111
HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Electrical and Computer Engineering in the Graduate College of the University of Illinois at Urbana-Champaign, 2014 Urbana, Illinois Doctoral Committee: Adjunct Associate Professor Eric Pop, Chair Professor Rashid Bashir Professor Jean-Pierre Leburton Professor Elyse Rosenbaum

HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

  • Upload
    others

  • View
    12

  • Download
    0

Embed Size (px)

Citation preview

Page 1: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND

MOLYBDENUM DISULFIDE

BY

VINCENT E. DORGAN

DISSERTATION

Submitted in partial fulfillment of the requirements

for the degree of Doctor of Philosophy in Electrical and Computer Engineering

in the Graduate College of the

University of Illinois at Urbana-Champaign, 2014

Urbana, Illinois

Doctoral Committee:

Adjunct Associate Professor Eric Pop, Chair

Professor Rashid Bashir

Professor Jean-Pierre Leburton

Professor Elyse Rosenbaum

Page 2: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

ii

ABSTRACT

My research focuses on the study of nanoscale transistor physics, particularly that of

atomically-thin two-dimensional (2D) crystals such as graphene and molybdenum disulfide

(MoS2). The excellent electrical and thermal properties of 2D materials like graphene have at-

tracted much attention for potential applications in integrated-circuit technology. Understanding

high-field transport in a semiconductor material is crucial not only from the perspective of fun-

damental device physics, but also for achieving practical device applications. Unfortunately,

many of the early measurements on these materials, especially graphene, were focused on low-

temperature and low-field physics.

Motivated by the above, we investigate electrical transport in graphene across a wide

range of temperatures (including near and above room temperature) up to high electric fields (> 1

V/µm) typical of modern transistors. For our measurements, we carefully engineered test struc-

tures to obtain uniform potential and heating along the channel, and we developed simple yet

practical models for heating and high-field drift velocity in graphene, including the roles of both

temperature and carrier density. We find that transport in supported-graphene devices does not

resemble that of ideal graphene, indicating that interactions with the underlying SiO2 play a role

in limiting high-field transport.

We sought to understand the intrinsic electrical and thermal properties of graphene, by

examining devices freely suspended across microscale trenches. We study the coupled electrical

and thermal transport in suspended graphene at high-fields, extracting both high-field drift veloc-

ity and thermal conductivity at breakdown of graphene up to higher temperatures than previously

possible (300-2200 K). We also directly measure the temperature rise due to self-heating in an

electrically biased suspended graphene device via Raman thermometry.

Page 3: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

iii

Lastly, we investigate the electron transport properties of few-layer MoS2 transistors. We

observe a strong temperature dependence of low-field mobility as well as strong self-heating ef-

fects during high-field operation. Interestingly, we observe high-field negative differential con-

ductance (NDC) at low temperature and high bias. Our high-field electrical measurements, com-

bined with detailed modeling and simulations, allow us to provide insight into the high-energy

band structure of MoS2.

As the scaling of transistor lengths approaches 10 nm, it becomes necessary to investigate

novel materials for future nanoscale electronics. The atomically-thin body of 2D materials makes

them robust against short-channel effects, which could enable scaling down to sub-10 nm

MOSFET channel lengths. Additionally, these materials may prove useful for new applications

that take advantage of their inherent 2D nature.

Page 4: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

iv

To my grandparents for emphasizing the importance of education,

and to my brothers for their continuing guidance

Page 5: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

v

ACKNOWLEDGMENTS

First, I would like to thank my advisor, Professor Eric Pop, for welcoming me into his re-

search group and teaching me how to be successful in graduate school. Your guidance, not only

in terms of solving challenging research problems, but also in terms of career and life decisions

has proved useful and invaluable. I am appreciative of Professor Kevin Kim for being my under-

graduate research advisor. You provided me with my first research opportunity, which motivated

me to attend graduate school and follow the path that has led me here. I thank Professor Jean-

Pierre Leburton for being my instructor for several advanced semiconductor courses as well as a

member of my doctoral committee. I am also grateful to Professor Elyse Rosenbaum for serving

on my doctoral committee. Professor Rosenbaum’s ECE 585 course was one of the most inter-

esting courses I attended in graduate school and I am sure the knowledge I gained will be useful

for many years to come. I thank Professor Rashid Bashir for serving on my doctoral committee

and also for initiating a collaborative research project between his group and the Pop lab.

I enjoyed my time spent with colleagues in the Pop lab and am appreciative of the friend-

ly and cooperative atmosphere maintained within the group. I would also like to thank Dr.

Myung-Ho Bae, who first taught me how to exfoliate graphene, perform e-beam lithography, and

many other experimental tasks in and out of the cleanroom. I also thank David Estrada and Al-

bert Liao; as senior members of the Pop lab you were always helpful and willing to give me ad-

vice. I thank Professor Kirill Bolotin and Hiram Conley of Vanderbilt University for providing

essential knowledge for the fabrication of suspended graphene devices. I thank Dr. Ashkan

Behnam for always being willing to help out with a project.

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 Mate-

Page 6: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

vi

rials Research Lab (MRL). I acknowledge funding support from a National Science Foundation

Graduate Research Fellowship, the Semiconductor Research Corporation (SRC), and all other

awards and fellowships I have received from UIUC and the Electrical and Computer Engineering

(ECE) Department.

Finally, I would like to thank all my family and friends. This thesis would not have been

possible without the support of so many people. My grandparents, Pap and Di, instilled in me at

a very young age the importance of education, and much of my success is attributed to their lov-

ing support. I thank Gramma for checking to make sure I was having fun in college and not stud-

ying too hard; you are missed. I thank my parents for encouraging my academic pursuits. My

older brothers, Pete and Greg, are amazing role models and have helped guide me throughout my

entire life. I thank my Aunt Steph for motivating me and always being willing to lend her exper-

tise and wisdom. I am grateful for the many friendships I have had throughout my academic life

and give special recognition to the PCS, Jackal, Jaja Village, and Haus crews. I thank Scott for

all the fun adventures we had; you made graduate school one of the best times of my life. Keisha,

thank you for your companionship and I look forward to our future together.

Page 7: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

vii

TABLE OF CONTENTS

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

1.1 Power Dissipation and Scaling in Integrated-Circuit Technology .................... 1

1.2 Graphene ............................................................................................................ 3

1.3 Molybdenum Disulfide ...................................................................................... 5

1.4 Device Fabrication of Graphene and MoS2 Field-Effect Transistors ................ 6

CHAPTER 2: REVIEW OF HIGH-FIELD TRANSPORT IN GRAPHENE AND

MOLYBDENUM DISULFIDE ...................................................................................................... 9

2.1 Practical Device Operation and High-Field Transport ...................................... 9

2.2 Review of High-Field Transport in Graphene ................................................. 12

2.3 Review of High-Field Transport in MoS2 ....................................................... 21

2.4 Ultimate Device Scaling and High-Field Transport ........................................ 21

CHAPTER 3: MOBILITY AND HIGH-FIELD DRIFT VELOCITY IN GRAPHENE ON

SILICON DIOXIDE ..................................................................................................................... 23

3.1 Introduction ..................................................................................................... 23

3.2 Charge Density Model ..................................................................................... 24

3.3 Low-Field Mobility ......................................................................................... 27

3.4 Thermal Analysis ............................................................................................. 32

3.5 High-Field Transport and Velocity Saturation ................................................ 34

3.6 Conclusions ..................................................................................................... 42

CHAPTER 4: HIGH-FIELD ELECTRICAL AND THERMAL TRANSPORT IN SUSPENDED

GRAPHENE ................................................................................................................................. 43

4.1 Introduction ..................................................................................................... 43

4.2 Electrostatics and Low-Field Mobility ............................................................ 45

4.3 High-Field Electrical Measurements and Drift Velocity at Breakdown ......... 48

4.4 Thermal Analysis ............................................................................................. 53

4.5 Raman Spectroscopy of Electrically Biased Suspended Graphene

Transistor ......................................................................................................... 60

4.6 Variability ........................................................................................................ 63

4.7 Suspended Graphene Nanoconstriction ........................................................... 63

4.8 Conclusions ..................................................................................................... 65

CHAPTER 5: VELOCITY SATURATION AND HIGH-FIELD NEGATIVE DIFFERENTIAL

CONDUCTANCE IN TWO-DIMENSIONAL MOLYBDENUM DISULFIDE

TRANSISTORS ............................................................................................................................ 67

5.1 Introduction ..................................................................................................... 67

5.2 Low-Field Mobility ......................................................................................... 68

5.3 High-Field Negative Differential Conductance and Transistor I-V Model ..... 72

5.4 Thermal Modeling and Raman Thermometry ................................................. 76

Page 8: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

viii

5.5 Boltzmann Transport Equation Model ............................................................ 80

5.6 Contact Resistance and High-Field NDC ........................................................ 83

5.7 Conclusions ..................................................................................................... 87

CHAPTER 6: CONCLUSIONS AND FUTURE WORK ............................................................ 88

REFERENCES ............................................................................................................................. 91

Page 9: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

1

CHAPTER 1: INTRODUCTION

1.1 Power Dissipation and Scaling in Integrated-Circuit Technology

The ability to consistently scale down the size of a transistor in integrated-circuit (IC)

technology has been the crowning achievement of the semiconductor industry for the last several

decades. This trend in scaling, commonly referred to as Moore’s Law [1], has resulted in the

doubling of the number of silicon metal-oxide-semiconductor field-effect transistors (MOSFETs)

for a given area approximately every two years [2], as shown in Fig. 1.1. The obvious result of

scaling is an increase in functionality due to the increased number of transistors per unit area, but

another benefit from scaling is the enhanced performance of the individual transistor.

However, as feature sizes of silicon transistors approach the nanometer scale, transistor

performance no longer scales in proportion with device dimensions, particularly channel length,

L. The trends of power supply voltage, VDD, and threshold voltage, VT, for transistors with short-

er channel lengths exemplify this issue, such that VDD no longer scales proportionally with L, and

VT no longer scales proportionally with VDD [3]. This occurs due to a tradeoff between circuit

speed versus leakage current when adjusting VDD and VT, where scaling down the threshold volt-

age may improve switching speed but will also cause an exponential increase in the “off” cur-

rent.

The inability to scale down VDD proportionally with L causes a problem with respect to

power dissipation. The static power of a transistor circuit may be defined by its leakage current

and supply voltage where PSTATIC = ILEAKVDD. The dynamic or switching power of a transistor

circuit is given by PDYN = CVDD2f, where C is the circuit’s total equivalent capacitance charged

and discharged in a clock cycle and f is the clock frequency. Both the static and dynamic power

increase with increasing VDD; thus. VDD not scaling with L has caused the power density of pro-

Page 10: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

2

cessors to increase exponentially over time, as shown in Fig. 1.2. Therefore, future scaling is lim-

ited by the rate at which heat can be removed from the circuit [4, 5]. Furthermore, the increasing

use of IC technology has made reducing power dissipation in integrated electronics essential, as

the U.S. information technology infrastructure uses up to 10% of national electricity today, a fig-

ure that may triple by 2025 [6]. In addition, PCs in corporate offices are responsible for CO2

emissions equivalent to that of approximately 5 million cars. Based on simple estimates, even a

2x more energy-efficient transistor would lower nation-wide power use by over 10 GW. Thus,

developing low-power nanoscale devices will affect energy consumption by society as a whole,

and has great environmental implications as well.

It is also important to point out that, as scaling has become more difficult with each new

process cycle, novel device structures are needed in order to improve short-channel performance.

The most prominent examples of novel device structures are multiple gate field-effect transistors

Figure 1.1: Scaling of MOSFET gate lengths through the decades in production-stage

ICs as of 2010 (solid red circles) and International Technology Roadmap for Semicon-

ductors (ITRS) targets (open red circles). The corresponding increase in the number of

transistors per processor (blue stars) is shown on the right axis [2].

Page 11: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

3

(FETs), often referred to as FinFETs [7] or Tri-Gate FETs [8, 9]. These device structures take

advantage of improved electrostatics in order to continue transistor scaling. This motivates re-

search that investigates the use of atomically thin materials as the channel material for transis-

tors, as 2D materials present the opportunity for ideal electrostatics and are optimal for the ulti-

mate ultra-thin body FET [10, 11]. This concept will be expanded upon further in Chapter 2.

1.2 Graphene

Graphene is a two-dimensional (2D) crystal of carbon atoms arranged in a hexagonal

structure (see Fig. 1.3a,b), where each atom is bonded to its nearest neighbor by a strong cova-

lent sp2 bond. Each atom also shares a π bond with its three nearest neighbors, which results in a

band of filled π orbitals (valence band) and a band of empty π* orbitals (conduction band). The

thickness of a single sheet of graphene is 0.34 nm. The electronic band structure of graphene is

shown in Fig. 1.3c,d. From the diagram, we see that the Dirac point (K or K´) is where the va-

Figure 1.2: Power density vs. time for computer processors manufactured by AMD, In-

tel, and Power PC over the past two decades [5]. The exponential trend in power density,

although flattened by the introduction of multi-core CPUs, is a limiting factor for the fu-

ture scaling of IC technology.

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 12: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

4

Figure 1.3: (a) Hexagonal lattice of graphene, which consists of two interpenetrating tri-

angular lattices [12] and (b) corresponding Brillouin zone. The distance between adjacent

carbon atoms is 1.42 Å and the distance between carbon atoms in different layers (i.e.,

graphene thickness) is 3.4 Å. (c) Band structure of graphene showing ab initio (solid

lines) and tight binding (dashed) calculations [13]. Graphene has six Dirac (K and K´)

points and a linear dispersion relationship around them. (d) Angle-resolved photoemis-

sion spectroscopy (ARPES) data along symmetry directions in Brillouin zone for gra-

phene [14]. (e) Phonon energy dispersion graphene determined theoretically (lines) from

ab initio calculations and experimentally (solid circles) from high-resolution electron en-

ergy-loss spectroscopy (HREELS) [15]. Phonon dispersion for bulk graphite is also

shown for comparison (open circles) [16].

Page 13: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

5

lence and conduction bands meet. Also, the bands are symmetrical around the Dirac point with a

linear dispersion relation described by E = ħvFk, where vF ≈ 108 cm/s is the Fermi velocity, ħ =

h/(2π) is the reduced Planck constant, and k is the 2D momentum. The symmetry of the conduc-

tion and valence bands around the Dirac point indicates that electrons and holes should have

equal mobilities, unlike in typical semiconductors like Si, Ge, or GaAs where mobilities are

asymmetric, with hole mobility being particularly low. Despite the attention gained by graphene

for future applications in integrated-circuit technology due to its excellent electrical [17, 18] and

thermal properties [19], as well as its impressive mechanical strength [20] and relatively high

optical phonon energies of ~180 meV (see Fig. 1.3e), the absence of a band gap makes it unsuit-

able for conventional digital transistors because of low on/off ratios.

1.3 Molybdenum Disulfide

The attractiveness of 2D materials for use in next-generation nanoelectronic devices has

resulted in the study of other materials besides graphene. For example, there has been a growing

interest in studying the electronic properties of 2D layered transition-metal dichalcogenides

(TMDs) of the form MX2 where M = metal and X = S, Se, or Te [21-23]. Like graphite [24], lay-

ered TMDs consist of stacked 2D atomic layers weakly bound by van der Waals forces, and thus,

like graphene from graphite, atomically-thin TMD layers may be isolated from bulk TMD crys-

tals. TMDs have been studied for decades but their behavior as single- and few-layer atomically

thin materials is new. Among the family of TMDs, molybdenum disulfide (MoS2) has received

special emphasis. Monolayer MoS2 with thickness t ≈ 6.5 Ǻ has a large direct band gap of ~1.8-

1.9 eV [25, 26] (see Fig. 1.4), whereas bulk MoS2 has an indirect band gap of 1.2 eV [27]. The

presence of a band gap in MoS2, unlike in graphene, has resulted in field-effect transistors

(FETs) with high on/off ratios (~108) and low sub-threshold swing (SS ≈ 70 mV per decade) [21,

Page 14: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

6

22]. Room-temperature mobility is typically of the order of 102 cm

2V

-1s

-1, which makes it com-

parable to mobility in ultra-thin Si [28]. Typical transistors from exfoliated MoS2 flakes only ex-

hibit n-type conduction, and as a result, in Chapter 5 we focus on electron transport in MoS2.

Although beyond the scope of this study, we point out that p-type conduction has been observed

in MoS2 [29] under certain conditions and in other TMDs [30].

1.4 Device Fabrication of Graphene and MoS2 Field-Effect Transistors

We fabricate graphene devices by two methods, one by mechanically exfoliating gra-

phene from natural graphite, the other by chemical vapor deposition (CVD) growth on Cu sub-

strates. With the standard “tape method,” graphene is mechanically exfoliated onto a substrate of

~300 nm (or ~90 nm in some cases) of SiO2 with a highly doped Si substrate (p-type, 5×10-3

Ω-

cm). The tape residue is then cleaned off by annealing at 400 °C for 120 min with a flow of

Ar/H2 (500/500 sccm) at atmospheric pressure. Monolayer graphene flakes are then identified

with an optical microscope and confirmed via Raman spectroscopy, as shown in Fig. 1.5a [32].

The process is similar for exfoliating MoS2 except that natural graphite is replaced with a mo-

lybdenite crystal. An example of the characteristic Raman spectrum for MoS2 is shown in Fig.

1.5b. Also, for MoS2 devices, unless stated otherwise, we use few-layer MoS2 flakes where flake

Figure 1.4: (a) Layered structure of MoS2. Thickness of a single layer is 6.5 Å. Geome-

try and atomic spacing [31] of MoS2 lattice is also shown. (b) Electronic band structure

of bulk and monolayer MoS2 calculated from density functional theory (DFT) [26].

Page 15: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

7

thickness is confirmed via atomic force microscopy (AFM) with t ≈ 4 nm (i.e., ~ 6 layers) being

typical.

Graphene growth by CVD is performed by flowing CH4 and Ar gases at 1000 °C and 0.5

Torr chamber pressure, which results primarily in monolayer graphene growth on both sides of

the Cu foil [34]. One graphene side is protected with a ~250 nm thick layer of polymethyl meth-

acrylate (PMMA) while the other is removed with a 20 sccm O2 plasma reactive ion etch (RIE)

for 20 seconds. The Cu foil is then etched overnight in aqueous FeCl3, leaving the graphene sup-

ported by the PMMA floating on the surface of the solution. The PMMA + graphene bilayer film

is transferred via a glass slide to a HCl bath and then to two separate deionized water baths.

Next, the film is transferred to the SiO2/Si substrate and left for a few hours to dry. The PMMA

is removed using a 1:1 mixture of methylene chloride and methanol, followed by a one hour

Ar/H2 anneal at 400 °C to remove PMMA and other organic residue.

After we have exfoliated, or transferred for the case of CVD graphene, onto the SiO2/Si

substrate, we pattern a rectangular graphene channel using e-beam lithography and an O2 plasma

etch. For MoS2 devices, instead of an O2 plasma etch we define rectangular channels using a

Figure 1.5: (a) Raman spectrum showing the G and 2D bands of monolayer graphene. A

single Lorentzian (red) is fitted to the 2D peak [32]. (b) Raman spectrum of MoS2 device

showing characteristic and peaks [33]. Scans here correspond to a 633 nm laser

line.

360 380 400 4200

250

500

750

1000

Raman Shift (cm-1)

Inte

nsit

y (

arb

. unit

s)

360 380 400 4200

250

500

750

1000

Raman Shift (cm-1)

Inte

nsit

y (

arb

. unit

s)

2400 2600 28000

500

1000

1500

2000

2500

Raman Shift (cm-1)

Inte

nsity (

arb

. units)

1400 1600 1800 20000

500

1000

1500

2000

2500

Raman Shift (cm-1)

Inte

nsit

y (

arb

. unit

s)

2D

G

2400 2600 28000

500

1000

1500

2000

2500

Raman Shift (cm-1)

Inte

nsity (

arb

. units)

12gE

1gAgraphene

(a) (b)

MoS2

Page 16: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

8

XeF2 etch consisting of 2 cycles at 60 s/cycle and XeF2 pressure = 3 Torr [35]. Another e-beam

lithography step is used to define the electrodes, which typically consist of ~0.5 nm of Cr and

~40 nm of Pd for supported graphene-based devices, ~1 nm of Cr and ~80 nm of Au for sus-

pended devices, and ~35 nm of Au (i.e., no underlying “sticking” layer) for MoS2-based devices.

For the suspended graphene study discussed in Chapter 4, the suspension process is as

follows. First, we etch away ~200 nm of the underlying 300 nm SiO2 [36]. The sample is then

placed in 50:1 BOE for 18 min followed by a deionized (DI) water bath for 5 min. Isopropyl al-

cohol (IPA) is squirted into the water bath while the water is poured out so that the sample al-

ways remains in liquid. After all the water has been poured out and only IPA remains, the sample

is put into a critical point dryer (CPD). Following the CPD process we confirm suspension using

scanning electron microscopy (SEM) or AFM. All electrical measurements for all studies dis-

cussed here, unless specified otherwise, are performed in vacuum (~10-5

Torr) at the stated back-

ground temperature (T0).

Page 17: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

9

CHAPTER 2: REVIEW OF HIGH-FIELD TRANSPORT IN GRAPHENE

AND MOLYBDENUM DISULFIDE

2.1 Practical Device Operation and High-Field Transport

Understanding high-field transport is not only important from a scientific point of view,

but it is essential for achieving practical applications. This can be further expounded upon by

comparing the long-channel model versus the short-channel model for a typical FET. The basic

long-channel MOSFET equation for calculating drain current (ID) is given by [37]

1

2D ox G T D D

WI C V V V V

L

(2.1)

where μ is the carrier mobility, L and W are the channel length and width, respectively, Cox is the

gate oxide capacitance, VG is the gate voltage, VT is the threshold voltage, and VD is the drain

voltage. Using the long-channel model, when VD ≥ VG – VT the inversion layer at the drain is ef-

fectively in “pinch-off” as shown in Fig. 2.1, and ID no longer rises with an increase in VD. The

saturation current (IDsat) is defined by the drain voltage at the pinch-off point (i.e., VD = VG – VT)

Figure 2.1: Schematic of a typical long-channel MOSFET, shown in this case at the on-

set of saturation such that the pinch-off point is at the drain side of the channel [3].

VGS > VT

VDS = VDSAT

Pinch-offDepletion regionp-Si

n+ n+

VBS

Page 18: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

10

and given by

2

2Dsat ox G T

WI C V V

L (2.2)

where we see that IDsat increases quadratically with overdrive voltage (VGT = VG – VT) and is in-

versely proportional to L. For short-channel MOSFETs, the long-channel analysis, which as-

sumes constant carrier mobility, is no longer applicable as much higher electric fields are present

in short-channel transistors. These higher fields result in the drift velocity (vd) of carriers in

short-channel devices approaching a limiting value known as the saturation velocity (vsat). This

leads to current saturation occurring in a short-channel transistor at much lower voltages than

one would predict if using the long-channel model [3, 38]. Consequently, for short-channel

MOSFETs the saturation current is given by

Dsat sat ox G T DsatI v WC V V V (2.3)

where VDsat is the drain voltage at the onset of current saturation. For increasingly shorter gate

lengths (L → 0), we can estimate the maximum drain current as

maxD sat ox G TI v WC V V (2.4)

where IDmax increases linearly, not quadratically, with overdrive voltage (VGT) and the term

Cox(VG – VT) is an approximation of the 2D carrier density (n) in the channel. Figure 2.2 shows a

comparison of the high-field behavior for long- and short-channel devices discussed here, and

further emphasizes that with increased scaling and higher electric fields in modern transistors,

improved understanding of high-field transport (i.e., the energy dissipation mechanisms that de-

termine vsat for a given material) is critical for enhancing practical device operation.

