ABSTRACT
Title of Document:
1/f NOISE AND LUTTINGER LIQUID
PHENOMENA IN CARBON NANOTUBES
David Andrew Tobias, Doctor of Philosophy,
2007
Directed By:
Associate Professor Michael S. Fuhrer and
Professor Christopher J. Lobb,
Department of Physics
Carbon nanotubes (CNTs) provide an ideal medium for testing the behavior of
one-dimensional electron systems and are promising candidates for electronic
applications such as sensors or field-effect transistors. This thesis describes the use of
low frequency resistance fluctuations to measure both the properties of the onedimensional electron system in CNTs, and the sensitivity of CNT devices to their
environment.
Low frequency noise was measured in CNTs in field effect transistor (FET)
geometry. CNTs have a large amount of surface area relative to their volume and are
expected to be strongly affected by their environment, leading to speculation that
CNTs should have large amounts of 1/f noise. My measurements indicate that the
noise level is in the same range as that of traditional FETs, an encouraging result for
possible electronic applications. The temperature dependence of 1/f noise from 1.2 K
to 300 K can be used to extract the characteristic energies of the fluctuators
responsible for the noise. The characteristic energies allows for the elimination of
structural and electronic transitions within the CNT itself as possible sources of 1/f
noise in CNTs, leaving the motion of defects in the gate dielectric, or possibly
strongly physisorbed species, as the likely culprits.
Another form of low frequency noise found in CNTs is random telegraph
signal (RTS), which manifests as the alternation between two current states at a stable
voltage bias. In CNTs, this phenomenon occurs due to the tunneling of electrons into
and out of the CNT from a nearby defect, and thus provides a way to probe the
tunneling density of states of the CNT itself. The tunneling density of states in turn
provides information on the strength of the electron-electron interaction in CNTs.
Due to the one-dimensional structure of CNTs their electronic state is expected to be
a Luttinger liquid, which should manifest as a power-law suppression of the tunneling
density of states at the Fermi energy. The power law exponent is measured in both
the temperature dependence and energy dependence of the tunneling rates. In
agreement with theory, the power-law exponent is significantly larger in
semiconducting CNTs than found in previous experiments on metallic CNTs. The
RTS can also be used as a “defect thermometer” to probe the electron temperature of
the CNT. The effect of the bias voltage on the electron temperature provides a means
to determine the energy relaxation length for the electrons in the CNT.
1/f NOISE AND LUTTINGER LIQUID PHENOMENA IN CARBON
NANOTUBES
By
David Andrew Tobias
Dissertation submitted to the Faculty of the Graduate School of the
University of Maryland, College Park, in partial fulfillment
of the requirements for the degree of
Doctor of Philosophy
2007
Advisory Committee:
Associate Professor Michael S. Fuhrer, Chair/Co-Advisor
Professor Christopher J. Lobb, Co-Advisor
Professor Richard Greene
Professor James R. Anderson
Professor Romel Gomez
© Copyright by
David Andrew Tobias
2007
Dedication
To my father and my uncle: James and William Tobias.
ii
Acknowledgements
I feel fortunate to have worked with wonderful people while at the University
of Maryland. My first year here I realized that you learn the most from solving the
homework with other students and a parallel certainly exists in the lab. I would like to
thank all of the students that have helped me while here at Maryland. This list is long
as many people have made Maryland a great place while I was here.
Christopher Lobb called me while I was still an undergrad at Case and asked
whether I was interested in coming to Maryland, so I asked if I could work in his lab
for the summer before starting classes in grad school. That was the beginning of
working for him and Frederick Wellstood in the sub-sub basement of the Center for
Superconductivity Research. Later I switched to nanotube research, where I was
fortunate to have Michael Fuhrer become my co-advisor. All three are both excellent
scientists and great advisors. Any graduate student would be lucky to work with them
and the wonderful environment in their labs starts with them. They have made me a
better scientist, speaker and writer. I would also like to thank my committee of Bob
Anderson, Richard Greene and Romel Gomez.
I first worked with Matthew Kenyon on the single electron transistor
microscope. He initiated me into the rites of working in the lab: from lithography to
plumbing to left coast thinking. Hadley Lawler and Hanhee Paik brought new
dimensions to working in the sub-basement and knowledge of the Korean language
(mostly from Hanhee). After switching to nanotubes, Tobias Durkop introduced me
to growing nanotubes and provided insight on how to always get your way for a lunch
destination. Todd Brintlinger taught me how to “find and wire” the tubes and taught
iii
me to play pool. Yung-Fu Chen was a constant source of information on what to
expect from my devices and how to understand them, and possessed a truly kind soul.
Gokhan Esen provided invaluable help with running the cryostat and taking
measurements, and was an excellent deskmate and commiserator. Finally, Masahiro
Ishigami joined me in investigating the low frequency noise in nanotubes, and added
energy and vocabulary to the second floor of the physics building.
Special thanks go to Paola Barbara and Alexander Tselev at Georgetown for
providing the samples that were used for the data in this thesis. No one can ever be
thanked enough for samples.
Doug Bensen and Brian Straughn helped with all sorts of experimental
difficulties. Belta Pollard and Margaret Lukomska helped with smoothing out
traveling and purchasing. Jane Hessing is a blessing to the grad students at Maryland
and has helped me more times than I can remember. Bob Dahms and Jesse Anderson
have also gone beyond the call many times. The electronics and machine shop have
also helped me fix my experiments.
I was fortunate to be a classmate with Samir Garzon. He taught me most of
what I know about quantum mechanics and I wouldn’t have passed the qualifier
without him. Joshua Higgins joined me in the Center after being a class mate and has
been a great friend since. Betsy Pugel has more fire than ten physicists need.
In addition to the people that worked directly with me on my projects, I was
blessed to work in the sub-basement with a cohesive group of people that made
physics a truly collaborative experience. The center is a great place to work due to the
opportunity to walk into any lab and find someone there to help you out. It also
iv
makes Maryland feel like a smaller place. Erin Fleet, Guy Chatraphorn, Gus Vlahacos
and Aaron Nielsen helped give me a push at the start with their SQUID microscopy
knowledge. Su-Young Lee was the most tenacious physicist and also the funniest.
Doug Strachan provided provocative conversation and fundamental knowledge of
superconductivity and dip testing. Branimir Vasilic showed me how to transfer
helium and what in physics was trivial. Matt Sullivan was a MacGyver of
experimental physics, and always there to throw the disc. Monica Lilly was the
greatest undergraduate to ever work in a lab.
The quantum computing team helped me despite their need to constantly fight
terrorism. Andrew Berkley was a post-doc from birth and pushed me to know more
about what I was doing. Huizhong Xu followed that up by somehow being an even
more amazing scientist and a great hot dogger. Sudeep Dutta showed me how great
our lab was and how fortunate I am to be at Maryland. Rupert Lewis and Tauno
Palomaki were even able to crack jokes while carrying the burden of fighting crime.
After switching to Fuhrer’s lab I was fortunate to continue my streak of good
co-workers. Stephanie Getty helped to make me a part of the group and Anthony
Ayari proved that French people can have a good sense of humor, too. Dan Lenski
and Enrique Cobas have answered too many questions about computers. Adrian
Southard got me onto a soccer field again. Tarek Ghanem introduced me to Tabir.
Alexandra Curtin was great for all sorts of arguments and for frisbee weekends.
Sungjae Cho reminded me of my beginning as a grad student. Finally Kristen Burson
made sure these acknowledgements got written.
v
Thanks to Jianhao Chen for making samples that were used in the low
frequency noise measurements done by Masa. Winston Yang was a wonderful
summer addition to the lab and a motivated youngster with a bright future. Elba
Gomar Nadal brought an outsider’s view to the physics department and was a great
lunch, hiking and beer-drinking partner.
I also had many great roommates as a student at Maryland. Aaron Nielsen
brought me into a “luxury” townhouse where I lived with DJ Patil, David Sweet,
Yao-Chin Chao, Sarah Boettcher and Evelyn Boettcher over my grad school years.
Living with someone creates a kind of friendship that can not be developed any other
way and these are great friends that have lead to salmon fests, bhangra and to a trip to
the other side of the world.
I would like to thank my family for supporting my decision to get a PhD. If I
ever doubt my luck in life, I remember them.
Finally, I would like to thank Silvia for putting up with me through the stress
of finishing my degree.
vi
Table of Contents
Dedication ..................................................................................................................... ii
Acknowledgements...................................................................................................... iii
Table of Contents........................................................................................................ vii
List of Tables ................................................................................................................ x
List of Figures .............................................................................................................. xi
Chapter 1 Introduction to carbon nanotubes................................................................. 1
1.1 Carbon nanotube overview ................................................................................. 1
1.2 Electronic band Structure.................................................................................... 3
1.3 Electronic device properties................................................................................ 6
Chapter 2 Sample fabrication and measurement ........................................................ 10
2.1 Growth methods................................................................................................ 10
2.2 Chemical vapor deposition ............................................................................... 11
2.3 Locate and contact ............................................................................................ 13
2.4 Georgetown technique ...................................................................................... 18
2.5 Experimental Setup........................................................................................... 19
Chapter 3 1/f noise ...................................................................................................... 22
3.1 Introduction to noise ......................................................................................... 23
3.2 Semiconductors and transistors......................................................................... 26
3.3 Basic noise model ............................................................................................. 27
3.4 Theory for temperature dependence ................................................................. 30
3.5 Previous results in nanotubes............................................................................ 32
vii
Chapter 4 Temperature dependence of 1/f noise in carbon nanotube transistors ....... 35
4.1 Noise signal....................................................................................................... 36
4.2 Data analysis techniques ................................................................................... 39
4.2.1 Power-law method ..................................................................................... 39
4.2.2 Inverse noise method ................................................................................. 40
4.2.3 Inverse noise plus telegraph method.......................................................... 42
4.2.4 “Show all the data” method ....................................................................... 44
4.3 Gate dependence ............................................................................................... 46
4.4 Noise in semiconducting devices at room temperature .................................... 48
4.5 Temperature dependence .................................................................................. 51
Chapter 5 Introduction to 1-D physics and telegraph signal....................................... 57
5.1 Drude and Luttinger.......................................................................................... 57
5.2 Previous measurements in carbon CNTs .......................................................... 62
5.3 Hysteresis in CNTs ........................................................................................... 66
5.4 Random telegraph signals ................................................................................. 69
Chapter 6 Random telegraph signals in carbon nanotubes and their use as a defect
thermometer ................................................................................................................ 75
6.1 Random Telegraph Signals in Carbon Nanotube Transistors........................... 75
6.2 Defect thermometry .......................................................................................... 82
Chapter 7 Coherence and correlations in carbon nanotubes studied using random
telegraph signals.......................................................................................................... 85
7.1 1-D electron behavior ....................................................................................... 85
Abbreviations.............................................................................................................. 97
viii
Symbols....................................................................................................................... 98
Bibliography ............................................................................................................. 100
ix
List of Tables
Table 2-1. Typical recipe for CNT growth. ................................................................ 13
Table 4-1. Device Characteristics ............................................................................... 36
Table 7-1. Luttinger parameter α for the three RTS................................................... 96
x
List of Figures
Figure 1-1. Hexagon lattice for a graphene sheet. ........................................................ 3
Figure 1-2. A single walled CNT.................................................................................. 4
Figure 1-3. Metal CNT band structure.......................................................................... 5
Figure 1-4. Semiconducting CNT band structure. ........................................................ 5
Figure 1-5. Current vs. gate voltage for a CNT in field effect transistor geometry...... 7
Figure 2-1. CVD furnace for CNT growth in the Fuhrer laboratory. ......................... 12
Figure 2-2. E-beam lithography process..................................................................... 15
Figure 2-3. Electron micrograph of the square alignment marker pattern.................. 16
Figure 2-4. Close up image of a CNT under contacts................................................. 18
Figure 2-5. Schematic of the electronic setup used in the cryostat............................. 21
Figure 3-1. Noise circuit. ............................................................................................ 24
Figure 3-2. 1/f noise formation. .................................................................................. 30
Figure 3-3. Graph from Dutta and Horn ..................................................................... 32
Figure 4-1. Dependence of noise on current bias. ...................................................... 37
Figure 4-2. Current noise versus frequency. ............................................................... 40
Figure 4-3. Plot of 1/SI versus frequency. .................................................................. 41
Figure 4-4. A nonlinear fit for 1/noise versus frequency............................................ 43
Figure 4-5. Noise power versus frequency with a nonlinear fit.................................. 44
Figure 4-6. “Show all the data” plot. .......................................................................... 45
Figure 4-7. Reciprocal of the noise prefactor 1/A ...................................................... 47
Figure 4-8. Noise data from Ishigami et al. ................................................................ 49
Figure 4-9. Comparison of inverse noise amplitude................................................... 51
xi
Figure 4-10. Noise vs temperature.............................................................................. 53
Figure 5-1. Drude vs Luttinger. .................................................................................. 60
Figure 5-2. Tunneling density of states....................................................................... 61
Figure 5-3. Plot from Bockrath et al........................................................................... 64
Figure 5-4. Semiconducting CNT hysteresis. ............................................................. 67
Figure 5-5. Previous RTS experiment. ....................................................................... 71
Figure 5-6. Segment of an I-Vg curve taken on a CNT FET...................................... 74
Figure 6-1. Small section of the current versus voltage curve for a two level
fluctuator. .................................................................................................................... 76
Figure 6-2. RTS schematic. ........................................................................................ 78
Figure 6-3. The natural log of the ratio of the tunneling rates versus gate voltage .... 80
Figure 6-4. Electron temperature as a function of bias voltage at various substrate
temperatures................................................................................................................ 82
Figure 7-1. A typical plot of the individual transition rates versus gate voltage ........ 88
Figure 7-2. Degenerate rate vs. electron temperature. ................................................ 90
Figure 7-3. Depiction of the theoretical calculation of the Luttinger parameter α
versus gate voltage. ..................................................................................................... 94
Figure 7-4. Depiction of the theoretical calculation of the Luttinger parameter g
versus gate voltage. ..................................................................................................... 95
xii
Chapter 1 Introduction to carbon nanotubes
1.1 Carbon nanotube overview
Carbon nanotubes (CNTs) are hollow tube-shaped structures with all of the
carbon atoms bonded together by sp2 bonds in a honeycomb lattice identical to that of
graphene. Conceptually, the CNT may be thought of as a single sheet of graphite
(termed graphene) curved into a seamless cylinder. These structures were first
identified by Ijima[1]. Initially the CNTs were exclusively multi-walled, meaning that
they consisted of several concentric cylinders. Future researchers were able to
develop methods capable of producing single-walled CNTs of varying lengths[2-4],
and some CNTs have been grown that are centimeters in length[5, 6]. CNTs have
diameters from just under a nanometer[4] to dozens of nanometers[1] and hence have
very large aspect ratios.
CNTs are characterized by many impressive properties. Individually they have
extremely high mechanical rigidity and toughness, leading to many hypothetical
applications for ropes and fibers[7], including the famous (or infamous) space
elevator[8]. Networks of CNTs have been found to have numerous interesting
properties including the ability to form fire resistant material[9] and liquid crystal
suspensions[10]. CNTs also have a powerful Van der Waals attraction[11] with
surfaces allowing them to be used as an adhesive material between paint and
plastics[12].
1
One of the main interesting properties of CNTs chemically and electronically
is that every atom is a surface atom, and is electronically “in series” with every other
atom in the CNT. This could be useful for creating chemical sensors that are able to
detect very low amounts of contaminants[13]. This also leads to concerns about
fluctuations and noise in CNTs[14]. CNTs can also be functionalized by many
interesting molecules[15], including DNA[16], in the hope of enabling bottom-up
construction of micro- and nano-structures.
