ABSTRACT
Title of Document: HIGH FREQUENCY GENERATION FROM
CARBON NANOTUBE FIELD EFFECT
TRANSISTORS USED AS PASSIVE MIXERS
Andrew Jacob Tunnell, Doctor of Philosophy, 2012
Directed By: Professor Ellen D. Williams,
Department of Physics
The high mobilities, low capacitances (due to small sizes), and high current
densities of carbon nanotube field-effect transistors (CNT FETs) make them valid
candidates for high frequency applications. The high cost of high frequency
measurement equipment has been the largest hurdle to observing CNT transistor
behavior at frequencies above 50 GHz. One economic solution to this barrier is to use
an external harmonic mixer to convert high frequency signals to lower frequencies
where they can be detected by a standard spectrum analyzer. By using this detection
method, a new regime of high frequency CNT FET behavior is available for study.
In this dissertation, we describe the design and fabrication of CNT FETs on
quartz substrates using aligned arrays of CNTs as the device channel. The nonlinear
input voltage to output drain current behavior of the devices is explained and
approximated to the first order by using a Taylor expansion. For the high frequency
mixing experiments, two input voltages of different frequencies are sourced on the
gate of the devices without any device biasing. The input frequencies are limited to
100 kHz to 40 GHz by the signal generators used. The nonlinearities of the fabricated
CNT FETs cause the input frequencies to be mixed together, even in the absence of a
source-drain bias (passive mixing). The device output is the drain current, which
contains sum and difference products of the input frequencies. By using an external
harmonic mixer in combination with a spectrum analyzer to measure the drain
current, output frequencies from 75 to 110 GHz can be observed. Up to 11th order
mixing products are detected, due to the low noise floor of the spectrum analyzer.
Control devices are also measured in the same experimental setup to ensure that the
measured output signals are generated by the CNTs. The cutoff frequencies from
previous passive mixing experiments predict that our devices should stop operating
near 13 GHz, however our measurement setup extends and overcomes these cutoffs,
and the generation of high frequency output signals is directly observed up to 110
GHz. This is the highest output frequency observed in CNT devices to date.
HIGH FREQUENCY GENERATION FROM CARBON NANOTUBE FIELD
EFFECT TRANSISTORS USED AS PASSIVE MIXERS
By
Andrew Jacob Tunnell
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
2012
Advisory Committee:
Professor Ellen Williams, Chair
Professor John Cumings
Professor Michael Fuhrer
Professor Ichiro Takeuchi
Professor Manfred Wuttig, Dean?s Representative
? Copyright by
Andrew Tunnell
2012
ii
ACKNOWLEDGEMENTS
When I began writing this dissertation, I instinctively used the pronoun ?we?
instead of ?I? when referencing the work that was done and would be presented. The
use of a majestic plural pronoun was not motivated by the high esteem in which I
hold myself, but by the dishonesty involved in claiming this work as the fruits of my
labor alone. My nearly seven years at the University of Maryland have convinced me
of the benefits of collaborations, and I?d like to mention a few of the people who have
aided me along the way.
I first must thank my advisor, Ellen Williams, for accepting me into her
research group before I even arrived in Maryland. Ellen has provided guidance and
always made time for my questions, in spite of her many responsibilities. It is
tempting to grow discouraged with failed experiments and incomprehensible results,
but Ellen is a master of highlighting the positive sides and telling a clear story. When
Ellen?s responsibilities took her to London, John Cumings took over my weekly
advising. John helped me narrow the scope of my research to what is presented in this
dissertation. I have met many graduate students who grow to dislike their field and/or
the whole experience of being in graduate school. I very much enjoy my field and
being a researcher in this group. If not for John reminding me that my goal should be
graduation, I would still be pursuing every experimental dead-end that struck me as
interesting. John and Ellen, thank you for getting me here.
The first scientist with whom I began working was Dan Hines. My first day in
the lab, Dan began patiently training me on cleanroom equipment. A significant
portion of my time here has been devoted to device fabrication, and nearly all of my
iii
lithography skills can be traced to Dan. My first two publications came from our early
work together. As my focus shifted from organic electronics to carbon nanotubes,
Dan and I have worked less together, but he has always given practical advice and has
been very generous in the correction process for several chapters. For my most recent
experiments, I have worked most closely with Vince Ballarotto. For experimental
design, equipment acquisition, proper measurement techniques, data analysis, and
more, Vince has been essential every step of the way. He has an excellent sense of
when to motivate and when to encourage. Vince?s hard work in editing much of this
dissertation is very much appreciated. I also want to thank Steven Anlage for helping
design our high frequency mixing experiments.
Most of my fellow graduate students graduated ahead of me and offered me
encouragement from beyond the academic world. Mihaela Ballarotto was the senior
graduate student when I began. Hers was the first defense I attended, and even then
she assured me that I would be in her position soon enough. She has provided moral
support throughout the years. Tracy Moore and Vinod Sangwan have been allies in
the lab and good friends outside of it. Vinod trained me in chemical vapor deposition
(CVD) and helped me get up to speed on CNT devices. Major thanks to Enrique
Cobas for providing our first samples of aligned CNTs on quartz. He did the hard
work of fine-tuning the CVD recipes. Merijntje Bronsgeest contributed significantly
to the experiments on burning CNTs and charge trap measurements. Her eagerness to
investigate and understand models that aren?t directly related to her research path
both impressed and helped me. And my appreciation to Dan Lenski and Jia Huang for
iv
their optimism and advice, to John Bavier for his help with high frequency modeling,
and to Tommy Willis for being the best storyteller I?ve ever met.
Moving to Maryland was my first experience on the east coast, a big transition
from my hometown of Boring, Oregon. The one Oregonian I knew out here, Libby
Deal, has been an irreplaceable friend and has helped me acclimate to life in (just
outside) a real city. Special thanks to Tobi Palmer, Pearlette Merriweather, Christina
Yancey, and Graham Childs, a solid group of friends in DC that made me feel at
home. Cheers to Bart Herman for a regular game of pool at Town Hall every
Thursday for two years. Thank you Tifa Trefry for loving the same arts I do and Sara
Sohriakoff for easing the writing process by providing me with the perfect
background music. Thank you Salomeh Moadab for your timing. Lastly, my thanks
and love to my family: my brothers Mark, Matt, and Ross, my sisters Meridith,
Thanh, and Shannon, and my parents Ross and Parm (who helped proofread). I have
never lacked support from any of you, and I often take that for granted.