The 2D devices discussed in later chapters for the work presented here are in effect a hy-

brid of long-channel and short-channel transistors. Channel lengths are typically a few microns

Page 19: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

11

(i.e., longer than the sub-100 nm lengths commonly associated with short-channel transistors),

but by applying a sufficiently high gate bias to these devices, the channel does not approach

pinch-off when operating under high drain-source voltages. Thus, current saturation behavior is

due to velocity saturation effects as discussed above for short-channel devices. Additionally, we

are careful to avoid the formation of significant non-uniform fields along the channel.

Detailed knowledge of the coupling of high-field transport with self-heating is also nec-

essary when discussing practical device operation. As mentioned in Section 1.1, the ability to

effectively remove heat from ICs is a limiting factor for future scaling. At high fields, the charge

carriers (e.g., electrons in conduction band) accelerate and gain energy, or “heat up.” Mecha-

nisms that may limit electron transport include electrons scattering with other electrons, phonons,

interfaces, defects, and impurities. These scattering events not only determine vsat, but when elec-

trons scatter with phonons, electrons can lose energy to the lattice and effectively raise the tem-

perature of the lattice (i.e., Joule heating or self-heating) [4, 39, 40]. Consequently, it is im-

Figure 2.2: (a) Current-voltage characteristics for a MOSFET with L = 0.25 µm and W =

9.5 µm. The solid lines are experimental curves while the dashed line is the long-channel

approximation with velocity saturation effects ignored [38]. (b) Model predictions for

MOSFET saturation current versus channel length without velocity saturation (dashed)

and with velocity saturation (solid) [37].

Without velocity saturation

Dsat 1/I L

With velocity saturation

Dmax ox sat G T( )I = WC v V -V

Channel length L

Sa

tura

tion

Dra

in C

urr

ent I D

sa

t

to ∞

(b)(a)VDS (V)

I DS

(mA

)

Long-channel

behavior

0.25 µm

NMOSFETVG = 2.5 V

2.0 V

1.5 V

1.0 V

0.5 V

VG = 2.5 V

1 2 300

4

8

12

16

Page 20: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

12

portant to account for self-heating effects when investigating high-field transport, as the electron-

ic properties of a material may vary drastically with temperature. For example, the saturation ve-

locity in Si decreases with rising temperature due to an increase in phonon scattering [41]. Anal-

ysis of the temperature dependence of high-field transport in graphene and MoS2 will be given in

later chapters.

2.2 Review of High-Field Transport in Graphene

A review of previous studies of high-field transport in graphene is presented now, while

in Chapters 3 and 4 we continue this discussion and present our original work and analysis of

high-field behavior in graphene. Discussion is limited to monolayer graphene, ignoring bilayer

and few-layer graphene as well as graphene nanoribbons (GNRs), which are beyond the scope of

this study. We also only give attention to steady-state transport, ignoring high-frequency re-

sponse. First, we focus on simulation results for ideal graphene and then include analysis of sub-

strate effects and experimental data.

A detailed review of the current theoretical understanding of electron transport in gra-

phene has recently been presented by Fischetti et al. [42]. We will summarize some of the key

points from this study, as well as other theoretical works, as they pertain to our discussion of

high-field transport. We initially give attention to phonon-limited transport in ideal graphene, as

shown in Fig. 2.3. We point out that the electron drift velocity appears to peak at ~2×107 cm/s at

1 V/µm for n = 1013

cm-2

at 300 K. The onset of negative differential velocity (NDV) at 0.3-0.5

V/µm for lower carrier densities is an intriguing observation, and in this case, is attributed to

electrons gaining enough energy at high electric fields to populate the flatter regions of the Bril-

louin zone, causing a degradation of the drift velocity. The appearance of NDV has been pre-

sented in previous simulations as well.

Page 21: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

13

Akturk and Goldsman [43], using a Monte Carlo simulator, show peak drift velocities for

intrinsic graphene as high as 4.6×107 cm/s and slight NDV for fields above ~10 V/µm. Shishir

and Ferry [44] show similar peak velocities and NDV for n ≤ 2×1012

cm-2

at fields above ~0.5

V/µm. Chauhan and Guo [45] show carrier velocity as a function of electric field up to 1 V/µm

with vsat near 4×107 cm/s, but no NDV is observed. However, the carrier density is fixed at

5.29×1012

cm-2

in their simulations so it is uncertain if NDV would have been observed at lower

densities. The calculated drift velocity vs. electric field for these works is shown in Fig. 2.4. The

variation among different theoretical works for transport in ideal graphene may, at least in part,

be associated with the different choices made for deformation potentials, band structures, and

phonon dispersions. Therefore, drawing qualitative, rather than quantitative, conclusions may be

more appropriate from these simulation studies. For example, a clear observation in [42] and [44]

is the decrease in high-field drift velocity with increasing carrier density, which is an interesting

Figure 2.3: (a) Drift velocity vs. electric field in graphene at 300 K calculated using the

Monte Carlo method [42], and considering only phonon-limited electron transport, i.e.,

electron-electron scattering, substrate effects, and self-heating are ignored. The various

curves are distinguished by the electron density (n). (b) Corresponding electron energy

distribution function calculated for various values of the electric field (F) for the case of

n = 3×1012

cm-2

.

(a) (b)

Page 22: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

14

phenomenon considering typical non-degenerate semiconductors like Si have a constant vsat (e.g.,

~107 cm/s for Si at 300 K) [41]. Graphene has no band gap, and thus is a degenerate semiconduc-

tor where the dependence of saturation velocity on carrier density is due to the degeneracy of

carriers and the Pauli exclusion principle [46].

Continuing the discussion of the carrier-density dependent electron transport in graphene,

we note that Ferry in his simulations [47] used an impurity density that increased with carrier

density, indicating that this increasing impurity density was the only way to match the experi-

mental data in [32] and [48]. It is plausible that a change in impurity density may occur during

measurements due to the application of a high gate field and subsequent motion of impurities

within the oxide. Another explanation is that a more dynamic form of screening, as proposed in

[49], is necessary to properly understand the dependence of transport on carrier density. For fu-

ture analysis of experimental work, we note that a constant impurity density (independent of gate

voltage) is assumed.

Figure 2.4: Calculated drift velocity vs. electric field from Monte Carlo simulations from

(blue solid) Akturk and Goldsman [43], (red dashed) Shishir and Ferry [44], and (green

dotted) Chauhan and Guo [45]. The blue line corresponds to intrinsic graphene. The red

dashed lines vary in carrier density from 5×1011

to 1013

cm-2

from top to bottom. The

multiple green dotted lines have impurity densities of nimp = 0, 1011

cm-2

, and 1012

cm-2

from top to bottom.

10-2

10-1

100

101

102

10-1

100

101

Electric Field (V/um)

Dri

ft V

elo

cit

y (

107

cm

/s) Akturk [43]

Chauhan [45]

Shishir [44]

Page 23: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

15

Returning to the previously mentioned high-field NDV in graphene shown in several the-

oretical studies [42-44, 50], Fang et al. suggest that the inclusion of electron-electrons (e-e) scat-

tering actually removes the NDV effect [46]. Because high-energy electrons exchange their mo-

mentum and energy with low-energy electrons, e-e scattering weakens the backscattering effect

that would lead to NDV [46]. Another effect not yet discussed is carrier multiplication due to

interband tunneling and e-e scattering [51, 52]. If the Fermi level is near the Dirac point we

could expect Zener tunneling and/or impact ionization to generate electrons and holes, especially

at high fields, and increase the carrier concentration. Experimental works have shown that it is

very difficult to obtain saturation of high-field current in graphene transistors [32, 53-56], a char-

acteristic credited to the lack of a band gap in graphene that facilitates ambipolar transport and

the easy transition from electron-dominated to hole-dominate transport (or vice-versa) during

high-field operation. Transport models have been developed that use a phenomenological term to

account for the current up-kick associated with ambipolar transport [57, 58], but we emphasize

that if the Fermi level is near the Dirac point then tunneling and impact ionization effects may

also play a role in the observed current up-kick [51, 52]. For the original experimental work dis-

cussed in later chapters, we typically operate at high electron densities in a unipolar regime, thus

simplifying our transport analysis.

We now turn to substrate effects, a necessary discussion in light of the fact that a majority

of the experimental work on graphene (including Chapter 3 here) has been performed with gra-

phene on SiO2/Si substrates. Although it is beyond our focus here, we also note that graphene on

hexagonal boron nitride (h-BN) is gaining increasing attention due to the observation of im-

proved device performance [59]. An additional scattering process that limits electrical transport

in supported graphene is scattering with charged impurities [60]. Substrate impurities that remain

Page 24: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

16

after the fabrication of graphene transistors [61-63] play a significant role in limiting the low-

field mobility of graphene transistors [64], but do not significantly affect high-field transport [55,

65]. Additionally, we expect that if a graphene transistor is used in the future for nanoscale elec-

tronics, then it will most likely be a top-gated transistor with a high-κ insulator as the top-gate

dielectric [66, 67]. Detailed models have been provided for analyzing the effect of charged impu-

rity scattering and its dependence on the surrounding dielectric environment [68-71], but as we

are primarily concerned with high-field transport here, we will move on to a more relevant scat-

tering process.

For insulators like SiO2 and HfO2 there are bulk dipoles associated with the ionicity of the

metal-oxide bonds. These diploes generate fringing fields on the substrate surface such that the

frequencies of these dipoles are typically determined by the bulk longitudinal optical phonons of

the insulator. Electrons in close proximity to the substrate surface may interact with these surface

optical (SO) phonons via a process commonly referred to as remote-phonon scattering. For Si

inversion layers, remote-phonon scattering has been investigated [72, 73], where it was found to

have a small effect on the drift velocity in the regime just beyond Ohmic transport, but it was not

strong enough to affect the saturation velocity [74], which is determined by the bulk Si optical

phonons. In supported graphene, it is reasonable to consider electrons interacting strongly with

SO phonons considering electrons are essentially confined to the graphene sheet, and the gap be-

tween the graphene and substrate is small, ~0.3-0.4 nm [75]. Furthermore, recent work shows

that graphene may even emit longitudinal out-of-plane acoustic phonons into the substrate, rep-

resenting an energy dissipation mechanism in the vertical direction [76].

Page 25: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

17

When discussing remote-phonon scattering in graphene, Ong and Fischetti [49, 77, 78]

indicate the importance of considering dynamic screening and the charge density response of

graphene due to the electric field created by SO phonons. Figure 2.5 shows the remote-phonon-

limited mobility (µRP) in graphene as a function of carrier density for various supporting sub-

strates. The mobility is calculated by accounting for the hybrid interface phonon-plasmon (IPP)

modes formed from the hybridization of the SO phonons with graphene plasmons [49]. We see

from the plot that µRP is saturating and weakly dependent on n at high carrier densities since

most of the remote phonons are dynamically screened. The observed behavior agrees with the

findings of Zou et al. [67], who extract µRP showing a slight increase with carrier density for a

graphene transistor with HfO2 as the top-gate dielectric. More importantly for our discussion

here, we are interested in the effect of remote-phonon scattering on high-field transport.

The experimental works of Meric et al. [54, 55] suggest that current saturation in a gra-

Figure 2.5: Remote-phonon-limited mobility (µRP) vs. electron density (n) calculated

using theory accounting for hybrid interface phonon-plasmon (IPP) modes [49, 77]. Gra-

phene is supported by various substrates: SiO2 (blue circles), HfO2 (green squares), h-BN

(red diamonds), and Al2O3 (cyan triangles), where degradation in mobility with increas-

ing substrate dielectric constant is evident.

Page 26: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

18

phene transistor may be observed due to remote-phonon scattering, or more specifically, scatter-

ing with the low-energy SO phonon of SiO2 (ħωOP = 55 meV) [79]. We use the term “low-

energy” here for comparison to the graphene (zone-edge) optical phonon with energy ħωOP = 160

meV [80]. Furthermore, Barreiro et al. [53] also suggest that the calculated current (from the

analytic model described below) using a phonon energy of 149 meV overestimates the experi-

mentally observed current in the high-field limit. Unfortunately, these works do not account for

self-heating effects in their analysis, which we know is evident in graphene transistors at high

currents and high fields [81-85] and may be necessary to explain high-field transport in graphene

[56, 65, 86, 87]. Our original work in Chapter 3 will discuss this topic in more detail and help

elucidate the behavior of high-field transport in supported graphene.

Figure 2.6 shows the extracted drift velocity from experimental I-V data [32, 53, 55, 56]

as a function of electric field. Comparison of these curves is difficult as they correspond to dif-

ferent values of carrier density and temperature as well as different device structures. Neverthe-

less, we are able to highlight some key features of high-field transport in graphene. Saturation

velocity in graphene, even at high carrier densities >1013

cm-2

, appears to be larger than 107 cm/s,

(i.e., larger than vsat in Si at room temperature). Figure 2.6 also shows the decrease in high-field

drift velocity with increasing carrier density, as mentioned above and shown in previous theoret-

ical works. Similarly, the high-field drift velocity decreases with increasing temperature, which

is an expected trend if we assume high-field transport is limited by emission of optical phonons.

Further discussion concerning the behavior of high-field drift velocity in graphene is provided in

Section 3.5.

Page 27: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

19

Lastly, in the interest of providing a thorough review, we quickly introduce the analytic

models from Barreiro et al. [53], Freitag et al. [83], and Fang et al. [46]. These models assume

that current saturation, or rather velocity saturation, in graphene is caused by a single optical

phonon of energy ħωOP. The first model [53], as shown in Fig. 2.7a, approximates the high-field

electron distribution with two half-disks such that positive-kx electrons are populated to an ener-

gy ħωOP higher than negative-kx moving electrons. The assumption here is that the inelastic pro-

cess of emitting an OP causes the electron to instantly backscatter. The second model [83] shown

in Fig. 2.7b assumes that for a given electron density defined by n = kF2/π, the high-field

Figure 2.6: Experimentally extracted drift velocity vs. electric field corresponding to da-

ta from (blue triangles) Barreiro et al. [53] for T0 = 300 K at n ≈ 1.7×1012

cm-2

, (red

squares) DaSilva et al. [56] for T0 = 20 K at n ≈ 2.1×1012

cm-2

, (cyan crosses)

Meric et al. [55] for T0 = 300 K at n ≈ 1.1×1012

to 1.4×1012

cm-2

from top to bot-

tom, (magenta crosses) Dorgan et al. [32] and Section 3.5 here for T0 = 80 K at n

≈ 1.9×1012

to 5.9×1012

cm-2

from top to bottom, (tan stars) Dorgan et al. [32] and

Section 3.5 here for T0 = 300 K at n ≈ 2.8×1012

to 6.6×1012

cm-2

from top to bot-

tom, (green circles) Section 3.5 here for T0 = 80 K at n ≈ 2.9×1012

to

1.23×1013

cm-2

from top to bottom, and (black diamonds) Section 3.5 here for T0 =

450 K at n ≈ 2.7×1012

to 7.5×1012

cm-2

from top to bottom. Data from Meric et al.

[55] corresponds to a top-gated device with L = 130 nm and a relatively low low-

field mobility due to high impurity scattering. Consequently, velocity saturation

effects are observed at relatively higher fields than the other devices shown here,

which are all back-gated SiO2-supported graphene transistors.

10-1

100

101

10-1

100

101

Electric Field (V/m)

Dri

ft V

elo

cit

y (

107

cm

/s)

Barreiro [53]

Dorgan [32] and

Section 3.5 here

DaSilva [56]

Meric [55]

Page 28: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

20

transport regime simply consists of electrons within a ±ħωOP/2 window around the Fermi energy

EF, such that EF = ħvFkF = ħvF(πn)1/2

. The last model (Fig. 2.7c) considers a low-energy disk (kl)

and a high-energy disk (kh) defined by ωOP/vF = kh – kl, such that if an electron travelling along

the kx-direction reaches the high-energy circle, then it instantly emits an OP and backscatters.

The “streaming” model accounts for carrier degeneracy in graphene by assuming that the occu-

pation of the low-energy circle is so full that the Pauli exclusion principle prohibits an electron

with energy less than Eh = ħvFkh to emit an OP, which forces the distribution function to be

squeezed and elongated along the direction of the electric field.

We emphasize that these models are more suited for empirical fitting than for providing

physical insight into high-field transport in graphene. It is known that even small changes in the

Figure 2.7: Electron distribution in k-space for analytic models from (a) Barreiro et al.

[53] (also discussed in [32]), (b) Freitag et al. [83] (also discussed in [65]), and

(c) Fang et al. [46]. These models describe high-field transport in graphene as-

suming a single optical phonon (OP) with energy ħωOP is responsible for velocity

saturation. The equations for velocity saturation vsat and current saturation Isat

based on each model are provided as well.

ωOP/vF

kx

ky ky

ωOP/vF

(a) (b) (c)

kl

2

2

21

4

OP OPsat

F

vnvn

2

2 11

4

OP OPsat

F

Wq nI

n v

2 OPsatv

n

ωOP/(2vF)

kx

ky

khkF

2 OPsat

Wq nI

2

2 OP lsat

kv

n

2

2 OP lsat

Wq kI

kx

Page 29: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

21

electron distribution can significantly affect charge transport, and thus, these models most likely

oversimplify the distribution of carriers during high-field operation. Therefore, the apparent

agreement between these simple models and reported trends for high-field drift velocity in gra-

phene is fortuitous (accidental). The “streaming” model from Fang et al. [46] most closely re-

sembles the electron distribution in k-space from Monte Carlo simulations [45] so we choose to

use it in Section 3.5 to empirically fit our graphene high-field transport measurements.

2.3 Review of High-Field Transport in MoS2

High-field transport in MoS2 has received little attention from the community studying

electrical transport in 2D materials. While physical breakdown of monolayer MoS2 at high field

has been reported [88], a solid understanding of high-field transport in mono- and few-layer

MoS2 is presently unavailable. Fiori et al. provide a simple model for high-field transport in few-

layer MoS2 [89], but in this study contact resistance and self-heating effects are ignored. Addi-

tionally, they do not conduct measurements at high enough electric fields and carrier densities or

across a wide enough range of temperatures to provide insight into the high-energy band struc-

ture of MoS2. This insight may prove useful considering the few-layer MoS2 band structure is not

well known as evidenced by some disagreement among existing band structure calculations [26,

31, 90-95]. A thermal analysis of MoS2 field-effect transistors (FETs) would also be important,

particularly the effect of self-heating on high-field operation which is expected to play a signifi-

cant role in applications intended on flexible substrates [96, 97]. Our work in Chapter 5 will pro-

vide a detailed analysis of high-field transport in few-layer MoS2 transistors on SiO2.

2.4 Ultimate Device Scaling and High-Field Transport

Before concluding this review section, we comment on the future direction of nanoscale

electronics and the usefulness of high-field measurements. The continuation of scaling and de-

Page 30: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

22

velopment of short-channel transistors with gate lengths below 10 nm means that charge

transport in devices will approach a quasi-ballistic regime [98, 99]. Understanding of more con-

ventional diffusive transport physics may remain important [100, 101], but in terms of ballistic

transport, accurate knowledge of the band structure and effective mass(es) as well as which val-

leys are populated during device operation will be essential in determining the number of modes

of transport, and ultimately, the current carrying ability of these transistors [102]. Analysis of

high-field transport in 2D materials provides practical and essential insight into the high-energy

band structure of these materials, as exemplified in Chapter 5 for few-layer MoS2.

Additionally, the introduction of FinFETs into state-of-the-art integrated-circuit technol-

ogy has emphasized the need to improve electrostatics for highly-scaled devices. Electrons in

intrinsic 2D materials (e.g., graphene and MoS2) are confined by ionic potentials rather than

band discontinuities so they may offer the only solution to scaling the thickness of semiconduct-

ing bodies. We consider the scaling of double-gate (DG) MOSFETs and point out that in order to

avoid short-channel effects, the scale length (λ) cannot be smaller than the thickness of the chan-

nel (tch) or two times the thickness of the insulator (2ti) [3]. High-κ gate insulators are limited by

tunneling to ~2.5 nm, while monolayer graphene and monolayer MoS2 are only 0.34 nm and

0.65 nm thick, respectively. Consequently, the insulator thickness would be the limiting factor in

terms of ultimate scaling of a DG MOSFET when using an intrinsically 2D channel material.

The minimum λ is ~5 nm resulting in a minimum channel length of ~5-10 nm (Lmin ≈ 1.5λ), rep-

resenting the possibility of a sub-10 nm transistor with a 2D body.

Page 31: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

23

CHAPTER 3: MOBILITY AND HIGH-FIELD DRIFT VELOCITY IN

GRAPHENE ON SILICON DIOXIDE

3.1 Introduction

In this study, we measure mobility in the T = 80-500 K range and high-field drift velocity

at fields F ≈ 1 V/μm in monolayer graphene on SiO2, both as a function of carrier density. We

also introduce simple models including the proper electrostatics, and self-heating at high fields

[81]. As mentioned previously in Chapter 2, the inclusion of self-heating effects is necessary for

proper analysis of high-field transport in graphene transistors. We find that mobility and high-

field drift velocity decrease with rising temperature, and also show a slight decrease with rising

carrier density above 2×1012

cm-2

. The relatively straightforward approach presented can be used

for device simulations or extended to graphene on other substrates.

We fabricated four-probe back-gated graphene structures on SiO2 as described in Section

1.4. We focus on data from two devices with oxide thicknesses tox = 300 nm (L = 4 µm and W =

7 µm) and tox = 90 nm (L = 4 µm and W = 2.2 µm), where the thinner 90 nm oxide results in a

higher gate capacitance, allowing us to obtain a wider range of carrier densities experimentally,

and also resulting in improved heat dissipation that helps mitigate instability at high-fields. We

note that the fabrication of the second device with tox = 90 nm did not require the additional e-

beam lithography step of patterning the graphene channel with an O2 plasma since the original

graphene flake was already rectangular. Also, after numerous high-field measurements on this

device, an outside electrode was damaged, requiring the subsequent use of two-probe measure-

ments using the inner electrodes. We extract contact resistance RC by comparing the four-probe

measured resistance (R4pp) with the two-probe measured resistance (R2pp), i.e., RC = (R2pp –

R4pp)/2. We then determine a temperature-dependent contact resistance RC = 0.3T0 + 48 Ω, where

T0 is the background temperature and the linear dependence of RC on T0 agrees with a previous

Page 32: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

24

study [103]. For analysis of two-terminal measurements we simply redefine the bias across the

inner channel region as V23 = VD – 2IDRC (similarly VG relative to the ground electrode is reduced

by IDRC). Extracted transport properties from four-probe measurements and two-probe measure-

ments (accounting for contact resistance) are identical. A schematic and optical image of the de-

vice structures used here are shown in Figs. 3.1a,b.

3.2 Charge Density Model

To obtain mobility and drift velocity from conductivity measurements, we model the car-

rier density including gate-induced (ncv), thermally generated (nth) carriers, electrostatic spatial

inhomogeneity (n*) and self-heating at high fields. Previous mobility estimates using only ncv

could lead to unphysically high mobility (μ → ∞) near the Dirac voltage (VG = V0) at the mini-

mum conductivity point.