CNTs also have fascinating electrical properties that derive from their
graphene origin, as discussed below. CNTs are either semiconducting or metallic[1720], depending upon the relative direction of the CNT axis with respect to the
graphene lattice. Obviously this and their nanoscale size makes them a speculative
candidate for future electronics technologies, but as of now the difficulty of
producing, orienting and contacting the CNTs has slowed the realization of this
application. However, individual single electron transistors[21] , high mobility
transistors[22] and other electronic devices have been realized using CNTs.
In one dimension, electrostatic interactions between electrons are strong, and
the electrons form a correlated state termed the Luttinger liquid[23-25] (LL). This
state of the electrons differs in many ways from that found in bulk conductors where
the electrons are able to re-arrange themselves easily to reduce the energy of the
interaction amongst them. This state should only exist in 1-D materials and thus
CNTs offer an excellent opportunity to study LL physics.
2
1.2 Electronic band Structure
The ability of CNTs to form metallic and semiconducting devices derives
from the band structure of graphene[26-28]. As mentioned above, CNTs can be
thought of as strips of graphene sheets that are rolled up to form a seamless cylinder.
The CNT will have different properties depending upon its helicity and diameter.
Graphene is a honeycomb lattice, a two dimensional hexagonal Bravais lattice, with a
basis of two carbon atoms as seen in Fig. 1-1. The distance between nearest neighbors
in the carbon lattice is 0.142 nm.
Figure 1-1. Hexagon lattice for a graphene sheet. The basis vectors are indicated in
the bottom left and a rolling vector for cutting the sheet into a strip in the middle.
When rolled into a cylinder the strip would form a CNT.
Since the method of rolling the CNT up from a graphene strip determines the
properties of the resulting CNT, the vector that points from an atom to the atom it will
3
roll into is called the rolling vector, R = na1 + ma2, where n and m are integers and a1
and a2 are the graphene unit lattice vectors; this vector also defines the circumference
of the CNT. The strip is defined by the dashed lines perpendicular to the beginning
and ending of this vector as in Fig. 1-1. The result of rolling up the sheet is shown in
Fig. 1-2.
Figure 1-2. A single walled CNT. This CNT has (n,m) = (5,5), and is metallic (see
text). (courtesy R.E. Smalley)
The electronic structure of the CNT may be well approximated by starting
with the band structure for graphene[29] and quantizing it so that the electronic wavefunction is single valued around the circumference of the CNT. The quantization
condition is R∏k=2πi where i is an integer and k is the wave vector. The result is that
the two-dimensional band structure for graphene is cut along a series of equally
spaced parallel lines to form a number of one-dimensional subbands.
The graphene band structure itself may be approximated as linear[29]:
E (q) = EF ±
3γ 0 qa
2
(1.1)
4
Figure 1-3. Metal CNT band structure. Slices through the band structure for graphene
that determine the band structure for a metallic CNT. The lowest two subbands are
depicted in the band diagram at right.
Figure 1-4. Semiconducting CNT band structure. Slices through the band structure for
graphene that result in a semiconducting CNT. The lowest two subbands are depicted
in the band diagram at right.
5
where q = k - K is the wave vector measured from the K point, γ0 the nearest-neighbor tightbinding integral, and a the graphite lattice constant. Thus the band structure for graphene
looks like a grid of cones with the tips at the vertices of the hexagonal Brillouin zone
as show in Figs. 1-3 and 1-4. Since the bands only cross the Fermi surface at this
point, the K point, only those CNTs which have R∏K=2πi will be metallic. All other
CNTs will have a bandgap and be semiconducting. This condition can be expressed
as n - m = 3q where q is an integer.
The values calculated via this method, e.g. the Fermi velocity, band gap, etc.,
agree very well with experiment, especially for larger diameter (d > 1 nm) CNTs, so
we will use this approximation.
The dispersion relation of the lowest-lying subbands can be written:
E (k ) = ∆2 + ( v F k )
2
(1.2)
where ħ is Planck’s constant, vF the Fermi velocity of graphene, and ∆ = 0 for
metallic CNTs, and for semiconducting CNTs
E g = 2∆ =
2γ 0 a
3d
≈
830 meV
d [nm]
(1.3)
where d is the diameter of the CNT.
1.3 Electronic device properties
A field effect transistor (FET) may be constructed from a CNT by contacting
the CNT with two metallic (source and drain) electrodes, and employing a third
metallic electrode, separated from the CNT by a dielectric, as a gate (Chapter 2 will
6
discuss some fabrication methods for CNT FETs in detail). Experimentally, it is
difficult to determine the wrapping vector for an individual CNT. However, once a
FET is constructed from an individual CNT two types of behavior are observed,
which are identified with metallic and semiconducting CNTs[19, 30].
Figure 1-5. Current vs. gate voltage for a CNT in field effect transistor geometry.
Fig. 1-5 is a typical data curve from a semiconducting CNT FET. The current
depends strongly on gate voltage, being finite for negative gate voltage (p-type FET
behavior) and dropping to near zero for positive gate voltage. Metallic CNT FETs
7
show nearly constant conductivity vs. gate voltage. All of the work in this thesis was
done on semiconducting CNTs.
In my dissertation I will explore low-frequency electronic noise in CNT FET
devices and the insights it gives about the behavior of electrons in one-dimensional
systems. In Chapter 2 I will present the basics of CNT growth and device fabrication,
followed by an explanation of the experimental setup. In Chapter 3 I will review the
state of knowledge on 1/f noise in traditional systems and in CNTs. This will provide
the theoretical and experimental background needed for Chapter 4 where I will
present the work of our group on the temperature dependence of 1/f noise in CNT
FETs.
Chapters 5-7 will investigate a different type of low-frequency noise, the
random telegraph signal (RTS), in CNTs. RTS in CNTs resulting from the tunneling
of an electron between the CNT and a nearby defect will be used to study the
Luttinger liquid state in CNTs. Chapter 5 will introduce the expected Luttinger liquid
state for electrons in CNTs, followed by previous experimental work on Luttinger
liquids in metallic CNTs and then a description of previous work using RTS to extract
information on the correlated electron system of semiconducting materials. Chapter 6
presents the use of RTS to determine the temperature of the electron system and the
energy relaxation length of electrons in CNTs. Chapter 7 analyzes the gate-voltage
and temperature dependence of the RTS to extract information about the Luttinger
liquid state in semiconducting CNTs.
The information in Chapter 4 and in Chapters 6-7 is currently being prepared
in the form of two publications, respectively, to be submitted to peer-reviewed
8
scientific journals. The material on room temperature noise in Chapter 4 has been
published.[31]
9
Chapter 2 Sample fabrication and measurement
Carbon nanotubes (CNTs) grow robustly in many situations where high
temperature and carbon meet. CNTs have even been synthesized from heated plant
matter[32] (including hemp!). The details of nanotube growth are still not fully
understood but I will give a brief overview of the main techniques used to grow
carbon nanotubes and then a description of the chemical vapor deposition (CVD)
method that was used to produce all of the CNTs that are discussed in this thesis.
Device fabrication consists of growing the CNTs on SiO2/Si substrates, and
using photolithography or electron-beam lithography (EBL) to establish contact to the
CNTs with metal electrodes. Afterwards the completed CNT devices are placed in a
cryostat for measurement of their electrical properties at low temperature.
2.1 Growth methods
Synthesis methods for production of small-diameter (single- or few-walled)
CNTs share in common a source of atomic carbon, a nano-particle catalyst (typically
a transition metal or alloy of transition metals), and high temperature. Laser
ablation[3] and arc discharge[1, 4] both use graphite as the source of carbon atoms. In
the arc-discharge technique, a high current between a carbon cathode and an anode in
an inert gas, e.g. helium, creates carbon-containing plasma, and if catalyst metal is
added to the graphite electrodes, CNTs grow from tiny droplets of metal coalescing
from the plasma. The laser ablation technique involves striking a piece of graphite
10
with intense laser pulses. Again, the graphite is impregnated with transition-metal
catalyst to produce single-walled CNTs. One of the drawbacks of these techniques is
that the CNTs are generally produced in bundles as opposed to individual CNTs. This
is a major drawback if one hopes to investigate the electrical properties of an
individual CNT. The CNTs also must be removed from the growth chamber, put in a
liquid suspension or solution, and then spun onto a chip before electrical
measurements can be made. The CVD method allows CNTs to be grown directly on a
silicon chip.
2.2 Chemical vapor deposition
Chemical vapor deposition is initiated by creating nano-particles of a metal
catalyst on the surface of an oxidized silicon chip.[2, 33] In my work, iron nanoparticles were obtained by dipping a silicon chip in a ferric nitrate solution and then
into hexane to force the ferric nitrate to precipitate out on the surface of the chip. The
density of the ferric nitrate is important for determining the density of nanotubes that
will be present on the chip after growth, values can range from 0.1 - 100 µg/ml with
lower values typical for single CNT devices and higher values used to obtain dense
films of CNTs.
11
temperature controller
multi gas flow meter
coil
quartz tube
Figure 2-1. CVD furnace for CNT growth in the Fuhrer laboratory. The flow meters
on the wall control the amount of carbon containing feedstock gases through the
system. The silicon chips are placed inside the quartz tube and after the oven lid is
closed the coils heat the oven to 850± C for the nanotube growth. Image courtesy Y.
Chen.
After the chips are catalyzed, they are placed onto a quartz boat and set in a
quartz tube oven; the growth recipe that I used is outlined in Table 1. The oven (see
Fig. 2-1) is heated to 850± C while flowing argon through the tube. At this stage, or
shortly after introduction of hydrogen during growth, the ferric nitrate particles are
reduced to iron. Once the oven has reached its final temperature, carbon-containing
12
feedstock gases (methane, ethylene) and hydrogen are fed through the quartz tube.
The ratio and flow rate of the gases (see Table 1) are adapted from the Dai group[34]
and have been optimized to produce long CNTs. In general the nanotubes will be
oriented along the direction of the gas flow, but numerous exceptions can be found on
any given chip, including CNTs that form arcs, circles or are perpendicular to the gas
flow direction.
Table 2-1. Typical recipe for CNT growth.
Growth
Recipe
Purge
Gas
Ar
Flow
(sccm*)
730
Ar
730
Ar
730
Ar
730
Ar
730
Ar
730
Ar
730
Ar
730
Ar
730
H2
1900
Nanotube
CH4
1300
Growth
C 2 H4
86
Ar
730
Cool Down
*sccm=standard cubic centimeters per minute
Heat
Soak
Heat
Soak
Heat
Soak
Heat
Soak
Temp
(oC)
Room
Temperature
RT ö650
650
650ö730
730
730ö800
800
800ö850
850
850
Time
(minute)
5
850ö200
wait until cool
15
5
2
3
3
3
5
10
10
2.3 Locate and contact
The CVD growth method described above produces nanotubes distributed
randomly on the surface of the chip. The next step of the process is to create
alignment markers on the surface of the chips to serve as guides for locating and
13
creating contacts to the CNTs themselves. An alternative process[35], described in
the next section, uses patterned catalyst and photolithography aligned to the catalyst
locations. Photolithography is a more reliable and quicker process than EBL, but EBL
does have the advantage of allowing maximal flexibility in creating devices of any
length up to the length of the CNT.
A standard e-beam process, depicted in Fig. 2-2, was used to create the grids
on the surface of the chip and is depicted in Fig. 2-3. This involves spinning resist
layers on a chip followed by baking them on a hot plate. First the methylmethacrylate
(MMA) is spun on at 4500 rpm for 45s and baked for 5 min at 150± C, and then the
polymethylmethacrylate (PMMA) is spun at 6000 rpm for 45s and baked for 5 min at
150± C. A modified scanning electron microscope (SEM) is then used to create a
pattern in the resist layers; this is caused by the electron beam weakening the bonds in
the polymer structure of the resist. After writing, the chip is developed in
methylisobutylketone / isopropanol (MIBK/IPA) (3:1) for around 30 s to remove the
written areas. Finally the chip is placed in a vacuum chamber where the resist acts as
a stencil mask for the thermally-deposited metals used to create the pattern. The two
layers of resist help create undercut; the MMA develops faster creating a tiered
structure seen in Figure 2-2e. The undercut separates the metal on the surface of the
resist from the metal on the SiO2 surface, allowing the unwanted metal to be removed
cleanly during lift-off. MMA is also more soluble in acetone which allows for better
lift-off after deposition. Lift-off is accomplished by soaking the chip in acetone to
remove the remaining resist and the metal on top of it.
14
Figure 2-2. E-beam lithography process. (a) New Si/SiO2 chip. (b) Chip coated with
MMA and PMMA (c) Section of resist exposed in the SEM (d) After exposed section
is developed in MIBK. (e) Metal film deposited on chip (f) After lift-off. (Courtesy
Tobias Durkop)
15
A
B
Figure 2-3. Electron micrograph of the square alignment marker pattern (array of
small squares and symbols) and the electrodes (larger features leading off the edge of
the image) for the individual CNTs. The four large markers in the corners are used
during the second stage to allow for proper aligning of the two e-beam stages.
Nanotubes can be seen individually and forming star-shaped patterns radiating from
clumps of catalyst in the top middle of the image and the extreme lower right
indicated by letters A and B.
The alignment mark pattern can be used to locate CNTs to contact electrically.
The SEM can be used in this “find” step to locate nanotubes with reference to the
grid[36]. After the CNTs are located relative to the alignment marks, EBL resist must
16
then be spun and baked on the chip again and the computer assisted drawing (CAD)
program can be used to create electrode patterns that are referenced to the grid.
Typically the metal deposited on devices in our group is Cr/Au with 1nm of Cr and
30nm of Au.
The EBL technique allows for creating metallic electrodes at any given
separation (up to the length of the nanotube) whenever a CNT is found near the grid
markers, this can be seen in Fig. 2-3. The device should be re-imaged after the leads
are made to make sure that only one nanotube is in the junction area, as in Fig. 2-4,
due to the fact that CNTs are often more visible after being contacted by metal; some
CNTs may have been overlooked in the initial imaging.[36] At minimum, two
electrodes contact the CNT. The heavily doped silicon substrate under the SiO2 acts
as a third or “gate” electrode, creating a field-effect transistor (FET) geometry.
Satisfactory electrical contact to the gate can be made either by creating a scratch
somewhere on the surface of the chip and using an ultrasonic wire bonder to attach a
wire to the scratch or to silver paint placed on the scratch, or by contacting silver
paint that is touching the side of the chip. The wire bonder is also used to make
electrical connections to the lithographically-patterned electrodes on the chip.
17
Figure 2-4. Close up image of a CNT under contacts. The individual grid markers can
be seen as well as an extra tube that almost created a two-tube device. The devices
need to be imaged after creation to ensure that a single tube was contacted in the
junction.
2.4 Georgetown technique
The entire process outlined above is called the “find-and–wire” approach to
creating devices. If patterned catalyst is used, electrodes can be created aligned to the
catalyst, where one expects the CNTs to be, and then the chip can be inspected to see
if the CNTs were contacted by the electrodes. This has been called the “wire-and-
18
find” approach, and was used by the group of Paola Barbara at Georgetown
University[35] to create some of the devices studied in this thesis.
The wire-and-find technique utilizes a patterned growth method that allows
for the growth of nanotubes only at certain locations on the chip. The other attractive
feature of the process is that it is done solely with photolithography which removes
the need for an SEM (costly apparatus!). The first step involves creating a
photolithographic pattern for the catalyst islands. The chip is then immersed in
catalyst solution, which can only reach the surface of the chip through the
photolithographic mask. When the mask is removed with acetone, the catalyst is only
left in small islands on the chip’s surface. After growing CNTs, another
photolithography step is done by aligning to the first pattern. This step puts electrodes
down that are matched to the catalyst island locations. The electrodes for the CNT
samples used in this study were Pd/Nb metal. Pd has a high work function and a good
wetting interaction with the tubes[37] so it is a good choice for FET devices, the Nb
was deposited to allow for superconductivity measurements done by the Georgetown
group[38].
