[12] A. N. Grigorenko, M. Polini, and K. S. Novoselov, "Graphene plasmonics," Nat.

Photonics, vol. 6, pp. 749-758 (2012).

[13] K. Kim, J. Y. Choi, T. Kim, S. H. Cho, and H. J. Chung, "A role for graphene in silicon-

based semiconductor devices," Nature, vol. 479, pp. 338-344 (2011).

[14] O. V. Yazyev and Y. P. Chen, "Polycrystalline graphene and other two-dimensional

materials," Nat. Nanotechnol., vol. 9, pp. 755-767 (2014).

[15] F. Schwierz, "Graphene transistors," Nat. Nanotechnol., vol. 5, pp. 487-496 (2010).

[16] C. R. Dean, A. F. Young, I. Meric, C. Lee, L. Wang, S. Sorgenfrei, K. Watanabe, T.

Taniguchi, P. Kim, K. L. Shepard, and J. Hone, "Boron nitride substrates for high-quality

graphene electronics," Nat. Nanotechnol., vol. 5, pp. 722-726 (2010).

[17] B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, "Single-layer MoS2

Page 119: © 2014 Zuanyi Li

113

transistors," Nat. Nanotechnol., vol. 6, pp. 147-150 (2011).

[18] Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, "Electronics

and optoelectronics of two-dimensional transition metal dichalcogenides," Nat.

Nanotechnol., vol. 7, pp. 699-712 (2012).

[19] M. Chhowalla, H. S. Shin, G. Eda, L. J. Li, K. P. Loh, and H. Zhang, "The chemistry of

two-dimensional layered transition metal dichalcogenide nanosheets," Nat. Chem., vol. 5,

pp. 263-275 (2013).

[20] A. K. Geim and I. V. Grigorieva, "Van der Waals heterostructures," Nature, vol. 499, pp.

419-425 (2013).

[21] A. A. Balandin, "Thermal properties of graphene and nanostructured carbon materials,"

Nat. Mater., vol. 10, pp. 569-581 (2011).

[22] M. M. Sadeghi, M. T. Pettes, and L. Shi, "Thermal transport in graphene," Solid State

Commun., vol. 152, pp. 1321-1330 (2012).

[23] E. Pop, V. Varshney, and A. K. Roy, "Thermal properties of graphene: Fundamentals and

applications," MRS Bull., vol. 37, pp. 1273-1281 (2012).

[24] Y. Xu, Z. Li, and W. Duan, "Thermal and Thermoelectric Properties of Graphene," Small,

vol. 10, pp. 2182-2199 (2014).

[25] H. Yanagisawa, T. Tanaka, Y. Ishida, M. Matsue, E. Rokuta, S. Otani, and C. Oshima,

"Analysis of phonons in graphene sheets by means of HREELS measurement and ab initio

calculation," Surf. Interface Anal., vol. 37, pp. 133-136 (2005).

[26] M. Mohr, J. Maultzsch, E. Dobardzic, S. Reich, I. Milosevic, M. Damnjanovic, A. Bosak,

M. Krisch, and C. Thomsen, "Phonon dispersion of graphite by inelastic x-ray scattering,"

Phys. Rev. B, vol. 76, p. 035439 (2007).

[27] M. C. Schabel and J. L. Martins, "Energetics of Interplanar Binding in Graphite," Phys.

Rev. B, vol. 46, pp. 7185-7188 (1992).

[28] D. L. Nika, E. P. Pokatilov, A. S. Askerov, and A. A. Balandin, "Phonon thermal

conduction in graphene: Role of Umklapp and edge roughness scattering," Phys. Rev. B,

vol. 79, p. 155413 (2009).

[29] N. Mingo and D. A. Broido, "Carbon Nanotube Ballistic Thermal Conductance and Its

Limits," Phys. Rev. Lett., vol. 95, p. 096105 (2005).

[30] D. G. Cahill, K. Goodson, and A. Majumdar, "Thermometry and Thermal Transport in

Micro/Nanoscale Solid-State Devices and Structures," J. Heat Transf., vol. 124, pp. 223-

241 (2002).

[31] C. D. S. Brites, P. P. Lima, N. J. O. Silva, A. Millan, V. S. Amaral, F. Palacio, and L. D.

Carlos, "Thermometry at the nanoscale," Nanoscale, vol. 4, pp. 4799-4829 (2012).

[32] S. Ghosh, I. Calizo, D. Teweldebrhan, E. P. Pokatilov, D. L. Nika, A. A. Balandin, W.

Bao, F. Miao, and C. N. Lau, "Extremely high thermal conductivity of graphene:

Prospects for thermal management applications in nanoelectronic circuits," Appl. Phys.

Lett., vol. 92, p. 151911 (2008).

[33] A. A. Balandin, S. Ghosh, W. Z. Bao, I. Calizo, D. Teweldebrhan, F. Miao, and C. N. Lau,

Page 120: © 2014 Zuanyi Li

114

"Superior thermal conductivity of single-layer graphene," Nano Lett., vol. 8, pp. 902-907

(2008).

[34] S. Ghosh, D. L. Nika, E. P. Pokatilov, and A. A. Balandin, "Heat conduction in graphene:

experimental study and theoretical interpretation," New J. Phys., vol. 11, p. 095012

(2009).

[35] S. Ghosh, W. Bao, D. L. Nika, S. Subrina, E. P. Pokatilov, C. N. Lau, and A. A. Balandin,

"Dimensional crossover of thermal transport in few-layer graphene," Nat. Mater., vol. 9,

pp. 555-558 (2010).

[36] W. W. Cai, A. L. Moore, Y. W. Zhu, X. S. Li, S. S. Chen, L. Shi, and R. S. Ruoff,

"Thermal Transport in Suspended and Supported Monolayer Graphene Grown by

Chemical Vapor Deposition," Nano Lett., vol. 10, pp. 1645-1651 (2010).

[37] J. U. Lee, D. Yoon, H. Kim, S. W. Lee, and H. Cheong, "Thermal conductivity of

suspended pristine graphene measured by Raman spectroscopy," Phys. Rev. B, vol. 83, p.

081419 (2011).

[38] S. S. Chen, A. L. Moore, W. W. Cai, J. W. Suk, J. H. An, C. Mishra, C. Amos, C. W.

Magnuson, J. Y. Kang, L. Shi, and R. S. Ruoff, "Raman Measurements of Thermal

Transport in Suspended Monolayer Graphene of Variable Sizes in Vacuum and Gaseous

Environments," ACS Nano, vol. 5, pp. 321-328 (2011).

[39] S. S. Chen, Q. Z. Wu, C. Mishra, J. Y. Kang, H. J. Zhang, K. J. Cho, W. W. Cai, A. A.

Balandin, and R. S. Ruoff, "Thermal conductivity of isotopically modified graphene," Nat.

Mater., vol. 11, pp. 203-207 (2012).

[40] S. S. Chen, Q. Y. Li, Q. M. Zhang, Y. Qu, H. X. Ji, R. S. Ruoff, and W. W. Cai, "Thermal

conductivity measurements of suspended graphene with and without wrinkles by micro-

Raman mapping," Nanotechnology, vol. 23, p. 365701 (2012).

[41] C. Faugeras, B. Faugeras, M. Orlita, M. Potemski, R. R. Nair, and A. K. Geim, "Thermal

Conductivity of Graphene in Corbino Membrane Geometry," ACS Nano, vol. 4, pp. 1889-

1892 (2010).