First, we note the gate voltage imposes a charge balance relationship as

0 /cv ox Gn p n C V q (3.1)

where Cox = εox/tox is the capacitance per unit area (quantum capacitance may be neglected here

[104, 105]), εox is the dielectric constant of SiO2, q is the elementary charge, and VG0 = VG–V0 is

Figure 3.1: (a) Schematic of graphene sample (L = 4 μm, W = 7 μm, tox = 300 nm) con-

nected to four-probe electrodes; graphene colorized for clarity. Thermal resistance model

is used to calculate average temperature rise at high bias. (b) Optical image of graphene

device on 90 nm oxide (L = 4 μm, W = 2.2 μm).

(a)

W

L

SiO2

Si (gate)

R B

R ox

R Si

T

T0

tox

12

3

4

1

2

34

(b)W

L

Page 33: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

25

the gate voltage referenced to the minimum conductivity point. Next, we define an average Fer-

mi level EF such that η = EF/kBT, leading to the mass-action law [104]

1 12

2

1 0thpn n

(3.2)

where nth = (π/6)(kBT/ħvF)2 is the thermal carrier density, vF ≈ 10

8 cm/s is the Fermi velocity, and

j(η) is the Fermi-Dirac integral with 1(0) = π2/12. We also point out that in the T → 0 K limit,

the Fermi level may be approximated as EF = ±ħvF(πCox|VG0|/q)1/2

where the upper (lower) sign

corresponds to the electron-doped, VG0 > 0, (hole-doped, VG0 < 0) regime.

Next, we account for the spatial charge (“puddle”) inhomogeneity of graphene due to

substrate impurities [61-63]. The surface potential can be approximated [105] as a periodic step

function whose amplitude ±Δ is related to the width of the minimum conductivity plateau [64] as

given by the residual carrier puddle density (n*) due to charged impurities in the SiO2 (nimp). The

charged impurity density at the SiO2 surface is determined based on the approach discussed in

Adam et al. [64] and given by nimp = BCox|dσ/dVG0|-1

where B = 5×1015

cm-2

is a constant deter-

mined by the screened Coulomb potential in the random phase approximation (RPA) [106] and

dσ/dVG0 is the slope of the low-field conductivity σ = (L/W)(I14/V23). The slope is determined by

a linear fit to σ around the maximum of |dσ/dVG0| as shown in Fig. 3.2. The value of nimp used

here for each sample is based on conductivity measurements at 80 K, where low-field mobility is

limited by Coulomb scattering [60]. From Eq. (10) of [64], we determine n*

≈ 0.297nimp ≈ 2.63×1011

cm-2

and n* ≈ 0.287nimp ≈ 2.89×1011

cm-2

in the samples with tox =

300 nm and 90 nm, respectively, (averaged over the electron- and hole-doped regimes), where n*

is the residual carrier puddle density representing the width of the minimum conductivity plat-

eau. From n*, we obtain a surface potential variation Δ ≈ ħvF (πn*)1/2

of 59 and 63 meV and ΔV0

≈ qn*/Cox of 3.66 and 1.21 V for our two samples, respectively. The extracted values for the sur-

Page 34: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

26

face potential variation are similar to a previous study (~54 meV) [105] and to scanning tunnel-

ing microscopy results (~77 meV) [63].

The total carrier density can be determined numerically by averaging Eqs. (3.1) and (3.2)

for the regions of ±Δ, but this does not yet yield an analytic expression. In order to simplify this,

we note that at low charge density (η → 0) the factor 1(η) 1(-η)/ 12(0) in Eq. (3.2) approaches

unity. Meanwhile, at large |VG0| and high carrier densities the gate-induced charge dominates, i.e.

ncv ≫ nth when η ≫ 1. Finally, we add a correction for the spatial charge inhomogeneity dis-

cussed above, resulting in a minimum carrier density of n0 = [(n*/2)2 + nth

2]1/2

. Consequently,

solving Eqs. (3.1) and (3.2) with the approximations given here results in an explicit expression

for the concentration of electrons and holes

2

0

2 42

1, nnnpn cvcv (3.3)

where the lower (upper) sign corresponds to electrons (holes). Equation (3.3) can be readily used

Figure 3.2: Conductivity vs. back-gate voltage at 80 K for devices of interest with (a) L = 4 µm, W = 7

µm, and tox = 300 nm and (b) L = 4 µm, W = 2.2 µm, and tox = 90 nm. The red dashed lines correspond to

linear fits to the interval around the maximums in |dσ/dVG0|. The slopes of these lines are used to calculate

nimp in accordance with [64].

-20 -10 0 10 200

0.5

1

1.5

2

VG0

(V)

(

mS

)

-20 -10 0 10 200

0.5

1

1.5

2

(

mS

)

VG0

(V)

T0 = 80 K

linear fit regions

( ΔVG = 2 V)

T0 = 80 K

linear fit

regions

( ΔVG = 3 V)

tox = 300 nm

(b)

tox = 90 nm

(a)

Page 35: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

27

in device simulations and is similar to a previous empirical formula [54], but derived here on

more rigorous grounds. We note Eq. (3.3) reduces to the familiar n = CoxVG0/q at high gate volt-

age, and to n = n0 (puddle regime) at VG ≈ V0. Figure 3.3 displays the role of thermal and “pud-

dle” corrections to the carrier density at 300 K and 500 K. These are particularly important near

VG0 = 0 V, when the total charge density relevant to transport (n + p) approaches a constant de-

spite the charge neutrality condition imposed by the gate (n – p = 0). At higher temperatures (kBT

≫ Δ) the spatial potential variation becomes less important due to thermal smearing and larger

nth.

3.3 Low-Field Mobility

Using the aforementioned model for calculating the carrier density in a graphene device,

low-field (~2 mV/µm) mobility is obtained as μ = σ/q(p + n), where the conductivity is given by

σ = (L/W)I14/V23. Figure 3.4 shows measurements of conductivity vs. VG0 for temperatures rang-

Figure 3.3: Calculated carrier density vs. gate voltage at 300 K and 500 K in electron-

doped regime (n > p). Solid lines include contribution from electrostatic inhomogeneity

n* and thermal carriers nth (both relevant at 300 K), dotted lines include only nth (signifi-

cant at 500 K). Dashed line shows only the contribution from the gate-induced carrier

density ncv. Note that the oxide thickness is assumed to be 300 nm in this case.

0 5 10 150

0.5

1

1.5

VG0

(V)

n+

p (

10

12 c

m-2)

gate only

+n0

300 K

+n0

500 K

Page 36: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

28

ing from 80 to 450 K. Mobility is shown as a function of carrier density in Fig. 3.5 for both de-

vices of interest at various temperatures and for both the electron-doped and hole-doped regimes.

Mobility values shown here are similar to previously reported Hall mobility values for graphene

[71]. Although not universal, the mobility at 300 K and above appears to peak at ~2×1012

cm-2

and then decrease at higher carrier densities. A universally observed effect is the decrease in mo-

bility with increasing temperature at high carrier densities [105], as shown in Fig. 3.6. The de-

pendence of mobility on carrier density and temperature suggests the dominant scattering mech-

anism changes from Coulomb to phonon scattering at higher densities and temperatures [105].

Inspired by empirical approximations for Si device mobility [107], the electron mobility

can be fit as (dashed line in Figs. 3.5a and 3.6a)

0 1

( , )1 / 1 / 1ref ref

n Tn n T T

(3.4)

where μ0 = 4650 cm2V

-1s

-1, nref = 1.1×10

13 cm

-2, Tref = 300 K, α = 2.2 and β = 3 for the tox = 300

Figure 3.4: Measure conductivity vs. back-gate voltage for background temperatures T0

= 80–450 K for device with dimensions, L = 4 µm, W = 2.2 µm, and tox = 90 nm. Note

that measurements here are two-terminal measurements, such that the mobility extracted

in Figs. 3.5c,d accounts for contact resistance as described in the text above.

-50 -25 0 25 500

1

2

3

4

VG0

(V)

(

mS

)

80 K

200 K

300 K

400 K

120 K

160 K

350 K

tox ≈ 90 nm

450 K

Page 37: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

29

nm device. Not all samples measured displayed the dip in mobility at low charge density, so we

did not force a mobility fit below ~2×1012

cm−2

. We also point out that the empirical Eq. (3.4)

Figure 3.5: Mobility vs. carrier density in the (a) electron-doped regime (VG0 > 0, n > p)

and (b) hole-doped regime (VG0 < 0, p > n), obtained from conductivity measurements at

T0 = 300–500 K for the device with L = 4 µm, W = 7 µm, and tox = 300 nm. (c) and (d)

correspond to mobility vs. carrier density in the electron-doped regime and hole-doped

regime, respectively, for the device with L = 4 µm, W = 2.2 µm, and tox = 90 nm. The

temperature varies from T0 = 80–450 K for the conductivity measurements of this device.

The qualitative dependence on charge density is similar to that found in carbon nanotubes

[108]. Dashed line in (a) shows fit of Eq. (3.4) with T = 400 K.

1 3 5 7 9 11

2000

3000

4000

5000

6000

7000

n+p (1012 cm-2)

p (

cm

2V

-1s

-1)

1 2 3 4 5 62500

3000

3500

4000

4500

5000

n+p (1012 cm-2)

n (

cm

2V

-1s

-1)

1 2 3 42500

3000

3500

4000

4500

5000

n+p (1012 cm-2)

p (

cm

2V

-1s

-1)

1 3 5 7 9

2000

3000

4000

5000

6000

7000

n+p (1012 cm-2)

n (

cm

2V

-1s

-1)

80 K

200 K

300 K

400 K

120 K

160 K

350 K

n > p

tox ≈ 90 nm

450 K

80 K

200 K

300 K

400 K

120 K

160 K

350 K

p > n

tox ≈ 90 nm

450 K

300 K

450 K

500 K

350 K

400 K

p > n

tox ≈ 300 nm

(c) (d)

n > p

tox ≈ 300 nm

(a) (b)300 K

450 K

500 K

350 K

400 K

Eq. (3.4)

Page 38: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

30

only applies to mobility at or above room temperature. Before concluding our analysis of low-

field mobility, we comment on sample-to-sample variation and uncertainty in the extracted mo-

bility, especially at low charge densities. A possible cause of sample-to-sample variation is the

spatial inhomogeneity of the “puddle” regime that is not accurately predicted by the simple ±Δ

potential model. With respect to uncertainty in our method for mobility extraction, Adam et al.

[64] noted that for the extracted nimp the predicted plateau width is approximate within a factor of

two. This leads to uncertainty in determining ΔV0, and thus uncertainty in the charge density and

extracted mobility values. However, this uncertainty is only notable around the Dirac point,

where the potential ripple contributes to the total carrier density. This limits the “confidence re-

gion” in Figs. 3.5, where we only show carrier densities greater than 0.85×1012

cm-2

. In Fig. 3.7

we estimate the mobility uncertainty at lower charge densities around the Dirac point, such that

the upper and lower bounds result from a potential ripple of ΔV0/2 and 2ΔV0, respectively.

An additional source of uncertainty arises from our use of inner voltage probes that span

the width of the graphene sheet. The advantage of such probes is that they sample the potential

Figure 3.6: Electron mobility vs. temperature at (a) n = 2×1012

(top), 3.5×1012

(middle),

and 5×1012

cm-2

(bottom) for device with dimensions L = 4 µm, W = 7 µm, and tox = 300

nm and at (b) n = 4×1012

(top), 6×1012

(middle), and 8×1012

cm-2

(bottom) for device with

dimensions L = 4 µm, W = 2.2 µm, and tox = 90 nm. Dashed line in (a) shows fit of Eq.

(3.4) with n = 3.5×1012

cm-2

.

100 200 300 400 5002000

3000

4000

5000

Temperature (K)

n (

cm

2V

-1s

-1)

300 400 5002000

3000

4000

5000

Temperature (K)

n (

cm

2V

-1s

-1)

(b)

Eq. (3.4)

5x1012 cm-2

n = 2x1012 cm-2

8x1012 cm-2

n = 4x1012 cm-2(a)

Page 39: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

31

uniformly across the entire graphene width, unlike non-invasive edge-probes which may lead to

potential non-uniformity particularly at high fields. However, full-width probes themselves in-

troduce a few uncertainties in our measurements, which are minimized through careful design as

described here. One challenge may be that, even under four-probe measurements, the current

flowing in the graphene sheet could enter the edge of a voltage probe and partially travel in the

metal. To minimize this effect, we used very narrow inner probes of 300 nm, which is narrower

than the typical charge transfer length between graphene and metal contacts (~1 μm) [109]. In

addition, we employed very “long” devices, with L = 4 μm between the inner probes. Thus, the

resistance of the graphene between the inner electrodes is much greater than both the resistance

of the graphene under the metal contact and that across the narrow metal contact itself.

Another challenge stems from the work function mismatch between the Cr/Pd electrodes

and that of the nearby graphene. This leads to potential and charge non-uniformity in the gra-

phene near the contacts. Previous theoretical work has calculated a potential decay length of ~20

nm induced by Pd contacts on graphene [110]. Experimental photocurrent studies have estimated

that doping from Ti/Pd/Au contacts can extend up to 0.2-0.3 μm into the graphene channel [111],

Figure 3.7: Detail of mobility at 300 K around the Dirac point, with error bars account-

ing for the uncertainty in ΔV0 (error > 15% for |n–p| < 4.5×1011

cm-2

). Data corresponds

to the device with dimensions L = 4 µm, W = 7 µm, and tox = 300 nm.

-0.4 -0.2 0 0.2 0.40

2000

4000

6000

8000

10000

n-p (1012 cm-2)

(

cm

2V

-1s

-1)

Page 40: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

32

although the study had a spatial resolution of only 0.15 μm. Regardless, the potential and charge

disturbance is much shorter than the total channel length of our devices (L = 4 μm). We estimate

at most a ~10% contribution to the resistance from charge transfer at our metal contacts. The er-

ror is likely to become smaller at higher charge densities (>2×1012

cm-2

) where the graphene

charge more strongly screens the contact potential, and at high fields where the graphene channel

becomes more resistive itself.

3.4 Thermal Analysis

Before turning our attention to high-field drift velocity measurements, we must present a

thermal analysis of the supported-graphene device structure. As discussed in Section 2.2, analy-

sis of high-field measurements in graphene poses challenges due to Joule heating effects that

must be taken into account. We estimate the average device temperature due to self-heating via

the thermal resistance network, as shown in Fig. 3.1a [5],

0 B ox SiT T T P R R R (3.5)

where P = I14V23, RB = 1/(hA), Rox = tox/(κoxA), and RSi ≈ 1/(2κSiA1/2

) with A = LW the area of the

channel, h ≈ 108 Wm

-2K

-1 the thermal conductance of the graphene-SiO2 boundary [112], κox and

κSi the thermal conductivities of SiO2 and the doped Si wafer, respectively. At 300 K for our ge-

ometry Rth ≈ 104 K/W, or ~2.8×10

-7 m

2K/W per unit of device area, where Rth is simply the total

thermal resistance calculated by summing the individual thermal resistance components in series.

The thermal resistance value of 104 K/W actually applies to both devices of interest here, since

although the device with tox = 90 nm benefits from improved heat dissipation through the thinner

oxide to the substrate, it has a smaller channel area that balances out this effect. The thermal re-

sistance of the 300 (90) nm SiO2 (Rox) accounts for ~84 (71)% of the total thermal resistance,

Page 41: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

33

while the spreading thermal resistance into the Si wafer (RSi) and the thermal resistance of the

graphene-SiO2 boundary (RB) account for only ~12 (19)% and ~4 (10)%, respectively. We note

that the roles of RSi and RB are more pronounced for the smaller device on a thinner oxide. The

thermal model in Eq. (3.5) can be used when the sample dimensions are much greater than the

SiO2 thickness (W, L ≫ tox) but much less than the Si wafer thickness [5].

The average graphene temperature during high-field measurement will be estimated by

Eq. (3.5). We note that the thermal resistance Rth depends on temperature through κox and κSi,

where we obtain κox = ln(Tox0.52

) – 1.687 and κSi = 2.4×104/T0 by simple fitting to the experi-

mental data in [113] and [114], respectively. Based on the thermal resistance model, we estimate

the average oxide temperature as Tox = (T0 + T)/2, and the temperature of the silicon substrate as

the background temperature T0. This allows for a simple iterative approach to calculating the

graphene temperature rise (ΔT) during a measurement.

Lastly, we discuss the temperature non-uniformity around the inner voltage probes,

which may act as local heat sinks. Analysis here focuses on the device with tox = 300 nm but can

be applied to the thinner oxide device as well. The thermal resistance “looking into” the metal

voltage probes can be estimated as RC,th = LT/κmA, where the thermal healing length LT =

(tmtoxκm/κox)1/2

= 685 nm, tm = 40 nm is the metal thickness, tox = 300 nm is the SiO2 thickness, κm

≈ 50 Wm-1

K-1

is the Pd metal thermal conductivity, κox ≈ 1.3 Wm-1

K-1

is the oxide thermal con-

ductivity (at 300 K), and A=tmWc is the cross-sectional area of the contact with WC = 300 nm. We

obtain RC,th ≈ 106

K/W based on the device geometry here (primarily due to the narrow inner

contacts being only ~300 nm wide), which is about two orders of magnitude greater than the

thermal resistance for heat sinking from the large graphene sheet through the oxide, i.e., Rth ≈

Page 42: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

34

104 K/W at 300 K. Thus, heat flow from the inner metal contacts is negligible.

3.5 High-Field Transport and Velocity Saturation

We now focus on high-field electrical measurements and analysis of high-field transport

in our graphene transistors. To minimize charge non-uniformity and temperature gradients along

the channel at high field [81], we bias the device at high |VG| and avoid ambipolar transport, i.e.,

VGS–V0 and VGD–V0 have the same sign [54]. For example, in the electron-doped regime we ap-

ply a negative bias across the outer electrodes (V14 < 0) such that the effective gate voltage in-

creases during the measurement, thus causing the electron density to increase. This bias setup,

which is opposite of a typical bias one would use for a practical n-type transistor when trying to

achieve pinch-off and current saturation, is appropriate here since we are investigating the unipo-

lar high-field transport physics of graphene. Figure 3.8 shows a comparison of |I14| vs. |V23| for

cases where the applied drain bias across the outer electrodes (V14) is negative (lines) and posi-

tive (crosses). As expected, we see relatively higher currents when V14 < 0 due to the increase in

carrier density during high-field operation. The analysis below pertains to data obtained with V14

< 0. We also indicate that to ensure no sample degradation due to high field stress and strong

self-heating, we checked that low-field σ–VG characteristics were reproducible after each high

bias measurement.

The drift velocity extracted from our measurements is vd = I14/(Wqn23) where n23 is the

average carrier density between terminals 2 and 3, the background temperature is held at T0 = 80

K or T0 = 300 K, and temperature in the channel is T = T0 + ΔT due to self-heating. Figure 3.9a

shows the high-field electrical measurements at 80 K for the device with dimensions L = 4 µm, W

= 2.2 µm, and tox = 90 nm. We estimate the temperature rise ΔT in Fig. 3.9b during high-field opera-

tion based on the thermal model discussed above. Figure 3.9c shows the extracted drift velocity

Page 43: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

35

versus electric field (F), whereas Fig. 3.9d shows vd versus both F and carrier density (n, where n

≫ p here). At high-fields (F > 0.5 V/µm) transport is no longer in the Ohmic regime since vd

does not linearly increase with F, representative of the carrier drift velocity beginning to exhibit

a tendency towards saturation. In Fig. 3.10 we further analyze the velocity-field relationship and

fit the drift velocity by

1/

( )

1d

sat

Fv F

F v

(3.6)

where μ is the low-field mobility and γ ≈ 1.5 provides a good fit (dashed lines in Fig 3.10) for the

carrier densities and temperatures in this work. The bounded shaded regions in Fig. 3.10 corre-

spond to upper and lower estimates for the drift velocity at high fields when using γ ≈ 1 and γ ≈

2, respectively, which have been used previously when fitting Eq. 3.6 to high-field transport data

in graphene [32, 54]. More specifically, we use γ ≈ 2 for fitting to our device with tox= 300 nm.

Figure 3.8: Comparison of |I14| vs. |V23| at T0 = 80 K for VG0 = 10–40 V for cases

where the applied drain bias across the outer electrodes (V14) is negative (lines)

and positive (crosses). Higher currents are observed for V14 < 0 due to the increase

in carrier density in the channel. Data here corresponds to device with dimensions L = 4 µm, W = 2.2 µm, and tox = 90 nm.

0 1 2 30

1

2

3

4

|V23

| (V)

|I14|

(mA

)VG0 = 10 V

0 1 2 30

0.5

1

1.5

2

2.5

3

3.5

4

|V23

| (V)

|I14| (

mA)

data1

data2

data3

data4

data5

data6

data7

data8

0 1 2 30

0.5

1

1.5

2

2.5

3

3.5

4

|V23

| (V)

|I14| (

mA)

data1

data2

data3

data4

data5

data6

data7

data8

V14 < 0

V14 > 0

40 V

20 V

30 V

Page 44: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

36

The larger value of γ = 2 that produces a good fit for the data corresponding to the sample

with tox = 300 nm could be representative of the device having a relatively stronger temperature

dependence for low-field mobility. Another possibility is the device undergoing relatively more

significant self-heating during high-field operation, which makes sense since we reached slightly

higher input powers for the larger device. For simplicity we let µ in Eq. (3.6) be constant for an

individual bias sweep, but we know from Fig. 3.6 that mobility varies with temperature and that

Figure 3.9: (a) High-field electrical measurement showing current vs. voltage at T0 = 80

K for VG0 = 10–50 V for device with dimensions L = 4 µm, W = 2.2 µm, and tox = 90

nm. (b) Estimated temperature rise (ΔT) during high-field operation using Eq. (3.5) for

the measurements in (a). (c) Corresponding velocity-field curves showing the transition

from Ohmic behavior (i.e., linear vd-F relationship) to the high-field regime where vd

trends towards saturation. (d) Drift velocity plotted vs. electron density and electric field,

where we observe a slight decrease in the high-field drift velocity (for equivalent lateral

field values) as the density increases.

2 5 8 11 00.4

0.80

0.5

1

1.5

2

|F| (V/m)n (1012 cm-2)

vd (

10

7 c

m/s

)

0 0.2 0.4 0.6 0.80

0.5

1

1.5

2

|F| (V/m)

vd (

10

7 c

m/s

)

0 1 2 30

1

2

3

4

5

|V23

| (V)

|I14|

(mA

)

VG0 = 10 V

40 V

20 V

30 V

50 V

0 1 2 30

50

100

150

|V23

| (V)

T (

K)

(b)(a)

(d)(c)

T0 = 80 K

VG0 = 10 V

40 V

20 V

30 V50 V

VG0 = 10 V

40 V

20 V

30 V

50 V

Page 45: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

37

the device heats up during high-field operation, and thus, it follows that a larger value of γ may

simply imply increased self-heating effects.