v
TABLE OF CONTENTS
Acknowledgements????????????????????????? ii
Table of Contents?????????????????????????... v
List of Tables???????????????????????????. vii
List of Figures??????????????????????????? viii
List of Abbreviations????????????????????????. xii
Chapter 1: Introduction???????????????????????.. 1
1.1 Introduction to Carbon Nanotubes????????????????. 1
1.2 Carbon Nanotube Electronic Properties??????????????. 5
1.3 Advantages of Carbon Nanotubes in High Frequency Applications???. 8
1.4 Overview of Dissertation???????????????????.. 11
Chapter 2: Carbon Nanotube Field Effect Transistor Behavior and Fabrication?.. 12
2.1 Field Effect Transistor Behavior????????????????? 12
2.1.1 Field Effect Transistor Basic Operation???????????... 12
2.1.2 Field Effect Transistor DC Modes of Operation????????... 15
2.1.3 CNT FET Nonlinearities?????????????????... 18
2.2 Device Fabrication??????????????????????. 24
2.2.1 Device Substrate Selection????????????????... 24
2.2.2 CVD Growth and Lithography???????????????. 27
2.2.3 Device Geometries???????????????????? 29
Chapter 3: Charge Traps??????????????????????? 32
3.1 Introduction to Charge Traps and Hysteresis in CNT FETs??????.. 32
3.2 Measurements of Traps Emptying Over Time???????????... 34
3.2.1 Dependence on Pulse Duration???????????????. 38
3.2.2 Dependence on Pulse Magnitude??????????????.. 41
3.3 Preisach Model of Hysteresis??????????????????. 44
3.4 Hysteresis Reducing Gate Waveforms??????????????.. 51
3.5 Summary?????????????????????????? 55
Chapter 4: Removal of Metallic CNTs from Mixed Arrays: Burning???? 57
4.1 On/Off Current Ratio?????????????????????. 57
4.2 Burning on SiO2???????????????????????. 59
4.2.1 Standard Burning Procedure on SiO2????????????... 59
4.2.2 Pulsed Burning on SiO2?.????????????????... 66
4.3 Burning on Quartz??????????????????????. 68
4.4 Summary?????????????????????????? 74
Chapter 5: HF CNT Measurement Techniques and Limitations???????... 76
vi
5.1 Active Devices: Current Gain Cutoff Frequency??????????... 77
5.2 Direct Observation of HF CNT Nonlinear Behavior in Literature???? 80
5.3 Tomonaga-Luttinger Liquid Theory???????????????.. 83
5.4 Our High Frequency Experimental Method????????????.. 85
Chapter 6: Carbon Nanotube Mixing Theory??????????????? 92
6.1 Generic Mixing???????????????????????.. 92
6.1.1 Taylor Expansion of Input to Output Nonlinearity???????... 92
6.1.2 Harmonic Generation??????????????????? 94
6.1.3 Two-Tone Mixing Products and Corresponding Nomenclature??... 95
6.2 Modified Taylor Expansion for Passive CNT Mixing????????... 97
6.3 Measurement of CNT DC Drain Current?????????????...100
6.4 Summary?????????????????????????? 105
Chapter 7: CNT Mixing Data????????????????????? 106
7.1 250 MHz to 25 GHz (SMA Regime) Measurements?????????.106
7.2 75 to 110 GHz (W Regime) Measurements with External Mixer????. 116
7.2.1 Experimental Setup???????????????????... 117
7.2.2 Two-Tone Mixing in W Regime with 100 kHz Separation????.. 118
7.2.2.1 Varying the Output Frequency from 75 to 110 GHz ????.. 119
7.2.2.2 Varying the Input Power ???????????????. 124
7.2.2.3 Examining the (1, 2) to (3, 0) Amplitude Difference????.. 130
7.2.3 Heterodyne Detection at 37 GHz LO Operation????????... 133
7.3 Summary?????????????????????????? 139
Chapter 8: Summary and Conclusions?????????????????... 141
Appendix A: Fabrication Procedure for Aligned CNT FETs on Si/SiO2 Substrate.. 147
Appendix B: Fabrication Procedure for Top-Gate Aligned CNT FET on PET with
Tri-Layer Dielectric and Vertical Integration??????????????? 153
Appendix C: Fabrication Procedure for Top Local Gated CNT FET on Quartz?... 158
Appendix D: Capacitance Calculations?????????????????. 166
Appendix E: Preisach Hysteresis Model Matlab Script???????????. 170
Appendix F: Recalibrating Two-Tone Input Power????????????.. 182
Bibliography???????????????????????????.. 190
vii
List of Tables
Table 3.1 Hysteresis data from the transfer curves in Figure 3.15. ?Id is the??... 52
Table 5.1 Summary of high frequency CNT device experiments that directly??. 80
Table 7.1 Channel widths, channel lengths, and gate lengths in micrometers??.. 132
Table D.1 The calculated capacitances for five fabricated parallel plate????.. 166
viii
List of Figures
Figure 1.1 Structural diagrams of the sp2 allotropes of carbon as formed from?? 2
Figure 1.2 a) Image of the honeycomb graphene lattice structure with the unit?... 4
Figure 1.3 Band structure of graphene. The conduction band (gray) and valence? 6
Figure 1.4 Band structures for a metallic and semiconducting carbon nanotube?.. 7
Figure 2.1 Diagrams of a top-gated p-type CNT FET with corresponding energy... 14
Figure 2.2 Plot of conduction (logarithmic scale) vs. gate voltage for a CNT?..... 14
Figure 2.3 Plots of conductance, dynamic conductance, and transconductance?... 18
Figure 2.4 Diagrams of a top-gated p-type CNT FET with corresponding???... 20
Figure 2.5 Diagrams of a top-gated p-type CNT FET with corresponding???... 22
Figure 2.6 Plots of a) drain current and b) dynamic conductance of a CNT FET?. 23
Figure 2.7 SEM image of aligned CNTs grown on a quartz substrate by????. 28
Figure 2.8 a) An optical image of a completed CNT FET in a high frequency??.30
Figure 2.9 Optical images of the open and through electrodes used for VNA??.. 31
Figure 2.10 a) Optical image and b) side-view diagram of fabricated parallel??. 31
Figure 3.1 Plot of a transfer curve taken at Vds = 5 V of a CNT FET ?????... 34
Figure 3.2 Diagram of the gate voltage waveform used to observe charge traps?.. 36
Figure 3.3 Plots of discharge data over the 100 s Toff interval for Ton = 1 s???... 37
Figure 3.4 Plot of discharge data over 100 s for Vgs Pulse = 1 V and several??... 39
Figure 3.5 Plots of a) the initial offset current values and b) the initial power??. 40
Figure 3.6 Plots of discharge data over 100 s for Ton = 1 ms for several values?... 42
Figure 3.7 Plots of a) the initial offset current values and b) the initial discharge? 43
Figure 3.8 Diagram of a hysteron relay: x is the input, y is the output, and ? and? 45
ix
Figure 3.9 Diagrams relating a single trap state in a CNT FET to a hysteron relay.. 46
Figure 3.10 Diagrams of a) ?-? parameter space with six hysterons and b)???.. 46
Figure 3.11 Plots of x-y planes from the Preisach hysteresis model using????48
Figure 3.12 Plots of x-y planes from the Preisach hysteresis model using ???...50
Figure 3.13 Waveforms of gate voltage vs. time used to reduce hysteresis in??.. 51
Figure 3.14 Transfer curves of a top-gated aligned CNT FET (W = 100 ?m ??... 52
Figure 3.15 Drain current from a) the Preisach model and b) measurements??... 54
Figure 3.16 Plot of the resulting hysteresis height (?VT) and width (?Id) for??.. 55
Figure 4.1 a) Diagram of FET geometry on Si/SiO2 with aligned CNTs in the?... 61
Figure 4.2 Drain current vs. source-drain voltage plot of three burning sweeps?... 62
Figure 4.3 IV curves of a CNT FET on SiO2. a) and b) correspond to the???? 63
Figure 4.4 Plot of normalized on-state drain current vs. on/off ratio for 36???.. 65
Figure 4.5 Gate voltage and source-drain voltage waveforms used to perform?? 67
Figure 4.6 Plot of normalized on-state drain current vs. on/off ratio for a CNT?... 68
Figure 4.7 Diagram of the external Si/SiO2 top-gate burning method on a quartz? 70
Figure 4.8 Optical image of joined and extended CNT contact electrodes on??.. 71
Figure 4.9 a), c) Output and b), d) transfer curves for 16 CNT FETs measured in.. 72
Figure 4.10 Optical images of the lithographic steps of isolating and completing? 73
Figure 4.11 a) Output and b) transfer curves for a CNT FET on quartz?????74
Figure 5.1 Plot of published current gain cutoff frequencies on a log scale???.. 79
Figure 5.2 a) Diagram of CNT FET with the input signal on the source and??? 81
Figure 5.3 a) Diagram of CNT FET with the input signal on the gate, the???... 87
Figure 5.4 Output power drop as a function of frequency for a low-pass filter?? 90
x
Figure 6.1 Output power spectrum read by an Agilent MXG analog signal???. 97
Figure 6.2 Diagram of the experimental setup used to measure the DC mixing?... 100
Figure 6.3 Conductance vs. gate voltage for a top-gated CNT FET on?????. 102
Figure 6.4 Plot of the measured drain current and the derivative of the total??? 103
Figure 6.5 Plot of the measured drain current and the derivative of the total ??... 104
Figure 7.1 Diagram of the experimental setup used to perform mixing????? 107
Figure 7.2 Optical images and side profile views of the different control DUTs?.. 109
Figure 7.3 Plot of output power attenuation for four different device types vs?..... 110
Figure 7.4 Output spectra captured from a spectrum analyzer from two-tone??.. 111
Figure 7.5 Output spectra captured at three times the input frequency from a??.. 113
Figure 7.6 Mixing product amplitudes vs. output frequency for a) a CNT FET?... 115
Figure 7.7 Diagram of the experimental setup used to perform mixing????? 117
Figure 7.8 Output spectra at a) 75.5 GHz and b) 109.5 GHz of a CNT FET ??... 120
Figure 7.9 Mixing product amplitudes vs. output frequency for a CNT FET........... 121
Figure 7.10 Mixing produce amplitudes plotted against two input frequency??.. 122
Figure 7.11 Output spectra at 75.5 GHz of a CNT FET used to mix two input?? 125