[42] I. Vlassiouk, S. Smirnov, I. Ivanov, P. F. Fulvio, S. Dai, H. Meyer, M. F. Chi, D. Hensley,

P. Datskos, and N. V. Lavrik, "Electrical and thermal conductivity of low temperature

CVD graphene: the effect of disorder," Nanotechnology, vol. 22, p. 275716 (2011).

[43] V. A. Ermakov, A. V. Alaferdov, A. R. Vaz, A. V. Baranov, and S. A. Moshkalev,

"Nonlocal laser annealing to improve thermal contacts between multi-layer graphene and

metals," Nanotechnology, vol. 24, p. 155301 (2013).

[44] K. F. Mak, C. H. Lui, and T. F. Heinz, "Measurement of the thermal conductance of the

graphene/SiO2 interface," Appl. Phys. Lett., vol. 97, p. 221904 (2010).

[45] Y. K. Koh, M.-H. Bae, D. G. Cahill, and E. Pop, "Heat Conduction across Monolayer and

Few-Layer Graphenes," Nano Lett., vol. 10, pp. 4363-4368 (2010).

[46] P. E. Hopkins, M. Baraket, E. V. Barnat, T. E. Beechem, S. P. Kearney, J. C. Duda, J. T.

Robinson, and S. G. Walton, "Manipulating Thermal Conductance at Metal-Graphene

Contacts via Chemical Functionalization," Nano Lett., vol. 12, pp. 590-595 (2012).

Page 121: © 2014 Zuanyi Li

115

[47] C. W. Zhang, W. W. Zhao, K. D. Bi, J. Ma, J. L. Wang, Z. H. Ni, Z. H. Ni, and Y. F.

Chen, "Heat conduction across metal and nonmetal interface containing imbedded

graphene layers," Carbon, vol. 64, pp. 61-66 (2013).

[48] Z. Chen, W. Jang, W. Bao, C. N. Lau, and C. Dames, "Thermal contact resistance

between graphene and silicon dioxide," Appl. Phys. Lett., vol. 95, p. 161910 (2009).

[49] J. H. Seol, I. Jo, A. L. Moore, L. Lindsay, Z. H. Aitken, M. T. Pettes, X. Li, Z. Yao, R.

Huang, D. Broido, N. Mingo, R. S. Ruoff, and L. Shi, "Two-Dimensional Phonon

Transport in Supported Graphene," Science, vol. 328, pp. 213-216 (2010).

[50] J. H. Seol, A. L. Moore, L. Shi, I. Jo, and Z. Yao, "Thermal Conductivity Measurement of

Graphene Exfoliated on Silicon Dioxide," J. Heat Transf., vol. 133, p. 022403 (2011).

[51] Z. Q. Wang, R. G. Xie, C. T. Bui, D. Liu, X. X. Ni, B. W. Li, and J. T. L. Thong,

"Thermal Transport in Suspended and Supported Few-Layer Graphene," Nano Lett., vol.

11, pp. 113-118 (2011).

[52] M. M. Sadeghi, I. Jo, and L. Shi, "Phonon-interface scattering in multilayer graphene on

an amorphous support," Proc. Natl. Acad. Sci. USA, vol. 110, pp. 16321-16326 (2013).

[53] M. T. Pettes, I. S. Jo, Z. Yao, and L. Shi, "Influence of Polymeric Residue on the Thermal

Conductivity of Suspended Bilayer Graphene," Nano Lett., vol. 11, pp. 1195-1200 (2011).

[54] W. Y. Jang, Z. Chen, W. Z. Bao, C. N. Lau, and C. Dames, "Thickness-Dependent

Thermal Conductivity of Encased Graphene and Ultrathin Graphite," Nano Lett., vol. 10,

pp. 3909-3913 (2010).

[55] M.-H. Bae, Z. Li, Z. Aksamija, P. N. Martin, F. Xiong, Z.-Y. Ong, I. Knezevic, and E.

Pop, "Ballistic to diffusive crossover of heat flow in graphene ribbons," Nat. Commun.,

vol. 4, p. 1734 (2013).

[56] W. Jang, W. Bao, L. Jing, C. N. Lau, and C. Dames, "Thermal conductivity of suspended

few-layer graphene by a modified T-bridge method," Appl. Phys. Lett., vol. 103, p.

133102 (2013).

[57] X. F. Xu, L. F. C. Pereira, Y. Wang, J. Wu, K. W. Zhang, X. M. Zhao, S. Bae, C. T. Bui,

R. G. Xie, J. T. L. Thong, B. H. Hong, K. P. Loh, D. Donadio, B. W. Li, and B. Ozyilmaz,

"Length-dependent thermal conductivity in suspended single-layer graphene," Nat.

Commun., vol. 5, p. 3689 (2014).

[58] V. E. Dorgan, A. Behnam, H. J. Conley, K. I. Bolotin, and E. Pop, "High-Field Electrical

and Thermal Transport in Suspended Graphene," Nano Lett., vol. 13, pp. 4581-4586

(2013).

[59] H. Q. Xie, L. F. Chen, W. Yu, and B. Q. Wang, "Temperature dependent thermal

conductivity of a free-standing graphene nanoribbon," Appl. Phys. Lett., vol. 102, p.

111911 (2013).

[60] Y. J. Yu, M. Y. Han, S. Berciaud, A. B. Georgescu, T. F. Heinz, L. E. Brus, K. S. Kim,

and P. Kim, "High-resolution spatial mapping of the temperature distribution of a Joule

self-heated graphene nanoribbon," Appl. Phys. Lett., vol. 99, p. 183105 (2011).

[61] M. E. Pumarol, M. C. Rosamond, P. Tovee, M. C. Petty, D. A. Zeze, V. Falko, and O. V.

Page 122: © 2014 Zuanyi Li

116

Kolosov, "Direct Nanoscale Imaging of Ballistic and Diffusive Thermal Transport in

Graphene Nanostructures," Nano Lett., vol. 12, pp. 2906-2911 (2012).

[62] F. Menges, H. Riel, A. Stemmer, C. Dimitrakopoulos, and B. Gotsmann, "Thermal

Transport into Graphene through Nanoscopic Contacts," Phys. Rev. Lett., vol. 111, p.

205901 (2013).

[63] R. R. Nair, P. Blake, A. N. Grigorenko, K. S. Novoselov, T. J. Booth, T. Stauber, N. M. R.

Peres, and A. K. Geim, "Fine structure constant defines visual transparency of graphene,"

Science, vol. 320, pp. 1308-1308 (2008).

[64] L. Yang, J. Deslippe, C. H. Park, M. L. Cohen, and S. G. Louie, "Excitonic Effects on the

Optical Response of Graphene and Bilayer Graphene," Phys. Rev. Lett., vol. 103, p.

186802 (2009).

[65] K. S. Kim, Y. Zhao, H. Jang, S. Y. Lee, J. M. Kim, K. S. Kim, J. H. Ahn, P. Kim, J. Y.

Choi, and B. H. Hong, "Large-scale pattern growth of graphene films for stretchable

transparent electrodes," Nature, vol. 457, pp. 706-710 (2009).

[66] V. G. Kravets, A. N. Grigorenko, R. R. Nair, P. Blake, S. Anissimova, K. S. Novoselov,

and A. K. Geim, "Spectroscopic ellipsometry of graphene and an exciton-shifted van

Hove peak in absorption," Phys. Rev. B, vol. 81, p. 155413 (2010).