2.5 Experimental Setup
This thesis concerns electrical measurements made on CNT devices in the
field effect transistor (FET) configuration. The conductance measurements are made
in two-probe configuration which is acceptable given the devices’ high resistances
(>100 KΩ). It is also difficult to make four probe measurements of nanotube devices
19
due to their complicated interaction with contacts[39]. Commonly the electrons will
completely leave the CNT and enter the contact, making the flow of current through a
four-probe device more like three two-probe devices in series.
The DC drain and gate voltages are sourced from a National Instruments
BNC-2090 data acquisition device (DAQ), as depicted in Fig. 2-5. The source current
is measured by an Ithaco 1201 current preamplifier which converts the current to a
voltage that is measured by either a National Instruments board for telegraph signal
measurements or a Stanford Research Systems SR785 spectrum analyzer for lowfrequency noise measurements. The control of the DC voltages for the device and
gate bias and the spectrum analyzer and A/D board were accomplished using
programs created in LABVIEW.
All of the measurements were done inside of a Desert Cryogenics 4He flow
cryostat. The accessible temperature range is 1.2 K to 325 K.
20
Figure 2-5. Schematic of the electronic setup used in the cryostat. The computer
controls the bias and gate voltages as well as the data collection parameters. For lowfrequency noise the spectrum analyzer is triggered and the frequency span is set by
the computer, for telegraph signals the computer takes in time series from the A/D
board. Vbias is the bias voltage and IS-D in the current flowing from the source to the
drain.
21
Chapter 3 1/f noise
1/f noise is the part of the spectrum at low frequencies where the noise power
versus frequency approximately exhibits 1/f dependence. 1/f noise is found in such a
wide range of physical phenomenon that it cries out to many people for a universal
cause. How can traffic flow and ocean tides and symphonic music all have 1/f noise
inside? However we are left with the simple fact that despite its bizarre universal
presence no universal theory can account for it.[40] Even more discouragingly,
instances where theory and experiment are able to merge the most satisfactorily occur
where the mechanism clearly is of a non-universal nature[40].
For years 1/f noise in condensed matter systems was not commonly studied.
This attitude prevailed up until the 1970’s when some people noted that, despite the
advances of solid state physics, it was difficult to explain the noise that appeared in a
truly simple circuit comprised of a metal film and a voltage bias.[40] The spectral
current noise power of such a circuit has a well-understood frequency-independent
contribution from the thermal or Johnson noise that dominates at high frequencies,
however at low frequencies the noise typically exhibits a 1/f spectrum.
In this chapter I will explain the basics of noise and introduce 1/f noise. Then I
will present a theoretical framework within which the temperature dependence of the
1/f noise can be related to the energy spectrum of the fluctuators that are responsible
for the 1/f noise.
22
3.1 Introduction to noise
We first consider the circuit in Fig. 3-1 which depicts a sample to be tested
connected to a current amplifier.
When the voltage bias is zero the frequency spectrum of the voltage noise
across the resistor will be white (frequency independent) and have a magnitude
proportional to the magnitude of the resistance. This noise is called thermal or
Johnson noise and is caused by the thermally distributed velocities of the charge
carriers. It is found in any resistive element and in many other systems that can be
thought of as involving energy loss to a random process (e.g. water flow through a
pipe). Johnson noise is given by
S I ( f ) = 4k b T / R
(3.1)
Here kb is the Boltzmann constant and SI is the current noise power per unit frequency
(A2/Hz), T is the temperature and R is the resistance.
23
Figure 3-1. Noise circuit. This schematic illustrates the basic setup used for noise
measurements. The op-amp is shown in a trans-impedance configuration which
converts the current to a voltage for the spectrum analyzer to perform a fast-fourier
transform (FFT) on. The bottom graph is an illustration of the three main types of
noise to be expected in such circuits.
Once the voltage is non-zero two other types of noise appear. One is called
shot noise and is caused by the finite size of the electrical charge, which leads to
statistical fluctuations in the current crossing a junction, for example, electrons
moving through the leads connected to the nanotube sample. This noise is also white
and is commonly given by
24
S I ( f ) = 2eI
(3.2)
where e is the electron charge and I is the current through the device. The dominant
noise at low frequencies, however, will have a 1/f frequency dependence and its
nature is still poorly understood for the majority of systems.
Unlike Johnson noise or shot noise, there is no equation that is derived from
physical principles that can predict the magnitude of the 1/f noise in a conductor. A
strictly phenomenological equation that I will frequently use as a tool was proposed
by Hooge[41, 42]
SV ( f ) = AV 2+ β / f
z
(3.3)
In this equation, known as Hooge’s law, SV is the voltage noise power (V2/Hz), A is
the noise magnitude (and is dimensionless as long as β=0 and z=1), V is the bias
voltage and f is again the frequency where the noise is being measured. The quadratic
dependence of the noise on voltage indicates that the fluctuations are not current
driven but are actually due to 1/f fluctuations in the value of the resistance[40]. Hooge
set A= ζ /N, where N is the number of carriers, to attempt to create a universal
parameter, ζ, for 1/f noise where the size of the sample led to different values of A. If
β is zero and z is one, ζ is dimensionless.
Hooge proposed that ζ was a universal quantity that would describe many 1/f
processes in simple metals and semiconductors. Initial analysis was heartening:
Many semiconductors showed values of ζ near 2 x 10-3. Unfortunately it was found
that the value can vary greatly even amongst samples fabricated in the same
25
batch.[42] Choosing the value of N to use in the equation can also be difficult; i.e. it is
not clear how many of the carriers are participating in the noise process. It is also
hard to separate out the contact-dependent portion of the 1/f noise. However the
quadratic voltage dependence and the inverse dependence of the noise on the number
of carriers (noise inversely proportional to the volume of the system) are commonly
observed in 1/f systems. Bulk conductors commonly have a noise that is inversely
proportional to their volume and gated transistors can be seen to have a noise that
varies with the gate voltage, indicating that the number of charge carriers is
determining the noise magnitude.[43]
It should be noted that Hooge’s law is strictly phenomenological and that
many exceptions are known: 1/f noise processes can be current driven in some
systems, z may differ from one, and the noise can be a surface effect in some systems,
removing the N dependence[40]. Furthermore it is obvious at some very low
frequency the noise must stop following this behavior or the total noise power
integrated over all frequencies will diverge, which is unphysical. Hooge’s law does
however prove to be a useful tool in many situations and will be referred to
frequently.
3.2 Semiconductors and transistors
Pure homogeneous semiconductor materials have been studied extensively. In
semiconductors ζ values vary from 10-3 to 10-6 for Si, Ge, GaAs and other common
semiconductors that have been measured.[42] Unfortunately the uncertainty for any
26
given material is usually about one order of magnitude due to variability between
samples. Assuming that β@0 and z@1, we can see that the equation from Hooge, Eq.
3.3, becomes
S I ( f ) = AI 2 f =
ζ
N
I2 f
(3.4)
which is a powerful way to parameterize the noise for comparison purposes.
Measurements taken at different biases, currents or frequencies can be used for
comparing the magnitude of the noise.
In transistor devices a good deal of work has been done to discern whether or
not the source of the noise is fluctuations in carrier number or carrier mobility.[43]
Since the observable is conductance fluctuations and conductance is σ=µne, where µ
is the mobility and n is the number of carriers; it is not immediately clear which is the
source (or if both are the source), but each assumption makes a different prediction
for the way the noise should change with the gate voltage. For number fluctuations
the value of ζ should vary with gate voltage but for mobility fluctuations it should
remain constant. It appears that for many semiconductors, n type transistors exhibit
number fluctuations and p type transistors have mobility fluctuations.[43] In the next
chapter data and discussion will be presented for nanotube transistors.
3.3 Basic noise model
Dutta and Horn present a theory that connects the most commonly used model
for 1/f noise with the energy spectrum of the fluctuators responsible for the noise.[40]
27
I will start off by presenting the relevant beginning model and then will show the
modifications to it to make it more physically plausible. The process is outlined
pictorially in Fig. 3-2.
First I begin with the most basic of fluctuators, the two-level system. A twolevel fluctuator will have a Lorentzian spectrum[44]
S (ω ) ∝
τ
ω τ +1
(3.5)
2 2
where ω is the angular frequency and τ is the characteristic time of the process. If we
integrate this function over a distribution of two-level fluctuators we get
S ( w) ∝ ∫
τ
D (τ )dτ
ω τ 2 +1
(3.6)
2
where D(τ) is the density of states for the fluctuators. To clear up some confusing
notation,
dn
dn
, D' ( E ) ≡
dτ
dE
dE
D' ( E ) ×
= D(τ )
dτ
D(τ ) ≡
(3.7)
where n is the density of electrons. The density of states is an operator that takes a
derivative with respect to E or τ, this should not be taken to mean that E=τ.
Unfortunately by varying the distribution of fluctuators this equation can be
used to produce many kinds of frequency spectrums. Assuming that the fluctuators
are inhomogeneous, and in particular the they are distributed as
D (τ ) ∝ τ −1
(3.8)
leads to a noise spectrum that is
28
S (ω ) ∝ ω −1
(3.9)
We are now left with the problem of justifying this distribution of fluctuators. A first
step is to think of the fluctuators as being caused by a thermally activated process.
Then τ =τoexp(E/kBT) and the required energy distribution would be D (E) = const for
all energies. For example if noise in a semiconductor were caused by trapping and
detrapping in the oxide, which modulated the carrier density, we would expect this
kind of thermally activated process. Thus if over a wide range of energies all trap
energies were equally probable, we would have a consistent explanation.
Unfortunately these assumptions lead to a linear dependence of the spectral
noise power on temperature, which is usually not seen in semiconductors (or many
other conductors, namely most metals). The flat distribution of the energies of the
traps also cannot extend to arbitrarily low and high energies, which will be the topic
of the next section.
29
Figure 3-2. 1/f noise formation. This schematic depicts how the noise from many
two-level systems can be summed to create a 1/f spectrum. The transform of the
telegraph signal from a two level system is a Lorentzian, it has a flat spectrum
followed by a knee and then a 1/f2 dependence. If these are summed over an
appropriate distribution of energies the result is a 1/f spectrum.
3.4 Theory for temperature dependence
A better model for noise would not force the distribution of energies to be flat
for all values, a clearly non-physical requirement. To correct the unphysical nature of
the assumption of an infinitely wide and flat distribution of fluctuators being
30
responsible for the noise, Dutta and Horn inserted a spectrum of two level fluctuators
that is limited in energy distribution. This alteration removes the strictly 1/f nature of
the noise by creating an exponent that should range between 0.8 and 1.4 and should
also vary slowly with temperature and frequency[45, 46]. The most significant result
is that the spectrum of the fluctuators can be directly related to the noise power,
S (ω , T ) ∝
k BT
ω
D' ( E ′)
(3.10)
where E’=-kTln(ωτ0). This is the first term of a Taylor series expansion of Eq. 3.6,
with D (E ) allowed to be a smoothly varying function of E . This allows for the
observation of the energy of the fluctuator that corresponds to the noise features at a
given temperature through
E ' p ≈ −k BTp ln(ωτ 0 )
(3.11)
This shows that a maximum at any temperature value, Tp, is correlated with a
maximum in the energy of the fluctuators, E p. τo is the characteristic attempt time for
the fluctuators of order 10-14 s (i.e. inverse of a typical phonon frequency). The ln
term is of the order 30 for frequencies between 0.1 and 100Hz. The exponent of the
1/f noise also varies with frequency and temperature, but the deviation from unity is
proportional to 1/ln(2πfτo) and is therefore small and hard to measure experimentally.
Data from Dutta and Horn[40] is shown in Fig. 3-3 and illustrates the peak in noise
and then extracts a peak energy for the fluctuators that are responsible for that noise
in Fig. 3-3c.
31
Figure 3-3. Graph from Dutta and Horn[40] illustrating the temperature dependence
of the noise in a Ag metal film. The line is the theoretical prediction for the data and
the points are the experimental values. (a) The data exhibits a peak in noise at a
temperature around 400 K. (b) The theory also predicts a small shift in the power of
the 1/f dependence with temperature. The predicted shift is from 1/f 1.1 and 1/f 0.8. The
y-axis is the exponent, denoted as z in this chapter (c) This is the calculated density of
states corresponding to the noise data in (a). The noise peak corresponds to a peak in
the noise at an energy around 0.9 eV.
3.5 Previous results in nanotubes
Nanotubes present an interesting medium for studying 1/f noise for several
reasons.[13, 14, 31, 47-49] The strong sp2 bonded carbon atoms in the nanotube
32
lattice should not be able to move around easily, eliminating a source of noise
commonly considered in typical bulk metals. The nanometer diameter of the material
presents the first straightforward opportunity to measure the phenomenon of 1/f noise
in a 1-D conductor. The nanotube also has all of its atoms as surface atoms, which has
led to the prediction that one-dimensional systems should intrinsically have more 1/f
noise than their higher dimensional counterparts[14]. Nanotubes also have a onedimensional current, so any contaminant or adsorbate that interacts with an atom on
the tube is interrupting the flow of current. In contrast, in 3-D systems the removal of
a single atom in the bulk of the material will have a negligible impact on the
conductivity of the device. Finally 1/f noise is considered to be a bulk effect in most
materials due to its 1/N dependence on the number of carriers in the system - the most
famous paper in the field is titled “1/f Noise is no surface effect”[41] - but nanotubes
can be viewed as a material that is all surface. This has led to several papers on the
magnitude of the noise in individual carbon nanotube devices and some of the
relevant past work will be discussed here.
The first work on nanotube 1/f noise was from the Zettl group.[14] Their data
indicated that the noise in the devices was strongly connected with the total device
resistance for samples including bulk collections of CNTs (3D), “mats” or thin films
of CNTs (2D) and devices constructed from individual or perhaps small bundles of
CNTs (1D). They determined that A/R=10-11Ω-1 which for typical single tube devices
gives a value for A of 10-7. This value is extremely high, four to ten orders of
magnitude higher than that for most typical resistors. This led the group to conclude
that nanotubes may indeed be fulfilling the prediction that 1-D conductors would be
33
unusually noisy due to all of the atoms being surface atoms. The paper attempted to
create an estimate for γ by taking the number of carriers as the number of atoms. This
led to a value for γ =0.2 which is 100 times as high as Hooge proposed for
semiconductors, and up to 10,000 times as high as is commonly seen in high quality
semiconductor devices. This was further validation for the view that nanotubes (and
perhaps all 1-D systems) are exceptionally noisy, but the calculation of the number of
carriers as being equal to the number of carbon atoms in the device is probably
inappropriate; it would certainly overestimate the carrier number in semiconducting
CNTs.
In the next chapter I will present our group’s results on noise in nanotube
transistors, first at room temperature and then as a function of temperature.
34
Chapter 4 Temperature dependence of 1/f noise in carbon
nanotube transistors
This chapter will present the results of measurements performed on individual
semiconducting CNTs in FET geometry. The most extensive measurements were
taken on two FET devices provided by our Georgetown collaborators; the fabrication
procedure for these devices is described in Chapter 2.
In the first three sections I will present the analysis techniques used to extract
the noise parameters. In section 4, I will describe the results of the initial work done
by our group on CNTs in FET geometry at room temperature and all the data will be
from Ishigami et al[31]. In the last section I will discuss the behavior of the noise
parameter γ at temperatures from 1.2 to 300 K, and the implications the data has for
the origin of the 1/f noise in the devices.
As discussed in the last chapter, perfect 1/f noise would require a perfectly flat
spectrum of fluctuators D(E) at all energies E. If this were true the pre-factor γ in
equation 3.4 would show a linear dependence on temperature. However, if the
spectrum of fluctuators D(E) is smoothly varying, it results in a temperature
dependence for the 1/f noise with the same functional form as D(E), as indicated by
equation 3.10.