Figure 3.11 exemplifies the temperature dependence of vd, and ultimately vsat, by showing

the extracted velocity-field values (circles), along with the fits (dashed lines) using Eq. (3.6),

from measurements taken at T0 = 80 K (blue) and 450 K (red). As expected, we clearly see a de-

crease in the carrier velocity at higher temperatures. Therefore, we introduce a temperature-

dependent factor to the single-phonon-limited models for saturation velocity discussed in Section

2.2 [46, 53, 83], such that vsat(T) = vsat,T=0/(1 + NOP), where NOP = 1/[exp(ħωOP/kBT) – 1] is the

OP occupation. This factor is qualitatively similar to that in Si [41, 115, 116], and due to the OP

scattering (emission) rate being proportional to (NOP + 1) [108]. If the graphene zone-edge OP with

ħωOP = 160 meV [80] is assumed dominant, then the model only predicts a decrease in vsat of

~2% between ~200 K and ~500 K. However, if we use assume an OP with lower energy plays a

role in limiting high-field transport, e.g., the substrate OP in SiO2 with energy ħωOP = 55 meV

[79], then a ~24% decrease in vsat is predicted between ~200 K and 500 K. The latter is more

Figure 3.10: Electron drift velocity vs. field where circles correspond to the extracted vd

as in Fig. 3.9c and the dashed lines correspond to fits using Eq. (3.6) with good agree-

ment found when using γ = 1.5. The shaded regions represent the range of predicted vd

when using γ = 1 (upper bound) and γ = 2 (lower bound) in Eq. (3.6).

0 0.25 0.5 0.75 10

0.5

1

1.5

2

2.5

|F| (V/m)

vd (

10

7 c

m/s

)

VG0 = 10 V

20 V

30 V

40 V

50 V

T0 = 80 K

Page 46: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

38

agreeable with the data in Fig. 3.11. We expect transport in supported-graphene devices to be

non-deal and affected by the surrounding environment. Thus, if high-field transport is not only

limited by the intrinsic phonons of graphene, we must consider other dissipation mechanisms to

also play a role, such as remote-phonon scattering (see Section 2.2).

As mentioned previously, we expect vsat in graphene to decrease with increasing carrier

density, an effect due to graphene’s unique energy band structure [42, 46]. In Figure 3.12, we

plot vHF vs. electron density, where vHF is used here to denote a drift velocity value at an electric

field that is 0.5 V/µm or greater. In Fig. 3.12a we show the extent of our experimental measure-

ments for both devices in this study by displaying the extracted vHF at the highest fields obtaina-

ble such that the device integrity was maintained and low-field σ–VG characteristics were repro-

ducible. In order to compare carrier velocities at equivalent fields, Fig. 3.12b plots the corre-

sponding vHF at F = 1 V/µm using the fits obtained with Eq. (3.6). We caution against extrapolat-

ing further than F = 1 V/µm because the increase in electron density at higher fields must also be

estimated. Combined with strong self-heating effects, this makes estimates at higher fields diffi-

Figure 3.11: Drift velocity vs. field at background temperatures of (blue) T0 = 80 K and

(red) 450 K for VG0 = 20 V, device with dimensions L = 4 µm, W = 2.2 µm, and tox = 90

nm. Circles (extracted vd) and dashed lines (fits) are as in Fig. 3.10. Estimated tempera-

ture (T = T0 + ΔT) at the peak extracted vd values are 175 K and 490 K for the T0 = 80 K

and 450 K curves, respectively.

0 0.25 0.5 0.75 10

0.5

1

1.5

2

|F| (V/m)

vd (

10

7 c

m/s

)

T = 450 K + ΔT

T = 80 K + ΔTVG0 = 20 V

Page 47: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

39

cult and erroneous.

Figure 3.13 compares the results from Fig. 3.12b with the analytic model by Fang et al.

[46] discussed in Section 2.2 and shown in Fig. 2.7c. The model estimates saturation velocity as

a function of carrier density assuming a “streaming” electron distribution, as shown in the inset

of Fig. 3.13. This model for vsat is limited by inelastic emission of OPs, and thus, we consider

two dominant phonon mechanisms in Fig. 3.13. As before, we consider ħωOP = 55 meV (lower

dashed, SiO2 substrate OP [79]) and ħωOP = 160 meV (upper dashed, graphene OP [80]) and as-

sume T0 = 300 K for both (see discussion of temperature-dependent vsat above). The model lim-

ited by SiO2 phonons slightly underestimates vHF while the model with graphene OPs overesti-

Figure 3.12: High-field drift velocity (vHF) vs. electron density (n). (a) Each symbol cor-

responds to a separate measurement. At T0 = 80 K (blue circles) and 450 K (green circles)

the experimentally extracted drift velocity values are shown where F ranges from 0.53 to

0.67 V/µm (right to left) and 0.5 to 0.55 V/µm (right to left), respectively for the device

with dimensions L = 4 µm, W = 2.2 µm, and tox = 90 nm. Similarly for the device with

dimensions L = 4 µm, W = 7 µm, and tox = 300 nm, at T0 = 80 K (blue squares) and 300

K (red squares) the experimentally extracted drift velocity values are shown where F

ranges from 0.77 to 0.93 V/µm (right to left) and 0.54 to 1 V/µm (right to left), respec-

tively. (b) Using the corresponding fits obtained with Eq. (3.6), we estimate drift velocity

and carrier density at a single electric field value, F = 1 V/µm, and plot for comparison.

The error bars for the blue circles correspond to the bounds of the shaded regions in Fig.

3.10 when varying the power law coefficient from γ = 1 (upper bound) to γ = 2 (lower

bound). Electron vsat for Si and Ge are shown for reference, and we also note that they are

largely independent of carrier density [41, 115, 116].

0 3 6 9 12 150

1

2

3

vH

F (

10

7 c

m/s

)

n (1012 cm-2)

0 3 6 9 12 150

1

2

3

vH

F (

10

7 c

m/s

)

n (1012 cm-2)

T0 = 80 K

T0 = 300 K T0 = 450 K

experimentally

extracted vHF

for F > 0.5 V/µm

vHF from Eq. 3.6

for F = 1 V/µm

T0 = 80 K

T0 = 300 KT0 = 450 K

(b)(a)

vsat,Ge

vsat,Si

vsat,Ge

vsat,Si

Page 48: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

40

mates the measured vHF. Considering the remote-phonon-limited mobility from Fig. 2.5 is shown

to be slightly increasing with increasing carrier density [42], while the extracted vHF decreases

with increasing carrier density, it appears remote-phonon scattering contributes to limiting vsat

but may not be the dominant scattering mechanism. We provide a reminder that this simple

“streaming” model for vsat should be viewed as an empirical equation, considering it is likely an

oversimplification of the electron distribution in the high-field regime. However, it is still useful

for graphene device simulations if calibrated to experimentally extracted values. For example,

the data from Fig. 3.13 can easily be fit using an intermediate OP energy, ħωOP ≈ 75 meV. A

summary of the high-field drift velocity values shown in Figs. 3.12 and 3.13, along with values

Figure 3.13: Comparison of the extracted vHF in Fig. 3.12b with the calculated graphene

saturation velocity (dashed lines) using the model from [46]. The lower and upper dashed

lines correspond to ħωOP = 55 meV (SiO2) and 160 meV (graphene), respectively. This

suggests there is room for improvement with respect to high-field transport for supported-

graphene devices. The inset shows the “streaming” electron distribution used in the ana-

lytic model [46]. We note the temperature-dependent term for vsat discussed above is used

here with T0 = 300 K [32].

Page 49: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

41

for vsat obtained from fits to Eq. (3.6), is provided in Table 3.1.

Before concluding, we estimate from the graphene band structure [42] that the energy-

dispersion relationship around the Dirac point remains linear up to at least ~1 eV. If a significant

proportion of the electron population during our high-field measurements is reaching the flatter

non-linear region, where electrons move slower, then we could attribute some of the velocity

saturation effects observed in our study to this. Using the typical quadratic expression for esti-

mating electron temperature (Te) [117], where Te = T[1+(F/FC)2] with FC corresponding to the

critical field at which hot electron effects begin, we calculate the electron energy to be less than

Page 50: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

42

50 meV for measurements from Fig. 3.9. Here we use FC = 0.5 V/µm, which is a relatively con-

servative estimate [52]. Therefore, it is unlikely that many electrons in our transport measure-

ments approach the flatter non-linear region of the band structure, indicating that other scattering

mechanisms (see above) are most likely responsible for the observed high-field behavior. Never-

theless, it is plausible that electrons at high fields in graphene devices on more ideal substrates

(e.g., hexagonal boron nitride [59]), or suspended graphene devices, may be able to gain enough

energy to populate the non-linear regime and contribute to velocity saturation.

3.6 Conclusions

In summary, we examined mobility and high-field drift velocity in graphene on SiO2, in-

cluding the roles of carrier density and temperature. The results indicate that high-field transport

in graphene on SiO2 does not exhibit exclusively intrinsic behavior, suggesting that additional

scattering mechanisms (e.g., remote-phonon scattering with the substrate) contribute to limiting

high-field drift velocity in graphene. Similarly, charged impurities on the substrate play a role in

limiting low-field mobility, especially at low carrier densities. Improving the graphene-substrate

interface is a necessary step to improve graphene device performance and reach the intrinsic

transport limits of this 2D material. Lastly, the models introduced here are simple yet practical,

and can be used in future simulations of graphene devices operating up to high fields.

Page 51: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

43

CHAPTER 4: HIGH-FIELD ELECTRICAL AND THERMAL

TRANSPORT IN SUSPENDED GRAPHENE

4.1 Introduction

As discussed in Chapters 2 and 3, the transport properties of atomically-thin graphene are

affected by interactions with adjacent materials. Therefore, in order to understand the intrinsic

electrical and thermal properties of graphene, it is necessary to study devices freely suspended

across microscale trenches [118-125]. Compared to devices on typical substrates like SiO2, the

intrinsic mobility of electrons and holes in graphene is increased by up to a factor of ten [32,

118-120], and the thermal conductivity by about a factor of five when suspended. However, pre-

vious electrical transport studies of suspended graphene have only examined low-field and low-

temperature conditions. Also in light of the fact that high-field measurements carried out on sus-

pended carbon nanotubes (CNTs) had previously revealed a wealth of new physical phenomena,

including negative differential conductance [126, 127], thermal light emission [128], and the

presence of non-equilibrium optical phonons [126, 129], we have sufficient motivation to inves-

tigate high-field behavior in suspended graphene. Furthermore, since we expect even more sig-

nificant self-heating in suspended devices than supported devices due to the lack of a heat dissi-

pation pathway to an underlying substrate, we give particular attention to the coupled high-field

electrical and thermal transport.

The approach used here, where we examine suspended graphene devices at high fields as

well as at high temperatures, enables us to extract both the high-field drift velocity of charge car-

riers and the thermal conductivity of graphene up to higher temperatures than previously possible

(>1000 K) [130]. Our systematic study includes experimental analysis of 15 samples combined

with extensive simulations, including modeling of coupled electrical and thermal transport and

modeling of graphene-metal contact effects. We uncover the important role that thermally gener-

Page 52: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

44

ated carriers play in such situations, and also discuss important high-field transport properties at

elevated temperatures up to device breakdown.

A schematic of a typical suspended graphene device is shown in Fig. 4.1a. To assess the

broadest range of samples, we fabricated such devices using both mechanically exfoliated gra-

phene and graphene grown by chemical vapor deposition (CVD). The suspension process is as

Figure 4.1: (a) Schematic of suspended graphene device. The color scale indicates the

temperature of a suspended device during high-field current flow in vacuum (here calcu-

lated for the sample corresponding to Fig. 4.2a,b, with applied power P = 1.2 mW).

Scanning electron microscopy (SEM) images of: (b) suspended graphene grown by

chemical vapor deposition (CVD), and (c,d) suspended exfoliated graphene samples. All

SEM images taken at a 70° tilt with respect to the substrate. (e) AFM image of suspended

exfoliated graphene where the dashed blue line corresponds to the (f) height vs. distance

trace. The initial SiO2 thickness is 300 nm, of which approximately 200 nm is etched dur-

ing the suspension process.

Page 53: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

45

described in Section 1.4. After fabrication we confirm suspension via scanning electron micros-

copy (SEM) and atomic force microscopy (AFM), as shown in Figs. 4.1b-f. In order to avoid

damaging the graphene from e-beam irradiation during SEM [131], we only image “dummy”

devices or use low acceleration voltages (~1 kV) to conduct SEM prior to making measurements.

Some devices show a small amount of “wrinkling” (see, e.g., Fig. 4.1d), possibly leading to some

of the sample-to-sample variability described in Section 4.6. We note that the CVD graphene

samples underwent a vacuum anneal at 200 °C after the suspension process. Typically, device

performance improved after this anneal but we also noticed several devices would break. Due to

the fragility and relatively limited number of exfoliated graphene devices, we did not perform

this additional annealing step with the exfoliated graphene samples.

4.2 Electrostatics and Low-Field Mobility

Figure 4.2a shows a typical resistance (R) vs. back-gate voltage (VG) measurement for a

suspended exfoliated graphene device with length L ≈ 1.5 μm and width W ≈ 850 nm, in vacuum

(~10-5

Torr) at room temperature. For all measurements, we limit the back-gate voltage to |VG| ≤

10 V to avoid collapsing the suspended channel [132, 133]. The as-fabricated devices do not

immediately exhibit the “clean” electrical behavior one might associate with freely suspended

graphene [118-120]. Although the channel is not in contact with the substrate, some residue from

processing remains on the device, and this can be removed or minimized through a current an-

nealing technique [118, 120]. In this process, we sweep the drain voltage (VD) to increasingly

higher values until the Dirac voltage (V0) appears within the narrow usable VG window (|VG| ≤ 10

V) and the electrical characteristics of the device stabilize (Fig. 4.2a).

We note that increases or decreases in resistance that do not always correspond to chang-

es in V0 are often seen during the annealing process. Although we expect the removal of impuri-

Page 54: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

46

ties to affect the graphene resistivity, we believe these shifts are often caused by a change in the

contact resistance. For cases when the device resistance increases substantially we find that the

graphene channel has undergone partial breakdown (see Sections 4.6 and 4.7). We took care to

avoid using such devices when extracting transport properties.

Figure 4.2b displays the room temperature effective mobility (μ0) of the device from Fig.

4.2a. There is some uncertainty in the mobility extraction because the carrier density is not well

known at low-field, in part due to limited knowledge of the residual doping density n*. Thus, Fig.

Figure 4.2: (a) Measured resistance vs. gate voltage VG at room temperature, before and

after current annealing for a suspended exfoliated graphene device (L = 1.5 μm, W = 0.85

μm, VDS = 50 mV). (b) Effective mobility for the data shown in (a), assuming three dif-

ferent carrier densities as labeled. The contact resistance was estimated using the transfer

length method (TLM) for devices of different channel lengths. (c) Calculated total carrier

density (n + p) vs. gate voltage at increasing temperatures. In such suspended devices the

carrier density becomes only a function of temperature (due to thermal carrier generation,

nth) [32] and independent of gate voltage at temperatures >600 K. This corresponds to all

high-field transport cases studied in this work (Fig. 4.4).

-10 -5 0 5 103.5

4

4.5

5

5.5

VG (V)

R (

k

)

0 2 4 6 8 1010

11

1012

1013

VG0

(V)

n+

p (

cm

-2)

(a) 2000 K

300 K

600 K

1500 K

1000 K

-10 -5 0 5 101

1.5

2

2.5

3

VG0

(V)

(

10

4 c

m2/V

s)

(b)

(c)

n* = 109 cm-2

2 1011 cm-2

1011 cm-2

n* = 2 1011 cm-2

before current

annealing

after current annealing

300 K

thermal

carriers

dominate

VG – V0 (V) VG – V0 (V)

n +

p (

cm

-2)

1013

1012

1011

μ0

(10

4cm

2/V∙s)

Page 55: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

47

4.2b displays the effective mobility at several residual doping levels [84] n* = 10

9, 10

11, and

2×1011

cm-2

. Nevertheless, the estimated mobility is above 15,000 cm2V

-1s

-1 at room tempera-

ture, consistent with previous work [120], which pointed out such values are limited by flexural

phonons in suspended graphene at all but the lowest temperatures (T ≥ 10 K).

Interestingly, we note that significantly above room temperature the carrier density in the

suspended graphene channel becomes dominated by thermally generated carriers (nth) and inde-

pendent of gate voltage as shown in Fig. 4.2c. The gate capacitance (CG) is determined by the

series combination of the air gap (tair ≈ 200 nm) and the remaining SiO2 (tox ≈ 100 nm), where

Cair = εairε0/tair, Cox = εoxε0/tox, and CG = (Cair-1

+ Cox-1

)-1

. We obtain CG ≈ 4 nF/cm2, which means

at VG0 = VG − V0 = 10 V we estimate that the gate voltage induces ncv = CGVG0/q ≈ 2.5×1011

cm−2

carriers; however, at 1000 K the total population of thermally generated carriers (electrons plus

holes) is dominant, 2nth = 2(π/6)(kBT/ħvF)2 ≈ 1.8×10

12 cm

−2, and increasing quadratically with

temperature, where kB is the Boltzmann constant, ħ is the reduced Planck constant, and vF ≈ 108

cm/s is the graphene Fermi velocity. These trends are illustrated in Fig. 4.2c at temperatures

ranging from 300 to 2000 K, calculated using the charge-density model described in Section 3.2.

Before discussing high-field measurements, we must first address the issue of contact re-

sistance in our devices. Contact resistance in the CVD graphene devices is determined using the

transfer length method (TLM). In Fig. 4.3 we plot R·W versus L and extract RCW ≈ 1100 Ω-um

from the linear fit (dashed line). We use the resistance value at breakdown for extracting contact

resistance. The spread in R·W values that results in a poor linear fit in Fig. 4.3 may be associated

with the difficulty in determining W after breakdown, along with other causes of sample-to-

sample variation discussed below. Therefore, we let RCW = 200 to 2000 Ω-μm (typical values for

“good” and “bad” contacts respectively) for extracting the upper and lower bounds of average

Page 56: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

48

carrier velocity and thermal conductivity shown later in Figs. 4.5 and 4.7. We use TLM to extract

contact resistance for the exfoliated devices, but due to the limited amount of breakdown data,

we use resistance values from low-field measurements of devices of varying length from the

same sample. The average RCW for the exfoliated graphene devices in this work is ~1800 Ω-μm.

We vary RCW by 50% to provide bounds in Figs. 4.5 and 4.7 similar to the case with CVD gra-

phene. A possible reason for the exfoliated graphene having a larger contact resistance than that

of the CVD graphene is the additional anneal in vacuum at 200 °C that the CVD graphene sam-

ples underwent, but the exfoliated graphene did not.

4.3 High-Field Electrical Measurements and Drift Velocity at Breakdown

We now turn to our high-field transport measurements of suspended graphene devices.

Figure 4.4a displays the measured current density (I/W) as a function of average electric field

along the channel F ≈ (VD − 2IRC )/L, up to irreversible electrical breakdown of suspended exfo-

liated (red) and CVD-grown (blue) graphene devices, in vacuum (~10−5

Torr) at room tempera-

ture. Some devices show linearly increasing current at low fields followed by saturation-like be-

havior at high fields, while other devices show linear (and sometimes superlinear) current

Figure 4.3: R·W vs. L for multiple CVD graphene devices taken at the breakdown point

in vacuum. The dashed line is a linear fit where the y-intercept corresponds to 2RCW ≈

2200 Ω-um.

0 1 2 30

5

10

15

20

25

L (m)

RW

(k

-m

)

2RCW

Page 57: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

49

throughout.

To understand this behavior, in Fig. 4.4b we use our self-consistent electrical−thermal

simulator of graphene [81, 82, 134] to model I/W versus F up to suspended graphene device

breakdown. The model used to provide the simulations shown is based on applying the sus-

pended device geometry to the models developed in previous works [32, 135]. First, the gate ca-

pacitance CG is calculated as described above. Second, we simulate a suspended channel region

by setting the thermal conductance to the substrate to zero, but we still allow for heat loss to the

substrate underneath the contacts. Third, we adjust our model of drift velocity saturation in order

to more accurately represent a suspended graphene sheet. Substrate effects should no longer limit

transport at high fields so we assume the saturation velocity (vsat) is determined by the zone-edge

optical phonon (OP), ħωOP = 160 meV [80]. We also use a Fermi velocity that is dependent on

carrier density due to changes in the linear energy spectrum near the neutrality point for sus-

pended graphene, vF(n) ~ v0[1 + ln(n0/n)/4], where v0 = 0.85×106 m/s and n0 = 5×10

12 cm

-2 [136].

Fourth, the thermal generation of carriers, nth = α(π/6)(kBT/ħvF)2, is given a slightly stronger than

T2 dependence above room temperature (although it decays back to a T

2 dependence at high T)

by introducing ( ) √ , where T0 = 300 K. This empirical fit is used to

account for the possible electron-hole pair generation from optical phonon decay [85] and/or Di-

rac voltage shifting that may occur during high-field device operation. Lastly, we add a contribu-

tion to the carrier density near the contacts (nC) to account for the modification of the graphene

electronic structure by the metal contacts, such that nC = 1011

cm-2

at the contact but exponential-

ly decreases away from the contact with a decay length of ~200 nm [111].

Page 58: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

50

For the simulated curves in Fig. 4.4b we varied the room-temperature low-field mobility

(μ0) from 2,500 to 25,000 cm2V

-1s

-1 along with the temperature dependence of the mobility (μ ~

T-β

where β varies from 1.5 to 2.5). This is meant to represent the range of “dirty” to “clean” de-

vices that we measured experimentally. We accordingly vary the room-temperature thermal con-

ductivity (κ0) from 2000 to 3000 Wm-1

K-1

and the breakdown temperature (TBD) from 1420 to

2860 K, respectively. Thus, we are able to replicate the different types of curves observed exper-

imentally. We find that devices showing saturation-like behavior at high fields typically have

higher μ0 and stronger mobility dependence on temperature (i.e., μ(T) ~ T−β

where β ≈ 2.5), being

essentially “cleaner” and less disordered. Conversely, devices not showing saturation-like behav-

ior have relatively low μ0 and a weaker temperature dependence (β ≈ 1.5), which most likely cor-

responds to higher residual doping and disorder [118, 137, 138]. The superlinear current rise in

such devices is due to the sharp increase in thermally generated carriers (nth) as the device heats

up. We note that overall the I-V behavior of suspended graphene devices is markedly different

from that of suspended carbon nanotubes [126-129]. Suspended carbon nanotubes show negative

Figure 4.4: (a) Measured current density (I/W) vs. average electric field up to breakdown

in vacuum of suspended graphene devices (VG = 0). Exfoliated graphene (red) and CVD-

grown graphene (blue) devices. The average electric field is F = (VDS – 2IRC)/L, account-

ing for the electrical contact resistance RC. (b) Simulated I/W vs. F with varying low-field

mobility (μ0) from 2,500 to 25,000 cm2V

-1s

-1. The simulations are based on our electro-

thermal self-consistent simulator [81, 82, 134], adapted here for suspended graphene.

Suspended devices which reach higher, saturating current and break down at lower volt-

age are expected to be representative of cleaner, less disordered samples.

0 1 2 30

0.2

0.4

0.6

0.8

1

F (V/m)

I/W

(m

A/

m)

0 1 2 30

0.2

0.4

0.6

0.8

1

F (V/m)

I/W

(m

A/

m)

μ0 = 2,500

cm2/V/s

μ0 = 25,000cm2/V/s

(a) (b)cleaner

more disordered

Page 59: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

51

differential conductance (NDC) and a current drop at high-fields due to strong one-dimensional

(1D) phonon scattering as the device heats up. By contrast, the linearly increasing density of

states in 2D graphene leads to enhanced thermal carrier generation as the device heats up, which

prevents the appearance of NDC.