[67] K. F. Mak, J. Shan, and T. F. Heinz, "Seeing Many-Body Effects in Single- and Few-

Layer Graphene: Observation of Two-Dimensional Saddle-Point Excitons," Phys. Rev.

Lett., vol. 106, p. 046401 (2011).

[68] A. C. Ferrari and D. M. Basko, "Raman spectroscopy as a versatile tool for studying the

properties of graphene," Nat. Nanotechnol., vol. 8, pp. 235-246 (2013).

[69] L. Shi, D. Y. Li, C. H. Yu, W. Y. Jang, D. Kim, Z. Yao, P. Kim, and A. Majumdar,

"Measuring thermal and thermoelectric properties of one-dimensional nanostructures

using a microfabricated device," J. Heat Transf., vol. 125, pp. 881-888 (2003).

[70] A. L. Moore and L. Shi, "On errors in thermal conductivity measurements of suspended

and supported nanowires using micro-thermometer devices from low to high

temperatures," Meas. Sci. Technol., vol. 22, p. 015103 (2011).

[71] P. Kim, L. Shi, A. Majumdar, and P. L. McEuen, "Thermal Transport Measurements of

Individual Multiwalled Nanotubes," Phys. Rev. Lett., vol. 87, p. 215502 (2001).

[72] C. H. Yu, L. Shi, Z. Yao, D. Y. Li, and A. Majumdar, "Thermal conductance and

thermopower of an individual single-wall carbon nanotube," Nano Lett., vol. 5, pp. 1842-

1846 (2005).

[73] C. W. Chang, A. M. Fennimore, A. Afanasiev, D. Okawa, T. Ikuno, H. Garcia, D. Y. Li,

A. Majumdar, and A. Zettl, "Isotope effect on the thermal conductivity of boron nitride

nanotubes," Phys. Rev. Lett., vol. 97, p. 085901 (2006).

[74] C. W. Chang, D. Okawa, H. Garcia, A. Majumdar, and A. Zettl, "Breakdown of Fourier's

law in nanotube thermal conductors," Phys. Rev. Lett., vol. 101, p. 075903 (2008).

[75] M. T. Pettes and L. Shi, "Thermal and Structural Characterizations of Individual Single-,

Double-, and Multi-Walled Carbon Nanotubes," Adv. Funct. Mater., vol. 19, pp. 3918-

Page 123: © 2014 Zuanyi Li

117

3925 (2009).

[76] J. K. Yang, Y. Yang, S. W. Waltermire, T. Gutu, A. A. Zinn, T. T. Xu, Y. F. Chen, and D.

Y. Li, "Measurement of the Intrinsic Thermal Conductivity of a Multiwalled Carbon

Nanotube and Its Contact Thermal Resistance with the Substrate," Small, vol. 7, pp. 2334-

2340 (2011).

[77] D. Y. Li, Y. Y. Wu, P. Kim, L. Shi, P. D. Yang, and A. Majumdar, "Thermal conductivity

of individual silicon nanowires," Appl. Phys. Lett., vol. 83, pp. 2934-2936 (2003).

[78] A. I. Boukai, Y. Bunimovich, J. Tahir-Kheli, J. K. Yu, W. A. Goddard, and J. R. Heath,

"Silicon nanowires as efficient thermoelectric materials," Nature, vol. 451, pp. 168-171

(2008).

[79] A. I. Hochbaum, R. Chen, R. D. Delgado, W. Liang, E. C. Garnett, M. Najarian, A.

Majumdar, and P. Yang, "Enhanced thermoelectric performance of rough silicon

nanowires," Nature, vol. 451, pp. 163-167 (2008).

[80] R. Chen, A. I. Hochbaum, P. Murphy, J. Moore, P. Yang, and A. Majumdar, "Thermal

Conductance of Thin Silicon Nanowires," Phys. Rev. Lett., vol. 101, p. 105501 (2008).

[81] J. W. Lim, K. Hippalgaonkar, S. C. Andrews, A. Majumdar, and P. D. Yang, "Quantifying

Surface Roughness Effects on Phonon Transport in Silicon Nanowires," Nano Lett., vol.

12, pp. 2475-2482 (2012).

[82] C. Y. Ho, R. W. Powell, and P. E. Liley, "Thermal Conductivity of the Elements," J. Phys.

Chem. Ref. Data, vol. 1, pp. 279-421 (1972).

[83] E. Pop, D. Mann, Q. Wang, K. E. Goodson, and H. J. Dai, "Thermal conductance of an

individual single-wall carbon nanotube above room temperature," Nano Lett., vol. 6, pp.

96-100 (2006).

[84] D. L. Nika, A. S. Askerov, and A. A. Balandin, "Anomalous Size Dependence of the

Thermal Conductivity of Graphene Ribbons," Nano Lett., vol. 12, pp. 3238-3244 (2012).

[85] N. Mingo and D. A. Broido, "Length dependence of carbon nanotube thermal conductivity

and the "problem of long waves"," Nano Lett., vol. 5, pp. 1221-1225 (2005).

[86] G. A. Slack, "Anisotropic Thermal Conductivity of Pyrolytic Graphite," Phys. Rev., vol.

127, pp. 694-701 (1962).

[87] C. N. Hooker, Ubbelohd.Ar, and D. A. Young, "Anisotropy of Thermal Conductance in

near-Ideal Graphite," Proc. R. Soc. Lon. Ser.-A, vol. 284, pp. 17-31 (1965).

[88] R. Taylor, "Thermal Conductivity of Pyrolytic Graphite," Philos. Mag., vol. 13, pp. 157-

166 (1966).

[89] J. N. Hu, S. Schiffli, A. Vallabhaneni, X. L. Ruan, and Y. P. Chen, "Tuning the thermal

conductivity of graphene nanoribbons by edge passivation and isotope engineering: A

molecular dynamics study," Appl. Phys. Lett., vol. 97, p. 133107 (2010).

[90] L. Lindsay, D. A. Broido, and T. L. Reinecke, "Phonon-isotope scattering and thermal

conductivity in materials with a large isotope effect: A first-principles study," Phys. Rev.

B, vol. 88, p. 144306 (2013).

[91] T. R. Anthony, W. F. Banholzer, J. F. Fleischer, L. H. Wei, P. K. Kuo, R. L. Thomas, and

Page 124: © 2014 Zuanyi Li

118

R. W. Pryor, "Thermal-Diffusivity of Isotopically Enriched C-12 Diamond," Phys. Rev. B,

vol. 42, pp. 1104-1111 (1990).

[92] R. Berman, "Thermal-Conductivity of Isotopically Enriched Diamonds," Phys. Rev. B,

vol. 45, pp. 5726-5728 (1992).

[93] J. D. Wood, S. W. Schmucker, A. S. Lyons, E. Pop, and J. W. Lyding, "Effects of

Polycrystalline Cu Substrate on Graphene Growth by Chemical Vapor Deposition," Nano

Lett., vol. 11, pp. 4547-4554 (2011).

[94] J. C. Koepke, J. D. Wood, D. Estrada, Z. Y. Ong, K. T. He, E. Pop, and J. W. Lyding,

"Atomic-Scale Evidence for Potential Barriers and Strong Carrier Scattering at Graphene

Grain Boundaries: A Scanning Tunneling Microscopy Study," ACS Nano, vol. 7, pp. 75-

86 (2013).