All the data presented in this chapter (except section 4.4 which is from a
separate set of devices) were taken on two CNT devices from the Georgetown group.
35
The data for all the plots in this chapter was taken on a 3 um long CNT with a
diameter of 1.4 nm and will be referred to in the thesis as Sample 1. Data from the
second device only appears in the final results in section 4.5 and is also 3um long
CNT with a diameter of 1.9 nm and will be referred to in the thesis as Sample 2.
Table 4-1. Device Characteristics
Device
Diameter (nm)
Contact Metal
Device Length
(µm)
Sample 1
1.4
Pd/Nb 3.5 nm / 215 nm
3
Sample 2
1.9
Pd/Nb 3.5 nm / 215 nm
3
4.1 Noise signal
Several methods are available to determine the noise parameter associated
with a given noise spectrum. As a reminder from the last chapter we want to
determine A where
SV ( f ) = AV 2 f
S I ( f ) = AI 2 f
z
(4.1)
z
if β=0. The two equations demonstrate the fluctuations can be measured as a function
of either electrical parameter, in this thesis the current noise is always being
measured. All of the data were generated on a spectrum analyzer that simply performs
an analog-to-digital conversion of the incoming signal and then performs a fast
Fourier transform (FFT) on the digital signal.
36
Figure 4-1. Dependence of noise on current bias. Numerator of the Hooge equation
plotted vs. current to check on the expected squared dependence of the noise. The
slope is 2.03 ± 0.03 on data from sample 1.
As a first check to see if our data follows Hooge’s law, I will show that the
noise spectral power is indeed quadratic in current, which is expected for resistance
fluctuations.
1
= f AI 2 = Bf
SI
(4.2)
37
Eq. 4.2 is a simplified version of Hooge’s law (Eq. 3.3) where β = 0 and z = 1. A plot
of 1/B vs. current should display an I2 dependence. This is done in Fig. 4-1 for sample
1 at 260 K and Vg = -8 V.
It is also useful to think about a form of Hooge’s law more suitable to
transistors,
z
S I ( f ) = AI 2 f =
ζ
N
I2 f z =
ζ
CGVG e
I2 f
z
(4.3)
where the last step is applicable if the bias voltage is held constant and the gate is
varied in the linear regime of the transistor, so that N=CGVG/e.
Next, I will turn to a careful examination of the frequency dependence; i.e.
can the frequency dependence be described as 1/f z? If z varies from unity then the
constant A is no longer dimensionless, which means variation in A will depend on the
frequency of the measurement. In many of the spectra I have taken at the same bias
temperature and bias voltage the exponent varies from 0.9 to 1.1 between scans at
different gate voltages, making it difficult to decide whether it is acceptable to ignore
the variation when trying to determine the prefactors A or ζ. Furthermore the data is
sometimes influenced by the presence of telegraph signal whose spectrum is
Si ( f ) =
bI 2
f
1 +
fo
2
(4.4)
where b is a constant. Hopefully the magnitude of this noise is small, or a region in
frequency space can be found where its effects are negligible. I will go through the
38
data analysis methods that I used to try to gain confidence that the trends observed in
the variation of A with temperature and carrier number are real.
4.2 Data analysis techniques
I attempted to analyze the noise spectra by several methods described below.
Each method assumes a certain functional form for the spectra, and so may introduce
systematic errors in the dependence of the noise pre-factor on temperature. I will
discuss the advantages and drawbacks of each method.
4.2.1 Power-law method
A first way to think about extracting the value of A for a given noise spectrum
is to plot the noise power vs. frequency on a log-log plot as in Fig. 4-2. Then the
slope d(lnSI)/d(lnf) gives the value of z, and the value of AI2 is the given by SI at f = 1
Hz. Since the value of z varies for different noise plots, the constant A is no longer
unitless. This may cause problems for the comparison of different spectra; for
example, the temperature dependence would in principle depend on the measurement
frequency. The value of A is also very sensitive to the errors in the slope.
39
Figure 4-2. Current noise versus frequency. The log-log plot allows for a linear fit to
extract the values of A and z for the spectrum. Here A = 2.67x10-8 and z = -1.11
4.2.2 Inverse noise method
Another useful way to display the data is to plot the reciprocal of the noise
power versus frequency as in Fig. 4-3. Since this should now be a proportional
relation if the exponent z = 1, the parameter A is straightforward to extract; the
reciprocal of the slope is AI2. The advantage of fitting the data with a line is that it
forces all of the plots to have z = 1 and therefore have a dimensionless A. This means
that comparison should be on a more equal footing. The difference between this and
40
the power law method, and a crucial factor to consider in fits in general, is that the
data will be weighted differently in these different methods. The default for most
fitting programs is to assume that there is a constant percentage uncertainty in the
data entered into the routine. Taking the reciprocal of the data without altering the
uncertainty will lead to different values for the fits.
Figure 4-3. Plot of 1/SI versus frequency. Here the frequency dependence becomes
linear and fitting a proportional relation to the data forces z =1. For this spectrum A =
1.36x10-8.
41
4.2.3 Inverse noise plus telegraph method
Some of the noise spectra appear to have a Lorentzian-like component, which
could be due to telegraph noise from a single fluctuator. If this is the case then it
would be desirable to account for the telegraph contribution by fitting a sum of the
telegraph noise spectrum and the 1/f spectrum to the data set[31], i.e.
AI 2
+
Si ( f ) =
f
BI 2
f
1 +
fo
2
(4.5)
I attempted this in the following manner. In the fit the value of the knee frequency in
the telegraph Lorentzian and the magnitude of the 1/f and telegraph noise are allowed
to vary but the value of z is fixed at 1. This is for two reasons: If the value of z is
allowed to vary the fitting process commonly fails to converge and the data needs to
be successfully fit over a large range of data for the exponent’s deviation from one to
be fit accurately. There is an additional difficulty that, in introducing an additional
component of the noise in the fit, that there will be a systematic reduction in the
magnitude of the 1/f noise obtained in the fit (because the best fit to the noisy data set
will always include some positive Lorentzian term). Figs. 4-4 and 4-5 show this
technique being applied to the reciprocal of the noise power and to the noise plotted
directly.
42
Figure 4-4. A nonlinear fit for 1/noise versus frequency. The added telegraph term
has a knee at 10 Hz and its effect is best seen by the fits bend at low frequencies.
43
Figure 4-5. Noise power versus frequency with a nonlinear fit. Same data as for the
reciprocal fit Fig. 4-4.
4.2.4 “Show all the data” method
Because of the difficulties discussed for the methods above, I developed a
new method that uses each data point of the spectrum as an independent measure of A
that is shown in Fig. 4-6. For each point I determine a value of 1/A = I2/fSI; this way
each data point in the spectrum produces a value for 1/A instead of the spectrum as a
whole. Then I can examine the dependence of 1/A at a particular frequency on gate
44
voltage and temperature, and determine whether there is a significant dependence on
the frequency.
Figure 4-6. “Show all the data” plot. a) Spectrum of 1/f noise from a CNT FET at a
bias voltage of 100 mV, a gate voltage of -8 V, and a temperature of 150 K, shown on
linear-linear scale (main panel) and log-log scale (inset). The solid line in the inset
indicates a slope of -1. b) Presentation of the noise spectra with the values
recalculated to give the value of the constant 1/A = (I2/fSI) at each frequency, as
discussed in the text. Colors indicate the value of the frequency the data is taken at
and the points are separated by 1 Hz.
45
4.3 Gate dependence
The observed gate dependence of the nanotube noise is best plotted as 1/A
versus gate voltage. Since the graph is linear it is easy to extract the value of γ from
z
AI 2 f =
ζ
N
I 2 f z ⇒ 1/ A =
CGVG e
ζ
(4.6)
This dependence indicates that the transistor is in the linear regime where the gate
voltage linearly increases the number of carriers in the nanotube. By comparing the
1/A vs. gate curve with the current vs. gate curve in Fig. 4-7 it is clear that the current
is not linear with gate voltage while the inverse of the noise magnitude is. This is an
indication that the data is described by Hooge’s law, as seen in Eq. 4.6 it is expected
that 1/A ∝ N. This is also strong evidence that the fluctuations are in mobility, not
carrier number.
Specifically, a model of random potential fluctuations[50] that has been used
to explain the noise in short, Schottky-barrier-dominated CNTs[51] predicts a much
stronger dependence of 1/A on Vg. Thus we can eliminate charge fluctuations in the
dielectric as a source of noise in our CNT devices, at least in the linear regime.
Data were taken at different temperatures to determine the evolution of ζ.
Another benefit of our use of ζ is that it also compensates for changes in the threshold
voltage at different temperatures and for the change in the Fermi energy versus gate
voltage; this is due to using the rate of change of 1/A vs Vg as seen in Eq. 4.6.
46
Figure 4-7. Reciprocal of the noise prefactor 1/A = I2/fSI (colored squares) and current
(filled squares) versus gate voltage for sample 1 at 150 K. Current data are taken with
drain voltage of 100mV. The 1/A data are color-coded according to frequency as in
Fig. 4-6. The open squares indicate the mean values of 1/A at each gate voltage, and
the dotted line is a linear fit to these points. The standard deviation of the mean for
these points is smaller than the size of the squares used to indicate the mean value.
Note that larger 1/A values correspond to less noise. A benefit of plotting the data
using this technique is that all of the data from the spectra are presented.
The main benefit of the show-all-the-data method is that all the data from the
noise spectra can be displayed in a 1/A versus gate voltage graph, and is all used to
47
determine the value of ζ. The method does assume implicitly that z = 1, however the
value is predicted to only vary slightly even for a non-constant D(E) as discussed in
Chapter 3. The particular frequency range does not significantly alter the magnitude
of the value of ζ. This can be seen by observing the color coding of the data. The
frequency range chosen does not affect the value of the slope obtained from the 1/A
vs Vg plot.
4.4 Noise in semiconducting devices at room temperature
Two papers on 1/f noise in individual semiconducting devices came out
concurrently in 2006, one from Maryland[31] and another from the IBM group[51].
Both papers revealed several significant aspects of the noise in CNTs and both found
that the reciprocal of the noise amplitude is linear with gate voltage. In this section I
will explain the results of the Maryland[31] paper and both of the plots, Figs. 4-8 and
4-9, are taken from that paper.
Assuming that the transistor is in the linear regime we again use Eq. 4.6. We
calculate the capacitance Cg = cgL, where cg is the gate capacitance of the device per
unit length, L is the length of the CNT and Vg is the applied gate voltage. For our
CNT devices, c g ≅
2πε 0ε sub
with εsub the dielectric constant of the oxide, t the
ln (2t / d )
thickness of the oxide and d is diameter of the CNT.
48
Figure 4-8. Noise data from Ishigami et al.[31] Measurements taken on devices at a
range of lengths to demonstrate that the source of the noise is the channel resistance
not the contact resistance. If the noise was being sourced at the contacts it would be
expected that the low length limit would be dominated by a contact term, while the
longer CNTs would be dominated by noise from the CNT. However, the behavior is
linear over the entire range indicating that the main source of noise is the CNT.
After seeing that the noise parameter 1/A varies linearly with the gate voltage
as I have also shown for my data in Fig. 4-7, the experiments verified that the noise
being measured in the two contact geometry is dominated by noise created in the
CNT and not noise from the contacts. This was done by plotting the quantity D =
49
cgL/eγ vs. length for several different devices with lengths ranging from 2 to 30 µm,
shown in Fig. 4-8. The linear behavior verifies that the noise is coming from the
fluctuations of the length-dependent resistance of the CNT. The value for ζ obtained
from this is 9.3 10-3, comparable to traditional FET devices. This means that
nanotubes are not excessively noisy; but since they do have far fewer carriers than
normal semiconductors individual CNT devices will have larger A values. All of this
data was taken in ultra-high vacuum (UHV).
Data were taken on the same device in UHV and ambient pressure to test the
prediction that physisorption of gases was a possible cause of noise in CNT. The
results of this indicate the reverse phenomenon from that expected if physisorption
were the source of noise: The CNT is actually noisier in UHV as shown in Fig. 4-9. It
is important to remember when looking at the graph that 1/A is the reciprocal of the
1/f noise magnitude, so larger values indicate less 1/f noise.
50
Figure 4-9. Comparison of inverse noise amplitude 1/A vs gate voltage Vg − Vth for
the same semiconducting CNT device in UHV and in air from Ishigami et al.[31] at
room temperature. The amplitude of the 1/f noise in air is three times smaller than in
UHV.
4.5 Temperature dependence
Fig. 4-10a shows the temperature dependence of ζ for two CNT devices.
Device 1 has a diameter of 1.4 nm, and Device 2 has a diameter of 1.9 nm. The
51
Hooge’s constant ζ, where A=ζ /N, has an exponential dependence on temperature
from 1.2 K to 150 K, with a change of about an order of magnitude, and is much less
temperature dependent at temperatures greater than 150 K. As described in the
previous chapter, we can use the temperature dependence of ζ to gain information
about the density of states of the fluctuators that are causing the 1/f noise. This is
done by using the Dutta and Horn result that Ep = -ln(2πfτ0)kB(T) ≈ 0.4 eV for f = 1
Hz and τ0 = 10-14 s (this value is introduced in Ch. 3 and Eq. 3.11)) and T =150 K.
The noise versus temperature data then indicated that the fluctuators responsible for
the 1/f noise are mostly at and above 0.4 eV. The Dutta and Horn model also connects
the exponential dependence of the noise to an exponential rise in the density of states
responsible for the 1/f noise.
52
Figure 4-10. Noise vs temperature (a) Temperature dependence of the Hooge
parameter ζ for two CNT devices. The data points are calculated using the slope from
<1/A> vs Vg, as shown in Fig. 4-7. The significant upward trend between 1.2 K and
about 150 K is seen in both samples. (b) Distribution of activation energies of the
fluctuators D(E) responsible for 1/f noise, calculated as described in text. Filled
squares and circles correspond to Device 1 and Device 2 respectively, in both (a) and
(b)
For another way to plot the data that allows for an easier identification of the
peak we can use a formula from the previous chapter,
S (ω , T ) ∝
k BT
ω
D′( E ′)
(3.10)
53
which means that ζ /T is proportional to the density of states, since ζ is also a measure
of the magnitude of the noise. This is plotted in Fig. 4-10b. This is the same data set
as in Fig. 4-10a, but this more clearly shows the location of the peak.
It is not surprising to see a spread between the two traces for the different
samples, even though they were prepared identically, with the same contacts and
similar lengths. Individual defect contributions for the two devices could be very
different, as random structural defects could vary greatly between the tubes. The local
density of defects in the oxide should also play a strong role in determining the
strength of the 1/f noise.
The main feature of Fig. 4-10 is the peak in D(E) at E @ 0.4 eV. This feature
is responsible for the majority of the room-temperature noise. The characteristic
energy scale allows us to rule out some possibilities for the source of the noise. The
energy scale is comparable to the bandgap (@ 0.5 eV and @ 0.37 eV for Devices 1 and
2 respectively) and therefore we can rule out electronic excitations (e.g. defect
ionization, etc.) within the CNT itself as the major noise source; such mechanisms
should have characteristic energies less than or equal to half the bandgap.
As
discussed above, we also rule out potential fluctuations due to the motion of charged
defects in the dielectric. Structural fluctuations of defects in the CNT lattice itself are
also ruled out, as they have very high characteristic energies. The energy is @ 10 eV
for Stone-Wales defect formation[52, 53] which involves one of the hexagons of the
lattice losing a carbon atom to become a pentagon.