Next, in Fig. 4.5 we extract the carrier drift velocity from our high-field transport data in

Fig. 4.4a, near the physical device breakdown. As discussed previously, the carrier density for

our suspended devices has little to no gate dependence at high temperature reached at high fields,

due to device self-heating. The maximum carrier drift velocity at the breakdown (BD) point is vd

= IBD/(qWntot), where ntot = 2[(n*/ 2)2 + nth

2]1/2

is the carrier density [32] including residual dop-

ing (n* ~ 2×1011

cm−2

) and thermal carrier generation, nth. The latter dominates at the elevated

temperatures in the middle of the channel. We note that the temperature profile, and thus, the car-

Figure 4.5: Charge carrier high-field drift velocity at breakdown and high temperature

(Tavg ≈ 1200 K) along the suspended graphene channel. The dashed line is a simple esti-

mate of vsat in intrinsic graphene at such temperature and average carrier density (see

text). Exfoliated (red) and CVD (blue) graphene devices. Samples are ordered by increas-

ing (IBD/W)/FBD, representative of increasingly “cleaner” devices as shown in Fig. 4.4b.

Lower bounds, symbols, upper bounds correspond to models with breakdown tempera-

tures of 2860, 2230, and 1420 K respectively. The estimation for the breakdown tempera-

ture is described below with our thermal analysis. Some uncertainty also comes from im-

precise knowledge of the device width W, which is also discussed later in the text.

0 2 4 6 8 10 12 140

1

2

3

4

5

v (

10

7 c

m/s

)

Sample #

0 2 4 60

0.2

0.4

0.6

0.8

1

Count

(

Wm

-1K

-1)

0 2 4 60

0.2

0.4

0.6

0.8

1

Count

(

Wm

-1K

-1)

0 2 4 60

0.2

0.4

0.6

0.8

1

Count

(

Wm

-1K

-1)

0 2 4 60

0.2

0.4

0.6

0.8

1

Count

(

Wm

-1K

-1)2.7x107 cm/s

Page 60: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

52

rier density and drift velocity, varies strongly along the channel near the BD point. However, be-

cause we cannot precisely model this profile for every device measured, we instead estimate the

average drift velocity at breakdown, which is evaluated for an average carrier density along the

channel, ⟨ntot⟩ ≈ 4×1012

cm−2

and Tavg ≈ 1200 K, based on the thermal analysis discussed below.

Figure 4.5 displays the drift velocity at breakdown from all samples, arranged in increas-

ing order of (IBD/W)/FBD, which our simulations (Fig. 4.4b) suggest will rank them from most to

least disordered. These data represent the high-field velocity (i.e., near breakdown and approach-

ing the saturation velocity) in intrinsic graphene, as all measurements reached fields greater than

1 V/μm. We also point out that the transport regime observed here is not identical to our previous

work in Chapter 3 on SiO2-supported graphene. Here, the high-field measurement can only be

carried out in an ambipolar, thermally intrinsic regime (n ≈ p ≈ nth/2), whereas previous work for

substrate-supported devices focused on unipolar (e.g. n ≫ p) transport.

The maximum velocity values at breakdown (vBD) seen here for sample numbers 13−15

are very close to the saturation velocity predicted by the previously discussed (see Sections 2.2

and 3.5) simple model by Fang et al. [46], when transport is only limited by graphene optical

phonons (OPs) with energy ħωOP = 160 meV, that is, vsat ≈ 2.7×107 cm/s for the average charge

density and temperature estimated here (4×1012

cm−2

and 1200 K, respectively). However, we

point out that the analytic model is based on unipolar, not ambipolar transport; as such the value

for vsat given here should be seen more as a guide than a fixed fundamental value. Nevertheless,

similar saturation velocities have been predicted for clean, intrinsic graphene by extensive nu-

merical simulations at comparable fields and carrier density [44, 46, 139]. However, the average

high-field velocity at breakdown observed across our suspended samples is lower, vBD = (1.7 +

0.6/-0.3)×107 cm/s, similar for exfoliated and CVD-grown graphene, at the average carrier densi-

Page 61: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

53

ties and temperatures reached here. The average value remains a factor of two higher than the

saturation velocity at elevated temperature in silicon (vSi = 8×106 cm/s at ~500 K) [41], but we

suspect that variability between our samples is due to the presence of disorder and some impuri-

ties [45], which also affect the low-field mobility. In addition, depending on the level of strain

built-in to these suspended samples and how the strain evolves at high temperature, flexural pho-

nons [120, 138] may also play a role in limiting high-field transport. It is apparent that future

computational work remains needed to understand the details of high-field transport in graphene

under a wide variety of temperatures and conditions, including ambipolar versus unipolar

transport, impurities, and disorder.

4.4 Thermal Analysis

Now we discuss the thermal analysis of our suspended graphene devices during high-field

operation. While the breakdown temperature of graphene in air is relatively well-known as TBD,air

≈ 600 °C = 873 K (based on thermogravimetric analysis [140, 141] and oxidation studies [142]),

the breakdown temperature of graphene in probe station vacuum (10−5

Torr) was not well under-

stood before the start of this study. To estimate this, we compare similar devices taken up to

electrical breakdown in air and in vacuum conditions. In both cases, we can assume heat

transport is diffusive in our suspended devices at high temperature based on the subsequent dis-

cussion. Following Pop et al. [135], we can estimate the phonon mean free path at 1000 K as λ =

(2κ/)/(Gb-κ/L) ≈ 21 nm, where κ = 310 Wm-1

K-1

, Gb = 9.6 GWK-1

m-2

is the ballistic

conductance per cross-sectional area (at 1000 K), and L = 1 to 3 m is the range of our sample

lengths. Thus, we always have λ ≪ L and phonon transport is diffusive at these high

temperatures. Also indicative of diffusive phonon transport is the fact that we do not observe a

dependence of κ on sample size. Now we can write the heat diffusion equation as

Page 62: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

54

2

02

20

d T P gT T

dx LWt t (4.1)

where κ is the thermal conductivity and t = 0.34 nm is the thickness of graphene, g = 2.9 × 104

Wm−2

K−1

is the thermal conductance per unit area between graphene and air [124], and g ≈ 0 in

vacuum. Here the power dissipation is P = I(V − 2IRC) within the suspended graphene channel,

and the temperature of the contacts at x = ±L/2 is assumed constant, T0 ≈ 300 K (the small role of

thermal contact resistance is discussed below). Assuming that κ is a constant (average) along the

graphene channel, we now compare breakdowns in air and vacuum and estimate the graphene

device breakdown temperature.

In Fig. 4.6a-c we show examples of the electrical breakdown of suspended graphene de-

vices in vacuum (~10-5

Torr), air, and O2 environments. Device failure in vacuum occurs instan-

taneously corresponding to a sudden drop in current over a very narrow range of voltage. How-

ever, breakdown in air and O2 is a more gradual process where the current degrades over a rela-

tively wide voltage range. We expect that the very low O2 partial pressure in vacuum allows for

the suspended device to reach higher temperature without oxidation degrading the device. This is

consistent with previous studies of substrate-supported carbon nanotubes [143] and graphene na-

noribbons [144]. Conversely, in air and O2 oxidation may occur sporadically at lower tempera-

tures (< 1000 K) due to the much greater availability of O2. We also note that the breakdown lo-

cation observed by SEM (Figs. 4.6d-f) is in the center of the graphene channel, corresponding to

the position of maximum temperature predicted by our thermal model here.

To estimate the breakdown temperature of suspended graphene devices in vacuum, we

compare breakdowns in air and vacuum. In air, solving for T(x) from the heat diffusion equation

above results in

Page 63: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

55

0 2

cosh( )( ) 1

cosh / 2

P mxT x T

m LWt mL

(4.2)

such that the breakdown temperature is given by

, 0 2

BDBD air

PT T

m LWt

(4.3)

where m = (2g/κt)1/2

and ζ = 1 – 1/cosh(mL/2). In vacuum, heat loss due to convection and radia-

tion is negligible. With respect to heat loss due to radiation, the power loss is estimated by Prad =

σεA(T4– T0

4), where σ = 5.67×10

-8 Wm

-2K

-4 is the Stefan-Boltzmann constant, ε ~ 2.3 % is the

emissivity of graphene and assumed equal to the absorption from [145], A = 2 μm2 is an upper

limit for the area of the graphene channel, and T = 1300 K is an upper limit for average

temperature along the channel. We estimate Prad ≈ 7.4 nW, which is several orders of magnitude

less than the electric power that is dissipated. Consequently, we use g = 0 which results in

Figure 4.6: ID vs. V

D and corresponding SEM images, taken at a 70° tilt with respect to

the substrate, of suspended graphene broken in (a,d) vacuum, (b,e) air, and (c,f) O2.

Breakdown in vacuum is relatively sudden and less gradual than breakdown in air or O2.

Page 64: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

56

2

0

2( ) 1

8

PL xT x T

Wt L

(4.4)

and

, 08

BDBD vac

P LT T

Wt . (4.5)

Although electrical breakdown of graphene in air is often gradual and consisting of a series of

partial breaks, as shown above in Fig. 4.6b, we can carefully choose the breakdown points from

our measurements, and assuming κair ≈ κvac, we estimate TBD,vac ≈ 2230 K.

We acknowledge some uncertainty in assuming κair ≈ κvac, but note that the devices used

for this comparison underwent similar processing. We expect graphene in vacuum to be relative-

ly “clean”, especially after current annealing, and have a higher thermal conductivity than that of

graphene in air, but devices in vacuum operate at a higher average temperature, which would

cause a decrease in thermal conductivity. Quantitatively evaluating these competing effects is

difficult, particularly the cleanliness of a sample, thus we aim to provide upper and lower bounds

for our estimate of TBD,vac. We estimate a lower bound for TBD,vac by using g = 0 as a lower limit

for the heat transfer coefficient in air [124], and TBD,air ≈ 400 °C to account for the tendency of

partial breakdown in air. We estimate an upper bound for TBD,vac by using g = 105 Wm

-2K

-1 in air,

the theoretical upper limit based on kinetic theory [124], and the typical TBD,air ≈ 600 °C. Thus,

the extreme lower and upper bounds for TBD,vac are 1420 and 2860 K respectively. The lower

limit appears to be a conservative estimate since the breakdown power scaled with device dimen-

sions is typically about three times higher in vacuum than in air. The upper limit is comparable to

the 2800 °C (i.e., 3073 K) previously estimated as the breakdown temperature of suspended gra-

phitic nanoribbons under Joule heating [146]. Also, suspended CVD graphene has been reported

to be thermally stable up to at least 2600 K [147]. However, we note these previous studies were

Page 65: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

57

performed in a transmission electron microscope (TEM), which is capable of obtaining lower

vacuum levels than our probe station, accounting for samples likely reaching higher tempera-

tures.

Having estimated a range for TBD,vac, we turn to more detailed thermal modeling in order

to extract κ(T) from the electrical breakdown data. In general, we expect the thermal conductivity

decreases with increasing temperature above 300 K, consistent with the case of carbon nano-

tubes, graphite and diamond [135]. Therefore, we write κ = κ0(T0/T)γ above room temperature

and solve the one-dimensional heat diffusion equation, obtaining

1

2 1

1

0

0 0

(1 ) 2( ) 1

8

PL xT x T

T Wt L

(4.6)

where κ0 and γ are fitting parameters, κ0 being the thermal conductivity at T0 = 300 K. (A similar

analytic solution was previously proposed for suspended carbon nanotubes, albeit with a differ-

ent functional form of the thermal conductivity [148].) The breakdown temperature is maximum

in the middle of the suspended graphene, at x = 0

1

11

, 0

0 0

(1 )

8

BDBD vac

P LT T

T Wt

(4.7)

where for TBD,vac ≈ 2230 K we obtain γ ≈ 1.9 and γ ≈ 1.7 for our exfoliated and CVD-grown gra-

phene samples, respectively.

Figure 4.7 shows the extracted thermal conductivity of each sample at T = 1000 K for de-

vices measured in vacuum. The lower bounds, circles, and upper bounds are based on TBD,vac =

2860, 2230, and 1420 K, respectively, the widest range of breakdown temperatures in vacuum

estimated earlier. The average thermal conductivities at 1000 K of the exfoliated and CVD gra-

phene samples are similar, κ = (310 + 200/−100) W m−1

K−1

. Of this, the electronic contribution

Page 66: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

58

is expected to be <10%, based on a Wiedemann−Franz law estimate [144]. The result suggests

that lattice phonons are almost entirely responsible for heat conduction in graphene even at ele-

vated temperatures and under high current flow conditions. The average values of thermal con-

ductivity found here are slightly lower than those of good-quality highly oriented pyrolytic

graphite (534 W m−1

K−1

at T = 1000 K) [149], but the latter is consistent with the upper end of

our estimates. The extrapolation of our model to room temperature yields an average κ0 ≈ 2500

W m−1

K−1

at 300 K, as shown in Fig. 4.8 and consistent with previous work on freely suspended

graphene [121-125].

Figure 4.8 plots thermal conductivity as a function of temperature, showing that previous-

ly reported studies [121-125] (most near room temperature) fall on the same trend as this work at

high temperature. However, our model of high-temperature thermal conductivity of suspended

graphene suggests a steeper decrease (~T−1.7

weighed between exfoliated and CVD samples) than

that of graphite (~T −1.1

). The difference is likely due to the flexural phonons of isolated gra-

phene, which could enable stronger second-order three-phonon [150] scattering transitions (~T2

Figure 4.7: Thermal conductivity at ~1000 K for exfoliated (red) and CVD (blue) sus-

pended graphene devices. As in Fig. 4.5, samples are ordered by increasing (IBD/W)/FBD,

representative of increasingly “cleaner” devices. Lower bounds, symbols, and upper

bounds correspond to models with breakdown temperatures of 2860, 2230, and 1420 K

respectively (in 10-5

Torr vacuum).

0 2 4 6 8 10 12 140

200

400

600

800

1000

(

Wm

-1K

-1)

Sample #

0 2 4 60

0.2

0.4

0.6

0.8

1

Count

(

Wm

-1K

-1)

0 2 4 60

0.2

0.4

0.6

0.8

1

Count

(

Wm

-1K

-1)

0 2 4 60

0.2

0.4

0.6

0.8

1

Count

(

Wm

-1K

-1)

0 2 4 60

0.2

0.4

0.6

0.8

1

Count

(

Wm

-1K

-1)

0 2 4 60

0.2

0.4

0.6

0.8

1

Count

(

Wm

-1K

-1)

534 W/m/K

Page 67: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

59

scattering rate at high temperature) in addition to common first-order Umklapp phonon-phonon

transitions (~T scattering rate). Second-order three-phonon processes involve the virtual

combination of two phonons into a virtual state, and the splitting of the virtual phonon into two

new phonons. For this reason, this process is also sometimes referred to as four-phonon

scattering. Similar observations as shown here were also made for silicon [151], germanium

[152], and carbon nanotubes [153, 154] at high temperatures but not in isolated graphene until

now.

Now we return to our thermal modeling to discuss the issue of thermal contact resistance

Figure 4.8: Suspended graphene thermal conductivity above room temperature estimated

from this work (lines) and that of previous studies (symbols). Shaded regions represent

the average ranges of values for exfoliated (red) and CVD (blue) graphene from this

work. The weighted average thermal conductivity for our samples is ~2500 Wm-1

K-1

at

room temperature and ~310 Wm-1

K-1

at 1000 K, with a steeper drop-off than graphite, at-

tributed to second-order three-phonon scattering (see text).

Page 68: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

60

and estimate the temperature rise at the contacts [134] due to Joule heating during high-field cur-

rent flow. The thermal resistance for heat flow from the suspended graphene channel into the

metal contacts can be approximated by [155]

,

1coth C

C th

T C T

LR

hL W L

(4.8)

where h is the thermal interface conductance per unit area for heat flow from the graphene into

the SiO2 or Au, WC is the width of the graphene under the contact, and LC is the metal contact

length. The thermal transfer length LT = (κt/h)1/2

corresponds to the distance over which the tem-

perature drops by 1/e within the contact [134]. Typical contact lengths are on the order of mi-

crons while LT ≈ 50 nm; thus, we have LC ≫ LT and we can simplify RC,th ≈ (hLTWC)-1

(WC)-1

(hκt)-1/2

. Heat dissipation at the contacts consists of parallel paths to the underlying SiO2

substrate and the top metal contact, where hg-ox ≈ 108 Wm

-2K

-1 and hg-Au ≈ 4×10

7 Wm

-2K

-1 [156,

157], respectively. The total thermal resistance for one contact is given by RC,th = (RC,ox-1

+

RC,Au-1

)-1

. The temperature rise at the contacts is estimated as ΔTC = TC – T0 = RC,thPBD/2 (where

PBD is the input electrical power at the breakdown point), which is only 10s of Kelvin. Including

thermal contact resistance for the analysis when estimating TBD,vac above would change the ex-

tracted values by less than 4%.

4.5 Raman Spectroscopy of Electrically Biased Suspended Graphene Transistor

Prior to now, we have inferred the temperature of graphene devices primarily through

simulations of our device structures and comparison with electrical data. Therefore, for thor-

oughness and to supplement our thermal modeling, we briefly discuss experimental work where

we directly measured the temperature of functioning graphene devices using a thermometry

method based on Raman spectroscopy [83, 158]. Raman thermometry relies on the temperature-

Page 69: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

61

dependent shift of graphene 2D and G bands (see Fig. 4.9a), which downshift with increasing

temperature as a result of anharmonic phonon coupling [159]. This technique has been used in

the past to measure the thermal conductivity of suspended graphene devices [121, 123-125] and

power dissipation in SiO2-supported graphene devices [84, 85]. However, no studies presently

exist on the Raman thermometry of suspended graphene devices heated by electrical bias.

We have calibrated this temperature dependence by placing graphene devices in a tem-

perature-controlled stage and measuring the peak position of the 2D band as a function of tem-

Figure 4.9: (a) Raman spectrum of graphene at T = 323, 413, and 573 K (blue, red, and

green lines respectively) for a 633 nm laser line. (Inset) Experimental observation of the

2D peak position decreasing with increasing temperature. (b) Peak position of Raman 2D

band in graphene measured from 300 K to 573 K going up in temperature (blue circles)

and then going down in temperature (red squares). Assuming a linear relation (dashed),

we obtain χ2D ≈ -0.047 cm-1

/K. (c) Temperature increase (ΔT) in an electrically biased

suspended graphene device (L = 2 μm, W = 1 μm) as a function of input power. Maxi-

mum power of 1.8 mW corresponds to F = 2 V/μm. Error bars correspond to the tempera-

ture hysteresis observed in the calibration from (b). Contact resistance is accounted for as

before in Figs. 4.5 and 4.7.

0 0.5 1 1.5 20

100

200

300

400

500

P (mW)

T (

K)

2400 2500 2600 27000.2

0.4

0.6

0.8

1

Raman Shift (cm-1)

Inte

nsity (

arb

. units)

2,615 2,675 2,655

300 400 500 6002635

2640

2645

2650

2655

T (K)

2D

peak p

ositio

n (

cm

-1)

(b) (c)

2D

T = 323, 413, 573 K

(a)

Page 70: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

62

perature, as shown in Fig. 4.9b. For the temperature regime discussed here, we assume a linear

relation [159] between peak position and temperature such that χ2D ≈ -0.047 cm-1

/K. Figure 4.9c

shows the increasing temperature in an electrically biased suspended graphene device as the

power is increased. We note that the experimental setup for performing Raman spectroscopy

while operating the device requires ambient conditions. Therefore, in Fig. 4.10 we use Eq. (4.2)

to extract the thermal conductivity of the device from our Raman thermometry data. We observe

a slight decrease in thermal conductivity from κ ≈ 800 to 500 Wm-1

K-1

as the temperature in-

creases from ~300 K to 650 K. The extracted κ here is on the lower end when compared to the

data shown in Fig. 4.8 for suspended graphene, not surprisingly, considering these Raman ther-

mometry measurements were performed in ambient conditions and we were unable to implement

Figure 4.10: Thermal conductivity (κ) vs. temperature extracted using Eq. (4.2) and Ra-

man thermometry data of the electrically biased suspended graphene device from Fig.

4.9c. The shaded region represents the range of extracted κ where the upper

(χ2D = -0.054 cm-1

) and lower (χ2D = -0.040 cm-1

) dashed lines correspond to the tempera-

ture hysteresis in the calibration from Fig. 4.9b. Symbols correspond to χ2D = -0.040 cm-1

.

As before with the extraction of TBD, we let the thermal conductance per unit area be-

tween graphene and air vary such that g = 2.9×104 (symbols), 0 (upper), and 10

5 (lower)

Wm-2

K-1

[124].

Page 71: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

63

the standard practice of annealing in vacuum and performing subsequent measurements in vacu-

um, without exposure to air. Nevertheless, we are able to directly measure the temperature in-

crease in a device through self-heating effects and obtain reasonable values for graphene thermal

conductivity with the Raman thermometry technique discussed here.

4.6 Variability

We now comment more on the observed variability between samples and on the role of

graphene-metal contact resistance. First, even after high-temperature current annealing, polymer

residue from processing may remain on the samples, increasing the scattering of both charge and

heat carriers [160]. This sample-to-sample variation can contribute to the spread in extracted car-

rier velocity and thermal conductivity in Figs. 4.5 and 4.7, respectively. In this respect, it is likely

that our samples yielding higher carrier velocity and higher thermal conductivity (e.g., sample

no. 13 in Figs. 4.5 and 4.7) are the “cleanest” ones, most closely approaching the intrinsic limits

of transport in suspended graphene. Second, some edge damage occasionally seen after high-

current annealing affects our ability to accurately determine the sample width, W. The value used

for W influences all our calculations, and its uncertainty is incorporated in our error analysis

here. Third, several of the CVD graphene samples in the study had W < 200 nm, and edge scat-

tering effects are known to limit transport in narrow ribbons [144, 161]. However, we saw no

obvious dependence of thermal conductivity, carrier velocity, or breakdown current density

(IBD/W) on sample width when comparing “wide” and “narrow” devices. We thus expect that

variation among samples due to edge scattering is smaller than other sources of variation.

4.7 Suspended Graphene Nanoconstriction

We recently mentioned above that a source of variability may be edge damage after high-

current annealing. We see from the SEM image after device breakdown in vacuum (Fig. 4.6d)

Page 72: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

64

that the edges of the graphene channel are damaged (i.e., twisted or burned away) during the

breakdown process. Burning away of the edges may also occur before breakdown during the cur-

rent annealing process and result in the formation of a graphene nanoconstriction [162]. Here we

show the behavior of a suspended CVD graphene device (L = 1 μm) with a nanoconstriction

formed by excessive current annealing. We note that this device was not used for extracting data

in previous parts of this study.

Figure 4.11a displays the measured ID-VG of a nanoconstriction device at T = 80–300 K

for VD = 200 mV, showing high on/off > 103 at room temperature and >10

9 at T = 80 K. In Fig.

4.11b we vary the drain bias at T = 150 K to show that the effective band gap and on/off is di-

Figure 4.11: (a) Measured current vs. gate voltage at T = 80–300 K of a suspended CVD

graphene nanoconstriction formed by current annealing. (b) Electrical transport meas-

urements at T = 150 K under varying bias VD = 50–1000 mV. High on/off > 106 is ob-

served for VD ≤ 200 mV, while an increase in bias to VD = 1000 mV results in on/off <

10. (c) Temperature dependence of the minimum current at VD = 200 mV. Effective

bandgap of Eg ~ 0.35 eV is extracted assuming thermal activation, Imin ~ exp(-Eg/2kBT).