Figure 7.12 Plot of the output power of six mixing products vs. the output power.. 125
Figure 7.13 Plots of the output power of six mixing product peaks vs. the output... 129
Figure 7.14 Output spectra at 75.5 GHz of a CNT FET used to mix two input?.... 130
Figure 7.15 Output amplitude difference between the (2,1) and (3,0) mixing??.. 133
Figure 7.16 Output power of the mixing product from the 37 GHz LO mixing?... 135
Figure 7.17 Output power of the mixing product from the 37 GHz LO mixing?... 136
Figure 7.18 Predicted attenuation from the CNT FET circuit model in Chapter 5... 137
xi
Figure 7.19 Output power of the mixing product from the 37 GHz LO mixing?... 138
Figure 8.1 Electrode geometries for a long channel length device that could??... 145
Figure B.1 Diagram of completed integrated circuit using CNT FETs with???. 153
Figure B.2 a) Optical image of completed integrated CNT FET circuit on PET?.. 154
Figure C.1 a) side-view and b) top-view diagram of a completed CNT FET on...... 158
Figure C.2 Diagram of the CNT FET device geometry with the iron catalyst??.. 160
Figure C.3 Diagram of the CNT FET device geometry with the source-drain??.. 161
Figure C.4 Diagram of the CNT FET device geometry with the nanotube???... 162
Figure C.5 Diagram of the CNT FET device geometry with the dielectric???... 164
Figure C.6 Diagram of the CNT FET device geometry with the gate lithography... 164
Figure D.1 Gate-CNT capacitance (CGCNT) plotted as a function of frequency?? 168
Figure D.2 Gate-drain capacitance (CGD) plotted as a function of frequency??... 169
Figure F.1 Plot of the output power of the (0, 3) mixing product vs. the output?... 183
Figure F.2 Diagram of our HF two-tone mixing experimental setup showing the? 184
Figure F.3 Plot of the output power of the (3, 0) mixing product peak vs. the??..186
Figure F.4 Plot of the output power of the (0, 3) mixing product peak vs. the?......187
Figure F.5 Plots of the output power of six mixing product peaks vs. the output?. 189
xii
List of Abbreviations
AC ? alternative current
AFM ? atomic force microscope
W ? channel width
CGCNT ? gate-nanotube capacitance
CGD ? gate-drain capacitance
CGS ? gate-source capacitance
CNT ? carbon nanotube
DC ? direct current
DUT ? device under test
GSG ? ground-signal-ground
GPIB ? general purpose interface bus
LO ? local oscillator
dBm ? decibel milliwatts
EF ? Fermi energy
FET ? field-effect transistor
CVD ? chemical vapor deposition
G ? gate length
HF ? high frequency
IV ? current voltage
SCCM ? standard cubic centimeters per minute
SSP ? single-side polished
Id ? drain current
xiii
IPA ? isopropyl alcohol
L ? channel length
PET ? polyethylene terephthalate
PMMA ? poly(methyl methacrylate)
RFID ? radio-frequency identification
RIE ? reactive ion etcher
SEM ? scanning electron microscope
SMA ? SubMiniature Version A
Vds ? drain-source voltage
Vgs ? gate-source voltage
1
Chapter 1: Introduction
1.1 Introduction to Carbon Nanotubes
Carbon is the basic building block of life on earth. In its neutral electronic
state, it has four valence electrons that allow up to four covalent chemical bonds to be
formed with adjacent atoms. In organic compounds, carbon atoms form the molecular
skeleton to which other elements attach to functionalize the molecule and account for
most of the chemical behavior. Hydrogen, oxygen, and nitrogen are the most common
functional elements in organic chemistry. Plastics, pharmaceuticals, fats, proteins, and
DNA are all composed of organic compounds.
Carbon can also exist as a pure solid in multiple allotropes. The kinds of
bonds that each carbon atom forms with its neighbor will determine the bulk
structure. When each carbon atom is sp3 hybridized, it forms four covalent bonds with
other carbon atoms. This allotrope of carbon is diamond. With all valence electrons
held in covalent bonds, no electrons are free to move through the bulk material, so
diamond is an insulator.
When each carbon atom is sp2 hybridized, it forms three bonds with other
carbon atoms which produces a honeycomb pattern of carbon atoms in a flat sheet.
This material is called graphene. Since only three of the four valence electrons are
used in bonds, each carbon atom contributes a free electron that is delocalized in two
pi bond orbitals above and below the graphene plane. Unlike diamond, these free
electrons allow graphene to conduct electricity.
2
Being only one atom thick, graphene is thought of as a two-dimensional
material. When layers of graphene are stacked on top of each other, the three-
dimensional material graphite is formed. When one sheet of graphene is rolled into a
tube, a carbon nanotube is formed, which is thought of as a one-dimensional material
and the material investigated in this dissertation. Molecules called fullerenes, which
are thought of as zero- dimensional materials due to their small size, can also be made
from a sheet of graphene. A diagram showing the structure of a fullerene (C-60), a
carbon nanotube, and graphite from a graphene sheet is shown in Figure 1.1. This
illustrates the material structure only and not how each material is made.
Figure 1.1 Structural diagrams of the sp2 allotropes of carbon as formed from the theoretical
shaping of a section of graphene. When wrapped into a ball it forms the fullerene C60; when
wrapped into a cylinder it forms a carbon nanotube. When stacked into sheets it forms
graphite.
3
Experimentally fabricated carbon nanotubes (CNTs) were first observed by
transmission electron microscopy in 1991.1 The observed ?helical microtubules of
graphitic carbon? were micrometers in length but had diameters of a few nanometers.
The longest measured CNTs so far have been 18.5 cm long.2 This large aspect ratio
(108) provides an interesting bridge between nanoscale quantum behavior and
macroscopic electronic devices.
Nanotubes can exist as single cylinders (single-walled carbon nanotubes) or as
multiple concentric cylinders (multi-walled carbon nanotubes). Since many different
cylindrical geometries can be formed from the sp2 honeycomb graphene lattice, each
nanotube geometry is specified by a chiral vector (Ch) that forms the nanotube
circumference. The chiral vector is specified by integer multiples of the graphene unit
vectors a1 and a2 and is written as Ch = na1 + ma2 = (n, m). To avoid multiple
vectors describing identical nanotube geometries, the convention 0 ? |m| ? n is used.
While n and m could theoretically be any positive integer, single- walled carbon
nanotube diameters are rarely seen to exceed 10 nm,3, 4 so the number of different
CNT geometries are experimentally limited to a few hundred. The unit vectors and
examples of different chiral vectors are shown in the graphene lattice in Figure 1.2a.
Nanotube geometries fall into three main geometric types. When the first
chiral index is zero (0, m), the resulting nanotube circumference makes a zigzag
pattern. These are called zigzag nanotubes. Nanotubes that have equal chiral indices
(n = m), are called armchair nanotubes, again due to the carbon pattern around the
circumference. Both zigzag and armchair nanotubes are achiral, meaning that they are
the same as their mirror image along the circumference. All other nanotube
4
geometries are chiral, not the same as their mirror images. An example of a zigzag,
armchair, and chiral nanotube are shown in Figure 1.2 b, c, and d.
Figure 1.2 a) Image of the honeycomb graphene lattice structure with the unit vectors (a1 and
a2) and chiral vectors shown. A chiral vector defines the circumference of a nanotube.
Examples of a zigzag, armchair, and chiral nanotube are shown in b), c), and d). Image
courtesy of Professor Susan Sinnott, UFL
5
Since the chiral vector forms the circumference of the CNT, the tube diameter
is the magnitude of the chiral vector divided by pi, shown in EQ 1.1. a is the graphene
unit vector length, which is equal to 2.49 ?.
nmmn
aC
d ???? 22
??
?
EQ 1.1
1.2 Carbon Nanotube Electronic Properties
The electronic band structure of CNTs comes from the band structure of
graphene with boundary conditions added from the cylindrical symmetry. The
graphene band structure was derived by Wallace using the tight binding model in
1947.5 The energy dispersion relation is shown in EQ 1.2, where EF is the Fermi
energy, ?0 is the transfer integral of overlapping orbitals, and kx and ky are the
momentum wave vector.
? ? ?
?
?
?
?
?
??
?
?
?
?
?
?
?
?
?
?
?
???
2
cos4
2
cos
2
3
cos41, 20
akakak
EkkE
yyx
Fxx ? EQ 1.2
The square root quantity is equal to zero for six k vectors (kx, ky):
?
?
?
?
?
?
?
a3
4
,0
?
and
??
?
?
??
?
?
??
aa 3
2
,
3
2 ?? . At these points, called K points, the conduction and valence
bands meet. Since they meet but do not overlap, graphene is considered to be a zero-
bandgap semiconductor. A diagram of this band structure is given in Figure 1.3.The
structure near each K point is conical for small energies above and below the Fermi
level.
6
Carbon nanotube band structures deviate from graphene because not all k
vectors are allowed. The finite circumference sets a periodic boundary condition on
momentum vectors parallel to the chiral vector. That boundary condition is given by
EQ 1.3, where p is an integer.
pmknkCk yxh ?2????