[95] L. G. Cançado, K. Takai, T. Enoki, M. Endo, Y. A. Kim, H. Mizusaki, A. Jorio, L. N.

Coelho, R. Magalhães-Paniago, and M. A. Pimenta, "General equation for the

determination of the crystallite size La of nanographite by Raman spectroscopy," Appl.

Phys. Lett., vol. 88, p. 163106 (2006).

[96] A. Y. Serov, Z.-Y. Ong, and E. Pop, "Effect of grain boundaries on thermal transport in

graphene," Appl. Phys. Lett., vol. 102, p. 033104 (2013).

[97] Z.-Y. Ong and E. Pop, "Effect of substrate modes on thermal transport in supported

graphene," Phys. Rev. B, vol. 84, p. 075471 (2011).

[98] J. Chen, G. Zhang, and B. W. Li, "Substrate coupling suppresses size dependence of

thermal conductivity in supported graphene," Nanoscale, vol. 5, pp. 532-536 (2013).

[99] Z. P. Xu and M. J. Buehler, "Heat dissipation at a graphene-substrate interface," J. Phys.

Condens. Matter, vol. 24, p. 475305 (2012).

[100] Z. X. Guo, J. W. Ding, and X. G. Gong, "Substrate effects on the thermal conductivity of

epitaxial graphene nanoribbons," Phys. Rev. B, vol. 85, p. 235429 (2012).

[101] B. Qiu and X. Ruan, "Reduction of spectral phonon relaxation times from suspended to

supported graphene," Appl. Phys. Lett., vol. 100, p. 193101 (2012).

[102] E. Pop, D. A. Mann, K. E. Goodson, and H. J. Dai, "Electrical and thermal transport in

metallic single-wall carbon nanotubes on insulating substrates," J. Appl. Phys., vol. 101, p.

093710 (2007).

[103] A. J. Schmidt, K. C. Collins, A. J. Minnich, and G. Chen, "Thermal conductance and

phonon transmissivity of metal-graphite interfaces," J. Appl. Phys., vol. 107, p. 104907

(2010).

[104] M.-H. Bae, S. Islam, V. E. Dorgan, and E. Pop, "Scaling of High-Field Transport and

Localized Heating in Graphene Transistors," ACS Nano, vol. 5, pp. 7936-7944 (2011).

[105] A. Behnam, A. S. Lyons, M.-H. Bae, E. K. Chow, S. Islam, C. M. Neumann, and E. Pop,

"Transport in Nanoribbon Interconnects Obtained from Graphene Grown by Chemical

Vapor Deposition," Nano Lett., vol. 12, pp. 4424-4430 (2012).

[106] S. Islam, Z. Li, V. E. Dorgan, M.-H. Bae, and E. Pop, "Role of Joule Heating on Current

Saturation and Transient Behavior of Graphene Transistors," IEEE Electron Device Lett.,

Page 125: © 2014 Zuanyi Li

119

vol. 34, pp. 166-168 (2013).

[107] M. H. Bae, Z. Y. Ong, D. Estrada, and E. Pop, "Imaging, Simulation, and Electrostatic

Control of Power Dissipation in Graphene Devices," Nano Lett., vol. 10, pp. 4787-4793

(2010).

[108] P. M. Norris, J. L. Smoyer, J. C. Duda, and P. E. Hopkins, "Prediction and Measurement

of Thermal Transport Across Interfaces Between Isotropic Solids and Graphitic

Materials," J. Heat Transf., vol. 134, p. 020910 (2012).

[109] R. J. Stoner, H. J. Maris, T. R. Anthony, and W. F. Banholzer, "Measurements of the

Kapitza Conductance between Diamond and Several Metals," Phys. Rev. Lett., vol. 68, pp.

1563-1566 (1992).

[110] K. C. Collins, S. Chen, and G. Chen, "Effects of surface chemistry on thermal

conductance at aluminum-diamond interfaces," Appl. Phys. Lett., vol. 97, p. 083102

(2010).

[111] G. D. Mahan and J. O. Sofo, "The best thermoelectric," Proc. Natl. Acad. Sci. USA, vol.

93, pp. 7436-7439 (1996).

[112] K. Saito, J. Nakamura, and A. Natori, "Ballistic thermal conductance of a graphene sheet,"

Phys. Rev. B, vol. 76, p. 115409 (2007).

[113] E. Watanabe, S. Yamaguchi, J. Nakamura, and A. Natori, "Ballistic thermal conductance

of electrons in graphene ribbons," Phys. Rev. B, vol. 80, p. 085404 (2009).

[114] K. C. Fong, E. E. Wollman, H. Ravi, W. Chen, A. A. Clerk, M. D. Shaw, H. G. Leduc,

and K. C. Schwab, "Measurement of the Electronic Thermal Conductance Channels and

Heat Capacity of Graphene at Low Temperature," Phys. Rev. X, vol. 3, p. 041008 (2013).

[115] S. Yigen, V. Tayari, J. O. Island, J. M. Porter, and A. R. Champagne, "Electronic thermal

conductivity measurements in intrinsic graphene," Phys. Rev. B, vol. 87, p. 241411

(2013).

[116] S. Yigen and A. R. Champagne, "Wiedemann-Franz Relation and Thermal-Transistor

Effect in Suspended Graphene," Nano Lett., vol. 14, pp. 289-293 (2014).

[117] J. Hone, M. Whitney, C. Piskoti, and A. Zettl, "Thermal conductivity of single-walled

carbon nanotubes," Phys. Rev. B, vol. 59, pp. R2514-R2516 (1999).

[118] Y. M. Zuev, W. Chang, and P. Kim, "Thermoelectric and Magnetothermoelectric

Transport Measurements of Graphene," Phys. Rev. Lett., vol. 102, p. 096807 (2009).

[119] P. Wei, W. Z. Bao, Y. Pu, C. N. Lau, and J. Shi, "Anomalous Thermoelectric Transport of

Dirac Particles in Graphene," Phys. Rev. Lett., vol. 102, p. 166808 (2009).

[120] J. G. Checkelsky and N. P. Ong, "Thermopower and Nernst effect in graphene in a

magnetic field," Phys. Rev. B, vol. 80, p. 081413 (2009).

[121] C. R. Wang, W. S. Lu, and W. L. Lee, "Transverse thermoelectric conductivity of bilayer

graphene in the quantum Hall regime," Phys. Rev. B, vol. 82, p. 121406 (2010).

[122] S. G. Nam, D. K. Ki, and H. J. Lee, "Thermoelectric transport of massive Dirac fermions

in bilayer graphene," Phys. Rev. B, vol. 82, p. 245416 (2010).

Page 126: © 2014 Zuanyi Li

120

[123] S. G. Nam, E. H. Hwang, and H. J. Lee, "Thermoelectric Detection of Chiral Heat

Transport in Graphene in the Quantum Hall Regime," Phys. Rev. Lett., vol. 110, p. 226801

(2013).

[124] D. Q. Wang and J. Shi, "Effect of charged impurities on the thermoelectric power of

graphene near the Dirac point," Phys. Rev. B, vol. 83, p. 113403 (2011).

[125] X. F. Liu, Z. S. Ma, and J. Shi, "Derivative relations between electrical and thermoelectric

quantum transport coefficients in graphene," Solid State Commun., vol. 152, pp. 469-472

(2012).