Unfortunately the characteristic energy @ 0.4 eV does not provide enough
information to pinpoint what is causing the noise. However, the fact that the noise
54
magnitude is comparable to conventional MOSFETs suggests that the noise may in
fact result from similar processes in CNT-FETs, i.e. motion of defects in the dielectric
or at the dielectric/CNT (or dielectric/vacuum) interface. Still, other processes, such
as binding and unbinding of strongly physisorbed species cannot be ruled out; the
binding energies for CO2 and H2O, for example, lie in this range[51, 54]. Though our
measurements are carried out in helium gas with an extremely low partial pressure of
atmospheric components, it is possible that previously-adsorbed water is still present
on the SiO2 surface and could be responsible for the noise.
To summarize, we have measured the Hooge parameter ζ(Τ) at temperatures
between 1.2 K and 300 K. The room temperature value, ζ(300 Κ) ~ 10-3, we observe
is comparable to α(Τ = 300 Κ) found in traditional FETs indicating that CNT-FETs
are not afflicted by inherently large noise at room temperature. I use α(T) to estimate
the distribution of activation energies of the fluctuators D(E) responsible for the
noise; D(E) shows two features: a rise at low energy with no characteristic energy
scale, and a broad peak at energy of order 0.4 eV. By using the theory presented in
Chapter 3, I determined that the latter feature is responsible for the room temperature
noise. Electronic excitations and structural fluctuations within the CNT itself can be
ruled out as the source of this feature. Fluctuations within, or at the surface of, the
amorphous dielectric are likely responsible for the room temperature 1/f noise in
CNT-FETs on SiO2, though some physisorbed species (e.g. H2O, CO2) have similar
binding energies[54] and could be responsible for the room-temperature noise.
To further test whether the noise is coming from the oxide, the oxide layer
under a CNT could be etched away. A particularly illuminating experiment would
55
create four leads on one CNT, measure the noise in the two devices created this way,
and then etch out the oxide under one of the devices. This would eliminate tube to
tube variation and make sure that the etching process didn’t damage the CNT. Other
possibilities include testing devices on different substrates and treating the surface
with chemicals that should passivate the traps in the substrate.
As a result of the work here at Maryland[31] and IBM[51] it is now clear that
semiconducting CNT devices have a noise level very similar to that of traditional
semiconductors. The value for γ is in line with many other materials and devices, and
the high values for A obtained by early experiments[14] was merely an indicator of
the small number of electrons in the material, not an indicator of an extraordinarily
noisy material.
56
Chapter 5 Introduction to 1-D physics and telegraph signal
CNTs are an ideal laboratory for studying one-dimensional (1-D) electron
behavior. This behavior is expected to differ from that in three-dimensional systems
due to the inability of electrons to re-arrange themselves to minimize electronelectron interactions.[55-57] This chapter will begin by motivating the need for a new
description of the electron state in CNTs. Then it will introduce the technique we
intend to use to study the phenomenon in semiconducting CNTs. The chapter will
conclude by explaining how the hysteresis in CNTs makes it possible to uncover the
state of the electrons in semiconducting CNTs.
5.1 Drude and Luttinger
The initial successful description of electrons in solids was produced by
Drude[58, 59]. This model for electron behavior assumes that the electrons do not
interact at all with each other, termed the independent electron approximation. In fact
it is assumed that the electrons only interact with the ion cores through hard core
scattering processes, resulting in a characteristic scattering time and length for a given
conductor. This assumption proves to be very good for most metals where the
distance between electron-electron scattering events can be in the millimeter
range.[59] This is the origin of the term electron gas, since the electrons are behaving
like gas molecules in the ideal gas model. The addition of another electron, for
example through tunneling, to the electron gas is possible at the energy of the highest
57
occupied state, as the other electrons will easily be able to re-arrange themselves to
eliminate interactions. This means that there are excited states present just above the
highest occupied electron state; there is no energy gap in the density of states for
adding an electron, termed the tunneling density of states (TDS). Superconductors
are a good example where electron-electron interactions result in a correlated electron
state which exhibits an energy cost to add another electron to the system; adding an
unpaired electron to the superconductor requires giving the electron an additional
energy above the energy of the highest occupied state.[60] It is important to keep in
mind the distinction between the density of states D(E) for the system and the
tunneling density of states (TDS). The TDS measures the energy distribution of
excited states for the sudden addition of one electron to a system initially containing
N electrons, i.e. a transition from N → N+1 electrons, while D(E) corresponds to the
energy distribution of single-particle states of a system with N electrons. For noninteracting electron systems the two densities of states are equivalent, but for
interacting systems they can be very different. This chapter will deal will almost
exclusively with the TDS and not D(E) for electrons already in a system.
It may seem obvious that there will be situations in which ignoring the
electron-electron scattering is no longer feasible. Shrinking the number of dimensions
in the system should begin to cause problems to the concept of the electrons being
free from interacting with each other. A first modification of this theory is the FermiLiquid theory[59, 61, 62]. In this theory the electrons do have some small interaction
with each other; however, it is assumed that this interaction can be treated as a small
perturbation to the original free electron gas states. This results in the requirement of
58
using an effective mass to calculate the new wave-vectors and energies, but the TDS
is still free-electron like; there is a one-to-one correspondence between noninteracting electron states and the new interacting “quasiparticle” states.
In one dimension it seems reasonable that this approach should fail. The
electrons will have to interact strongly with each other as they are confined to stay
along the line defined by the 1-D conductor. In this situation Tomonaga and Luttinger
predicted that the electrons would form an interacting electron state where the
perturbative terms used in the Fermi-Liquid theory would diverge[63], this state is the
Luttinger liquid state[55-57] (Tomonaga actually originated the concept of the new
state, but only for a restricted set of conditions, Luttinger showed that it should occur
in any arbitrarily weakly interacting 1-D electron state). The Luttinger state was first
described in the 60’s but no experimental attempts to measure the signatures of this
state were successful (or at least published) until 1995 for constricted AlGaAs/GaAs
heterostructures[64-68] followed closely in 1999 for CNTs[69, 70].
As opposed to the independent electron assumption, perhaps visualized as a
few ping-pong balls bouncing around the Grand Canyon, the Luttinger liquid model
could be thought of as the executive desk toy, where each electron knows exactly
what the rest are doing, as depicted in Fig. 5-1. This will obviously create a different
TDS spectrum than in the previous model, as each electron will have formed a
coordinated lowest energy with all the other electrons in the system and each electron
will have to be disturbed in order for an extra electron to enter the system. The result
is that a finite energy is required to add an electron to the system; at T = 0 it is
59
impossible to add an extra electron to the system at exactly the Fermi level, and the
TDS has a power law behavior.[63]
Figure 5-1. Drude vs Luttinger. This is a visual depiction of the electron behavior in
the two different models. In the Fermi gas the electrons act independently from one
another, but in the Luttinger model electron-electron interactions should cause a
bosonic state to form that would alter the physics of the system.
This power-law behavior, TDS(E) ∂ (E-Ef)α is depicted in Fig. 5-2. This can
be measured experimentally in a tunnel junction between the Luttinger liquid and a
Fermi liquid or another Luttinger liquid; such a junction shows a power-law behavior
60
for the zero-bias conductivity versus temperature and differential conductance bias
voltage[63], G(T) ∂ Tα and dI/dV ∂ V α.
Figure 5-2. Tunneling density of states. The left graph depicts the availability of TDS
just above the Fermi energy in a system described by Fermi statistics. On the right is
a graph of the TDS for a Luttinger system, with its characteristic dip at energies near
the Fermi Energy. This TDS leads to tunneling observables for the zero bias
conductivity versus temperature G(T)~Tα, and for the conductivity versus bias
voltage, dI/dV~Vα.
There is a further complication for determining the TDS. The unitless
parameter α that describes the experimentally measurable effects is determined by the
61
interaction parameter g, a unitless variable that describes the amount of electron
interaction in the system. g should always be the same for a given one-dimensional
system, but α depends on the geometry of the experiment, i.e. whether the electron is
tunneling into the “end” of the one dimensional system or the “bulk”. This will be
illustrated in the next section where I describe the initial experiment on the Luttinger
state in CNTs. g ranges from 1 to 0 with smaller values indicating stronger
interactions.
5.2 Previous measurements in carbon CNTs
The original measurements performed on the Luttinger state in CNTs were
performed by Bockrath et al.[69] on metallic CNTs in two different geometries: metal
electrodes on top of or below the CNT. The significance of doing this is that different
tunneling behaviors are observed for the two contact situations. When metal leads are
first placed on the chip and then CNTs are placed on top of the leads, the device is
referred to as having bottom contacts. This contact geometry usually results in higher
contact resistance due to a weaker coupling between the CNT and the metal contact.
In essence the CNT is just resting on top of the metallic lead. This results in the
electrons having the opportunity to tunnel into any part of the CNT that is lying above
the contact, or the “bulk” of the CNT.
When the CNTs are first placed on the chip and then metal leads are created
on top of the CNTs, as in the devices used in this thesis, the devices are said to have
top contacts. In this situation the presence of the metal electrode “cuts” the CNT
62
electronically; electrons in the CNT impinging on the metal electrode have essentially
zero probability of continuing under the electrode in the CNT. Thus the electronic
current from electrode to CNT essentially remains entirely within the metallic lead
until it is forced at the last moment to exit the lead and tunnel into the CNT, and the
geometry approximates tunneling into the “end” of the Luttinger liquid. This picture
is born out by low temperature measurements of the charging energy of devices in
both configurations[71, 72]; CNTs with top contacts have energies determined by the
length of the CNT between the leads while CNTs with bottom contacts have a
charging energy determined by the entire length of the CNT. The equations for the
exponent in the two different geometries are[73, 74]
α end
α bulk
1
−1
g
=
4
1
+g−2
g
=
8
(5.1)
As a result of performing both temperature dependent (zero bias conductivity
vs. T), as in Fig. 5-3, and bias voltage dependent (dI/dV vs. Vbias) measurements on
CNTs of both geometries; Bockrath et al. were able to extract values for the
exponents in both geometries αbulk = 0.3-0.4 and αend = 0.5-0.7. The values are in good
agreement with theory which predicts g ≈ 0.28 and αbulk = 0.24 and αend = 0.65. Fig 53. is a plot from Bockrath et al.[69] depicting the power law behavior of the
conductivity, which allows for the extraction of the values for α and g. Later
experiments were able to see this behavior in crossed metallic CNTs[25, 75],
providing another example of bulk tunneling. Another experiment with a kinked
63
metallic CNT[25] saw behavior of end-end tunneling from one 1-D system to
another.
Figure 5-3. Plot from Bockrath et al.[69] showing the Luttinger liquid dependence of
the conductance G against temperature T. The plot on the left shows tunneling into
the bulk with the leads under the CNT and the plot on the right shows the opposite
scenario with the leads on top. The effects of the lead placement are discussed in the
text. The log-log plot shows the power-law dependence expected for Luttinger
liquids, with the solid lines representing the data and the dashed lines taking into
account a correction for Coulomb charging at low temperatures. Open circles in the
inset indicate α values for end contacted samples and crosses indicate values for bulk
contacted samples.
64
It might not be obvious why attaching 3-D metallic leads to a metallic CNT
results in a tunnel junction even at high temperatures. The exact nature of the barrier
in these and many other experiments on CNTs remains unclear. However, the fact
that some processes can produce contacts with no (or almost no) barrier[37, 76, 77]
indicates that the barrier is an extrinsic property of metal-CNT junctions. It is
fortuitous that this accidental barrier has good properties for studying the energydependent tunneling into CNTs; the barrier itself must have relatively energyindependent transmission.
Unfortunately semiconducting CNTs can’t be studied using the same
techniques. Semiconducting CNTs form contacts which are more complicated than
metallic CNTs, although a direct measurement in the vein of Bockrath et al. has been
tried on multi-walled CNTs[78]. Schottky barriers[79] may form for semiconducting
CNTs and have temperature, bias-voltage, and gate-voltage dependences of their
own. Semiconducting CNTs also can be doped by nearby contaminants. This doping
level will also have its own temperature dependence. These effects will mask the
possible Luttinger effects on tunneling dependence. The rest of this chapter will set
up a path to avoid the need to consider metallic contacts for probing the Luttinger
liquid in semiconducting CNTs.
Other experiments have also measured the Luttinger parameter with
photoemission studies on bundles of CNTs[70]. What is missing is a direct method of
measuring the tunneling in a single CNT device without mixing in the effects of the
65
contacts. In the next two sections I will outline how telegraph noise and hysteresis in
CNTs allow us to observe the tunneling of individual electrons into the CNT.
5.3 Hysteresis in CNTs
Many semiconducting CNTs in an FET geometry show hysteresis in the
current versus gate voltage curve.[80] This effect has been used to make memory
elements from CNT devices and, since it involves the transfer of electrons from traps
to the CNT, it is a useful tool to study tunneling into the electron system of the CNT.
First I will discuss how it was used as a memory device and how that indicated it
could be used for my purposes.
66
Figure 5-4. Semiconducting CNT hysteresis. This is a depiction of the stages of
hysteresis in the I-Vg curve for a CNT-FET. The charge traps around the CNT-FET
act as an additional gate for the circuit, and their long life creates a memory effect for
the current state of the device. As can be seen if the gate voltage is swept to a
negative value, holes will be present in nearby traps. If the gate voltage is returned to
zero then the current will be suppressed by the field created by these traps. The
opposite effect is seen if the gate voltage is swept to positive gate voltages.
As can be seen in Fig. 5-4, there is a large hysteresis in the I-Vg graph for
CNTs. To think about what the source of this might be it is useful to think about the
67
strength of the electric fields near the CNT when gate voltages of -10V are applied to
one of the devices.
E = V g KRt ln( ρ g ρ t )
(5.2)
Where ρt ~ 1 to 2 nm is the CNT radius, ρg
500 nm is the dielectric thickness, and
K = 3.9 is the dielectric constant of SiO2. This gives us field strengths in the range of
0.1 to 1 V/nm which is on the order of the breakdown field of SiO2 (about 0.2 V/nm,
but varies depending upon quality and growth technique for the SiO2). This suggests
that a likely explanation for the hysteresis is charge reordering from traps near the
CNT to the CNT.[80] It might be thought the charge rearrangement occurs between
two traps near the CNT and not actually with the CNT itself. However, this would
result in the hysteresis loop having the opposite sign: Positive gate voltage increasing
the threshold voltage indicates that the electrons are actually entering and leaving the
CNT. If the hysteresis were due to charges moving from trap to trap in the substrate
we would expect the opposite sign for the hysteresis loop.[80]
In effect, the moving charge is acting as an extra gate voltage, meaning that
the field the CNT is affected by is not just that applied through the gate, but also that
created by the charge dislocated from the traps. This means that the hysteresis is
caused by a number fluctuation. If instead the charges moving around created a
mobility change by altering the scattering process in the CNT we would expect a
completely different type of behavior to be seen in the I-Vg curves. Instead of having
a horizontal shift of the curves, the threshold voltage would remain constant and the
conductivity would shift up and down as the moving charges altered the mobility of
the device.
68
This can be used as a memory device by placing the gate voltage at Vg=0 since
the current state will depend upon whether the gate voltage was ramped up or down
to get to that point.[80] The high current and low current states are both very long
lived (~10,000s) another important attribute for a memory bit. The state can be
written, erased, read and rewritten repeatedly.
The important part of the story for my thesis is that this hysteresis implies that
there is tunneling occurring between two different number states of the CNT, and that
if this tunneling happens on an appropriate time scale we can measure the tunneling
rates to gauge the TDS in the CNT. This will be an indication of whether the
electrons in the semiconducting CNT are following the Luttinger liquid or Fermi gas
model.
5.4 Random telegraph signals
If the bias voltage and gate voltage are left constant, the same tunneling that
results in hysteresis can instead give rise to a random telegraph signal (RTS) as in
Fig. 5-5. This means that the system switches back and forth between (hopefully two)
discrete states. These sorts of signals have been used in the past to understand the
behavior of other novel electron systems.[81] Here I will discuss the concepts
necessary to proceed from the observation of a two-state RTS to an understanding of
the TDS in the CNT.