?? EQ 1.3
Figure 1.3 Band structure of graphene. The conduction band (gray) and valence band (green)
lie above and below the Fermi level respectively. Image plotted in Maple from EQ 1.2.
This boundary condition divides the graphene band structure into a discrete
number of allowed momentum vectors for each CNT chirality. Only when the
difference between the chiral indices (n ? m) is divisible by 3 do the allowed
momentum vectors pass through the K points. In those cases, there is no bandgap and
the CNT is metallic. Armchair nanotubes meet this condition and are all metallic. In
reality, the curvature of CNTs induce deviations to this band structure, and many of
the nanotubes that meet this condition contain small band gaps ( ~10 meV) and are
7
semimetallic.6 For most measurements, such a small band gap will be
indistinguishable from a metal, so we treat semimetallic CNTs as metallic CNTs for
the rest of this dissertation. For all the CNTs where the difference in chiral indices is
not divisible by 3, a bandgap will exist, and the CNT will be semiconducting. This
results in twice as many possible semiconducting CNTs as metallic CNTs. Diagrams
of the band structures for a metallic and semiconducting nanotube are given in Figure
1.4.
Figure 1.4 Band structures for a metallic and semiconducting carbon nanotube, shown as
lines of allowed momentum wave vectors from the periodic boundary condition EQ 1.3.
Image provided by Michael Fuhrer.
8
1.3 Advantages of Carbon Nanotubes in High Frequency Applications
Carbon nanotubes have many impressive properties. They are the stiffest
material known, with a Young?s modulus ranging from 1 to 5 TPa.[7, 8] Diamond, the
hardest naturally occurring mineral, has a Young?s modulus of 1.22 TPa,[9] and steel
has a Young?s modulus of 200 GPa. CNTs are also excellent thermal conductors,
with experimental room temperature thermal conductivities near 3000 W/mK,[10]
while diamond is between 900 and 2320 W/mK[11] and copper is 400 W/mK. But it is
the extraordinary electronic properties of carbon nanotubes and their potential for
high frequency applications that is the motivation and focus of this dissertation.
The quantity that best describes a semiconductor?s high frequency
performance is its charge carrier mobility (?), which defines how quickly charges can
move through a material in response to an electric field. It is related to the drift
velocity through a material by EQ 1.4.
EVd ??
EQ 1.4
The larger the drift velocity of a charge carrier, the faster it can respond to a
change in electric field. When an AC signal is input on a semiconductor, the electric
field changes at the input frequency, so the higher the input frequency, the less time
carriers will have to travel through the semiconductor before the field changes and
they reverse direction. The larger the carrier mobility, the greater distance carriers
will cover in each period. When carriers stop moving, a semiconductor can no longer
function. Carbon nanotubes have a measured room temperature carrier mobility of
100,000 cm2/Vs[12], which is 70 times greater than the room temperature mobility of
silicon (1,400 cm2/Vs for electrons).
9
Silicon has dominated the semiconductor industry due to its ease of growth
into high purity wafers, its elemental abundance on earth, and its robust thermal
oxide. The miniaturization of silicon devices has doubled the operating frequency of
silicon transistors every 18 months, following Moore?s Law. This trend must
eventually come to an end as silicon devices approach atomic barriers. To continue
increasing semiconductor operational frequencies, other materials with larger
mobilities will have to be used.
For semiconductor operation at frequencies beyond what silicon can provide,
gallium arsenide and indium antimonide have been used, which have room
temperature electron mobilities of 8,500 and 77,000 cm2/Vs respectively. Neither of
these semiconductors have threatened silicon?s dominance for two reasons. First, they
do not produce a high quality thermal oxide which is a useful feature when
constructing field effect transistors. Second, they are much more expensive to
manufacture. The source minerals are less abundant on earth, and high quality wafers
require slow and expensive fabrication processes such as molecular beam epitaxy.
The high fabrication costs have limited their incorporation into all but a few
expensive high frequency components.
Carbon nanotubes do not produce a natural thermal oxide either, requiring the
deposition of dielectric layers that may not form an ideal interface with the nanotube
surface. These issues are discussed in Chapter 4. However, nanotubes are very
inexpensive to produce. Large volumes of CNTs can be synthesized from arc
discharge1, laser ablation13, and chemical vapor deposition14, 15, all cheaper processes
than those used to produce gallium arsenide and indium antimonide. Also, the carbon
10
source materials are much more abundant on earth than gallium, arsenic, indium, or
antimony.
One other limitation to a semiconductor?s performance at high frequencies
comes from its resistance and capacitance. At frequencies above 1/2?RC, more than
half of the input power is lost through that capacitance, so the smaller the resistance
and capacitance of the nanotube, the faster it will be able to operate efficiently. Due
to its nanometer scale diameter, the quantum capacitance from a CNT is very small,
with a theoretical minimum value of 100 aF / ?m of channel length.16 The conduction
through a CNT is quantized, with the ideal quantum conductance per CNT equal to
155 ?S. That corresponds to a impedance of 6.45 k?.16 If a nanotube device were
able to be made with these minimum values, the RC cutoff frequency would scale
with channel length as EQ 1.5.
? ?
??
LCR
f
QQ
RC
/2
1
?
250 GHz
Real devices will include contact resistances and capacitances between the
CNT with other device electrodes, but the scale of this approximation demonstrates
the appeal of using carbon nanotubes in high frequency devices. One barrier to
integrating CNTs into everyday electrical components is the high impedance of
individual tubes. To preserve power transmission down multiple electronic
components, each is typically impedance matched to 50 ?. The minimum impedance
of a single CNT is 130 times that impedance. For this reason, our devices include
hundreds of CNTs in our channels, with the goal of coming closer to matching the
impedance of the other electrical equipment used in our experiments.
11
1.4 Overview of Dissertation
The goal of this research is to fabricate carbon nanotube field-effect
transistors to be used as mixers in a frequency regime where their behavior has not
yet been directly observed. In Chapter 2 we introduce carbon nanotube field-effect
transistors, how they are expected to behave in our experiments, and how they were
fabricated. Chapter 3 discusses the problems associated with imperfect dielectric
layers in our CNT FET devices, most notably the existence of charge traps visible as
hysteresis in transfer curves. A measurement technique utilizing short voltage pulses
is presented to reduce the effects of trapped charges, eliminating hysteresis in transfer
curves. In Chapter 4 we discuss the removal of metallic CNTs from our as-grown
networks. Though common on standard silicon substrates, this had not been done
before on our quartz device substrates. Chapter 5 summarizes the published
experimental results of other research groups and clarifies how our experiments differ
to overcome expected frequency cutoffs. Chapter 6 describes how passive frequency
mixing occurs in field-effect transistors by exploiting the nonlinear input voltage
output drain current behavior of CNT FETs. Two-tone mixing experiments are
performed in two frequency regimes (1 to 25 GHz and 75 to 110 GHz) in Chapter 7.
CNT devices are compared with controls to confirm that the observed output is due to
the nanotubes. Chapter 8 summarizes and concludes the dissertation, with some
suggestions of future work.
12
Chapter 2: Carbon Nanotube Field Effect Transistor Behavior and
Fabrication
In this chapter, we describe the general operation of a field-effect transistor.
We present the expected drain current dependence on input voltages in the
operational regime in which we will be taking our high frequency measurements. The
nonlinear relationship between the drain current and the source-drain input voltage in
carbon nanotube field-effect transistors (CNT FETs) is described and demonstrated
through DC measurements. Lastly, the design and fabrication of our CNT FETs is
briefly described.
2.1.1 Field Effect Transistor Basic Operation
A field-effect transistor (FET) is a three-terminal device composed of metal
electrodes, a semiconducting channel, and an insulating dielectric layer. The
electrodes contacting the semiconductor are called the source and drain, while the
third electrode that is not in electrical contact with the channel is called the gate.
When a bias is set between the source and drain, current flows through the channel.
The conductivity of the channel is tuned by the voltage on the gate electrode, which
sets the Fermi level in the semiconductor relative to the conduction and valence
bands, controlling the number of available carriers. Carbon nanotube field-effect
transistors (CNT FETs) use semiconducting CNTs as the channel. Metal contacts and
oxide dielectric layers typically dope CNT FETs p-type, meaning that the Fermi level
lies closer to the valence band (where holes are carriers) than the conduction band
13
(where electrons are carriers). A negative voltage applied to the gate electrode of a p-
type CNT FET will lower the Fermi level closer to the valence band, allowing more
holes to flow through the channel. A positive gate voltage will move the Fermi level
up into the band gap, decreasing conduction. Some CNT FETs demonstrate
ambipolarity, in which a large gate voltage moves the Fermi level close enough to the
conduction band that electrons begin to be conducted. An illustration of these gating
situations along with energy levels for a top-gated p-type CNT FET are shown in
Figure 2.1. Figure 2.2 plots conduction vs. gate voltage for a CNT FET,
demonstrating the different gating conditions shown in Figure 2.1.