0.004 0.007 0.0110

-16

10-14

10-12

10-10

10-8

1/T (K-1)

I min

(A)

-10 -5 0 5 1010

-15

10-13

10-11

10-9

10-7

10-5

VG0

(V)

I D (

A)

-10 -5 0 5 1010

-13

10-11

10-9

10-7

10-5

VG0

(V)

I D (

A)

VD = 50 mV

1000 mV

500 mV

200 mV

T= 300 K

200 K

150 K

100 K

80 KVD = 200 mV T = 150 K

Eg ~ 0.35 eV

VG – V0 (V) VG – V0 (V)(a) (b)

(c)

Page 73: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

65

minished at high fields. We observe high on/off > 106 for VD ≤ 200 mV, but a low on/off < 10

when we increase the bias to VD = 1 V. At low bias and low temperature we observe discrete

conductance peaks (e.g., VD = 50 mV and T = 150 K in Fig. 4.11b). It has been suggested that the

regions of the graphene channel that connect the narrow nanoconstriction to the wider graphene

sheet may actually be confined longitudinally (i.e., along the length of the channel) and behave

as quantum dots in series [162]. Thus, the conductance peaks correspond to resonant tunneling

through the quantized energy levels of these quantum dots. In Fig. 411c we assume thermal acti-

vation, Imin ~ exp(-Eg/2kBT), to extract an effective band gap of Eg ~ 0.35 eV and corresponding

width of ~12 nm, where W = 2πħvF/Eg [104]. This width extraction may be an underestimate

since the aforementioned quantum dot regions may increase the effective band gap. Unfortunate-

ly, the device broke before we were able to image and measure the nanoconstriction width.

4.8 Conclusions

In summary, we fabricated suspended graphene devices and carefully analyzed their

high-field electrical and thermal transport. The electrical transport is entirely dominated by ther-

mally generated carriers at high temperatures (>1000 K), with little or no control from the sub-

strate “gate” underneath such devices. The maximum high-field breakdown velocity recorded is

>3×107 cm/s, consistent with theoretical predictions for intrinsic transport limited only by gra-

phene optical phonons. However, average extracted drift velocities are lower (although remain-

ing a factor of two greater than in silicon at these temperatures), due to sample-to-sample varia-

tion. We estimated the breakdown temperature of graphene in 10−5

Torr vacuum to be ~2230 K,

which combined with our models, yields an average thermal conductivity of ~310 W m−1

K−1

at

1000 K for both exfoliated and CVD-grown graphene. The models show thermal conductivity

dependence as ~T−1.7

above room temperature, a steeper drop-off than that of graphite, suggest-

Page 74: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

66

ing stronger effects of second-order three-phonon scattering. The Raman thermometry technique

implemented here provided a direct measurement of the temperature rise of a suspended gra-

phene device due to self-heating at high bias. Our study also highlights remaining unknowns that

require future efforts on electrical and thermal transport at high field and high temperature in

graphene.

Page 75: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

67

CHAPTER 5: VELOCITY SATURATION AND HIGH-FIELD NEGATIVE

DIFFERENTIAL CONDUCTANCE IN TWO-DIMENSIONAL

MOLYBDENUM DISULFIDE TRANSISTORS

5.1 Introduction

We continue to emphasize that understanding high-field electrical transport in nano-

materials is essential for many applications including transistors and photodetectors, especially

when we consider that the analysis of high-field electrical measurements can provide insight into

the high-energy band structure of a material. As discussed in Section 2.3, despite the increasing

amount of interest given to studying the electrical and optical properties of 2D materials [23,

163], little is known about high-field transport in 2D MoS2 [88, 89], particularly as a function of

temperature and carrier density. We also note that some properties of 2D MoS2 films are more

similar to Si than graphene, both in terms of low-field electron mobility [32, 164, 165] and the

presence of an energy band gap (~1.2 eV for bulk and few-layer [25]) which is desirable for both

digital transistor and photodetector applications [166]. Demonstrations of MoS2 for potential use

in future nanoscale electronics include the fabrication of integrated circuits based on bilayer

MoS2 transistors [167] as well as MoS2 transistors fabricated on flexible substrates [96, 97]. In

addition, the thinness of MoS2 (1 layer = 6.5 Å) could enable scaling transistors to sub-10 nm

dimensions [168].

A thermal analysis of MoS2 field-effect transistors (FETs) is important, specifically the

effect of self-heating on high-field operation, which is expected to play a significant role in ap-

plications intended on flexible substrates [96, 97]. Given that mobility and saturation current

densities in MoS2 are relatively low compared to graphene, it is possible others have assumed

that power densities in MoS2 FETs do not reach high enough values to cause significant heating.

However, our work here shows that Joule heating may play an important role in high-field

transport in MoS2.

Page 76: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

68

In this study we investigate MoS2 transistors from 80-500 K as a function of carrier den-

sity and, by comparison with theoretical models, we uncover new transport physics at high elec-

tric fields. We examine device channel lengths from L = 0.4-7 μm and thicknesses tch ≈ 0.7-6 nm.

At a combination of low temperature and high carrier density, we observe negative differential

conductance (NDC) at high field, i.e. the current reaches a peak and then decreases at higher

voltages. This behavior is repeatable across multiple devices and sweep conditions. Comparison

with our theoretical models suggests that NDC in 2D MoS2 is caused by a combination of self-

heating and complex band structure effects enabling transport through more than one conduction

band.

The schematic of a typical measured device is shown in Fig. 5.1a, along with an SEM

image in Fig. 5.1b. As described in Section 1.4, all devices here are based on MoS2 flakes me-

chanically exfoliated onto ~90 nm of SiO2. Metal contacts consist of 35 nm thick Au electrodes.

In order to remove residues potentially left over from processing and improve device behavior,

we perform a post-fabrication anneal in vacuum (~10-5

Torr) at 300 °C for 1 hour and then allow

devices to cool overnight. Figure 5.1c shows drain current (ID) versus gate voltage (VG) before

and after a vacuum anneal. Additional characterization by AFM is shown in Figs. 5.1d,e.

5.2 Low-Field Mobility

Before exploring high lateral field transport, we carefully establish the low-field behavior

of the devices of interest. Figure 5.2 shows the measured ID versus VG for a MoS2 FET in vacu-

um at various T0. As before with our studies of graphene, we use T0 to denote background tem-

perature, and T for the higher device temperature during high-field operation. In the electron-

accumulation regime (ID linear with VG and VG > VT, where VT is the threshold voltage estimated

as shown in Fig. 5.1c) we observe decreasing ID as T0 rises, consistent with stronger phonon scat-

Page 77: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

69

tering. This point is further emphasized by the temperature dependence of the electron field-

effect mobility (µFE) shown in Fig. 5.3. Accounting for contact resistance RC, µFE =

gm′L/(WCoxVD

′) where gm

′ = dID/dVGS in the accumulation regime, Cox ≈ 38 nF/cm

2 is the oxide

capacitance, VD′ = VD – 2IDRC, and VGS = VG – VS with the true source potential VS = IDRC. The

estimated RCW ≈ 2.6 kΩ⋅µm is based on device fitting with our simulations discussed below and

is comparable to previous results for MoS2 with Au contacts [168]. The error bars in Fig. 5.3

arise from uncertainty in RC, consistent with measurements on multiple devices. At 80 K the con-

tact resistance accounts for ~25% of the total device resistance, but this contribution decreases at

higher temperatures (e.g., ~10% and <5% at 300 and 500 K, respectively) and high fields. Addi-

tional discussion of contact resistance is provided in Section 5.3.

Figure 5.1: (a) Schematic of few-layer MoS2 transistor. (b) Scanning electron microsco-

py image of fabricated device with L = 1.5 μm, W = 1.3 μm. (c) Measured current ID ver-

sus gate voltage VG for a MoS2 device (L = 3 µm, W = 1.8 µm, VD = 200 mV) at room

temperature before (blue) and after (red) 300 °C vacuum anneal. The threshold voltage VT

is determined by the x-intercept of the linear part of the ID-VG curve. (d) AFM image of

exfoliated MoS2 device from (b). (e) Height profile corresponding to blue line in (d)

shows tch = 5.2 nm (~8 layers of MoS2). Note that the device of interest in Fig. 5.4 is ~6

layers thick (tch ≈ 4 nm) but otherwise has the same L = 1.5 µm and W = 1.3 µm dimen-

sions.

Page 78: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

70

The strong temperature dependence of mobility observed in Fig. 5.3, particularly above

~150 K, is caused by increased phonon scattering at higher temperatures. These observations are

consistent with studies of bulk MoS2 [164] and multilayer MoS2 FETs of thickness about an or-

der of magnitude greater than ours [22]. Fitting the extracted data against our Boltzmann

transport equation (BTE) model described in Section 5.5, we can estimate the contributions to

mobility from charged-impurity (μimp), acoustic phonon (μAC), and optical phonon (μOP) scatter-

ing mechanisms. For our device of interest (blue circles and line in Fig. 5.3), μOP (~ T-2.4

) appears

dominant near and above room temperature, although other samples may have a slightly different

temperature dependence ostensibly due to different impurity or defect densities. We note that μOP

includes analysis of both homopolar and polar optical phonons (HOP and POP). Below ~150 K

the mobility peaks and becomes limited by charged-impurity scattering (μimp ~ T) in this model

and others [22].

Figure 5.2: Measured current ID vs. gate voltage VG for a MoS2 device (L = 1.5

μm, W = 1.3 μm, tch ≈ 4 nm, VD = 0.2 V) at background temperatures T0 = 120 K

(blue, solid), 300 K (red, dashed), and 500 K (green, dotted). Data is shown for

both log (left) and linear (right) axes.

-30 -15 0 15 3010

-3

10-2

10-1

100

101

ID (A)

VG (V)

-30 -15 0 15 300

3

6

9

12

15

120 K

300 K

500 K

Page 79: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

71

The second set of experimental data (magenta squares) in Fig. 5.3 is from the device

shown in Fig. 5.1b,d. We have already mentioned that for the first device the mobility falls as

~T-2.4

at high temperatures due to scattering with optical phonons (OPs), and similar to data for

bulk MoS2 [164]. However, for the second device we see a slightly weaker temperature depend-

ence (~T-1.6

), similar to a previous result on monolayer MoS2 [170]. Sample-to-sample variation

is most likely the cause of this discrepancy, as varying crystal quality and processing residue

have previously been shown to affect transport in 2D materials [130]. Specifically for MoS2, de-

fects in the crystal (e.g., sulfur vacancies [169, 171]) may play a role in affecting charge

transport. Nonetheless, the stronger temperature dependence is consistent with measurements

Figure 5.3: Low-field electron mobility vs. temperature, experimental data (symbols)

and theoretical BTE model (lines). Data correspond to μFE from devices both with L =1.3

µm, W = 1.5 µm and with tch ≈ 4 nm (blue circles) and tch = 5.2 nm (magenta squares).

The solid line is the mobility model with contributions from impurity scattering (μimp),

OP scattering (μOP including both HOP and POP), and AC phonon scattering (µAC). Error

bars arise from uncertainty in our device contact resistance (the error for T ≥ 300 K is

smaller than the symbol size). Extracted data shows the mobility near and above room

temperature with a dependence of µ~T-2.4

(blue solid line) and µ~T-1.6

(magenta solid line)

for the two devices, respectively. Sample-to-sample defect and impurity variation is the

most likely cause of the difference in observed temperature dependence of mobility

[169].

100 200 300 400 50010

1

102

103

104

Temperature (K)

Mobilit

y (

cm

2 V

-1 s

-1)

μ ~ T-1.6experimental

extraction (µFE)

100 200 300 400 50010

1

102

103

104

T (K)

0 (

cm

2/(Vs

)

experiment

data2

data3

data4

data5

μOP ~ T-2.4

100 200 300 400 50010

1

102

103

104

Temperature (K)

Mobility (

cm

2 V

-1 s

-1)

data1

data2

data3

data4

data5

data6

data7

data8

data9

µAC ~ T-1.3

μimp ~ T

calculated total µ

Page 80: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

72

performed on bulk MoS2 crystals, and thus have been correlated with the intrinsic phonon-

limited properties of this material [164].

5.3 High-Field Negative Differential Conductance and Transistor I-V Model

We now turn to high-field measurements of our MoS2 devices. In Fig. 5.4a we show

Figure 5.4: Measurements and modeling of negative differential conductance (NDC) in

MoS2 transistors. (a) Measured ID/W vs. VD at ambient T0 of 80 to 500 K from top to bot-

tom and VGT = VG – VT ≈ 30 V (corresponding to electron density ~7×1012

cm-2

). Forward

and reverse ID-VD sweeps are plotted for all data sets (arrows indicate sweep direction),

showing repeatability with minimal hysteresis. (b) Transistor model (dashed lines) using

Eq. (5.3) shows good fit to the measured data [subset of data in (a) replotted as circles].

(c) Model (dashed lines) and measured data (circles) at T0 = 80 K for varying overdrive

VGT = 28, 23, 18 V. Inset shows the average temperature rise of the MoS2 channel due to

self-heating ΔT, calculated using Eq. (3) at VGT = 28 V.

0 2 4 6 80

50

100

150

200

250

300

VD (V)

ID/W

(A/m

)

0 2 4 6 80

50

100

150

200

250

300

VD (V)

ID/W

(A/m

)

0 2 4 6 80

50

100

150

200

250

300

VD (V)

ID/W

(A/m

)

(a) (b)

(c)

80 K

200 K

300 K

500 K

120 K160 K

400 K

VGT ≈ 30 V

T0 = 80 K

VGT = 28 V

23 V

18 V

0 2 4 6 80

50

100

T (

K)

VD (V )

VGT = 28 V

Page 81: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

73

measured ID/W vs. VD also has a strong dependence on T0 for the first device discussed in Fig.

5.3 (blue circles and lines). At T0 = 80 K we obtain peak ID/W = 280 µA/µm (at VD = 6 V for our

device L = 1.5 µm), while at 300 and 500 K the peak currents are 138 and 56 µA/µm (at VD = 9

V), respectively. Interestingly, below 200 K we observe NDC at high bias, i.e. the current in-

creases with drain voltage until it peaks, then decreases with further increasing voltage. The data

in Fig. 5.4a show highly repeatable, forward and reverse ID-VD sweeps (as indicated by arrows),

eliminating hysteresis as a possible cause of the observed NDC. We ensured no sample degrada-

tion due to high-field stress by checking that low-field ID-VG characteristics were reproducible

after each high bias measurement.

In order to understand the intriguing NDC feature, we first set up a simple transistor

model using a modified long-channel approximation [3]. Because the MoS2 body is only a few

layers thick (tch < 6 nm), we assume the entire body is in accumulation (i.e., electron-doped) for

large VGT. Note that we are only probing the accumulation regime in this work. This is the case

because we are interested in high-field transport of MoS2 when the channel is conductive and

highly populated with electrons. For the device of interest, we choose a minimum VGT ≈ 20 V

such that we operate in the accumulation regime and avoid pinch-off effects during high-field

measurements. Most data in Fig. 5.4, is shown at VGT ≈ 30 V, where NDC is more apparent due

to the high carrier density, n ≈ CoxVGT/q = 7×1012

cm-2

.

The electron density in the channel is calculated as ( ) ( ) Bq x V x k T

ch Dn t N e , where ND ≈

1016

cm-3

is the intrinsic doping of the MoS2 based on previous capacitance-voltage measure-

ments [22] for MoS2 flakes obtained from crystals from the same manufacturer (carrier density is

dominated by the gate-induced accumulation layer so the estimate for the intrinsic doping is not

important, i.e., VGT ≫ qtchND/Cox even for ND ≈ 1018

cm-3

), x is the direction along the channel

Page 82: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

74

with x = 0 and x = L being the positions of the source and drain edge, respectively, ψ(x) is the

amount of band bending relative to the flat band condition (VGS = VFB), and V(x) is the drain-

induced potential along the channel with V(x = L) – V(x = 0) = VD – 2ID·RC. Equivalently, the

electron density can be determined by n = Cox[VGS – VFB – ψ(x)]/q + toxND, where the oxide ca-

pacitance Cox = εox/tox ≈ 38 nF/cm2 and VGS = VG – VS with the source potential defined as VS =

IDRC. Combining the two equations for carrier density above allows us to calculate

( )

( ) ( ) ln 1ox GS FBB

ch D

C V V xk TV x x

q qt N

(5.1)

where the drain current is defined by

( )

D

S

D

W dVI q n x d

L d (5.2)

and dV/dψ is found from Eq. (5.1). We can now obtain

2 21

2

D ox GS T D S D S

WI C V V

L (5.3)

where ψD and ψS are the potential of the MoS2 channel at the drain and source, respectively, µ =

vd(F)/F is the mobility accounting for electron drift velocity effects at high electric fields F, and

the threshold voltage is defined as ch DB

T FB

ox

qt Nk TV V

q C

. We note that the mobility μ can

be taken out of the integral in Eq. (5.2) because we assume an effective mobility for average gate

and drain fields [3], such that vd(F) is described by

( )

1

FE sat C

d

C

F v F Fv F

F F

(5.4)

where µFE is the low-field mobility described above, vsat is the saturation velocity discussed be-

low, and γ = 2 and the critical field (FC) are fitting parameters (FC ranges from 3.2 to 20 V/µm

Page 83: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

75

from T0 = 80 to 500 K) that provide reasonable agreement with the experimental data. The func-

tional form of Eq. (5.4) is the same as that used for GaAs in [172], although here we use γ = 2

instead of γ = 4, which is representative of NDC in MoS2 being less severe than in GaAs. We

repeat that all measurements shown here are carried out in strong electron accumulation condi-

tions (VGT > VD > 0, where VGT = VG – VT is the gate voltage overdrive) such that we do not reach

pinch-off. Thus, the observed current saturation is associated with velocity saturation effects, not

channel depletion near the drain.

Returning to Eq. (5.3), we are able to determine ψD and ψS in our simulations by equating

( ) ( )( ) 1Bq x V x k T

ox GS FB ch DC V V x qt N e

with input parameters VG and VD and letting

VFB ≈ VT. We can assume VFB ≈ VT since VGT ≫ qtchND/Cox, as mentioned above, and the thermal

voltage kBT/q ≪ VGS (i.e., we are only concerned with VFB and VT relative to the gate voltage

such that VGT ≈ VGS –VFB). Lastly, we emphasize that an iterative, self-consistent approach is

used to incorporate contact resistance effects such that the potential drop across the channel, V(x

= L) – V(x = 0) = VD – 2ID·RC, and the source potential, VS = IDRC, are continuously updated for

each bias condition until a stable solution is found. This iterative, self-consistent approach also

applies to the thermal modeling and self-heating effects discussed later in Section 5.4. In Figs.

5.4b,c we show the results of the transistor model (dashed lines) that fit our measured data (sym-

bols).

The lumped transistor I-V model assumes a constant field in the channel, which is a valid

assumption if the carrier density does not vary significantly from source to drain. Typically for

high-field measurements the drain voltage is large enough to cause the carrier density near the

drain to significantly decrease (VGD ≪ VGT). However, as stated above VGT > VD for all our

measurements such that we do not reach pinch off. We further validate our constant field approx-

Page 84: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

76

imation by comparing the simple transistor model to a self-consistent, electro-thermal MoS2

simulator, which is based on a finite-element drift-diffusion (DD) approach adapted from our

previous simulator used to model graphene [87]. The DD simulation calculates the voltage, field,

electron density, and temperature at each point along the channel in the direction of current flow,

while assuming the device is uniform in the width direction perpendicular to current flow. For

the lumped transistor model, once we have already calculated ID we are also able to estimate

these properties along the channel by letting [3]

( ) .x

S

D

dVI x q W n x d

d

(5.5)

Figure 5.5 shows good agreement between the two models and validates our use of a constant

field in the transistor I-V model.

5.4 Thermal Modeling and Raman Thermometry

Next, we briefly discuss the thermal analysis of our MoS2 devices. The ID-VD simulations

described above were calculated self-consistently with a thermal model of the 2D devices. We

estimate the temperature profile using the heat diffusion equation [82, 130]

2

020

ch ch

d T gT T

dx Wt Wt

p

(5.6)

where T(x) is the temperature at position x along the channel, κ ≈ 52 Wm-1

K-1

is the in-plane

thermal conductivity of few-layer MoS2 [173], p′ = IDF(x) is the Joule heating rate per unit

length, and g ≈ 14.6 Wm-1

K-1

is the thermal conductance to the substrate per unit length (esti-

mated here using the thermal resistance model discussed in Section 3.4), at 300 K. The inset of

Fig. 5.4c shows an example of the calculated average temperature rise (ΔT) during high-field op-

eration such that ΔT > 100 K for VD > 7 in this case, at ambient T0 = 80 K. Also, a temperature

Page 85: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

77

Figure 5.5: Comparison between lumped transistor I-V model (blue, dashed) and DD

MoS2 simulator (red, solid). Device dimensions are identical to the device of interest (L =

1.5 μm, W = 1.3 μm, tch ≈ 4 nm) and calculations here correspond to T0 = 80 K and VGT =

28 V. (a) Simulated ID vs. VD shows good agreement. At VD ≈ 5 V, we plot (b) channel

potential V(x), (c) electric field F(x), (d) electron density n(x), and (e) temperature T(x).

The blue dotted lines in (c) and (e) correspond to the average value along the channel

calculated with the transistor model. Results show good agreement between both simula-

tions and validate our use of a constant field in the transistor I-V model.

-0.5 0 0.550

100

150

200

x (m)

T(x) (

K)

-0.5 0 0.50

2

4

6

8

x (m)

n(x) (

10

12 c

m-2)

-0.5 0 0.50

1

2

3

4

x (m)

V(x) (

V)

0 1 2 3 4 50

50

100

150

200

250

300

VD (V)

ID (A/m

)

-0.5 0 0.50

1

2

3

x (m)

F(x) (

V/m

)

-0.5 0 0.50

1

2

3

x (m)

Fx (

V/m

)

-0.5 0 0.50

2

4

6

8

x (m)

nx (

10

12 c

m-2)

0123450

50

100

150

200

250

300

VD (V)

ID/W

(A/

m)

data1

data2

lumped transistor model

DD simulator

(c)(b)

(a)

(e)(d)

Page 86: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

78

profile T(x) is shown in Fig. 5.5e for a device at VD 5 V. When evaluating Eq. (5.4) above we

include the temperature dependence of µFE, which is shown in Fig. 5.3 and discussed previously.

We also include the temperature dependence of saturation velocity as vsat = v0/(1 + NOP) where v0

= 2.2×106 cm/s and NOP = 1/[exp(ħωOP/kBT) – 1] is the OP occupation with ħωOP = 48 meV

[174]. The factor of 1/(1 + NOP) accounts for the decrease in vsat at higher temperatures due to

increased OP scattering, similar to previous models for graphene [32] and silicon [41].

Our study highlights the current dearth of knowledge with respect to MoS2 thermal prop-

erties, especially in-plane thermal conductivity κ for few-layer samples. Although in our work

we have used the estimate for MoS2 thermal conductivity from [173], we note it is obtained from

Raman studies of suspended few-layer MoS2 prepared using a high-temperature vapor-phase

method. However, our transistor simulations are not very sensitive to the in-plane κ of MoS2, be-

cause most heat flows into the substrate for devices of dimensions employed in this work. This is

also true for graphene devices with dimensions >0.3 μm [87] and even more applicable to MoS2

which has a significantly lower thermal conductivity than graphene. Additionally, we could not

find measurements in the literature for MoS2-SiO2 thermal boundary resistance (RB), hence we

used values determined for the graphene-SiO2 thermal coupling [112, 156].