In particular RTS has been used to discern the electronic state of a transistor
built from a two-dimensional electron gas (2DEG)[81], with an electron assumed to
69
be tunneling from a defect to the two-dimensional gas which allowed for the
confirmation of the theoretical prediction for the electronic state of the electron
system. This experiment by Cobden et al.[81] was a direct inspiration for the work in
the next chapters of this thesis, not only due to their use of telegraph signal to provide
insight into the electron system of a novel material, but also in terms of understanding
the evolution of the TDS with temperature. In this 2-D system the TDS at the Fermi
level exhibits a maximum and follows a power law behavior with an exponent of
about -0.8, i.e. the tunneling rate is proportional to (Ef-Ed)-0.8. This is very analogous
to the Luttinger liquid case where the TDS vanishes at the Fermi level as a power
law, so it provides an immediate starting point for the theoretical analysis for the
semiconducting CNT situation.
70
Figure 5-5. Previous RTS experiment. Taken from Cobden et al.[81] γ1 and γ2 are the
rates for each state, Ed is the donor level and Ef is the Fermi energy. The upper left
image depicts the effect of shifting the gate voltage on the relative energy between the
defect and the Fermi energy and the semiconductor. The upper right image depicts the
telegraph signal for a single gate voltage, the average time in each state is used to
determine the transition rate. The data at the bottom depicts the rates for the switcher
at two temperatures, as described in text.
71
In Fig. 5-5 the lower left plot shows the data for a device at 1.2 K. The
detailed balance equation for a system at thermal equilibrium is
γ1
[−(E
=e
γ2
d
−E f
) kT ]
(5.3)
Here γ1 and γ2 are the rates for each state (the reciprocal of the mean lifetime for each
telegraph state), Ed is the defect energy, Ef is the Fermi energy and T is the
temperature. The straight line in the log plot shows the ratio of the rates for the two
states, this indicates that the system is at thermal equilibrium with the defect and that
the gate voltage is shifting the energy of the defect. Above this is the plot of the
individual rates for each of the states.
γ 1 = (2π / ) D∆2 f ( E d )
(5.4)
γ 2 = (2π / ) D∆2 [1 − f ( E d )]
Here D is the density of states and ∆ is the tunneling matrix. The data is fit using
Fermi statistics and Fermi’s golden rule shown in Eq. 5.4. The lower right plot shows
data at 0.5 K where the system is no longer obeying Fermi statistics and there is an
enhancement of tunneling at the Fermi energy. This allows for the extraction of
information about the amount of interaction of the electron system in the
semiconductor by fitting the data to a theory for the behavior of electrons in a 2DEG.
For nanotubes, I will insert a theory for tunneling into a Luttinger liquid in Chapter 7.
In CNTs telegraph signals have been observed by numerous groups[80, 82]. It
is not essential that these traps are located in the oxide as depicted in the figure; they
72
could also be in contaminants adsorbed on the surface of the chip, although the field
strength is perhaps circumstantial evidence suggesting that they are in the oxide. As
suggested above there are two reasons to suggest that the tunneling is actually
occurring between the CNT and the defect: first, the electric field is strongest closest
to the CNT, and second, the sign of the hysteresis loop indicates tunneling to and
from the CNT. In this experiment we observe gate voltage dependent tunneling rates
that follow that predicted for transition of an electron between two states in thermal
equilibrium.
To make this study we need a sample with a defect with energy close to the
Fermi level of the CNT and only one such defect. If there are several active defects
the switching will be amongst many states and become much more difficult to
interpret.
To discover whether a fluctuator can be isolated the device is cooled to the
base temperature, 1.2 K, where the fewest defects should be active and the gate
voltage is swept slowly. As the potential of the gate, Vg is varied, the defect energy,
Ed, is also varied with respect to the Fermi energy, Ef, of the CNT. At some gate
voltage switching of the current between two discrete states may be observed, as seen
in Fig. 5-6. These gate voltages cannot be chosen ahead of time since the technique
relies upon defects that are intrinsic to the device; they are not designed by the
experimenter.
73
Figure 5-6. Segment of an I-Vg curve taken on a CNT FET. Since the device was
swept in both directions the hysteresis is visible. There are two regions in the image
indicated by the arrows where two-level switchers are active, the left one of these is
further investigated in the later chapters.
74
Chapter 6 Random telegraph signals in carbon nanotubes
and their use as a defect thermometer
This chapter will characterize random telegraph signals in CNTs and show
that the signals may be used to extract the electron temperature in the CNT. The
dependence of the electron temperature on bias voltage is used to extract the energy
relaxation length lε in the CNT.
6.1 Random Telegraph Signals in Carbon Nanotube Transistors
The initial goal is to find a region of gate voltage in which the current displays
clear switching between two and only two states; i.e. that appears to be influenced by
only a single two-level fluctuator that can be studied in isolation over a range of bias
voltages and hopefully temperatures. The best fluctuator I was able to find showed
consistent two-level behavior from 1 mV to 100 mV in bias voltage and from
temperatures from 1.2 K to over 80 K. This was on sample 1 from the 1/f noise
section 4.5 and has a length of 3 µm and a diameter of 1.4 nm.
Fig. 6-1 depicts a section of an I-VG curve indicating the presence of a two
level fluctuator. At more negative Vg, it can be seen that the system prefers the high
current state with occasional switching events to the low current state. At
intermediate gate voltages (-8.22 < Vg < -8.18) both states are nearly equal in
occupation probability. As the gate voltage is swept more positive the lower-current
state is favored.
75
Figure 6-1. Small section of the current versus voltage curve for a two level
fluctuator. To visualize how the data was taken would require a third axis for time to
be shown at many of the gate voltage locations. It can obviously be seen that the
system is switching from preferring one state to the other with a section in the center
where both states are nearly equal in occupation probability.
This behavior suggests that the gate voltage controls the defect energy with
respect to the Fermi energy of the CNT, which affects the probability of finding the
system in one state or the other. This allows us to develop a model for the gated
defect-CNT system, as depicted in Figure 6-2 A-C. In this model, the defect lies in
76
the gate dielectric between the CNT and gate electrode, close enough to the CNT for
tunneling to occur. In Fig. 6-2A, when Vg is lowered (corresponding to higher
electron energy, or a rise in Fig. 6-2A), the defect chemical potential Ed is raised
relative to the CNT chemical potential EF. Likewise, when the Vg is raised, Ed is
lowered relative to EF. Thus the gate electrode controls the difference in chemical
potential of the defect and CNT:
E d − E f = −ηe(V g − V g 0 )
(6.1)
where Vgo is the gate voltage where Ed equals EF and η is the dimensionless gate
efficiency which represents the ability of the gate voltage to move the defect potential
with relation to the Fermi level in the CNT. The gate efficiency is less than unity due
to the capacitive coupling of the CNT and defect to the gate, source and drain.
Within this model, we identify the switching events between two states as the
stochastic process of electron tunneling between the defect and CNT. We can
analyze this process by recording the current as a function of time. Fig. 6-2D-F I
shows data for the same RTS depicted in Fig. 1, but now the gate voltage is kept fixed
while the current is recorded as a function of time. In Fig. 6-2D, Vg - Vgo is negative,
so the defect chemical potential is higher than the chemical potential of the CNT; this
corresponds to the diagram in Fig. 6-2A. Here the time trace of the RTS shows that
the system spends most of its time in the higher current state. Fig. 6-2E the defect is
at the Fermi energy so the system spends an equal amount of time in both states. Fig.
6-2F shows the opposite situation of 2D where the system now spends more time in
the other state since Vg - Vgo is positive.
77
Figure 6-2. RTS schematic. (A-C) Schematic of the band diagrams for the
semiconducting CNT, defect (in the SiO2), and gate electrode, for the conditions (A)
Ed - Ef > 0, (B) Ed - Ef = 0, and (C) Ed - Ef < 0. The defect is shown as being located
in the oxide but that is not essential to the physics. The thin solid lines indicate the
spectrum of excited states at zero temperature in a Lutinger liquid. The arrows
indicate the direction of the largest tunneling rate. (D-F) Time series of the current
through the CNT at three gate voltages which correspond to the diagrams (A-C). The
current fluctuates between two discrete states. As the gate voltage is changed the
relative tunneling rates between the two states change, resulting in the system
spending more or less time in the respective states. This is reflected in the time series
becoming more dominated by one current state or the other.
78
From time traces of the RTS as shown in Fig 2D-F, we can define two
tunneling rates γ1 and γ2 corresponding to an electron tunneling into and out of the
defect. We determine these rates by calculating the mean time spent in each state
(high current or low current) <t1>, <t2>, then γ1,2 = <t1,2>-1.
The experimental procedure is as follows. Once an isolated RTS fluctuator is
found, several time traces are taken. At a given temperature the voltage bias is set and
then time traces of 30 to 150s are taken at a constant gate and bias voltage. Then the
gate voltage is incrementally increased to take further time traces, with the fluctuator
slowly changing from predominantly one state to the other. This entire process is
then repeated at different bias voltages and different temperatures. As a reminder, the
bias voltage, gate voltage and temperature are all constant while the data is being
recorded. (This was also true for the 1/f noise experiments.) From the time traces the
average times <t1,2> spent in each state are calculated, along with the number of
switching events to gauge the statistical uncertainty. The reciprocal of the average
times <t1,2>-1 determines the switching rates γ1,2.
The first point to verify is whether the data satisfies the detailed balance
condition for a two level system:
γ1
[−(E
=e
γ2
d
−E f
) kT ]
,
(6.2)
,Using Eq. 6.1 above, we have:
γ1
[ηe (V −V ) kT ]
=e
.
γ2
g
g0
(6.3)
79
As will be shown below, once the gate efficiency, η, is known, Eq. 6-3 may also
be used to determine the electron temperature of the system.
Figure 6-3 shows the natural logarithm of the ratio of the rates ln(γ2/ γ1) as a
function of gate voltage for the same RTS studied in Figs. 6-1 and 6-2. To compare
with the exponential behavior predicted by Eq. 6.3, I plot the natural log of the ratio
of the two rates. The linear behavior of ln(γ2/ γ1) vs. Vg indicates Eq. 6-3 is obeyed.
From the slope of ln(γ2/ γ1) vs. Vg we extract the exponential prefactor -ηe/kT.
Figure 6-3. The natural log of the ratio of the tunneling rates versus gate voltage for
the same RTS as Figs. 6-1 and 6-2.
80
ln(
γ1
ηe
)=−
(Vg − Vg 0 )
kT
γ2
(6.4)
It is tempting to simply use the base temperature of the cryostat as T, and
therefore extract the gate efficiency η from the slope -ηe/kT in Figure 6-3. However,
it is necessary to ensure that the electron gas is in thermal equilibrium with the
substrate before performing this calculation. To measure whether the electrons are
being heated by the bias voltage, the slope is measured at many different bias
voltages and several base temperatures. At each bias voltage, I measure ln(γ2/ γ1) vs.
Vg and set the slope equal to -ηe/kT to extract a temperature T which I identify with
the electron temperature of the CNT. The gate efficiency η is chosen such that the
extracted T tends to the cryostat base temperature at low bias for cryostat
temperatures of 20 K, 40 K, and 80 K; this determines η = 0.053. The electron
temperature as a function of bias voltage is shown in Fig. 6-4. This demonstrates that
it is indeed important to consider the effect of heating of the electron system by the
bias voltage.
81
6.2 Defect thermometry
Figure 6-4. Electron temperature as a function of bias voltage at various substrate
temperatures. Electron temperature is determined from the logarithmic slope of the
tunneling rate ratio as a function of gate voltage as in Fig. 6-3. The gate voltage
efficiency η = 0.053 is chosen such that the low-bias electron temperature
extrapolates to the substrate temperature (solid colored lines). The rate of the
switchers drops for lower electron temperature data, limiting the range where data can
be taken.
As seen above in Fig. 6-4, the low-bias limit of the slope -ηe/kT taken at
different temperatures can be used to extract the gate efficiency η. However, the
82
slope -ηe/kT at higher bias can be used to determine the rise in temperature of the
electron system due to the influence of the bias voltage (and transport current). The
RTS acts as a “defect thermometer”; such a thermometer has been used previously to
study heating of the electron gas in metal wires under conditions of charge
transport[83-85].
As the CNT electron temperature deviates from the substrate temperature, the
main source of thermal resistance between the CNT electron system and the substrate
can be determined. The CNT electrons equilibrate to the substrate through two
effective thermal resistances in series: Λep, the thermal resistance of the CNT
electrons interacting with the phonons, and Λsub, the thermal resistance of the CNT
phonons interacting with the substrate. Different behaviors will result if one
resistance is dominant. For Λep < Λsub, Joule heating of the CNT should cause the
device temperature to rise above the substrate temperature, this should result in a ∆T
~ V2. However, as is typical in metals at low temperature, this behavior is not
observed in Fig. 6-5; the slope of T vs. V on this log-log plot is 1, implying T ∝ V1.
Interestingly, this indicates that the electrons do not achieve thermal equilibrium with
the phonons in the CNT at moderate biases (~40 mV) even at high temperatures (40
K). This is an indication of the very small electron-phonon coupling in CNTs, which
is partially due to their one-dimensional nature[86, 87].
This implies that the electron-phonon process is the bottleneck for thermal
transport from the electron system to substrate (i.e. Λep > Λsub); then the temperature
dependence typically exhibits a power law in voltage[83]. The electrons will gain
83
energy from the electric field over a distance called the energy loss length, lε and
Boltzmann transport theory predicts that the temperature of the electron system is[83]
kT = 0.780eElε
(6.5)
at high electric field, i.e. eElε >> kTsub where Tsub is the substrate temperature.
For constant energy relaxation length, the temperature rise of the electron
system is linear in bias voltage. I can then use the slope of T vs. V in Fig. 6-5 to
extract lε = 280 nm. I currently do not understand why the energy relaxation length is
constant; typically the energy relaxation length varies as a power-law in the electron
temperature lε ∝ Ta, where p = 2, 3, or 4 has previously been calculated[88]
depending on the dimensionality of the electron system (2 or 3) and phonon system (2
or 3). However, I am not aware of any calculations of the energy relaxation length
for CNTs or other 1-D systems.
The energy relaxation length lε may be used to extract an energy relaxation
time τε = lε/vF, where vF is the Fermi velocity. For a heavily doped semiconducting
CNT, vF approaches the value for a metallic CNT, 9.3 x 107 cm/s[29, 89, 90]. Then τε
~ 300 fs. This time is an upper bound to the coherence time for electrons in the CNT
(at least under the transport conditions probed in our experiment), so has implications
for use of CNTs in any quantum-coherent applications.
84
Chapter 7 Coherence and correlations in carbon nanotubes
studied using random telegraph signals
Chapter 5 described the previous work on Luttinger liquids (LL) and
discussed the techniques that have been used to study the LL state in carbon
nanotubes (CNTs) and similar correlations in other electron systems. Chapter 6
showed that random telegraph signals (RTS) in semiconducting CNT transistors
result from tunneling of an electron between CNT and a defect, probably located in
the gate dielectric. The gate voltage can be used to control the energy of this defect
relative to the Fermi energy of the CNT. The ratios of the tunneling rates as a
function of gate voltage were used to extract the electron temperature of the CNT
using the detailed balance relation. In this chapter I will study the gate voltage
dependence of the individual rates from random telegraph signals in CNTs and
analyze the data to arrive at a value for the Luttinger liquid interaction parameter g.
7.1 1-D electron behavior
The previous chapter analyzed the RTS in a CNT and examined only the ratio
of the tunneling rates as the gate voltage was swept. Here we will examine how the
individual tunneling rates change with gate voltage, which will allow us to test
whether the electrons are obeying Fermi gas behavior or if the electron system of onedimensional CNTs is better described by Luttinger liquid theory.