[126] X. S. Wu, Y. Hu, M. Ruan, N. K. Madiomanana, C. Berger, and W. A. de Heer,

"Thermoelectric effect in high mobility single layer epitaxial graphene," Appl. Phys. Lett.,

vol. 99, p. 133102 (2011).

[127] A. V. Babichev, V. E. Gasumyants, and V. Y. Butko, "Resistivity and thermopower of

graphene made by chemical vapor deposition technique," J. Appl. Phys., vol. 113, p.

076101 (2013).

[128] X. D. Xu, N. M. Gabor, J. S. Alden, A. M. van der Zande, and P. L. McEuen, "Photo-

Thermoelectric Effect at a Graphene Interface Junction," Nano Lett., vol. 10, pp. 562-566

(2010).

[129] K. L. Grosse, M.-H. Bae, F. Lian, E. Pop, and W. P. King, "Nanoscale Joule heating,

Peltier cooling and current crowding at graphene-metal contacts," Nat. Nanotechnol., vol.

6, pp. 287-290 (2011).

[130] J. P. Small, K. M. Perez, and P. Kim, "Modulation of thermoelectric power of individual

carbon nanotubes," Phys. Rev. Lett., vol. 91, p. 256801 (2003).

[131] M. Cutler and N. F. Mott, "Observation of Anderson Localization in an Electron Gas,"

Phys. Rev., vol. 181, pp. 1336-1340 (1969).

[132] S. Cho, S. D. Kang, W. Kim, E. S. Lee, S. J. Woo, K. J. Kong, I. Kim, H. D. Kim, T.

Zhang, J. A. Stroscio, Y. H. Kim, and H. K. Lyeo, "Thermoelectric imaging of structural

disorder in epitaxial graphene," Nat. Mater., vol. 12, pp. 913-918 (2013).

[133] J. Park, G. W. He, R. M. Feenstra, and A. P. Li, "Atomic-Scale Mapping of

Thermoelectric Power on Graphene: Role of Defects and Boundaries," Nano Lett., vol. 13,

pp. 3269-3273 (2013).

[134] F. OuYang, B. Huang, Z. Li, J. Xiao, H. Wang, and H. Xu, "Chemical functionalization of

graphene nanoribbons by carboxyl groups on Stone-Wales defects," J. Phys. Chem. C,

vol. 112, pp. 12003-12007 (2008).

[135] B. Huang, Z. Li, Z. Liu, G. Zhou, S. Hao, J. Wu, B.-L. Gu, and W. Duan, "Adsorption of

gas molecules on graphene nanoribbons and its implication for nanoscale molecule

sensor," J. Phys. Chem. C, vol. 112, pp. 13442-13446 (2008).

[136] J. Li, Z. Li, G. Zhou, Z. Liu, J. Wu, B.-L. Gu, J. Ihm, and W. Duan, "Spontaneous edge-

defect formation and defect-induced conductance suppression in graphene nanoribbons,"

Phys. Rev. B, vol. 82, p. 115410 (2010).

[137] F. Lin, G. Zhou, Z. Li, J. Li, J. Wu, and W. Duan, "Molecular and atomic adsorption of

Page 127: © 2014 Zuanyi Li

121

hydrogen on TiO2 nanotubes: An ab initio study," Chem. Phys. Lett., vol. 475, pp. 82-85

(2009).

[138] B. Huang, Z. Li, Y.-W. Son, G. Kim, W. Duan, and J. Ihm, "Multiple localized states and

magnetic orderings in partially open zigzag carbon nanotube superlattices: An ab initio

study," J. Chem. Phys., vol. 133, p. 084702 (2010).

[139] D. Sim, D. Y. Liu, X. C. Dong, N. Xiao, S. A. Li, Y. Zhao, L. J. Li, Q. Y. Yan, and H. H.

Hng, "Power Factor Enhancement for Few-Layered Graphene Films by Molecular

Attachments," J. Phys. Chem. C, vol. 115, pp. 1780-1785 (2011).

[140] N. Xiao, X. C. Dong, L. Song, D. Y. Liu, Y. Tay, S. X. Wu, L. J. Li, Y. Zhao, T. Yu, H.

Zhang, W. Huang, H. H. Hng, P. M. Ajayan, and Q. Y. Yan, "Enhanced Thermopower of

Graphene Films with Oxygen Plasma Treatment," ACS Nano, vol. 5, pp. 2749-2755

(2011).

[141] S. Hong, E. S. Kim, W. Kim, S. J. Jeon, S. C. Lim, K. H. Kim, H. J. Lee, S. Hyun, D.

Kim, J. Y. Choi, Y. H. Lee, and S. Baik, "A hybridized graphene carrier highway for

enhanced thermoelectric power generation," Phys. Chem. Chem. Phys., vol. 14, pp.

13527-13531 (2012).

[142] C. A. Hewitt and D. L. Carroll, "Extrinsic properties affecting the thermoelectric power

output of few layer graphene/polyvinylidene fluoride composite thin films," Synthetic

Met., vol. 162, pp. 2379-2382 (2012).

[143] B. Huang, Q. Yan, Z. Li, and W. Duan, "Towards graphene nanoribbon-based

electronics," Front. Phys. China, vol. 4, pp. 269-279 (2009).

[144] A. D. Liao, J. Z. Wu, X. Wang, K. Tahy, D. Jena, H. Dai, and E. Pop, "Thermally Limited

Current Carrying Ability of Graphene Nanoribbons," Phys. Rev. Lett., vol. 106, p. 256801

(2011).

[145] F. Miao, S. Wijeratne, Y. Zhang, U. C. Coskun, W. Bao, and C. N. Lau, "Phase-coherent

transport in graphene quantum billiards," Science, vol. 317, pp. 1530-1533 (2007).

[146] X. Wang, Y. Ouyang, L. Jiao, H. Wang, L. Xie, J. Wu, J. Guo, and H. Dai, "Graphene

nanoribbons with smooth edges behave as quantum wires," Nat. Nanotechnol., vol. 6, pp.

563-567 (2011).

[147] Y. Yang and R. Murali, "Impact of size effect on graphene nanoribbon transport," IEEE

Electron Device Lett., vol. 31, pp. 237-239 (2010).

[148] Z. Li, B. Huang, and W. Duan, "The Half-Metallicity of Zigzag Graphene Nanoribbons

with Asymmetric Edge Terminations," J. Nanosci. Nanotechnol., vol. 10, pp. 5374-5378

(2010).

[149] Z. Li, H. Qian, J. Wu, B.-L. Gu, and W. Duan, "Role of symmetry in the transport

properties of graphene nanoribbons under bias," Phys. Rev. Lett., vol. 100, p. 206802

(2008).

[150] T. S. Tighe, J. M. Worlock, and M. L. Roukes, "Direct thermal conductance

measurements on suspended monocrystalline nanostructures," Appl. Phys. Lett., vol. 70,

pp. 2687-2689 (1997).

Page 128: © 2014 Zuanyi Li

122

[151] K. Schwab, E. A. Henriksen, J. M. Worlock, and M. L. Roukes, "Measurement of the

quantum of thermal conductance," Nature, vol. 404, pp. 974-977 (2000).

[152] S. Ryu, J. Maultzsch, M. Y. Han, P. Kim, and L. E. Brus, "Raman Spectroscopy of

Lithographically Patterned Graphene Nanoribbons," ACS Nano, vol. 5, pp. 4123-4130

(2011).