14
EV
EC
EF
Vgs < 0 Vgs > 0
Hole Conduction Band Gap Electron Conduction
Vgs ? 0
EC ? eVgs
EV ? eVgs
EC ? eVgs
EV ? eVgs
EC ? eVgs
EV ? eVgs
a) b) c) d)
Figure 2.1 Diagrams of a top-gated p-type CNT FET with corresponding energy levels for
different gate voltages. a) The Fermi level in a p-type FET will lie closer to the valence band
than the conduction band. b) With a negative gate voltage, the Fermi level moves down (or
the other bands move up) increasing the number of holes available. c) A positive gate voltage
moves the Fermi level up, further into the band gap, decreasing conduction. d) Large enough
positive gate voltages move the Fermi level close to the conduction band where electrons can
act as carriers.
Figure 2.2 Plot of conduction (logarithmic scale) vs. gate voltage for a CNT FET. For
negative gate voltages, the Fermi level lies in the valence band, and holes are charge carriers.
For positive gate voltages, the Fermi level lies in the band gap and conduction is suppressed.
For large positive gate voltages, the Fermi level approaches the conduction band, allowing
electrons to conduct. The measurement was taken at 1 V source-drain bias.
15
By convention, the source electrode is typically grounded and is considered
the charge carrier source (hole source for p-type FETs). The gate voltage and source-
drain voltage are usually referenced relative to the source. In this dissertation, they
will be referred to as Vgs and Vds. The drain current is the standard device output
quantity.
2.1.2 Field Effect Transistor DC Modes of Operation
DC operation of a CNT FET falls into three modes. The cutoff or subthreshold
region corresponds to when the Fermi level lies in the band gap. The only conduction
in this regime comes from a Boltzmann distribution, depending exponentially on the
gate voltage. The gate voltage where the channel conduction begins to approach zero
is called the threshold voltage. When the gate voltage is greater than the threshold
voltage, the FET is considered off.
1) When the gate voltage is less than the threshold voltage, the CNT FET acts
like a variable conductor in which the number of charge carriers contributing to the
current through the channel is a function of gate voltage.
2) When the magnitude of the source-drain voltage is greater than the gate
voltage, all the carriers available for conduction are being used. Therefore, increasing
the source-drain voltage further, fails to increase the drain current. This region is
called saturation. Our CNT FETs rarely demonstrate saturation as the source drain
biases necessary can result in channel currents large enough to burn out the CNTs.
16
3) The last operational mode occurs when the gate voltage is less than the
threshold voltage, and the magnitude of the source-drain bias is less than the gate
voltage. This regime is called triode mode or the linear regime. It is in this regime that
most of our high frequency experiments were conducted. The equation for the device
drain current in this regime is given by EQ 2.1, where ? is the effective carrier
mobility, Cg is the gate capacitance per unit length, L is the channel length, and Vth is
the threshold voltage.
dsdsgsth
g
d VVVV
L
C
I ?
?
?
?
?
?
???
2
1? EQ 2.1
17
The conductance (G), dynamic conductance (GD), and transconductance (gm)
in this regime are given by EQ 2.2-4.
?
?
?
?
?
?
???? dsgsth
g
ds
d VVV
L
C
V
I
G
2
1? EQ 2.2
? ?dsgsth
g
ds
d
D VVV
L
C
V
I
G ???
?
?
?
? EQ 2.3
ds
g
gs
d
m V
L
C
V
I
g
?
??
?
?
?
EQ 2.4
The dynamic conductance dependence on gate voltage is equal and opposite
to the dependence on the source-drain voltage. The local charge distribution will be
equally be increased by a positive change in source-drain voltage or a negative
change in gate voltage. This equivalence is given in EQ 2.5 and is used as a
substitution in Chapter 5.
gs
D
ds
D
V
G
V
G
?
?
??
?
?
EQ 2.5
17
Our high frequency CNT FET experiments are done mostly in the absence of
a source-drain bias. To the first order in this low bias regime, the conductance and
dynamic conductance are equal, and the transconductance is zero, as shown in EQ 2.6
and 2.7.
? ?gsth
g
VdsDVds
VV
L
C
GG ???
??
?
00
EQ 2.6
0
0
?
?Vdsm
g
EQ 2.7
To confirm this behavior, measurements of a CNT FET were taken for all
three conductance regimes as a function of source-drain voltage. For each measured
source-drain voltage, the gate voltage was swept between ? 50 mV to extract the
transconductance. The conductance was calculated as the measured drain current
divided by the source-drain voltage, and the dynamic conductance was extracted as
the slope of a linear fit to the drain current over the source-drain voltage over five
points. The conductances are plotted in Figure 2.3.
18
Figure 2.3 Plots of conductance, dynamic conductance, and transconductance for a CNT FET
as a function of source-drain voltage. The lines correspond to linear fits to each conductance.
As expected at Vds = 0, the transconductance is zero, and the conductance and
dynamic conductance are equal. The slopes of the dynamic conductance and
transconductance should be equal and opposite. Their ratio from Figure 2.3 is -1.25.
The slope of the dynamic conductance should be twice that of the conductance. Their
ratio from Figure 2.3 is 2. The correspondence of the measured conductances to EQ
2.2-4 confirms the form of the drain current in this linear region.
2.1.3 CNT FET Nonlinearities
The dependence of conductance on gate voltage and source-drain voltage
makes the CNT FET?s drain current non-linear in both input voltages. It is the non-
linearity of CNT FETs that allows them to be used as mixers, as presented in Chapter
19
5. There are two physical sources for this nonlinearity. The first comes from the
source and drain contacts to the CNT. When materials with different work functions
come into contact, electrons flow from the higher work function material to the lower
work function material, accumulating in the interface until the Fermi levels are equal.
These charges create a dipole between the materials and are responsible for bending
of the conduction and valence bands. When the work function of a metal contact is
less than the work function of a p-type semiconductor, the dipole created impedes
hole transport from the metal to the semiconductor. The energy barrier associated
with this dipole is called a Schottky barrier (?B)
18, 19. The metal contact material used
in our devices is titanium, which has a work function of 4.33 eV. The work function
of a p-type CNT ranges from 4.7 to 5.1 eV, so small Schottky barriers should exist at
both ends of our channels. Ignoring the effect of the gate field, the conduction at low
source-drain bias should be reduced in both directions by the Schottky contacts.
When the bias is large enough, the Schottky barriers will be overcome and conduction
should increase. This small difference in conductance should be symmetric about
Vds = 0 and is the first source of CNT FET nonlinearity.
The larger nonlinear behavior of the CNT FET drain current comes from the
gate field. Since the gate fixes the Fermi level inside the semiconductor, the relative
biases of the other two electrodes will determine the band bending, which can
introduce barriers to conduction. When the gate, source, and drain are all grounded,
the Fermi level is constant through all three materials. The potential energy diagram
for this configuration is shown in Figure 2.4 a). Energy diagrams are typically drawn
for electrons, so hole energy increases in the negative y direction. When a bias is
20
applied to the drain relative to the source, the Fermi level is fixed from the source
through the CNT, with most of the bias-induced band bending occurring at the CNT-
drain interface. When a positive drain bias is applied (Figure 2.4 b) the drain Fermi
level is lowered. Holes will move from the drain to the source in this configuration.
There are energy barriers for holes conducting from the drain to the CNT and from
the CNT to the source, both equal to ?B. When a negative drain bias is applied
(Figure 2.4 c) the drain Fermi level is raised. Holes will move from the source to the
drain in this configuration. The energy barrier for holes conducting from the source to
the CNT is still equal to ?B, but the barrier for conduction from the CNT into the
drain has been reduced to ?B ? eVds. This difference in energy barriers with drain bias
causes an asymmetry about Vds = 0 in the channel conduction.
Vgs = 0
+Vds -Vds
Vgs = 0
Source P-type
CNT
Drain
?B ?B
a) b) c)
Eholes
?B
(?B ? eV)?B
?B
EC
EV
EF
Figure 2.4 ? Diagrams of a top-gated p-type CNT FET with corresponding potential energy
diagrams for different source drain biases. a) Unbiased, Schottky contacts exist at the source
and drain due to the work function difference of the metal contacts and p-type CNT. b) For a
grounded gate and source and positive source-drain bias, holes conduct from right to left and
energy barriers exist at both interfaces. c) For a grounded gate and source and negative
source-drain bias, holes conduct from left to right and the energy barrier from the CNT to
drain has been reduced.