85
Varying the gate voltage varies the energy difference between the defect level
and the Fermi level of the nanotube, Ed - Ef according to Eq. 6.1. As the gate voltage
alters this difference, the defect level acts as a probe of the occupation probability and
tunneling density of states (TDS) of the system at that energy. The rates for tunneling
into and out of the system predicted by Fermi gas theory combined with Fermi’s
golden rule are
γ 1 = (2π / ) D∆2 f ( E d )
(7.1)
γ 2 = (2π / ) D∆2 [1 − f ( E d )]
where D is the tunneling density of final states, f is the Fermi function, and
overlap integral between initial and final states. D and
2
2
the
are assumed not to vary
with energy. I have shown previously in Chapter 6 (see Fig. 6.3) that Ed - Ef is
linearly related to the gate voltage, i.e. E d − E f = −ηe(V g − V g 0 ).
Plotting the individual rates vs. gate voltage there is a simple way to check if
Fermi statistics are being obeyed. Comparing the rates when the defect energy is near
the Fermi energy of the nanotube and when the defect level is far away from the
Fermi energy of the nanotube, we have for Fermi statistics
γ 1 ( Ed << E f ) / γ 1 ( Ed = E f ) = γ 2 ( Ed >> E f ) / γ 2 ( Ed = E f ) = 2
(7.2)
Stated in words, the rate of the switching at the edges of Fig. 6-1 should be twice the
rate at the point where the two data sets cross (this is also where Ed - Ef).
However the data in Fig. 7-1 shows that the ratio clearly exceeds two. This
indicates that Fermi statistics are not sufficient to explain the tunneling behavior into
the CNT device. This indicates that the tunneling at Ed - Ef is suppressed compared to
86
its expected value, an indication that Luttinger liquid theory may better describe the
phenomenon.
From the TDS at zero temperature one can create a corresponding tunneling
rate for a Luttinger liquid at zero temperature[63]
γ 1 ~ θ ( E f − E d )( E f − E d )α
(7.3)
the tunneling exponent α differentiates the LL from the Fermi gas which has a
uniform density of states just above the Fermi energy. This must be extended to finite
temperature[88, 91]
γ 1, 2
Ef − Ed Γ[(α + 1) / 2 + i ( Ed − Ef ) 2πkT ]
= CT exp(±
)
2kT
Γ(α + 1)
2
α
(7.4)
where the sign in the exponential switches for the two different rates. By
using E d − E f = −ηe(V g − V g 0 ) we have for the Luttinger case
ηe(Vg − Vg 0 ) Γ[(α + 1) / 2 + iηe(Vg − Vg 0 ) 2πkT ]
= CT exp(±
)
.
Γ(α + 1)
2kT
2
γ 1, 2
α
(7.5)
which is fitted to the data in Fig. 7-1, where the Luttinger fit provides a superior fit to
the data. This is due to the ability of the Luttinger model to take into account the
reduced tunneling rate for the situation where the defect energy is close to the Fermi
energy of the CNT. The only undetermined parameter in the fit is the value of α; the
temperature, T, and the gate efficiency, η, are determined by the fit to the natural
logarithm of the ratio of the rates for the two states versus gate voltage as in Fig. 6-3
in chapter 6.
87
Figure 7-1. A typical plot of the individual transition rates versus gate voltage for a
two-level RTS. The black curves are the fit to Fermi gas theory and the blue curves
are the fit to the Luttinger liquid model described in text with α = 2.
As discussed in chapter 5 α reveals the strength of the interactions of the
electrons in the Luttinger liquid system. The relation between α and the LL parameter
g should be
α bulk
1
+ g − 2
g
=
8
(5.1)
88
for our system, since the defect is tunneling into the bulk of the tube. The value for α
is extracted for several fluctuators in the last section of this chapter. This value is
called arate to indicate that the value is obtained at one temperature and bias voltage.
Remember that g can range from 1 to 0 with smaller values indicating stronger
interactions amongst the electrons in the system.
The value for α can also be calculated by taking the switching rate at the point
where Ef = Ed (equivalent to Vg = Vg0) and plotting it versus temperature. Eq. 7.5
becomes
γ (E f = E d ) = CT
α
Γ[(α + 1) / 2]
Γ(α + 1)
2
(7.6)
so that a plot of log rate vs. log temperature will yield the value of α as the slope. I
will call this αtemp.
89
Figure 7-2. Degenerate rate vs. electron temperature. The rate of switching when the
defect is at the Fermi energy vs. the temperature of the electron system. The
temperature of the electron system is calculated using the detailed balance condition
for the ratio of the rates. The power fit is used to give the value for α. Here
αtemp=0.97. The red dots indicate points taken at low bias voltages, while the black
points are from points where the tube is being heated by the bias voltage. (see Chapter
6 for details)
The temperature dependence follows a power law, with an exponent of αtemp =
0.97 ± 0.1. Analysis of the temperature dependence of another RTS gives a power
law exponent of αtemp = 0.7 ± 0.1. Note that the expected behavior for tunneling into
90
a Fermi liquid would be an absence of temperature dependence; the strong
temperature dependence in Figure 7-2 is in itself evidence for non-Fermi liquid
behavior. The data of Figure 7-2 are poorly fit by an activated (Arrhenius)
temperature dependence, and such a fit results in an unphysically low activation
energy on order of 2 meV.
The theoretical value for g is given for any system by[69, 73, 74]
2U
g = 1 +
∆
−1 2
(7.7)
where U is the Coulomb charging energy and ∆ is the single particle level spacing.
For a semiconducting CNT the level spacing is a function of the Fermi energy. For a
metal the calculation yields a theoretical value of g = 0.28 and therefore αbulk = 0.24
[71, 72]. Therefore we need to determine how the level spacing for the
semiconducting CNT will vary versus gate voltage to replace the U/∆ ratio for the
metallic nanotubes.
Since the charging energy will be the same for either metallic or
semiconducting
U=
e2
2C g
(7.8)
and the single particle level spacing is
∆=
dE ∆N
1
for ∆N = 1
=
dn L
D( E ) L
(7.9)
to find g all that is now needed is the density of states for the two systems. The
energy of the electrons in a metallic CNT is E = ( vfk)2, where
is Planck’s constant,
vf = 9x107cm/s is the Fermi velocity and k = πn/4 is the wave vector, where n is the
91
number of carriers. So the density of states is D(E) = dn/dE=(dE/dn)-1=4/π vf. For a
metal the calculation yields a theoretical value of g = 0.28 and therefore α = 0.24. For
the semiconducting case the bandgap must be taken into account, and the new
dispersion relation is approximated as hyperbolic E2 = δ2 + ( vfk)2, where δ is the
bandgap. This makes the density of states
D ( E ) semi =
( v π)
2
f
(δ + ( v πn 4) )
2 1/ 2
2
16
f
n
(7.10)
where the bandgap is 0.59eV for a tube with a diameter of 1.4nm. If I assume that the
number of carriers is linear with the gate voltage, n = CgVg/e, I can plot how the
parameters g and α should vary with gate voltage for a semiconducting CNT by using
eqs. 5.1 and 7.10. This assumes that the only alteration required to Luttinger theory
when switching from metallic to semiconducting CNTs is to take into account the
new density of states.
Figure 7-3 plots the expected variation of the LL tunneling exponent α with
gate voltage, as well as my experimentally-determined values of α from analysis of
the gate-voltage dependence and temperature dependence of the individual tunneling
rates. Fig 7-4 is the corresponding plot for the value of g assuming bulk tunneling.
The values of α determined from experiment are significantly higher than the
expected values. There are several possible explanations for this. First, it is quite
possible that the simple analysis above overestimates g and underestimates α. A
more careful analysis by Egger and Gogolin[73] gives g = 0.18 for a 3 µm length
metallic CNT, corresponding to α = 0.46, in good agreement with photoemission
experiments on metallic CNTs[70]. This would result in a nearly doubled estimate of
92
the semiconducting α compared to the values plotted in Figure 7-3. Second, our
analysis also neglected any interaction between the electron system and the defect
itself, which seems reasonable, since the typical change in resistance upon charging
and discharging the defect is on order several kOhms, corresponding to a change in
transmission on order 1/2. However, the backscattering of electrons by the defect
itself may cause correlations in the electron system (this is the essence of the work by
Cobden et al.[81]). More theoretical work is needed to understand whether this is
relevant in the CNT case. Third, the interactions in semiconducting CNTs may
simply be stronger than expected, for reasons not yet elucidated.
93
Figure 7-3. Depiction of the theoretical calculation of the Luttinger parameter α
versus gate voltage. The red points are from two fluctuators on sample 1 and the
black dot is from a fluctuator on sample 2. The squares indicate values obtained from
fitting the individual rates vs gate voltage while the triangle points were obtained
from fitting the rate vs. temperature. (both methods explained in text above) The
details are given in the table below. The difference in the theoretical curves is due to
the different diameters of the tubes, which results in a different band gap.
94
Figure 7-4. Depiction of the theoretical calculation of the Luttinger parameter g
versus gate voltage. The red points are from two fluctuators on sample 1 and the
black dot is from a fluctuator on sample 2. The squares indicate values obtained from
fitting the individual rates vs gate voltage while the triangle points were obtained
from fitting the rate vs. temperature. (both methods explained in text above) The
horizontal line indicates the value for metallic CNTs. The difference in the theoretical
curves is due to the different diameters of the tubes, which results in a different band
gap.
95
Table 7-1. Luttinger parameter α for the three RTS.
Fluctuators
Diameter (nm)
Vthreshold-Vg0
αtemp
αrate
Sample 1
1.4
5.0
0.5 +/- 0.1
0.9 +/- 0.1
Sample 1
1.4
3.2
2.2 +/- 0.7
0.7 +/- 0.1
Sample 2
1.9
3.9
0.8 +/- 0.1
NA
Table 7-1 gives the details for the RTS signal studied in this chapter. The first
two fluctuators are on the same device but at different gate voltages, with the first
fluctuator having the largest range of observable fluctuations with respect to
temperature. The last fluctuator was not stable over a wide enough range of
temperatures to extract a fit for the change in rate vs. temperature.
In conclusion I have analyzed the temperature, bias voltage, and gate voltage
dependence of the random telegraph signal resulting from an electron tunneling
between a semiconducting carbon nanotube and a nearby defect. The RTS is used as
a sensitive probe of the tunneling density of states of the Luttinger liquid state of the
semiconducting CNT. We show that the tunneling rate is strongly suppressed at the
Fermi level, consistent with Luttinger liquid theory confirming the more strongly
interacting nature of electrons in semiconducting CNT relative to metallic CNT. Our
value of g<0.2 indicates that the electrons in semiconducting CNTS are interacting
more strongly than the electrons in metallic CNTs.
96
Abbreviations
1-D One dimensional
2DEG Two-dimensional electron gas
A/D Analog to digital
CAD Computer assisted drawing
CNT Carbon nanotube
CVD Chemical vapor deposition
DAQ Data acquisition
EBL Electron beam lithography
FET Field effect transistor
IPA isopropanol
LL Luttinger liquid
MIBK methylisobutylketone
MMA methylmethacrylate
MOSFET Metal-oxide-semiconductor field-effect transistor
PMMA polymethylmethacrylate
RTS Random telegraph signal
SEM Scanning electron microscope
TDS Tunneling density of states
UHV Ultra-high vacuum
97
Symbols
a
graphite lattice constant
a1,a2
graphene unit vectors
A
noise magnitude
β
correction to V2 dependence of noise
CG
gate capacitance
∆
bandgap
D( )
density of states operator
d
CNT diameter
e
electron charge
E
energy
γ0
tight-binding integral
γ1 , γ2
rates into and out of an RTS system
ħ
Planck’s constant
I
current
k
wave-vector
kb
Boltzmann constant
K
K point
f
frequency
n
density of electrons
N
number of electrons
98
q
wave vector from K point
R
rolling vector
R
resistance
SI
current noise power
SV
voltage noise power
T
temperature
τ
characteristic time of fluctuator
V
voltage
Vg
gate voltage
Vsd
source-drain voltage
Vth
threshold gate voltage (gate voltage where the device begins to conduct)
vf
Fermi velocity
ω
angular frequency
z
exponent for 1/fz noise (z close to 1)
ζ
Hooge noise parameter
99
Bibliography
[1]
Iijima, S., Helical Microtubules of Graphitic Carbon. Nature 354, 56-58
(1991).
[2]
Dai, H., A.G. Rinzler, P. Nikolaev, A. Thess, D.T. Colbert, and R.E. Smalley,
Single-wall nanotubes produced by metal-catalyzed disproportionation of
carbon monoxide. Chemical Physics Letters 260, 471-475 (1996).
[3]
Thess, A., R. Lee, P. Nikolaev, H.J. Dai, P. Petit, J. Robert, C.H. Xu, Y.H.
Lee, S.G. Kim, A.G. Rinzler, D.T. Colbert, G.E. Scuseria, D. Tomanek, J.E.
Fischer, and R.E. Smalley, Crystalline ropes of metallic carbon nanotubes.
Science 273, 483-487 (1996).
[4]
Bethune, D.S., C.H. Klang, M.S. de Vries, G. Gorman, R. Savoy, J. Vazquez,
and R. Beyers, Cobalt-catalysed growth of carbon nanotubes with singleatomic-layer walls. Nature 363, 605-607 (1993).
[5]
Zheng, L.X., M.J. O'Connell, S.K. Doorn, X.Z. Liao, Y.H. Zhao, E.A.
Akhadov, M.A. Hoffbauer, B.J. Roop, Q.X. Jia, R.C. Dye, D.E. Peterson,
S.M. Huang, J. Liu, and Y.T. Zhu, Ultralong single-wall carbon nanotubes.
Nature Materials 3, 673-676 (2004).
[6]
Hong, B.H., J.Y. Lee, T. Beetz, Y.M. Zhu, P. Kim, and K.S. Kim, Quasicontinuous growth of ultralong carbon nanotube arrays. Journal of the
American Chemical Society 127, 15336-15337 (2005).
100
[7]
Miaudet, P., S. Badaire, M. Maugey, A. Derre, V. Pichot, P. Launois, P.
Poulin, and C. Zakri, Hot-drawing of single and multiwall carbon nanotube
fibers for high toughness and alignment. Nano Letters 5, 2212-2215 (2005).
[8]
Yakobson, B.I. and R.E. Smalley, Fullerene nanotubes: C-1000000 and
beyond. American Scientist 85, 324-337 (1997).
[9]
Beyer, G., Flame retardancy of nanocomposites based on organoclays and
carbon nanotubes with aluminium trihydrate. Polymers for Advanced
Technologies 17, 218-225 (2006).
[10]
Baughman, R.H., A.A. Zakhidov, and W.A. de Heer, Carbon Nanotubes--the
Route Toward Applications. Science 297, 787-792 (2002).
[11]
Hertel, T., R.E. Walkup, and P. Avouris, Deformation of carbon nanotubes by
surface van der Waals forces. Physical Review B 58, 13870 (1998).
[12]
O. Breuer, U.S., Big returns from small fibers: A review of polymer/carbon
nanotube composites. Polymer Composites 25, 630-645 (2004).
[13]
Collins, P.G., K. Bradley, M. Ishigami, and A. Zettl, Extreme oxygen
sensitivity of electronic properties of carbon nanotubes. Science 287, 18011804 (2000).
[14]
Collins, P.G., M.S. Fuhrer, and A. Zettl, 1/f noise in carbon nanotubes.
Applied Physics Letters 76, 894-896 (2000).
[15]
Banerjee, I.A., L. Yu, and H. Matsui, Location-Specific Biological
Functionalization on Nanotubes: Attachment of Proteins at the Ends of
Nanotubes Using Au Nanocrystal Masks. Nano Lett. 3, 283-287 (2003).