Fortuitously, despite the lack of knowledge with respect to MoS2 thermal properties, we

find that at all temperatures of interest ≥ 90% of the thermal conductance from the device into

the substrate, g ≈ 1/[L(RB + Rox + RSi)], is determined by the ~90 nm SiO2 thermal resistance.

[Rox = tox/(Aκox) where A = L⋅W is the surface area of the channel and κox is the thermal conduc-

tivity of SiO2] and the Si substrate thermal resistance (RSi) [32]. For example, at 300 K we find

that RB ≈ 5×103 K/W, Rox ≈ 4×10

4 K/W, and RSi ≈ 4×10

3 K/W, here for tox = 90 nm, and device

area A = L⋅W = 1.95 µm2 where L = 1.5 µm and W = 1.3 µm. Therefore, our thermal analysis

Page 87: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

79

provides a good estimate for the temperature rise in the devices during operation. As described in

Section 3.4, we also account for the temperature dependence of the thermal resistance through

κox and κSi, where κSi is the thermal conductivity of Si. For example, g ≈ 14.6 Wm-1

K-1

at 300 K

but decreases to g ≈ 7.9 Wm-1

K-1

at 80 K (at 80 K, g is 94% dominated by the underlying SiO2

and Si thermal resistance). We note that the thermal resistance at the MoS2-Au contacts, although

not playing a significant role here, is accounted for by following the approach described in Sec-

tion 4.4.

Before concluding our thermal analysis, we briefly describe a recent attempt to directly

measure the temperature rise in a MoS2 transistor due to self-heating at high-fields by employing

the same Raman thermometry technique described in Section 4.5. The only difference here is

that we use a 488 nm laser line instead of a 633 nm laser line and that we are able to make elec-

trical measurements to the device while it is in a temperature-controlled stage at 5×10-2

Torr.

Figure 5.6a shows the temperature calibration of the A1g peak, resulting in χA1g = -4.9×10-3

cm-1

/K if we assume a linear relation [173]. Figure 5.6b shows the estimated temperature rise

(ΔT) in the device during electrical measurements at T0 = 80 K. We compared the extracted ΔT

from Raman thermometry (circles) to the calculated ΔT using our simple thermal resistance net-

work (dashed lines). The calculations underestimate the extracted values when we use the same

boundary thermal conductance value as for graphene-SiO2, i.e., h = 108 Wm

-2K

-1, but show

agreement when we let h = 107 Wm

-2K

-1. Lastly, we comment that the temperature sensitivity of

the A1g Raman peak position here is relatively small (an order of magnitude less sensitive than

that of the 2D Raman peak position in graphene) so further investigation is needed to clarify if

the peak shifts during device operation are solely due to self-heating, and thus, whether or not

they are accurate estimates of the lattice temperature rise in the MoS2 channel.

Page 88: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

80

5.5 Boltzmann Transport Equation Model

The temperature rise due to self-heating and its effect on electron transport help explain

our observed NDC behavior. Similar effects have been observed in suspended carbon nanotubes,

where NDC was attributed to strong self-heating and scattering of electrons by non-equilibrium

OPs [126]. However, it is apparent that self-heating is a necessary but not sufficient condition for

the measurable NDC in MoS2. Equation (5.4) and BTE simulations below suggests that a nega-

tive differential velocity is also needed in order to fully reproduce the experimental data. To un-

derstand the physical cause of such drift velocity behavior, we perform numerical simulations

using the BTE, summarized in Fig. 5.7. We calculate drift velocity as a function of field includ-

ing self-heating, the scattering mechanisms described below, and a simplified two-valley band

structure. We use the shifted Fermi-Dirac distribution function while balancing momentum and

energy gained from the field with that released by scattering [175]:

Figure 5.6: (a) Experimental observation and calibration of the MoS2 A1g peak position

decreasing with increasing temperature, χA1g ≈ -4.9×10-3

cm-1

/K assuming a linear rela-

tion (dashed line). (b) Temperature increase (ΔT) from Raman thermometry (circles) vs.

power (P) for a device (L = 4 µm, W = 2 µm, tch ≈ 4 nm) electrically biased at T0 = 80 K.

Comparison with calculated ΔT using the simple thermal resistance network model

(dashed lines) indicates a better fit with h = 107 Wm

-2K

-1 (green) than with h = 10

8

Wm-2

K-1

(red), where h is the boundary thermal conductance value for the MoS2-SiO2 in-

terface. Due to the low sensitivity of the peak position shift (i.e., small χA1g), future work

is needed to verify the accuracy of the extracted ΔT from Raman thermometry.

0 2 4 60

50

100

150

P (mW)

T (

K)

0 200 400 600402

403

404

405

406

T (K)

A1g P

eak P

osit

ion (

cm

-1)

T0 = 80 K

(b)(a)

Page 89: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

81

1

exp 1k x d F

B e

E k v Ef

k T

(5.7)

( ) n fq S k

k

F k (5.8)

· ( ) E Sqn fd k k

k

F v (5.9)

where Ek is the electron energy in the parabolic approximation, kx the quasi-momentum in the

direction of current flow, EF the Fermi level, Te the electron temperature, q the elementary

charge, n the carrier density, k the full electron quasi-momentum, and S(fk) the scattering inte-

gral. S(fk) includes impurity scattering calculated with Thomas-Fermi screening theory [176], AC

phonons with deformation potentials 4.5 eV for the lower valley (LV) and 2.8 eV for the upper

valley (UV) [91], HOPs with deformation potentials 8 eV/Å for intra-valley scattering in the LV

and 14 eV/Å for intra-valley and inter-valley scattering in the UV, and POP scattering assuming

Frölich deformation potential with dielectric constants ε0 = 7.6 and ε∞ = 7.3 [22]. Both HOP and

POP energies are ħωOP ≈ 48 meV, as in the transistor model described earlier. The effective mass

is assumed to be isotropic, mLV = 0.45m0 in the 2-fold degenerate LV and mUV = 0.7m0 in the 6-

fold degenerate UV (m0 is the bare electron mass), which is similar to theoretical first-principle

calculations [90, 91]. The energy difference between valleys is a fitting parameter in our calcula-

tions (ΔE = EUV – ELV ≈ 130 meV) and reasonable agreement could be obtained with slightly

different ΔE, given a corresponding change in effective mass.

Figure 5.7a displays the calculated dependence of vd on electric field (lines), comparing it

to values extracted from the experimental data in Fig. 5.4 (symbols). The role of the UV is eval-

uated in Fig. 5.7b, where we see that both valleys and self-heating are needed to obtain notice-

Page 90: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

82

ble negative differential velocity at T0 = 80 K. We do not observe negative differential velocity

for calculations without self-heating and calculations with only one valley. We have used a range

of phonon deformation potentials and different relative positions of the UV and LV in our simu-

lation (ΔE), but we were unable to obtain a reasonable agreement with the experimental transport

data without a 6-fold degenerate UV (with stronger scattering) above a 2-fold degenerate LV.

Figure 5.7: Effect of MoS2 band structure on high-field transport. (a) Electron drift ve-

locity (vd) vs. field F for ambient T0 = 80 to 500 K from top to bottom. Symbols are fit to

ID-VDS measurements using Eq. (5.4) and accounting for self-heating by Eq. (5.6). Dotted

lines correspond to Eq. (5.4) assuming no self-heating (T = T0). Solid lines are numerical

simulations based on our two-valley BTE model with self-heating. (b) In-depth look at vd

vs. F at T0 = 80 K with BTE simulations, separating the role of self-heating vs. that of

one-valley and two-valley conduction. Calculations with only one valley or without self-

heating do not show negative differential velocity. (c) (Top) Two-valley band structure

leading to the observed NDC and (bottom) calculated distribution function for electrons

at F = 3 V/µm and T0 = 80 K for (left) lower valley and (right) upper valley (white circle

indicates equilibrium Fermi surface as reference).

Page 91: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

83

Thus, the lower and upper valleys of our model appear to correspond to the conduction band val-

leys in MoS2 at the K and Q points, respectively [31, 91]. (The Q point is located approximately

halfway along the Γ-K axis.) Although this relative valley positioning has been previously ob-

tained in some theoretical work [93, 94], it is in contrast with other theoretical estimates [90, 92]

predicting the 6-fold degenerate Q valley being lower for few-layer MoS2. Figure 5.7c shows the

representative two-valley band structure of our model as well as the distribution function for

electrons in each valley, where in this case F = 3 V/µm and T0 = 80 K.

5.6 Contact Resistance and High-Field NDC

Before concluding our study, we discuss in more detail the role on contact resistance in

our devices and provide additional data on high-field NDC in MoS2. Figure 5.8a shows the

measured low-field ID versus VD in the accumulation regime (VGT ≈ 20 V) for the device of inter-

est from above. At VD below a few hundred millivolts we see nonlinear behavior such that the

resistance of the device is decreasing with increasing VD. This nonlinear behavior is more pro-

nounced below room temperature. We attribute the nonlinearity to small Schottky barriers at the

metal-MoS2 interface of the Au contacts. Assuming the contact resistance (RC) at large VGT is

dominated by thermally assisted tunneling (TAT) [177], increasing VD results in increased band

bending at the contacts, and thus, increased TAT. Below room temperature higher VD is needed

to increase the band bending and tunneling current until ID-VD appears linear. Above room tem-

perature the TAT is large enough to result in linear low-field ID-VD behavior and Ohmic contact

behavior. However, low-temperature measurements confirm the existence of a Schottky barrier

for Au contacts to MoS2, which is expected for n-type transport and high work function metals

[178]. Also, it is appropriate to point out that when extracting mobility for Fig. 5.3, we use VD ≤

1, i.e., sufficiently large VD for measurements below room temperature such that ID-VD is in the

Page 92: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

84

linear regime and the assumption of a constant contact resistance is more appropriate.

We note that although the TAT current increases with temperature, the resistance of the

MoS2 underneath the contact also increases with temperature due to a decreasing mobility. As

the temperature increases, the former mechanism causes a decrease while the latter causes an

increase in contact resistance, RC. The simple transistor model used in this work assumes the

temperature dependencies of these competing effects cancel out and lead to an approximately

constant RC, relatively independent of temperature which is in agreement with other recent ob-

servations [179]. Ongoing work in our group [180] is examining in more depth the different con-

tributions to RC and their dependence on temperature and metal work function.

With respect to our high-field measurements, contacts are not playing a role in causing

the observed NDC for several reasons. First, for all measured device channel lengths (L = 0.4-7

μm) the channel resistance at 80 K is two to more than 10 times larger than RC and dominates the

Figure 5.8: (a) Low-field ID vs. VD (VGT ≈ 20 V) for a MoS2 transistor (L = 1.5 μm, W =

1.3 μm, tch ≈ 4 nm) at background temperatures T0 = 120 (blue, solid), 300 (red, dashed),

and 500 K (green, dotted). A Schottky barrier-like contact resistance is observed at low

VD below room temperature. (b) Resistance at position of maximum drain current scaled

with device width, RImaxW, vs. channel length L for all devices measured in this work.

The dependence of RImaxW on L indicates the device resistance is dominated by the MoS2

channel, not the contacts, at high fields. Dashed line shows linear fit where the device

with L = 7 µm has relatively low mobility and is treated as an outlier.

0 2 4 6 80

50

100

150

200

250

L (m)

RpeakW

(k

-m

)

-0.4 -0.2 0 0.2 0.4-20

-10

0

10

20

VD (V)

ID (A)

120 K

VGT ≈ 20 V

300 K

500 K

024680

50

100

150

200

250

L (m)

RpeakW

(k

-m

)

data1

data2

VGT ≈ 30 V

VGT ≈ 25 V

(b)(a)

Page 93: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

85

total device resistance (see Fig. 5.7b). Second, as discussed above, we believe the resistance at

the metal-MoS2 interface to be determined by TAT current, and thus we expect more band bend-

ing at the contacts at higher fields to further decrease RC. Third, RC has been found to decrease

with increasing VGT [179, 180] (which renders the contacts more conductive) but our high-field

ID-VD measurements show NDC behavior (and thus a more resistive channel) only at high VGT.

Also, the electric field value where NDC starts (i.e., position of peak current) decreases with in-

creasing VGT (see Fig. 5.4c). Lastly, a plot of device resistances (adjusted for width) at the maxi-

mum current position (i.e., current value at the onset of NDC) shows a clear length dependence

in Fig. 5.8b. If contact resistance were dominant at high fields then we would expect little length

dependence.

Figure 5.9 shows all devices measured at high fields and T0 = 80 K ambient. Under these

conditions, all electrical data showed repeatable NDC. As in Fig. 5.4, we show forward and re-

verse ID-VD sweeps in order to highlight that hysteresis is negligible for most devices. Despite

some sample-to-sample variation, current densities at high drain bias (and high VGT) are compa-

rable among our devices, with all FETs showing ID > 200 µA/µm and a maximum measured val-

ue of ID = 291 µA/µm (see Fig. 5.9d). The devices corresponding to Figs. 5.9h and 5.9i show rel-

atively less pronounced NDC, which may be due to these samples having relatively larger chan-

nel areas (L×W). The larger area leads to improved heat dissipation to the substrate, which low-

ers the self-heating effect contribution to NDC.

We note that Fig. 5.9d is a monolayer MoS2 device with Ni contacts, showing high-field

NDC similar to that of the few-layer MoS2 devices with Au contacts. Observing comparable

NDC behavior for different contact metals provides additional support to our claim above that

contacts do not play a role in causing NDC. A monolayer sample showing NDC could indicate

Page 94: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

86

that the dominant mechanism causing NDC may be self-heating, which has little dependence on

channel thickness here. Another possibility is the structure of the conduction band for monolayer

MoS2 being fairly similar to that of few-layer MoS2. As stated above, more theoretical work may

be needed to settle the high-energy band structure of single- to few-layer MoS2.

Figure 5.9: (a-i) Measured current vs. drain voltage VD at T0 = 80 K for nine different

MoS2 devices, ordered by increasing channel length, L from 0.4 to 7 μm. All plots corre-

spond to few-layer samples with Au contacts except for (d), which corresponds to a mon-

olayer sample with Ni contacts. The gate overdrive voltages VGT correspond to the meas-

urements from top to bottom, respectively. Both forward and reverse ID-VD sweeps are

plotted (arrows indicate sweep direction) but most devices show little to no hysteresis.

We observe NDC in all devices, especially for large VGT.

0 5 10 15 20 250

50

100

150

200

250

VD (V)

ID (A/m

)

0 2 4 6 8 100

50

100

150

200

250

VD (V)

ID (A/m

)

0 10 20 30 400

50

100

150

200

250

VD (V)

ID (A/m

)

0 2 40

50

100

150

200

250

VD (V)

ID (A/m

)

0 2 4 60

50

100

150

200

250

300

VD (V)

ID (A/m

)

0 2 4 6 8 100

50

100

150

200

250

300

VD (V)

ID (A/m

)

0 2 4 6 8 100

50

100

150

200

250

300

VD (V)

ID (A/m

)

0 2 4 6 80

50

100

150

200

250

300

VD (V)

ID (A/m

)

0 2 4 60

50

100

150

200

250

300

VD (V)

ID (A/m

)

(c)(b)(a)

(f)(e)(d)

(i)(h)(g)

VGT = 20 V

15 V

10 V

VGT = 20 V

15 V

10 V

VGT = 20 V

15 V

10 V

VGT = 30 V

25 V

20 V

VGT = 25 V

20 V

15 V

10 V

VGT = 25 V

20 V

15 V

10 V

VGT = 30 V

25 V

20 V

15 V

10 V

VGT = 25 V

20 V

15 V

10 V

VGT = 35 V

30 V

monolayer

Ni contacts

0.4

1.4

L

W

0.6

1.4

L

W

0.9

1.4

L

W

1.0

2.8

L

W

2.0

1.0

L

W

1.5

1.35

L

W

6.0

3.5

L

W

2.0

1.35

L

W

7.0

2.0

L

W

Page 95: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

87

5.7 Conclusions

In conclusion, we experimentally and theoretically analyzed transport in MoS2 over a

wide range of temperatures (80-500 K) and up to high electric fields (6 V/μm). Low-field mobili-

ty shows a strong temperature dependence above room temperature due to scattering with OPs.

High-field measurements reveal NDC below room temperature and usual velocity saturation at

higher temperatures. We uncover that the NDC is due to a combination of self-heating and high-

energy band structure effects in 2D MoS2. Importantly, our results suggest that both the lower

(K) and upper (Q) valleys play important roles in all common transport regimes, and must be

carefully considered in MoS2 transport problems. While our two-valley model is sufficient to ex-

plain the experimental transport data in MoS2, future work remains needed in order to understand

the high-energy band structure of other similar 2D materials.

Page 96: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

88

CHAPTER 6: CONCLUSIONS AND FUTURE WORK

We have provided detailed studies of high-field transport in two-dimensional graphene

and MoS2 and investigated the energy dissipation mechanisms that limit transport, and ultimate-

ly, the current carrying ability of these materials. Substrate effects play a role in limiting

transport in supported-graphene devices, and thus, choosing optimal substrates for ideal interfac-

es is an essential task for the future development of 2D nanoelectronics. Our study of suspended

graphene allowed us to glimpse the intrinsic properties of graphene, but sample-to-sample varia-

bility associated with the challenging fabrication process complicated our analysis. Nevertheless,

the “extreme” case of suspended graphene allowed us to observe coupled electrical and thermal

transport at high-fields up to device breakdown.

The challenge posed by the absence of a band gap in graphene for future use in nanoelec-

tronics, especially digital electronics, is great. Although it was not discussed here, we briefly

comment on it now given that much of the motivation for studying MoS2 relative to graphene is

because MoS2 has a band gap. Small band gaps can appear in graphene nanoribbons (GNRs) due

to quantization of the width direction, but transport properties degrade due to edge roughness

[181]. Analogously, Si bodies with thicknesses down to 2 nm have been reported [182] but rough

interfaces cause their mobility to deteriorate and thickness fluctuations lead to unacceptable

threshold-voltage variations.

Moving on to our study of electrical transport in 2D MoS2 transistors, we find that few-

layer MoS2 has a mobility two orders of magnitude lower than that of graphene and its saturation

velocity is almost an order of magnitude less than that of bulk Si. With respect to ultrathin Si, the

transport properties of MoS2 may be more comparable, but we emphasize that if future develop-

ment of large-area growth techniques does not produce atomically-thin MoS2 (or other TMD)

Page 97: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

89

films with electronic properties better than those (at the present time) of exfoliated MoS2 flakes,

then TMDs like MoS2 may only prove suitable for applications where device performance is not

the top priority. One example of this is flexible electronics, where functionality may be sacrificed

to achieve arbitrary form factors and make electronic devices available to a wider range of envi-

ronments.

The observed phenomena of high-field negative differential conductance in few-layer

MoS2 may not be directly practical, considering the NDC is slight and only seen below room

temperature. However, the analysis of our high-field electrical data allowed us to probe the high-

energy band structure, revealing discrepancies among reported works and an unsettled picture of

the MoS2 band structure. A firm understanding of a material’s band structure is fundamental to

semiconductor technology research, and as such, we hope our work motivates other researchers

to investigate this matter further.

The issue of high-energy band structure becomes more important when we look into ul-

tra-short channel devices and the possibility of quasi-ballistic transport, where detailed

knowledge of the band structure is essential to properly understand device operation. If we take

the time to further consider ultimate-scaled transistors and fundamental limits to device perfor-

mance, research into 2D nanoelectronics must address the issue of large contact resistances. Alt-

hough contact resistance was accounted for in all of our analysis here, we did not go into detail

about the challenges of contacting a three-dimensional (3D) metal to a 2D transistor body. This

may prove to be the most challenging problem when trying to scale down 2D transistors, and be

the limiting factor in terms of current drive. Similar problems exist for designing contacts to ul-

trathin silicon [183].

To provide one last point, any near-future successful applications taking advantage of

Page 98: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

90

graphene or MoS2 (or any other 2D material) will most likely exist because the application is

novel, not just the material. Most of the work here (and in the 2D community in general) has

been conducted with respect to how we traditionally study and investigate semiconductor tech-

nology. Replacing entrenched technology is difficult, and at times, not plausible. However, we

hope that our specific work of improving the understanding of transport properties in these mate-

rials will not only advance us along the path to applicable traditional devices, but facilitate the

development of unforeseen 2D-based structures. Scaling down the transistor has propelled semi-

conductor technology for decades, and as the continuation of scaling becomes increasingly more

challenging and uncertain, many look to find new ways to achieve the same performance bene-

fits offered by scaling (either by using a new material or a new device structure). However, find-

ing innovative and original uses for IC technology, not just ways to improve existing IC technol-

ogy, may prove equally necessary for the continued growth of the semiconductor industry.

Page 99: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

91

REFERENCES

[1] G. E. Moore, in Electronics Magazine (McGraw-Hill, 1965), pp. 114.

[2] F. Schwierz, Nature Nanotechnology 5, 487, (2010).

[3] Y. Taur and T. H. Ning, Fundamentals of Modern VLSI Devices (Cambridge University

Press, New York, 2009), 2nd

edn., pp. 150-221.

[4] E. Pop, S. Sinha, and K. E. Goodson, Proceedings of the IEEE 94, 1587, (2006).

[5] E. Pop, Nano Research 3, 147, (2010).

[6] EPA Report on Server and Data Center Energy Efficiency. Available:

http://www.energystar.gov/index.cfm?c=prod_development.server_efficiency_study,

(2007).

[7] H. Xuejue et al., in International Electron Devices Meeting (Washington, D.C., 1999),

pp. 67.

[8] R. Chau, B. Doyle, J. Kavalieros, D. Barlage, A. Murthy, M. Doczy, R. Arghavani, and

S. Datta, in International Conference on Solid State Devices and Materials (Nagoya,

Japan, 2002), pp. 68.

[9] S. Damaraju, G. Varghese, S. Jahagirdar, T. Khondker, R. Milstrey, S. Sarkar, S. Siers, I.

Stolero, and A. Subbiah, in IEEE International Solid-State Circuits Conference Digest of

Technical Papers (San Francisco, CA, 2012), pp. 56.

[10] Y.-K. Choi, K. Asano, N. Lindert, V. Subramanian, K. Tsu-Jae, J. Bokor, and H.

Chenming, IEEE Electron Device Letters 21, 254, (2000).

[11] D. J. Frank, T. Yuan, and H. S. P. Wong, IEEE Electron Device Letters 19, 385, (1998).

[12] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim,

Reviews of Modern Physics 81, 109, (2009).

Page 100: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

92

[13] S. Reich, J. Maultzsch, C. Thomsen, and P. Ordejón, Physical Review B 66, 035412,

(2002).

[14] K. R. Knox, A. Locatelli, M. B. Yilmaz, D. Cvetko, T. O. Menteş, M. Á. Niño, P. Kim,

A. Morgante, and R. M. Osgood, Physical Review B 84, 115401, (2011).

[15] H. Yanagisawa, T. Tanaka, Y. Ishida, M. Matsue, E. Rokuta, S. Otani, and C. Oshima,

Surface and Interface Analysis 37, 133, (2005).