This modeling assumes that the gate voltage relative to the source voltage
(Vgs) is parameter responsible for the position of the Fermi level in the
21
semiconductor, which is the convention most commonly used with FETs. This means
that changing the voltage on the drain has no effect on the Fermi level inside the
CNT. However, in a symmetric device geometry, the labeling of source and drain is
arbitrary. If the gate voltage relative to the (Vgd) drain was used instead of Vgs, the
effective gate voltage would be changing as source-drain bias changed. The real
gating effect is a combination of Vgs and Vgd. It is more accurate to consider the gate
voltage relative to the voltage in the center of the CNT as responsible for setting the
Fermi level. This is only used in this section; the standard gating convention of Vgs is
used for the rest of this dissertation.
To evaluate the energy bands with this as the Fermi level reference, the
voltages are adjusted so that the center of the CNT is at ground. This is valid since it
is the potential differences that are relevant, not the absolute potential. Energy
diagrams with this labeling are shown in Figure 2.5. The gate and source are still at
the same potential. When a positive bias (V) is applied to the drain relative to the
source, the source has a potential of ?V/2 while the drain has a potential of +V/2
(Figure 2.5 a). Similar band bending near the contacts as in Figure 2.4 b) is shown.
However, since the gate potential is equal to the source potential, which is negative,
the Fermi level relative to conduction and valence bands is lowered, increasing the
number of carriers in the valence band and increasing conduction. When a negative
bias is applied to the drain relative to the source, the source has a potential of +V/2
while the drain has a potential of ?V/2 (Figure 2.5 b). Now the gate potential is
positive, which raises the Fermi level relative to the conduction and valence bands,
decreasing the number of carriers, and decreasing conduction.
22
-V/2
+V/2
a) b)
Eholes
EF
-V/2
+V/2
-V/2+V/2
eV
eV
Source Source
Drain
Drain
Figure 2.5 Diagrams of a top-gated p-type CNT FET with corresponding potential energy
diagrams for different source drain biases. a) For a positive bias on the drain relative to the
source, an effective negative potential exists at the gate, shifting the Fermi level down
relative to the conduction and valence bands. b) For a negative bias on the drain relative to
the source, an effective positive potential exists at the gate, shifting the Fermi level up
relative to the conduction and valence bands.
The size of the energy barrier at both contacts and the relative position of the
Fermi level both contribute to the nonlinear dependence of drain current on source-
drain voltage. These nonlinear effects are directly seen in our CNT FET devices.
Drain current vs. source-drain voltage measurements were made on a CNT FET with
the gate electrode left floating to observe the nonlinear effects from the Schottky
contacts and with the gate electrode grounded to observe the gate field induced
nonlinearity. Plots of the resulting current and dynamic conductance of both
measurements are shown in Figure 2.6.
23
Figure 2.6 Plots of a) drain current and b) dynamic conductance of a CNT FET with a
floating gate and with a grounded gate. When the gate is floating, the deviations in the
dynamic conductance with source-drain bias are due to Schottky contacts. When the gate is
grounded, the deviations in dynamic conductance with source-drain bias are due to the gate
field. The dynamic conductance was calculated from a linear fit over 100 mV.
The dynamic conductance plots demonstrate the asymmetry expected from the
potential energy diagrams. The effects of Schottky contacts are visible as a reduction
in conductance for low bias. A slight asymmetry in that curve is visible, likely due to
slightly different contact resistances for the source and drain. The contact resistance
can be extracted by comparing the resistance at zero bias where the contacts
contribute, to the resistance at ?2 V when the Schottky contacts have been overcome.
In this device, which was composed of hundreds of CNTs, the contact resistance is
near 4 k?. When the gate is fixed, the field it produces breaks the symmetry in the
dynamic conductance. The gate-induced nonlinearity in the CNT FET drain current
has a larger effect than the nonlinearity due to the Schottky contacts. It is this gate-
24
induced nonlinear behavior that will be used in the mixing experiments reported later
in this dissertation.
2.2.1 Device Substrate Selection
The most commonly used device substrate for micrometer-scaled electronics
is a silicon wafer with a silicon dioxide layer. The conductive silicon can act as a
global bottom gate in field-effect transistors (FETs), and the thermal oxide layer
provides a robust insulating dielectric. Lithography on silicon is well-established, so
the only new obstacles in fabricating carbon nanotube (CNT) FETs involves growing
and patterning the CNTs. The global silicon back gate is convenient in not requiring
the extra fabrication steps for adding a local top-gate, and its presence aids in the
removal of metallic CNTs from the network, as will be discussed in Chapter 4.
However, it is not always ideal. Integrated circuits with multiple FETs will require
local gating, with the output of one FET feeding into the gate of another. In such
cases, one could ignore the global back gate and add local top-gates in integrating
devices. For DC applications, that solution is appropriate. However, for AC
applications, especially as frequencies approach radio and microwave frequencies, the
presence of the conductive silicon layer will add unwanted parasitic capacitance.
Every metal electrode will be capacitively coupled to the silicon layer through the
thermal oxide, like a parallel plate capacitor, which is in turn capacitively coupled to
every other metal electrode. This extra capacitance will short every electrode together
as frequency is increased. One solution researchers use to overcome this problem is to
25
use highly resistive silicon substrates. Lower conductivity in the silicon will reduce
the capacitive coupling but not eliminate it. For our high frequency (HF) devices,
which are designed to operate at 1 GHz and above, we use insulating substrates.
Control devices fabricated on Si/SiO2 substrates are mentioned in Chapter 4. The step
by step fabrication procedure we follow for a CNT FET on a Si/SiO2 substrate is
given in Appendix A.
The last decade has shown a lot of interest in flexible electronics. Liberating
circuits from rigid substrates can have many advantages, such as less expensive
materials, adjustable sizes, such as a display that can roll up like a scroll, and
potentially fully transparent circuits, as many of the commonly used thermal plastic
substrates are optically transparent. It also allows for new applications, such as
smoothly integrating electronics into living anatomy20. Fabricating devices on plastics
presented new challenges since many of the chemical processes involved in silicon
lithography would damage or dissolve most plastic substrates. The method our group
developed to build devices on plastic was to fabricate each layer using standard
photolithography on silicon wafers first, and then transfer print each layer
sequentially to a plastic substrate. Our transfer printing process uses a nanoimprint
lithography machine to facilitate transfer from the hard silicon substrate to the soft
plastic one and is well documented in other publications.21-26 Alignment is the biggest
limitation in our ability to fabricate devices on plastic. Our alignment of subsequent
layers is limited to around 5 ?m resolution, which is insufficient for our HF FETs.
This is only a limitation of our equipment, not a fundamental limitation of plastic
electronics. Though some devices were transfer-printed into plastic, our HF
26
measurements were all performed on devices on quartz substrates. The step-by-step
fabrication procedure we follow for a CNT FET on a plastic substrate (polyethylene
terephthalate) is given in Appendix B.
One economic and simple way to grow single-walled carbon nanotubes is via
chemical vapor deposition (CVD). After depositing a ferritin catalyst solution onto a
substrate surface, the sample is heated in a tube furnace to nearly 1000 ?C and a
carbon-containing gas is introduced. The carbon nanotubes assemble themselves in
this environment. On most substrates there is no preferred growth direction, so the
result is a random network with CNTs going in all directions. While these random
networks can have useful bulk properties,27, 28 for most applications one would prefer
to have all the CNTs aligned along the same direction, with each CNT connecting the
source and drain. Since many of the CNTs in a random network will not bridge the
channel themselves, the current will pass from tube to tube as it passes through the
channel. Each tube crossing adds unwanted impedance and metallic CNTs will
inevitably be a part of the current pathway through channel. This will make it
impossible to completely remove all the metallic CNTs from the network, as will be
discussed in Chapter 4.
In 2005, John Roger's research group at Northwestern demonstrated aligned
CVD growth of CNTs, under certain conditions, on quartz substrates.29 The alignment
was due to using quartz wafers that had been ?miscut? at a small angle off the 011
plane. It was previously thought that the miscut produced periodically spaced atomic
steps, along which the single-walled CNTs grew, but more recently it has been
suggested that alignment is due to anisotropy in polarizability along the quartz surface
27
bond orientations.30 Whatever the mechanism, the ability to easily grow aligned
CNTs via CVD made quartz the substrate of choice for most of our devices. The lack
of a conductive back plane also made it a suitable substrate for our high frequency
devices.