[153] L. Xie, H. Wang, C. Jin, X. Wang, L. Jiao, K. Suenaga, and H. Dai, "Graphene

Nanoribbons from Unzipped Carbon Nanotubes: Atomic Structures, Raman Spectroscopy,

and Electrical Properties," J. Am. Chem. Soc., vol. 133, pp. 10394-10397 (2011).

[154] X. Wang and H. Dai, "Etching and narrowing of graphene from the edges," Nat. Chem.,

vol. 2, pp. 661-665 (2010).

[155] Y. K. Koh, M.-H. Bae, D. G. Cahill, and E. Pop, "Reliably Counting Atomic Planes of

Few-Layer Graphene (n > 4)," ACS Nano, vol. 5, pp. 269-274 (2010).

[156] A. C. Ferrari and J. Robertson, "Interpretation of Raman spectra of disordered and

amorphous carbon," Phys. Rev. B, vol. 61, pp. 14095-14107 (2000).

[157] N. Stojanovic, J. M. Berg, D. H. S. Maithripala, and M. Holtz, "Direct measurement of

thermal conductivity of aluminum nanowires," Appl. Phys. Lett., vol. 95, p. 091905

(2009).

[158] D. Cahill, "Thermal conductivity measurement from 30 to 750 K: the 3 method," Rev.

Sci. Instrum., vol. 61, pp. 802-808 (1990).

[159] P. E. Hopkins, J. R. Serrano, L. M. A. F. Phinney, S. P. Kearney, T. W. Grasser, and C. T.

Harris, "Criteria for Cross-Plane Dominated Thermal Transport in Multilayer Thin Film

Systems During Modulated Laser Heating," J. Heat Trans., vol. 132, p. 081302 (2010).

[160] J. Zhu, D. W. Tang, W. Wang, J. Liu, K. W. Holub, and R. G. Yang, "Ultrafast

thermoreflectance techniques for measuring thermal conductivity and interface thermal

conductance of thin films," J. Appl. Phys., vol. 108, p. 094315 (2010).

[161] H. C. Chien, D. J. Yao, M. J. Huang, and T. Y. Chang, "Thermal conductivity

measurement and interface thermal resistance estimation using SiO(2) thin film," Rev. Sci.

Instrum., vol. 79, p. 054902 (2008).

[162] O. W. Kading, H. Skurk, and K. E. Goodson, "Thermal Conduction in Metallized Silicon-

Dioxide Layers on Silicon," Appl. Phys. Lett., vol. 65, pp. 1629-1631 (1994).

[163] J. H. Kim, A. Feldman, and D. Novotny, "Application of the three omega thermal

conductivity measurement method to a film on a substrate of finite thickness," J. Appl.

Phys., vol. 86, pp. 3959-3963 (1999).

[164] S. M. Lee and D. G. Cahill, "Heat transport in thin dielectric films," J. Appl. Phys., vol.

81, pp. 2590-2595 (1997).

[165] T. Yamane, N. Nagai, S. Katayama, and M. Todoki, "Measurement of thermal

conductivity of silicon dioxide thin films using a 3ω method," J. Appl. Phys., vol. 91, pp.

9772-9776 (2002).

[166] G. Slack, "Thermal Conductivity of Pure and Impure Silicon, Silicon Carbide, and

Diamond," J. Appl. Phys., vol. 35, pp. 3460-3466 (1964).

Page 129: © 2014 Zuanyi Li

123

[167] M. Asheghi, K. Kurabayashi, R. Kasnavi, and K. Goodson, "Thermal conduction in doped

single-crystal silicon films," J. Appl. Phys., vol. 91, pp. 5079-5088 (2002).

[168] A. D. McConnell and K. E. Goodson, "Thermal conduction in silicon micro- and

nanostructures," Annual Reviews of Heat Transfer, vol. 14, pp. 129-168 (2005).

[169] G. S. Kumar, G. Prasad, and R. O. Pohl, "Experimental Determinations of the Lorenz

Number," J. Mater. Sci., vol. 28, pp. 4261-4272 (1993).

[170] E. Munoz, J. Lu, and B. I. Yakobson, "Ballistic Thermal Conductance of Graphene

Ribbons," Nano Lett., vol. 10, pp. 1652-1656 (2010).

[171] R. Prasher, "Thermal boundary resistance and thermal conductivity of multiwalled carbon

nanotubes," Phys. Rev. B, vol. 77, p. 075424 (2008).

[172] M. S. Shur, "Low ballistic mobility in submicron HEMTs," IEEE Electron Device Lett.,

vol. 23, pp. 511-513 (2002).

[173] J. Wang and M. Lundstrom, "Ballistic transport in high electron mobility transistors,"

IEEE Trans. Electron Devices, vol. 50, pp. 1604-1609 (2003).

[174] C. Jeong, S. Datta, and M. Lundstrom, "Full dispersion versus Debye model evaluation of

lattice thermal conductivity with a Landauer approach," J. Appl. Phys., vol. 109, p.

073718 (2011).

[175] C. Jeong, R. Kim, M. Luisier, S. Datta, and M. Lundstrom, "On Landauer versus

Boltzmann and full band versus effective mass evaluation of thermoelectric transport

coefficients," J. Appl. Phys., vol. 107, p. 023707 (2010).

[176] J. Haskins, A. Kınacı, C. Sevik, H. l. Sevincli, G. Cuniberti, and T. Cagın, "Control of

Thermal and Electronic Transport in Defect-Engineered Graphene Nanoribbons," ACS

Nano, vol. 5, pp. 3779-3787 (2011).

[177] Z. Aksamija and I. Knezevic, "Lattice thermal conductivity of graphene nanoribbons:

Anisotropy and edge roughness scattering," Appl. Phys. Lett., vol. 98, p. 141919 (2011).

[178] P. N. Martin, Z. Aksamija, E. Pop, and U. Ravaioli, "Reduced Thermal Conductivity in

Nanoengineered Rough Ge and GaAs Nanowires," Nano Lett., vol. 10, pp. 1120-1124

(2010).

[179] J. Sadhu and S. Sinha, "Room-temperature phonon boundary scattering below the Casimir

limit," Phys. Rev. B, vol. 84, p. 115450 (2011).

[180] A. Y. Goharrizi, M. Pourfath, M. Fathipour, H. Kosina, and S. Selberherr, "An Analytical

Model for Line-Edge Roughness Limited Mobility of Graphene Nanoribbons," IEEE

Trans. Electron Devices, vol. 58, pp. 3725-3735 (2011).

[181] M. Fujii, X. Zhang, H. Q. Xie, H. Ago, K. Takahashi, T. Ikuta, H. Abe, and T. Shimizu,

"Measuring the thermal conductivity of a single carbon nanotube," Phys. Rev. Lett., vol.

95, p. 065502 (2005).

[182] M. E. Siemens, Q. Li, R. Yang, K. A. Nelson, E. H. Anderson, M. M. Murnane, and H. C.

Kapteyn, "Quasi-ballistic thermal transport from nanoscale interfaces observed using

ultrafast coherent soft X-ray beams," Nat. Mater., vol. 9, pp. 26-30 (2010).