101
[16]
Kohli, P., C.C. Harrell, Z. Cao, R. Gasparac, W. Tan, and C.R. Martin, DNAFunctionalized Nanotube Membranes with Single-Base Mismatch Selectivity.
Science 305, 984-986 (2004).
[17]
Hamada, N., S. Sawada, and A. Oshiyama, New One-Dimensional
Conductors - Graphitic Microtubules. Physical Review Letters 68, 1579-1581
(1992).
[18]
Mintmire, J.W., B.I. Dunlap, and C.T. White, Are Fullerene Tubules Metallic.
Physical Review Letters 68, 631-634 (1992).
[19]
Tans, S.J., A.R.M. Verschueren, and C. Dekker, Room-temperature transistor
based on a single carbon nanotube. Nature 393, 49-52 (1998).
[20]
Odom, T.W., J.L. Huang, P. Kim, M. Ouyang, and C.M. Lieber, Scanning
tunneling microscopy and spectroscopy studies of single wall carbon
nanotubes. Journal of Materials Research 13, 2380-2388 (1998).
[21]
Postma, H.W.C., T. Teepen, Z. Yao, M. Grifoni, and C. Dekker, Carbon
nanotube single-electron transistors at room temperature. Science 293, 76-79
(2001).
[22]
Durkop, T., S.A. Getty, E. Cobas, and M.S. Fuhrer, Extraordinary mobility in
semiconducting carbon nanotubes. Nano Letters 4, 35-39 (2004).
[23]
Kane, C.L. and M.P.A. Fisher, Transport in a One-Channel Luttinger Liquid.
Physical Review Letters 68, 1220-1223 (1992).
[24]
Bockrath, M., D.H. Cobden, J. Lu, A.G. Rinzler, R.E. Smalley, T. Balents,
and P.L. McEuen, Luttinger-liquid behaviour in carbon nanotubes. Nature
397, 598-601 (1999).
102
[25]
Yao, Z., H.W.C. Postma, L. Balents, and C. Dekker, Carbon nanotube
intramolecular junctions. Nature 402, 273-276 (1999).
[26]
Souza, A.G., A. Jorio, J.H. Hafner, C.M. Lieber, R. Saito, M.A. Pimenta, G.
Dresselhaus, and M.S. Dresselhaus, Electronic transition energy E-ii for an
isolated (n,m) single-wall carbon nanotube obtained by anti-Stokes/Stokes
resonant Raman intensity ratio. Physical Review B 63, 241404 (2001).
[27]
Dresselhaus, M.S., G. Dresselhaus, and P.C. Eklund, Science of fullerenes and
carbon nanotubes. 1996, San Diego: Academic Press.
[28]
Saito, R., G. Dresselhaus, and M.S. Dresselhaus, Physical properties of
carbon nanotubes. 1998, London: Imperial College Press.
[29]
Wallace, P.R., The Band Theory of Graphite. Physical Review 71, 622-634
(1947).
[30]
Martel, R., T. Schmidt, H.R. Shea, T. Hertel, and P. Avouris, Single- and
multi-wall carbon nanotube field-effect transistors. Applied Physics Letters
73, 2447-2449 (1998).
[31]
Ishigami, M., J.H. Chen, E.D. Williams, D. Tobias, Y.F. Chen, and M.S.
Fuhrer, Hooge's constant for carbon nanotube field effect transistors. Applied
Physics Letters 88, 203116 (2006).
[32]
Zhenhui, K., W. Enbo, M. Baodong, S. Zhongmin, C. Lei, and X. Lin,
Obtaining carbon nanotubes from grass. Nanotechnology 16, 1192 (2005).
[33]
Kong, J., A.M. Cassell, and H.J. Dai, Chemical vapor deposition of methane
for single-walled carbon nanotubes. Chemical Physics Letters 292, 567-574
(1998).
103
[34]
Kim, W., H.C. Choi, M. Shim, Y.M. Li, D.W. Wang, and H.J. Dai, Synthesis
of ultralong and high percentage of semiconducting single-walled carbon
nanotubes. Nano Letters 2, 703-708 (2002).
[35]
Tselev, A., K. Hatton, M.S. Fuhrer, M. Paranjape, and P. Barbara, A
photolithographic process for fabrication of devices with isolated singlewalled carbon nanotubes. Nanotechnology 15, 1475 (2004).
[36]
Brintlinger, T., Y.F. Chen, T. Durkop, E. Cobas, M.S. Fuhrer, J.D. Barry, and
J. Melngailis, Rapid imaging of nanotubes on insulating substrates. Applied
Physics Letters 81, 2454-2456 (2002).
[37]
Javey, A., J. Guo, Q. Wang, M. Lundstrom, and H. Dai, Ballistic carbon
nanotube field-effect transistors. Nature 424, 654-657 (2003).
[38]
Zhang, J., A. Tselev, Y. Yang, K. Hatton, P. Barbara, and S. Shafraniuk, Zerobias anomaly and possible superconductivity in single-walled carbon
nanotubes. Physical Review B (Condensed Matter and Materials Physics) 74,
155414 (2006).
[39]
Gao, B., Y.F. Chen, M.S. Fuhrer, D.C. Glattli, and A. Bachtold, Four-Point
Resistance of Individual Single-Wall Carbon Nanotubes. Physical Review
Letters 95, 196802-4 (2005).
[40]
Dutta, P. and P.M. Horn, Low-frequency fluctuations in solids: 1/f noise.
Reviews of Modern Physics 53, 497 (1981).
[41]
Hooge, F.N., 1/f noise is no surface effect. Physics Letters A 29, 139-140
(1969).
104
[42]
Hooge, F.N., 1/f noise sources. Electron Devices, IEEE Transactions on 41,
1926-1935 (1994).
[43]
Vandamme, L.K.J., L. Xiaosong, and D. Rigaud, 1/f noise in MOS devices,
mobility or number fluctuations? Electron Devices, IEEE Transactions on 41,
1936-1945 (1994).
[44]
Machlup, S., Noise in Semiconductors: Spectrum of a Two-Parameter
Random Signal. Journal of Applied Physics 25, 341-343 (1954).
[45]
Dutta, P., P. Dimon, and P.M. Horn, Energy Scales for Noise Processes in
Metals. Physical Review Letters 43, 646 (1979).
[46]
Verleg, P.A.W.E. and J.I. Dijkhuis, Resistance fluctuations in hydrogenated
amorphous silicon: Thermal equilibrium. Physical Review B 58, 3904 (1998).
[47]
Snow, E.S., J.P. Novak, M.D. Lay, and F.K. Perkins, 1/f noise in singlewalled carbon nanotube devices. Applied Physics Letters 85, 4172-4174
(2004).
[48]
Ouacha, H., M. Willander, H.Y. Yu, Y.W. Park, M.S. Kabir, S.H.M. Persson,
L.B. Kish, and A. Ouacha, Noise properties of an individual and two crossing
multiwalled carbon nanotubes. Applied Physics Letters 80, 1055-1057 (2002).
[49]
Lin, Y.M., J. Appenzeller, J. Knoch, Z. Chen, and P. Avouris, Low-Frequency
Current Fluctuations in Individual Semiconducting Single-Wall Carbon
Nanotubes. Nano Lett. 6, 930-936 (2006).
[50]
Tersoff, J., Low-Frequency Noise in Nanoscale Ballistic Transistors. Nano
Lett. 7, 194-198 (2007).
105
[51]
Kingrey, D., O. Khatib, and P.G. Collins, Electronic Fluctuations in Nanotube
Circuits and Their Sensitivity to Gases and Liquids. Nano Lett. 6, 1564-1568
(2006).
[52]
Dumitrica, T. and B.I. Yakobson, Strain-rate and temperature dependent
plastic yield in carbon nanotubes from ab initio calculations. Applied Physics
Letters 84, 2775-2777 (2004).
[53]
Zhao, Q., M.B. Nardelli, and J. Bernholc, Ultimate strength of carbon
nanotubes: A theoretical study. Physical Review B 65, 144105 (2002).
[54]
Ulbricht, H., R. Zacharia, N. Cindir, and T. Hertel, Thermal desorption of
gases and solvents from graphite and carbon nanotube surfaces. Carbon 44,
2931-2942 (2006).
[55]
Tomonaga, S., Prog. Theor. Phys 5, 544 (1950).
[56]
Luttinger, J.M., Fermi Surface and Some Simple Equilibrium Properties of a
System of Interacting Fermions. Physical Review 119, 1153 (1960).
[57]
Luttinger, J.M., An Exactly Soluble Model of a Many-Fermion System.
Journal of Mathematical Physics 4, 1154 (1963).
[58]
Drude, P., Annalen der Physik 1, 566 (1900).
[59]
Ashcroft, N.W. and N.D. Mermin, Solid State Physics. 1976: Cornell
University.
[60]
Tinkham, M., Introduction to Superconductivity. 1975: Krieger Pub. Co.
[61]
Landau, L.D., Sov. Phys. JETP 3, 920 (1957).
[62]
Pines, S., P. Nozieres, and W.A. Benjamin, The Theory of Quantum Liquids I.
1966, Menlo Park, California.
106
[63]
Voit, J., One-dimensional Fermi liquids. Reports on Progress in Physics 58,
977 (1995).
[64]
Tarucha, S., T. Honda, and T. Saku, Reduction of quantized conductance at
low temperatures observed in 2 to 10 [mu]m-long quantum wires. Solid State
Communications 94, 413-418 (1995).
[65]
Yacoby, A., H.L. Stormer, N.S. Wingreen, L.N. Pfeiffer, K.W. Baldwin, and
K.W. West, Nonuniversal Conductance Quantization in Quantum Wires.
Physical Review Letters 77, 4612 (1996).
[66]
Milliken, F.P., C.P. Umbach, and R.A. Webb, Indications of a Luttinger liquid
in the fractional quantum Hall regime. Solid State Communications 97, 309313 (1996).
[67]
Chang, A.M., L.N. Pfeiffer, and K.W. West, Observation of Chiral Luttinger
Behavior in Electron Tunneling into Fractional Quantum Hall Edges. Physical
Review Letters 77, 2538 (1996).
[68]
Grayson, M., D.C. Tsui, L.N. Pfeiffer, K.W. West, and A.M. Chang,
Continuum of Chiral Luttinger Liquids at the Fractional Quantum Hall Edge.
Physical Review Letters 80, 1062 (1998).
[69]
Bockrath, M., D.H. Cobden, J. Lu, A.G. Rinzler, R.E. Smalley, L. Balents,
and P.L. McEuen, Luttinger-liquid behaviour in carbon nanotubes. Nature
397, 598-601 (1999).
[70]
Ishii, H., H. Kataura, H. Shiozawa, H. Yoshioka, H. Otsubo, Y. Takayama, T.
Miyahara, S. Suzuki, Y. Achiba, M. Nakatake, T. Narimura, M. Higashiguchi,
K. Shimada, H. Namatame, and M. Taniguchi, Direct observation of
107
Tomonaga-Luttinger-liquid state in carbon nanotubes at low temperatures.
Nature 426, 540-544 (2003).
[71]
Bockrath, M., D.H. Cobden, P.L. McEuen, N.G. Chopra, A. Zettl, A. Thess,
and R.E. Smalley, Single-Electron Transport in Ropes of Carbon Nanotubes.
Science 275, 1922-1925 (1997).
[72]
Tans, S.J., M.H. Devoret, H. Dai, A. Thess, R.E. Smalley, L.J. Geerligs, and
C. Dekker, Individual single-wall carbon nanotubes as quantum wires. Nature
386, 474-477 (1997).
[73]
Egger, R. and A.O. Gogolin, Effective Low-Energy Theory for Correlated
Carbon Nanotubes. Physical Review Letters 79, 5082 (1997).
[74]
Kane, C., L. Balents, and M.P.A. Fisher, Coulomb Interactions and
Mesoscopic Effects in Carbon Nanotubes. Physical Review Letters 79, 5086
(1997).
[75]
Gao, B., A. Komnik, R. Egger, D.C. Glattli, and A. Bachtold, Evidence for
Luttinger-Liquid Behavior in Crossed Metallic Single-Wall Nanotubes.
Physical Review Letters 92, 216804 (2004).
[76]
Liang, W., M. Bockrath, D. Bozovic, J.H. Hafner, M. Tinkham, and H. Park,
Fabry - Perot interference in a nanotube electron waveguide. Nature 411, 665669 (2001).
[77]
Yaish, Y., J.Y. Park, S. Rosenblatt, V. Sazonova, M. Brink, and P.L. McEuen,
Electrical Nanoprobing of Semiconducting Carbon Nanotubes Using an
Atomic Force Microscope. Physical Review Letters 92, 046401 (2004).
108
[78]
Kanda, A., K. Tsukagoshi, Y. Aoyagi, and Y. Ootuka, Gate-Voltage
Dependence of Zero-Bias Anomalies in Multiwall Carbon Nanotubes.
Physical Review Letters 92, 036801 (2004).
[79]
Heinze, S., J. Tersoff, R. Martel, V. Derycke, J. Appenzeller, and P. Avouris,
Carbon Nanotubes as Schottky Barrier Transistors. Physical Review Letters
89, 106801 (2002).
[80]
Fuhrer, M.S., B.M. Kim, T. Durkop, and T. Brintlinger, High-Mobility
Nanotube Transistor Memory. Nano Lett. 2, 755-759 (2002).
[81]
Cobden, D.H. and B.A. Muzykantskii, Finite-Temperature Fermi-Edge
Singularity in Tunneling Studied Using Random Telegraph Signals. Physical
Review Letters 75, 4274 (1995).
[82]
Radosavljevic, M., M. Freitag, K.V. Thadani, and A.T. Johnson, Nonvolatile
Molecular Memory Elements Based on Ambipolar Nanotube Field Effect
Transistors. Nano Lett. 2, 761-764 (2002).
[83]
Arai, M.R., A fundamental noise limit for biased resistors at low temperatures.
Applied Physics Letters 42, 906-908 (1983).
[84]
Chun, K. and N.O. Birge, Single-defect thermometer as a probe of electron
heating in Bi. Physical Review B 49, 2959 (1994).
[85]
Roukes, M.L., M.R. Freeman, R.S. Germain, R.C. Richardson, and M.B.
Ketchen, Hot electrons and energy transport in metals at millikelvin
temperatures. Physical Review Letters 55, 422 (1985).
109
[86]
McEuen, P.L., M. Bockrath, D.H. Cobden, Y.-G. Yoon, and S.G. Louie,
Disorder, Pseudospins, and Backscattering in Carbon Nanotubes. Physical
Review Letters 83, 5098 (1999).
[87]
Ando, T., T. Nakanishi, and R. Saito, Berry's Phase and Absence of Back
Scattering in Carbon Nanotubes. Journal of the Physical Society of Japan 67,
2857-2862 (1998).
[88]
Anderson, P.W. and G. Yuval, Exact Results in the Kondo Problem:
Equivalence to a Classical One-Dimensional Coulomb Gas. Physical Review
Letters 23, 89 (1969).
[89]
Souza Filho, A.G., A. Jorio, J.H. Hafner, C.M. Lieber, R. Saito, M.A.
Pimenta, G. Dresselhaus, and M.S. Dresselhaus, Electronic transition energy
E_{ii} for an isolated (n,m) single-wall carbon nanotube obtained by antiStokes/Stokes resonant Raman intensity ratio. Physical Review B 63, 241404
(2001).
[90]
Saito, R. and H. Kataura, Optical Properties and Raman Spectroscopy of
Carbon Nanotubes, in Carbon Nanotubes: Synthesis, Structure, Properties
and Application, M.S. Dresselhaus, G. Dresselhaus, and P. Avouris, Editors.
2001, Springer Berlin: Heidelberg. p. 213–247
[91]
Ohtaka, K. and Y. Tanabe, Golden-rule approach to the soft-x-ray-absorption
problem. III. The temperature dependence. Physical Review B 30, 4235
(1984).
110