2.2.2 CVD Growth and Lithography
For our device substrates, we used 3-inch-diameter single crystal quartz
wafers (0.5 mm thick, ST-cut 42?45?, single-side polished (SSP), from Hoffman
Materials). To prevent the CNTs from growing in all directions, producing a random
network, thermal annealing in air for 12 hours at 900 ?C is necessary before growth to
prepare the surface. The catalyst that we used for our nanotube growth was ferritin
purchased from Sigma Aldrich, a protein complex that contains iron nanoparticles.
Ferritin was deposited onto the sample surface by dipcoating, submerging the sample
in a dilute solution for twenty minutes. This distributes ferritin catalyst particles over
the whole sample surface. The sample is then rinsed with deionized water, blown dry
with nitrogen, and placed in a tube furnace for CVD growth. The sample is then
heated in air to 700 ?C to burn away the organic ferritin protein complex, leaving
behind iron nanoparticles. It is then cooled to room temperature and re-heated up to
925 ?C under hydrogen to prevent the particles from oxidizing. Methane is then
introduced at a flow rate of 1000 sccm to grow the CNTs. After 20 minutes the
methane is turned off and the sample is cooled to room temperature under hydrogen.
28
A scanning electron microscope (SEM) image of the resulting CNT array on the
quartz surface is shown in Figure 2.7.
Not all CNTs are perfectly aligned. This is due to the large number of ferritin
particles covering the surface, which will occasionally block the straight growth of a
CNT. This alignment is still preferred to a random network. The typical CNT density
for our samples ranged from 1 to 3 CNTs per micrometer.
Figure 2.7 SEM image of aligned CNTs grown on a quartz substrate by chemical vapor
deposition (CVD).
After the CNTs are on the quartz substrate, source and drain electrodes
consisting of 5 nm of titanium and 50 nm of gold are deposited via photolithography.
Alignment marks for the remaining layers are also patterned in this step. Since the
CNTs are grown over the entire surface, the channels of each device are covered with
29
photoresist and the uncovered CNTs are destroyed with an oxygen plasma. This step
defines the channel width. A 65 nm Al2O3 dielectric layer is then deposited via
atomic layer deposition (ALD) and lithographically patterned afterward with a
buffered oxide etch. Lastly, 100 nm gold local top-gates are added, again with a 5 nm
titanium adhesion layer. A detailed step-by-step procedure that we used to fabricate
our CNT FETs on quartz, including the CNT growth parameters and lithographic
recipes, is given in Appendix C.
2.2.3 Device Geometries
The field-effect transistor geometry used for our carbon nanotube devices is
shown in Figure 2.8. We utilize two carbon nanotube channels that share one drain
electrode and have separate outer source electrodes. The top local gates for each
channel are connected together at the gate contact pad. This device geometry was
chosen because of the ground-signal-ground (GSG) high frequency probes used for
AC measurements. The center electrode (signal) of one probe contacts the drain
electrode while the center electrode (signal) of the other touches the gate. The
grounded outer electrodes of both probes contact both source electrodes. The pitch of
the probes used was 150 ?m, so the spacing between each contact pad is also 150 ?m.
Devices were made with different channel widths (W), channel lengths (L), and gate
lengths (G). The channel width is set by etching the CNTs into different-sized strips,
producing channel widths of 100, 50, 25, 5, and 0 ?m. The channel lengths ranged
from 1 to 75 ?m. The gate lengths varied from overlapping the channel by 5 ?m to
30
leaving 5 ?m of the channel uncovered by a gate. The purpose of a gate electrode not
spanning the whole channel is to reduce the parasitic capacitance that comes from a
gate-to-source and gate-to-drain overlap. The standard device geometry used for most
measurements in this dissertation had a 100 ?m channel width, 10 ?m channel length,
and 8 ?m gate length.
Channel Length (L)
Gate Length (G)Gate
Drain
Source Source
Quartz
Ti/Au
Al2O3
Ch
annel W
idth (W
)
a)
b)
c)
d)
e)
Gates
Drain
source source
Figure 2.8 a) An optical image of a completed CNT FET in a high frequency probing
geometry. b) A top-view diagram of the same CNT FET geometry with the source, drain, and
gate electrodes labeled. c) A side-view diagram of the same CNT FET geometry showing the
substrate, source and drain electrodes, Al2O3 dielectric layer, and top-gate electrodes. d) An
SEM image of the nanotubes in a CNT FET dual channel before the dielectric layer and gate
are added. The channel width is labeled. e) An enlarged side-view diagram of a single
channel showing the gate length and channel length. Diagrams are not to scale.
Other components were also fabricated on each CNT FET device chip to be
used as calibration references and controls. Each included ground pads on each side
31
of the device in the same 150 ?m GSG spacing to enable probing in the same
configuration. The calibration references were open and through electrodes, used for
impedance calibration in vector network analyzer (VNA) measurements. They are
shown in Figure 2.9. All nanotubes near these electrodes were removed.
Figure 2.9 Optical images of the open and through electrodes used for VNA calibration.
Lastly, parallel plate capacitors with different plate sizes were also fabricated
to function as AC controls and to verify our capacitance measurement techniques.
Optical images of the fabricated capacitors and the side-view diagram showing the
dielectric and metal layers are shown in Figure 2.10. DC calculations and AC
measurements were used to extract the capacitances of these devices and of the CNT
FETs. The details of those calculations and measurements are given in Appendix D.
a)
b)
Figure 2.10 a) Optical image and b) side-view diagram of fabricated parallel plate capacitors.
32
Chapter 3: Charge Traps
3.1 Introduction to Charge Traps and Hysteresis in CNT FETs
In an ideal dielectric material, local charges are able to freely move within the
material in response to an external field inducing a polarization, but they cannot
conduct into adjacent materials. When the external field is removed, the charges
should go back to an electrically neutral arrangement. However, in real materials,
solids are rarely perfect crystals. Any variation in the atomic lattice, called a defect,
will produce a local change in the electric potential, which will attract or repel free
charges. These defects include the absence of an atom from a lattice point, leaving
behind dangling bonds, or an extra or different kind of atom at a lattice point.
Surfaces will also provide a local potential change because the bonds of an atom at
the surface of a material will be terminated differently than the atoms in the bulk of
the material. The interface joining two materials is a common location for defects
since different materials can have different lattice shapes and spacings. Also,
conduction between a dielectric and a conductor is possible through tunneling. When
an external field is applied through a real dielectric material, free charges may
become trapped at defect sites, such that when the field is removed, they remain
fixed, preventing the material from returning to an electronically neutral state. In a
field-effect transistor, the population of these trap states is strongly dependent on the
applied electric field (i.e. gate voltage) and is visible as an offset between electrical
measurements acquired with increasing, as opposed to decreasing, gate voltage. This
difference in current with sweep direction is due to the trap occupations that occurred
at previous gate voltages, the history of the electric field. Thus, it is referred to as
33
hysteresis and is commonly seen in carbon nanotube field-effect transistors (CNT
FETs) with SiO2 dielectric layers and is most often attributed to charge traps in the
dielectric layer or in water molecules on the dielectric surface if the CNTs are
exposed to air.31-42
The presence of trapped charges affects the CNT conduction by screening the
gate voltage from reaching the CNT. It can be thought of as follows: When the gate
voltage is negative, electron traps are being evacuated and leaving behind positive
holes, so when sweeping the gate voltage in the positive direction, the field reaching
the CNTs is more positive, so the drain current magnitude is decreased. For positive
gate voltages, electron traps are being filled, reducing the number of positive empty
traps, so when sweeping the gate voltage in the negative direction, the field reaching
the CNTs is less positive (more negative), so the drain current magnitude is increased.
A transfer curve of a CNT FET device is shown in Figure 3.1. The magnitude of the
hysteresis is quantified by the difference in threshold voltage (?Vgs) between the two
sweeps and by the difference in drain current at Vgs = 0 (?Id).
For devices with significant hysteresis, one cannot accurately predict what the
drain current will be at a particular source-drain and/or gate voltage; the output
depends on how the device arrived at those voltages, the device history. For some
applications, such as memory devices, hysteresis can be exploited to serve a
function,31, 43-45 but in typical field-effect transistors, it is problematic both for
extracting physical parameters from a device and for impedance matching when
integrating FETs into circuits. In order to reduce the problems associated with
34
hysteresis, the behavior of drain current as a function of time and gate voltage is
investigated.
Figure 3.1 Plot of a transfer curve taken at Vds = 5 V of a CNT FET (W = 100 ?m, L = 10 ?m,
G = 8 ?m) demonstrating hysteresis. The change in threshold voltage and drain current at
Vgs = 0 is shown.