[183] Z. H. Ni, H. M. Wang, Y. Ma, J. Kasim, Y. H. Wu, and Z. X. Shen, "Tunable stress and

Page 130: © 2014 Zuanyi Li

124

controlled thickness modification in graphene by annealing," ACS Nano, vol. 2, pp. 1033-

1039 (2008).

[184] Z. Aksamija and I. Knezevic, "Thermal transport in graphene nanoribbons supported on

SiO2," Phys. Rev. B, vol. 86, p. 165426 (2012).

[185] W.-P. Hsieh, A. S. Lyons, E. Pop, P. Keblinski, and D. G. Cahill, "Pressure tuning of the

thermal conductance of weak interfaces," Phys. Rev. B, vol. 84, p. 184107 (2011).

[186] Y. Wang, B. Qiu, and X. Ruan, "Edge effect on thermal transport in graphene

nanoribbons: A phonon localization mechanism beyond edge roughness scattering," Appl.

Phys. Lett., vol. 101, p. 013101 (2012).

[187] F. Volklein, H. Reith, T. W. Cornelius, M. Rauber, and R. Neumann, "The experimental

investigation of thermal conductivity and the Wiedemann-Franz law for single metallic

nanowires," Nanotechnology, vol. 20, p. 325706 (2009).

[188] M. C. Wingert, Z. C. Y. Chen, S. Kwon, J. Xiang, and R. K. Chen, "Ultra-sensitive

thermal conductance measurement of one-dimensional nanostructures enhanced by

differential bridge," Rev. Sci. Instrum., vol. 83, p. 024901 (2012).

[189] R. Sultan, A. D. Avery, G. Stiehl, and B. L. Zink, "Thermal conductivity of

micromachined low-stress silicon-nitride beams from 77 to 325 K," J. Appl. Phys., vol.

105, p. 043501 (2009).

[190] J. K. Yu, S. Mitrovic, D. Tham, J. Varghese, and J. R. Heath, "Reduction of thermal

conductivity in phononic nanomesh structures," Nat. Nanotechnol., vol. 5, pp. 718-721

(2010).

[191] P. Ferrando-Villalba, A. F. Lopeandia, A. Ll, J. Llobet, M. Molina-Ruiz, G. Garcia, M.

Gerbolès, F. X. Alvarez, A. R. Goñi, F. J. Muñoz-Pascual, and J. Rodríguez-Viejo, "In-

plane thermal conductivity of sub-20 nm thick suspended mono-crystalline Si layers,"

Nanotechnology, vol. 25, p. 185402 (2014).

[192] I. Jo, M. T. Pettes, J. Kim, K. Watanabe, T. Taniguchi, Z. Yao, and L. Shi, "Thermal

Conductivity and Phonon Transport in Suspended Few-Layer Hexagonal Boron Nitride,"

Nano Lett., vol. 13, pp. 550-554 (2013).

[193] M. M. Shulaker, J. Van Rethy, T. F. Wu, L. S. Liyanage, H. Wei, Z. Li, E. Pop, G. Gielen,

H. S. P. Wong, and S. Mitra, "Carbon Nanotube Circuit Integration up to Sub-20 nm

Channel Lengths," ACS Nano, vol. 8, pp. 3434-3443 (2014).

[194] K. Kurabayashi, M. Asheghi, M. Touzelbaev, and K. E. Goodson, "Measurement of the

thermal conductivity anisotropy in polyimide films," J. Microelectromech. Syst., vol. 8,

pp. 180-191 (1999).

[195] M. E. Itkis, F. Borondics, A. P. Yu, and R. C. Haddon, "Thermal conductivity

measurements of semitransparent single-walled carbon nanotube films by a bolometric

technique," Nano Lett., vol. 7, pp. 900-904 (2007).

[196] D. Estrada and E. Pop, "Imaging dissipation and hot spots in carbon nanotube network

transistors," Appl. Phys. Lett., vol. 98, p. 073102 (2011).

[197] S. W. Hong, T. Banks, and J. A. Rogers, "Improved Density in Aligned Arrays of Single-

Page 131: © 2014 Zuanyi Li

125

Walled Carbon Nanotubes by Sequential Chemical Vapor Deposition on Quartz," Adv.

Mater., vol. 22, pp. 1826-1830 (2010).

[198] C. A. Wang, K. M. Ryu, L. G. De Arco, A. Badmaev, J. L. Zhang, X. Lin, Y. C. Che, and

C. W. Zhou, "Synthesis and device applications of high-density aligned carbon nanotubes

using low-pressure chemical vapor deposition and stacked multiple transfer," Nano Res.,

vol. 3, pp. 831-842 (2010).

[199] M. C. McAlpine, H. Ahmad, D. W. Wang, and J. R. Heath, "Highly ordered nanowire

arrays on plastic substrates for ultrasensitive flexible chemical sensors," Nat. Mater., vol.

6, pp. 379-384 (2007).

[200] T. Prosen and D. K. Campbell, "Momentum conservation implies anomalous energy

transport in 1D classical lattices," Phys. Rev. Lett., vol. 84, pp. 2857-2860 (2000).

[201] G. Basile, C. Bernardin, and S. Olla, "Momentum conserving model with anomalous

thermal conductivity in low dimensional systems," Phys. Rev. Lett., vol. 96, p. 204303

(2006).

[202] A. Lippi and R. Livi, "Heat conduction in two-dimensional nonlinear lattices," J. Stat.

Phys., vol. 100, pp. 1147-1172 (2000).

[203] L. Wang, B. B. Hu, and B. W. Li, "Logarithmic divergent thermal conductivity in two-

dimensional nonlinear lattices," Phys. Rev. E, vol. 86, p. 040101(R) (2012).

[204] J. A. Johnson, A. A. Maznev, J. Cuffe, J. K. Eliason, A. J. Minnich, T. Kehoe, C. M. S.

Torres, G. Chen, and K. A. Nelson, "Direct Measurement of Room-Temperature

Nondiffusive Thermal Transport Over Micron Distances in a Silicon Membrane," Phys.

Rev. Lett., vol. 110, p. 025901 (2013).

[205] Y. S. Ju and K. E. Goodson, "Phonon scattering in silicon films with thickness of order

100 nm," Appl. Phys. Lett., vol. 74, pp. 3005-3007 (1999).

[206] K. Esfarjani, G. Chen, and H. T. Stokes, "Heat transport in silicon from first-principles

calculations," Phys. Rev. B, vol. 84, p. 085204 (2011).

[207] W. Li, N. Mingo, L. Lindsay, D. A. Broido, D. A. Stewart, and N. A. Katcho, "Thermal

conductivity of diamond nanowires from first principles," Phys. Rev. B, vol. 85, p. 195436

(2012).

[208] F. Yang and C. Dames, "Mean free path spectra as a tool to understand thermal

conductivity in bulk and nanostructures," Phys. Rev. B, vol. 87, p. 035437 (2013).

[209] K. T. Regner, D. P. Sellan, Z. H. Su, C. H. Amon, A. J. H. McGaughey, and J. A. Malen,

"Broadband phonon mean free path contributions to thermal conductivity measured using

frequency domain thermoreflectance," Nat. Commun., vol. 4, p. 1640 (2013).

[210] P. E. Blöchl, "Projector Augmented-Wave Method," Phys. Rev. B, vol. 50, pp. 17953-

17979 (1994).

[211] G. Kresse and J. Furthmüller, "Efficient iterative schemes for ab initio total-energy

calculations using a plane-wave basis set," Phys. Rev. B, vol. 54, pp. 11169-11186 (1996).