[16] C. Oshima, T. Aizawa, R. Souda, Y. Ishizawa, and Y. Sumiyoshi, Solid State

Communications 65, 1601, (1988).

[17] S. V. Morozov, K. S. Novoselov, M. I. Katsnelson, F. Schedin, D. C. Elias, J. A.

Jaszczak, and A. K. Geim, Physical Review Letters 100, 016602, (2008).

[18] Y. W. Tan, Y. Zhang, K. Bolotin, Y. Zhao, S. Adam, E. H. Hwang, S. Das Sarma, H. L.

Stormer, and P. Kim, Physical Review Letters 99, 246803, (2007).

[19] A. A. Balandin, Nature Materials 10, 569, (2011).

[20] C. Lee, X. Wei, J. W. Kysar, and J. Hone, Science 321, 385, (2008).

[21] B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, Nature

Nanotechnology 6, 147, (2011).

[22] S. Kim et al., Nature Communications 3, 1011, (2012).

[23] Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, Nature

Nanotechnology 7, 699, (2012).

[24] K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich, S. V. Morozov,

and A. K. Geim, Proceedings of the National Academy of Sciences of the United States

of America 102, 10451, (2005).

[25] K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, Physical Review Letters 105,

Page 101: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

93

136805, (2010).

[26] A. Kuc, N. Zibouche, and T. Heine, Physical Review B 83, 245213, (2011).

[27] K. K. Kam and B. A. Parkinson, Journal of Physical Chemistry 86, 463, (1982).

[28] J.-H. Choi, C. H. Park, and H.-S. Min, IEEE Electron Device Letters 16, 527, (1995).

[29] W. Bao, X. Cai, D. Kim, K. Sridhara, and M. S. Fuhrer, Applied Physics Letters 102,

042104, (2013).

[30] H. Fang, S. Chuang, T. C. Chang, K. Takei, T. Takahashi, and A. Javey, Nano Letters 12,

3788, (2012).

[31] E. S. Kadantsev and P. Hawrylak, Solid State Communications 152, 909, (2012).

[32] V. E. Dorgan, M.-H. Bae, and E. Pop, Applied Physics Letters 97, 082112, (2010).

[33] H. Li, Q. Zhang, C. C. R. Yap, B. K. Tay, T. H. T. Edwin, A. Olivier, and D. Baillargeat,

Advanced Functional Materials 22, 1385, (2012).

[34] X. Li et al., Science 324, 1312, (2009).

[35] Y. Huang et al., Nano Research 6, 200, (2013).

[36] K. I. Bolotin, K. J. Sikes, Z. Jiang, M. Klima, G. Fudenberg, J. Hone, P. Kim, and H. L.

Stormer, Solid State Communications 146, 351, (2008).

[37] R. S. Muller, T. I. Kamins, and M. Chan, Device Electronics for Integrated Circuits

(John Wiley & Sons, New York, 2003), 3rd

edn., pp. 429-461.

[38] Y. Taur, C. H. Hsu, B. Wu, R. Kiehl, B. Davari, and G. Shahidi, Solid-State Electronics

36, 1085, (1993).

[39] E. Pop, R. W. Dutton, and K. E. Goodson, Applied Physics Letters 86, 082101, (2005).

[40] E. Pop and K. E. Goodson, Journal of Electronic Packaging 128, 102, (2006).

[41] C. Jacoboni, C. Canali, G. Ottaviani, and A. Alberigi Quaranta, Solid-State Electronics

Page 102: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

94

20, 77, (1977).

[42] M. V. Fischetti, J. Kim, S. Narayanan, Z. Y. Ong, C. Sachs, D. K. Ferry, and S. J. Aboud,

Journal of Physics: Condensed Matter 25, 473202, (2013).

[43] A. Akturk and N. Goldsman, Journal of Applied Physics 103, 053702, (2008).

[44] R. S. Shishir and D. K. Ferry, Journal of Physics: Condensed Matter 21, 344201, (2009).

[45] J. Chauhan and J. Guo, Applied Physics Letters 95, 023120, (2009).

[46] T. Fang, A. Konar, H. Xing, and D. Jena, Physical Review B 84, 125450, (2011).

[47] D. K. Ferry, in IEEE Nanotechnology Materials and Devices Conference (Waikiki

Beach, HI, 2012), pp. 43.

[48] E. Kim, N. Jain, R. Jacobs-Gedrim, Y. Xu, and B. Yu, Nanotechnology 23, 5, (2012).

[49] Z.-Y. Ong and M. V. Fischetti, Physical Review B 86, 165422, (2012).

[50] R. S. Shishir, D. K. Ferry, and S. M. Goodnick, Journal of Physics: Conference Series

193, 012118, (2009).

[51] A. Girdhar and J. P. Leburton, Applied Physics Letters 99, 043107, (2011).

[52] L. Pirro, A. Girdhar, Y. Leblebici, and J. P. Leburton, Journal of Applied Physics 112,

093707, (2012).

[53] A. Barreiro, M. Lazzeri, J. Moser, F. Mauri, and A. Bachtold, Physical Review Letters

103, 076601, (2009).

[54] I. Meric, M. Y. Han, A. F. Young, B. Ozyilmaz, P. Kim, and K. L. Shepard, Nature

Nanotechnology 3, 654, (2008).

[55] I. Meric, C. R. Dean, A. F. Young, N. Baklitskaya, N. J. Tremblay, C. Nuckolls, P. Kim,

and K. L. Shepard, Nano Letters 11, 1093, (2011).

[56] A. M. DaSilva, K. Zou, J. K. Jain, and J. Zhu, Physical Review Letters 104, 236601,

Page 103: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

95

(2010).

[57] B. W. Scott and J. P. Leburton, in IEEE Conference on Nanotechnology (Seoul, Korea,

2010), pp. 655.

[58] B. W. Scott and J. P. Leburton, IEEE Transactions on Nanotechnology 10, 1113, (2011).

[59] I. Meric, C. R. Dean, N. Petrone, W. Lei, J. Hone, P. Kim, and K. L. Shepard,

Proceedings of the IEEE 101, 1609, (2013).

[60] J. H. Chen, C. Jang, S. Adam, M. S. Fuhrer, E. D. Williams, and M. Ishigami, Nature

Physics 4, 377, (2008).

[61] J. Martin, N. Akerman, G. Ulbricht, T. Lohmann, J. H. Smet, K. Von Klitzing, and A.

Yacoby, Nature Physics 4, 144, (2008).

[62] Y. B. Zhang, V. W. Brar, C. Girit, A. Zettl, and M. F. Crommie, Nature Physics 5, 722,

(2009).

[63] A. Deshpande, W. Bao, F. Miao, C. N. Lau, and B. J. LeRoy, Physical Review B 79,

205411, (2009).

[64] S. Adam, E. H. Hwang, V. M. Galitski, and S. Das Sarma, Proceedings of the National

Academy of Sciences 104, 18392, (2007).

[65] V. Perebeinos and P. Avouris, Physical Review B 81, 195442, (2010).

[66] S. Kim, J. Nah, I. Jo, D. Shahrjerdi, L. Colombo, Z. Yao, E. Tutuc, and S. K. Banerjee,

Applied Physics Letters 94, 062107, (2009).

[67] K. Zou, X. Hong, D. Keefer, and J. Zhu, Physical Review Letters 105, 126601, (2010).

[68] Z.-Y. Ong and M. V. Fischetti, Physical Review B 86, 121409, (2012).

[69] Z.-Y. Ong and M. V. Fischetti, Applied Physics Letters 102, 183506, (2013).

[70] F. Chen, J. Xia, and N. Tao, Nano Letters 9, 1621, (2009).

Page 104: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

96

[71] L. A. Ponomarenko et al., Physical Review Letters 102, 206603, (2009).

[72] K. Hess and P. Vogl, Solid State Communications 30, 797, (1979).

[73] B. T. Moore and D. K. Ferry, Journal of Applied Physics 51, 2603, (1980).

[74] J. P. Leburton and G. Dorda, Solid State Communications 40, 1025, (1981).

[75] M. Ishigami, J. H. Chen, W. G. Cullen, M. S. Fuhrer, and E. D. Williams, Nano Letters 7,

1643, (2007).

[76] I. J. Chen et al., Nano Letters 14, 1317, (2014).

[77] Z.-Y. Ong and M. V. Fischetti, Physical Review B 86, 199904, (2012).

[78] Z.-Y. Ong and M. V. Fischetti, Physical Review B 88, 045405, (2013).

[79] M. V. Fischetti, D. A. Neumayer, and E. A. Cartier, Journal of Applied Physics 90, 4587,

(2001).

[80] K. M. Borysenko, J. T. Mullen, E. A. Barry, S. Paul, Y. G. Semenov, J. M. Zavada, M. B.

Nardelli, and K. W. Kim, Physical Review B 81, 121412, (2010).

[81] M.-H. Bae, Z.-Y. Ong, D. Estrada, and E. Pop, Nano Letters 10, 4787, (2010).

[82] M.-H. Bae, S. Islam, V. E. Dorgan, and E. Pop, ACS Nano 5, 7936, (2011).

[83] M. Freitag, M. Steiner, Y. Martin, V. Perebeinos, Z. Chen, J. C. Tsang, and P. Avouris,

Nano Letters 9, 1883, (2009).

[84] S. Berciaud, M. Y. Han, K. F. Mak, L. E. Brus, P. Kim, and T. F. Heinz, Physical Review

Letters 104, 227401, (2010).

[85] D.-H. Chae, B. Krauss, K. von Klitzing, and J. H. Smet, Nano Letters 10, 466, (2009).

[86] X. Li, B. D. Kong, J. M. Zavada, and K. W. Kim, Applied Physics Letters 99, 233114,

(2011).

[87] S. Islam, L. Zuanyi, V. E. Dorgan, B. Myung-Ho, and E. Pop, IEEE Electron Device

Page 105: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

97

Letters 34, 166, (2013).

[88] D. Lembke and A. Kis, ACS Nano 6, 10070, (2012).

[89] G. Fiori, B. N. Szafranek, G. Iannaccone, and D. Neumaier, Applied Physics Letters 103,

233509, (2013).

[90] H. Peelaers and C. G. Van de Walle, Physical Review B 86, 241401, (2012).

[91] X. Li, J. T. Mullen, Z. Jin, K. M. Borysenko, M. Buongiorno Nardelli, and K. W. Kim,

Physical Review B 87, 115418, (2013).

[92] J. K. Ellis, M. J. Lucero, and G. E. Scuseria, Applied Physics Letters 99, 261908, (2011).

[93] T. Cheiwchanchamnangij and W. R. L. Lambrecht, Physical Review B 85, 205302,

(2012).

[94] F. Zahid, L. Liu, Y. Zhu, J. Wang, and H. Guo, AIP Advances 3, 052111, (2013).

[95] E. Cappelluti, R. Roldán, J. A. Silva-Guillén, P. Ordejón, and F. Guinea, Physical Review

B 88, 075409, (2013).

[96] H.-Y. Chang, S. Yang, J. Lee, L. Tao, W.-S. Hwang, D. Jena, N. Lu, and D. Akinwande,

ACS Nano 7, 5446, (2013).

[97] G.-H. Lee et al., ACS Nano 7, 7931, (2013).

[98] A. Lochtefeld and D. A. Antoniadis, IEEE Electron Device Letters 22, 95, (2001).

[99] K. Tatsumura, M. Goto, S. Kawanaka, and A. Kinoshita, ECS Transactions 28, 75,

(2010).

[100] M. S. Lundstrom, IEEE Electron Device Letters 22, 293, (2001).

[101] W. Lan, O. Mysore, and D. Antoniadis, IEEE Transactions on Electron Devices 59, 1263,

(2012).

[102] M. Luisier, M. Lundstrom, D. A. Antoniadis, and J. Bokor, in IEEE International

Page 106: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

98

Electron Devices Meeting (Washington, D.C., 2011), pp. 251.

[103] F. N. Xia, V. Perebeinos, Y. M. Lin, Y. Q. Wu, and P. Avouris, Nature Nanotechnology

6, 179, (2011).

[104] T. Fang, A. Konar, H. Xing, and D. Jena, Applied Physics Letters 91, 092109, (2007).

[105] W. Zhu, V. Perebeinos, M. Freitag, and P. Avouris, Physical Review B 80, 235402,

(2009).

[106] E. H. Hwang, S. Adam, and S. Das Sarma, Physical Review Letters 98, 186806, (2007).

[107] N. D. Arora and G. S. Gildenblat, IEEE Transactions on Electron Devices 34, 89, (1987).

[108] Z. Yang, A. Liao, and E. Pop, IEEE Electron Device Letters 30, 1078, (2009).

[109] A. Venugopal, L. Colombo, and E. M. Vogel, Applied Physics Letters 96, 013512,

(2010).

[110] P. A. Khomyakov, A. A. Starikov, G. Brocks, and P. J. Kelly, Physical Review B 82,

115437, (2010).

[111] T. Mueller, F. Xia, M. Freitag, J. Tsang, and P. Avouris, Physical Review B 79, 245430,

(2009).

[112] Z. Chen, W. Jang, W. Bao, C. N. Lau, and C. Dames, Applied Physics Letters 95,

161910, (2009).

[113] S.-M. Lee and D. G. Cahill, Journal of Applied Physics 81, 2590, (1997).

[114] M. Asheghi, K. Kurabayashi, R. Kasnavi, and K. E. Goodson, Journal of Applied Physics

91, 5079, (2002).

[115] F. Assaderaghi, D. Sinitsky, J. Bokor, P. K. Ko, H. Gaw, and H. Chenming, IEEE

Transactions on Electron Devices 44, 664, (1997).

[116] R. Quay, C. Moglestue, V. Palankovski, and S. Selberherr, Materials Science in

Page 107: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

99

Semiconductor Processing 3, 149, (2000).

[117] K. Hess, Advanced Theory of Semiconductor Devices (Wiley-IEEE Press, New York,

2000).

[118] K. I. Bolotin, K. J. Sikes, J. Hone, H. L. Stormer, and P. Kim, Physical Review Letters

101, 096802, (2008).

[119] X. Du, I. Skachko, A. Barker, and E. Y. Andrei, Nature Nanotechnology 3, 491, (2008).

[120] E. V. Castro, H. Ochoa, M. I. Katsnelson, R. V. Gorbachev, D. C. Elias, K. S. Novoselov,

A. K. Geim, and F. Guinea, Physical Review Letters 105, 266601, (2010).

[121] W. W. Cai, A. L. Moore, Y. W. Zhu, X. S. Li, S. S. Chen, L. Shi, and R. S. Ruoff, Nano

Letters 10, 1645, (2010).

[122] C. Faugeras, B. Faugeras, M. Orlita, M. Potemski, R. R. Nair, and A. K. Geim, ACS

Nano 4, 1889, (2010).

[123] J.-U. Lee, D. Yoon, H. Kim, S. W. Lee, and H. Cheong, Physical Review B 83, 081419,

(2011).

[124] S. S. Chen et al., ACS Nano 5, 321, (2011).

[125] 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, Nature Materials 11, 203, (2012).

[126] E. Pop, D. Mann, J. Cao, Q. Wang, K. Goodson, and H. Dai, Physical Review Letters 95,

155505, (2005).

[127] D. Mann, E. Pop, J. Cao, Q. Wang, and K. Goodson, Journal of Physical Chemistry B

110, 1502, (2006).

[128] D. Mann et al., Nature Nanotechnology 2, 33, (2007).

[129] M. Amer, A. Bushmaker, and S. Cronin, Nano Research 5, 172, (2012).

Page 108: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

100

[130] V. E. Dorgan, A. Behnam, H. J. Conley, K. I. Bolotin, and E. Pop, Nano Letters 13, 4581,

(2013).

[131] D. Teweldebrhan and A. A. Balandin, Applied Physics Letters 94, (2009).

[132] J. S. Bunch, A. M. van der Zande, S. S. Verbridge, I. W. Frank, D. M. Tanenbaum, J. M.

Parpia, H. G. Craighead, and P. L. McEuen, Science 315, 490, (2007).

[133] J. Sabio, C. Seoánez, S. Fratini, F. Guinea, A. H. C. Neto, and F. Sols, Physical Review B

77, 195409, (2008).

[134] K. L. Grosse, M. H. Bae, F. F. Lian, E. Pop, and W. P. King, Nature Nanotechnology 6,

287, (2011).

[135] E. Pop, V. Varshney, and A. K. Roy, MRS Bulletin 37, 1273, (2012).

[136] D. C. Elias et al., Nature Physics 7, 701, (2011).

[137] J. H. Chen, C. Jang, S. D. Xiao, M. Ishigami, and M. S. Fuhrer, Nature Nanotechnology

3, 206, (2008).

[138] D. K. Ferry, Journal of Computational Electronics 12, 76, (2013).

[139] X. Li, E. A. Barry, J. M. Zavada, M. Buongiorno Nardelli, and K. W. Kim, Applied

Physics Letters 97, 232105, (2010).

[140] I. W. Chiang, B. E. Brinson, A. Y. Huang, P. A. Willis, M. J. Bronikowski, J. L.

Margrave, R. E. Smalley, and R. H. Hauge, The Journal of Physical Chemistry B 105,

8297, (2001).

[141] K. Hata, D. N. Futaba, K. Mizuno, T. Namai, M. Yumura, and S. Iijima, Science 306,

1362, (2004).

[142] L. Liu, S. Ryu, M. R. Tomasik, E. Stolyarova, N. Jung, M. S. Hybertsen, M. L.

Steigerwald, L. E. Brus, and G. W. Flynn, Nano Letters 8, 1965, (2008).

Page 109: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

101

[143] A. Liao, R. Alizadegan, Z.-Y. Ong, S. Dutta, F. Xiong, K. J. Hsia, and E. Pop, Physical

Review B 82, 205406, (2010).

[144] A. D. Liao, J. Z. Wu, X. Wang, K. Tahy, D. Jena, H. Dai, and E. Pop, Physical Review

Letters 106, 256801, (2011).

[145] R. R. Nair, P. Blake, A. N. Grigorenko, K. S. Novoselov, T. J. Booth, T. Stauber, N. M.

R. Peres, and A. K. Geim, Science 320, 1308, (2008).

[146] X. T. Jia et al., Science 323, 1701, (2009).

[147] K. Kim, W. Regan, B. S. Geng, B. Aleman, B. M. Kessler, F. Wang, M. F. Crommie, and

A. Zettl, Physica Status Solidi-Rapid Research Letters 4, 302, (2010).

[148] X. Y. Huang, Z. Y. Zhang, Y. Liu, and L.-M. Peng, Applied Physics Letters 95, (2009).

[149] C. Y. Ho, R. W. Powell, and P. E. Liley, Journal of Physical and Chemical Reference

Data 1, 279, (1972).

[150] D. L. Nika, A. S. Askerov, and A. A. Balandin, Nano Letters 12, 3238, (2012).

[151] C. J. Glassbrenner and G. A. Slack, Physical Review 134, A1058, (1964).

[152] T. J. Singh and G. S. Verma, Physical Review B 25, 4106, (1982).

[153] N. Mingo and D. A. Broido, Nano Letters 5, 1221, (2005).

[154] E. Pop, D. Mann, Q. Wang, K. Goodson, and H. Dai, Nano Letters 6, 96, (2006).

[155] C. Yu, A. M. Cassell, B. A. Cruden, S. Saha, J. Zhou, Q. Ngo, J. Li, and L. Shi, Journal

of Heat Transfer 128, 234, (2005).

[156] Y. K. Koh, M.-H. Bae, D. G. Cahill, and E. Pop, Nano Letters 10, 4363, (2010).

[157] A. J. Schmidt, K. C. Collins, A. J. Minnich, and G. Chen, Journal of Applied Physics

107, (2010).

[158] I.-K. Hsu, M. T. Pettes, M. Aykol, L. Shi, and S. B. Cronin, Journal of Applied Physics

Page 110: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

102

108, 084307, (2010).

[159] N. Bonini, M. Lazzeri, N. Marzari, and F. Mauri, Physical Review Letters 99, 176802,

(2007).

[160] M. T. Pettes, I. Jo, Z. Yao, and L. Shi, Nano Letters 11, 1195, (2011).

[161] A. Behnam, A. S. Lyons, M.-H. Bae, E. K. Chow, S. Islam, C. M. Neumann, and E. Pop,

Nano Letters 12, 4424, (2012).

[162] M. W. Lin, C. Ling, Y. Y. Zhang, H. J. Yoon, M. M. C. Cheng, L. A. Agapito, N.

Kioussis, N. Widjaja, and Z. X. Zhou, Nanotechnology 22, 7, (2011).

[163] V. Singh, D. Joung, L. Zhai, S. Das, S. I. Khondaker, and S. Seal, Progress in Materials

Science 56, 1178, (2011).

[164] R. Fivaz and E. Mooser, Physical Review 163, 743, (1967).

[165] M. M. Chowdhury and J. G. Fossum, IEEE Electron Device Letters 27, 482, (2006).

[166] O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic, and A. Kis, Nature

Nanotechnology 8, 497, (2013).

[167] H. Wang et al., Nano Letters 12, 4674, (2012).

[168] H. Liu, A. T. Neal, and P. D. Ye, ACS Nano 6, 8563, (2012).

[169] H. Qiu et al., Nature Communications 4, 2642, (2013).

[170] B. Radisavljevic and A. Kis, Nature Materials 12, 815, (2013).

[171] B. H. Kim et al., RSC Advances 3, 18424, (2013).

[172] C. K. Williams, T. H. Glisson, J. R. Hauser, M. A. Littlejohn, and M. F. Abusaid, Solid-

State Electronics 28, 1105, (1985).

[173] S. Sahoo, A. P. S. Gaur, M. Ahmadi, M. J. F. Guinel, and R. S. Katiyar, Journal of

Physical Chemistry C 117, 9042, (2013).

Page 111: HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE … · HIGH-FIELD TRANSPORT IN TWO-DIMENSIONAL GRAPHENE AND MOLYBDENUM DISULFIDE BY VINCENT E. DORGAN DISSERTATION Submitted in partial

103

[174] A. Molina-Sánchez and L. Wirtz, Physical Review B 84, 155413, (2011).

[175] V. F. Gantmakher and Y. B. Levinson, Carrier Scattering in Metals and Semiconductors

(North-Holland, Amsterdam, 1987), pp. 143-145.

[176] J. H. Davies, The Physics of Low-Dimensional Semiconductors: An Introduction

(Cambridge University Press, New York, 1998), pp. 353-355.

[177] J. Appenzeller, M. Radosavljević, J. Knoch, and P. Avouris, Physical Review Letters 92,

048301, (2004).

[178] S. Das, H.-Y. Chen, A. V. Penumatcha, and J. Appenzeller, Nano Letters 13, 100, (2013).

[179] A. T. Neal, H. Liu, J. J. Gu, and P. D. Ye, in IEEE Device Research Conference

(University Park, PA, 2012), pp. 65.

[180] C. D. English, V. E. Dorgan, G. Shine, F. Xiong, K. C. Saraswat, and E. Pop, in MRS

Spring Meeting (San Francisco, CA, 2014).

[181] T. Fang, A. Konar, H. Xing, and D. Jena, Physical Review B 78, 205403, (2008).

[182] I. Aberg and J. L. Hoyt, IEEE Electron Device Letters 26, 661, (2005).

[183] J. Seger, P.-E. Hellström, J. Lu, B. G. Malm, M. von Haartman, M. Östling, and S.-L.

Zhang, Applied Physics Letters 86, 253507, (2005).