[212] J. P. Perdew and Y. Wang, "Accurate and Simple Analytic Representation of the Electron-

Gas Correlation-Energy," Phys. Rev. B, vol. 45, pp. 13244-13249 (1992).

Page 132: © 2014 Zuanyi Li

126

[213] Y. Baskin and L. Meyer, "Lattice Constants of Graphite at Low Temperatures," Phys.

Rev., vol. 100, pp. 544-544 (1955).

[214] P. Trucano and R. Chen, "Structure of Graphite by Neutron-Diffraction," Nature, vol. 258,

pp. 136-137 (1975).

[215] Y. X. Zhao and I. L. Spain, "X-Ray-Diffraction Data for Graphite to 20 Gpa," Phys. Rev.

B, vol. 40, pp. 993-997 (1989).

[216] A. Bosak, M. Krisch, M. Mohr, J. Maultzsch, and C. Thomsen, "Elasticity of single-

crystalline graphite: Inelastic x-ray scattering study," Phys. Rev. B, vol. 75, p. 153408

(2007).

[217] R. S. Pease, "Crystal Structure of Boron Nitride," Nature, vol. 165, pp. 722-723 (1950).

[218] W. Paszkowicz, J. B. Pelka, M. Knapp, T. Szyszko, and S. Podsiadlo, "Lattice parameters

and anisotropic thermal expansion of hexagonal boron nitride in the 10-297.5 K

temperature range," Appl. Phys. A, vol. 75, pp. 431-435 (2002).

[219] N. Wakabayashi, H. G. Smith, and R. M. Nicklow, "Lattice-Dynamics of Hexagonal

Mos2 Studied by Neutron-Scattering," Phys. Rev. B, vol. 12, pp. 659-663 (1975).

[220] W. J. Schutte, J. L. Deboer, and F. Jellinek, "Crystal-Structures of Tungsten Disulfide and

Diselenide," J. Solid State Chem., vol. 70, pp. 207-209 (1987).

[221] J. Serrano, A. Bosak, R. Arenal, M. Krisch, K. Watanabe, T. Taniguchi, H. Kanda, A.

Rubio, and L. Wirtz, "Vibrational properties of hexagonal boron nitride: Inelastic X-ray

scattering and ab initio calculations," Phys. Rev. Lett., vol. 98, p. 095503 (2007).

[222] R. Nicklow, H. G. Smith, and Wakabaya.N, "Lattice-Dynamics of Pyrolytic-Graphite,"

Phys. Rev. B, vol. 5, pp. 4951-4962 (1972).

[223] M. T. Yin and M. L. Cohen, "Theory of Lattice-Dynamical Properties of Solids -

Application to Si and Ge," Phys. Rev. B, vol. 26, pp. 3259-3272 (1982).

[224] A. Togo, F. Oba, and I. Tanaka, "First-principles calculations of the ferroelastic transition

between rutile-type and CaCl(2)-type SiO(2) at high pressures," Phys. Rev. B, vol. 78, p.

134106 (2008).

[225] G. Constantinescu, A. Kuc, and T. Heine, "Stacking in Bulk and Bilayer Hexagonal Boron

Nitride," Phys. Rev. Lett., vol. 111, p. 036104 (2013).

[226] T. Tohei, A. Kuwabara, F. Oba, and I. Tanaka, "Debye temperature and stiffness of

carbon and boron nitride polymorphs from first principles calculations," Phys. Rev. B, vol.

73, p. 064304 (2006).

[227] C. H. Ho, C. S. Wu, Y. S. Huang, P. C. Liao, and K. K. Tiong, "Temperature dependence

of energies and broadening parameters of the band-edge excitons of Mo1-xWxS2 single

crystals," J. Phys. Condens. Matter, vol. 10, pp. 9317-9328 (1998).

[228] V. N. Popov, "Low-temperature specific heat of nanotube systems," Phys. Rev. B, vol. 66,

p. 153408 (2002).

[229] Z. Li, M.-H. Bae, and E. Pop, "Substrate-supported thermometry platform for

nanomaterials like graphene, nanotubes, and nanowires," Appl. Phys. Lett., vol. 105, p.

023107 (2014).

Page 133: © 2014 Zuanyi Li

127

[230] P. L. Garrido, P. I. Hurtado, and B. Nadrowski, "Simple one-dimensional model of heat

conduction which obeys Fourier's law," Phys. Rev. Lett., vol. 86, pp. 5486-5489 (2001).

[231] Z. L. Wang, D. W. Tang, X. H. Zheng, W. G. Zhang, and Y. T. Zhu, "Length-dependent

thermal conductivity of single-wall carbon nanotubes: prediction and measurements,"

Nanotechnology, vol. 18, p. 475714 (2007).

[232] Z. L. Wang, D. W. Tang, X. B. Li, X. H. Zheng, W. G. Zhang, L. X. Zheng, Y. T. T. Zhu,

A. Z. Jin, H. F. Yang, and C. Z. Gu, "Length-dependent thermal conductivity of an

individual single-wall carbon nanotube," Appl. Phys. Lett., vol. 91, p. 123119 (2007).

[233] Q. W. Li, C. H. Liu, X. S. Wang, and S. S. Fan, "Measuring the thermal conductivity of

individual carbon nanotubes by the Raman shift method," Nanotechnology, vol. 20, p.

145702 (2009).

[234] E. K. Sichel, R. E. Miller, M. S. Abrahams, and C. J. Buiocchi, "Heat-Capacity and

Thermal-Conductivity of Hexagonal Pyrolytic Boron-Nitride," Phys. Rev. B, vol. 13, pp.

4607-4611 (1976).

[235] A. Simpson and A. D. Stuckes, "The Thermal Conductivity of Highly Orientated Pyrolytic

Boron Nitride," J. Phys. C: Solid St. Phys., vol. 4, pp. 1710-1718 (1971).

[236] R. S. Yan, J. R. Simpson, S. Bertolazzi, J. Brivio, M. Watson, X. F. Wu, A. Kis, T. F.

Luo, A. R. H. Walker, and H. G. Xing, "Thermal Conductivity of Monolayer

Molybdenum Disulfide Obtained from Temperature-Dependent Raman Spectroscopy,"

ACS Nano, vol. 8, pp. 986-993 (2014).

[237] C. Muratore, V. Varshney, J. J. Gengler, J. J. Hu, J. E. Bultman, T. M. Smith, P. J.

Shamberger, B. Qiu, X. Ruan, A. K. Roy, and A. A. Voevodin, "Cross-plane thermal

properties of transition metal dichalcogenides," Appl. Phys. Lett., vol. 102, p. 081604

(2013).

[238] Y. Xu, J. S. Wang, W. H. Duan, B. L. Gu, and B. W. Li, "Nonequilibrium Green's

function method for phonon-phonon interactions and ballistic-diffusive thermal transport,"

Phys. Rev. B, vol. 78, p. 224303 (2008).

[239] C. Dames and G. Chen, "Theoretical phonon thermal conductivity of Si/Ge superlattice

nanowires," J. Appl. Phys., vol. 95, pp. 682-693 (2004).

[240] Z. Y. Wei, J. K. Yang, W. Y. Chen, K. D. Bi, D. Y. Li, and Y. F. Chen, "Phonon mean

free path of graphite along the c-axis," Appl. Phys. Lett., vol. 104, p. 081903 (2014).