3.2 Measurements of Traps Emptying Over Time
The most commonly referenced analysis of hysteresis in CNT FETs was
published by Slava Rotkin in 2005.41 Charge traps were modeled as a single
cylindrical trap surrounding the CNT. The total potential (EQ 3.1) was a combination
of the gate voltage (Vgs), a fitting parameter (V-), the potential due to the cylindrical
trap (VT), and charge distribution on the CNT (?). The Fowler-Noidheim tunneling
mechanism was used to account for charges moving from the CNT into a trap state in
35
the dielectric layer (EQ 3.2), and the charge distribution was calculated from the CNT
density of states (?) and Fermi distribution at the total potential (EQ 3.3). The drain
current was then calculated as the source-drain voltage divided by the nanotube
resistance. There is no closed form solution for the drain current in this model, but
numerical simulations can be made. All the physical values used with this model are
given in the article.
? ?
? ?
G
Tgs
C
t
tVVV
,
,
???? ???? ?
EQ 3.1
? ?
? ?
?
?
?
?
?
?
?
?
?
?
?
? ?
?? t
e
J
N
C
R
tV FNT
T
T
T
???? exp12,
EQ 3.2
? ? ? ? ? ?? ???? dEeEfEet ???? ,
EQ 3.3
? ?
? ?teL
V
tI dsd
,/
,
???
? ?
EQ 3.4
Transfer curve hysteresis was analyzed as a function of both gate voltage
sweep rate and maximum/minimum voltages values. As expected from a tunneling
mechanism, the more slowly the gate voltage is swept and the larger the bounds, the
greater the hysteresis. Rather than analyze the hysteresis in transfer curves, as has
been done already, we chose to observe how drain current changes over time after a
change in gate voltage.
As mentioned in Chapter 2, Our CNT FETs on quartz substrates are
encapsulated in atomic layer deposited (ALD) Al2O3 dielectric layers. Encapsulation
eliminates the charge traps associated with water molecules in the air and
significantly increases the operational lifetime of the device. Most of the devices
36
measured throughout this dissertation were over one year old. ALD is known to
produce very pure, pin-hole free dielectric layers.46, 47 However, CNT walls are inert
to most ALD precursors,48, 49 producing an imperfect interface between the CNT and
the dielectric bulk, providing a possible source for charge traps.
Most of the following measurements utilize the gate voltage waveform shown
in Figure 3.2. The gate voltage was first held at ground, then a short voltage pulse of
duration Ton and magnitude Vgs was applied to the gate. The gate voltage was then
returned to ground for a duration of Toff at which time an opposite gate voltage (-Vgs)
pulse was applied. Data acquisition is performed only during the Toff time interval
where Vgs = 0. Toff = 100 s for all the data presented in this section. The source-drain
voltage is held fixed throughout all measurements.
Vgs
Vgs Puls
-Vgs Pulse
Time
Toff
Ton
Figure 3.2 Diagram of the gate voltage waveform used to observe charge traps empty after
short pulses on the gate. The gate is held at ground until it is pulsed with the magnitude Vgs
Pulse for Ton. Then the gate is returned to ground for Toff to allow the charge traps to empty.
Then the gate is pulsed with the magnitude of ?Vgs Pulse for Ton, and again returned to ground
for Toff.
Drain current data for a CNT FET (W = 100 ?m, L = 10 ?m, G = 8 ?m) for
both a positive and negative gate pulse, (Ton = 1 s, Toff = 100 s, Vgs Pulse = 1 V) are
shown in Figure 3.3. This type of data will be referred to as discharge data. t = 0 is set
when gate voltage is returned to zero during the Toff interval. The response from the
37
positive and negative gate pulses are plotted together on the same time axis for direct
comparison. Measurements were made with a Keithley 2612A sourcemeter, which is
limited to ~ 500 ?s pulses. The sourcemeter was controlled with Keithley?s Test
Script Building software, which uses the Lua programming language.
Figure 3.3 Plots of discharge data over the 100 s Toff interval for Ton = 1 s and Vgs Pulse = 1 V.
The black circles correspond to ?Vgs Pulse and the white circles correspond to + Vgs Pulse. a)
is plotted on a linear scale and b) is a linear-log plot. The red lines correspond to logarithmic
fits to the data.
Plotting using a linear time scale (Figure 3.3 a), the discharge data indicates
that the majority of the change in current after both gate voltage pulses occurs in the
first second. The same data is replotted on a logarithmic scale for the time axis
(Figure 3.3 b) and fit to the logarithmic function given in EQ 3.5. Eventually, the
drain current reaches an equilibrium value and stops following this trend, seen as
deviations from the red lines in Figure 3.3 b). This equilibrium current value (Ieq = I(t
= ?)) is included in EQ 3.5 to highlight the deviations from equilibrium over time.
38
The logarithmic dependence on time will only last until the current reaches
equilibrium, at which point it will remain constant.
?
?
?
?
?
?
???
0
0)(
t
t
ALOGIItI eqd
for t < teq EQ 3.5
In EQ 3.5, I0 is referred to as the initial drain current offset and is equal to the
drain current difference from Ieq at t = t0 after a change in gate voltage (during Toff). t0
was chosen to be 1 ms because it is near the limit of our equipment measurement
resolution. The slope of the equation, A, is referred to as the discharge rate because it
specifies the slope (on a linear-log plot) of the drain current approaching equilibrium.
In the next section, we acquire discharge data as functions of the two variables (Ton
and Vgs Pulse) in our gate voltage measurement waveform (Figure 3.2). That data is
then fit to the logarithmic equation given in EQ 3.5 and analyzed in terms of I0 and A.
The purpose of this analysis is to gain insight into methods for minimizing the effect
of hysteresis. This logarithmic dependence with time of the drain current was
observed in simulations following the previously mentioned Rotkin charge trap
model. However, the model did not yield any significant dependence on the gate
voltage pulse duration nor the voltage pulse magnitude, so a more accurate model,
including a distribution of trap depths, would be needed to compare with our
measurements.
3.2.1 Dependence on Pulse Duration
Discharge data measurements were performed on a CNT FET with Vgs Pulse
= 1 V, and different Ton durations are shown in Figure 3.4. The points at the top of the
39
graph (negative slope) correspond to positive gate voltage pulses, while the points at
the bottom (positive slope) correspond to negative gate voltage pulses.
Figure 3.4 Plot of discharge data over 100 s for Vgs Pulse = 1 V and several values of Ton on a
logarithmic time scale. The initial offset current values are seen to depend strongly on Ton.
For Ton = 100 ms, the charge traps that were filled during the gate pulse are
mostly unoccupied after the first second into Toff. For Ton > 1 s the effects of filled
traps are still evident after 100 s. The data from these measurements were fit to
logarithms with time, and the extracted parameters are plotted vs. Ton in Figure 3.5 for
Vgs Pulse = 1 and 5 V. Since deviations from the logarithmic behavior occur when the
drain current approaches equilibrium, fits were made over the first 100 ms of
discharge data. The discharge rate will be referred to as the initial discharge rate, A.
40
Figure 3.5 Plots of a) the initial offset current values and b) the initial discharge rates vs. Ton
for Vgs Pulse = 1 V and Vgs Pulse = 5 V. For both, Ton is plotted on a logarithmic scale. The
initial offset current is defined as the current 1 ms after the gate transition from ? Vgs Pulse to
ground. The initial discharge rate is derived from a logarithmic fit to the discharge data
during the first 100 ms.
For most values of Ton the initial offset drain current follows an additional
logarithmic trend with Ton, shown as lines fit to the data. This dependence on the
Logarithm of Ton demonstrates the advantages of using equipment capable of
sourcing short pulse widths. Ton must be reduced by orders of magnitude to
substantially reduce the initial offset current. This indicates that the longer the gate is
held at a non-zero bias, the more trap states are filled. The lesson from this trend is
that the key to not populating charge traps with long decay times is to only use short
pulses on the gate when sourcing non-zero voltages.
The initial discharge rate, A, initially also depends logarithmically on Ton
before reaching a maximum value. For the Vgs Pulse = 5 V data, leveling off occurs
near Ton = 10 s. Considering that our fitting regime was 1 ? 100 ms, this suggests that
even trap states filled after long periods of gating will begin to empty immediately
41
(within 1 ms). This trend eventually ends, and traps filled at very long periods of
gating will not contribute to the initial discharge rate.
Though the discharge rate increases with Ton, this increased slope does not
compensate for the larger initial offset drain current. Therefore, I0 is the more
informative parameter. Employing the lesson from these measurements, in the next
experiment, discharge data was taken for Ton = 1 ms as a function of the gate voltage
pulse magnitude.
3.2.2 Dependence on Pulse Magnitude
Discharge data measurements were performed on a CNT FET with Ton = 1 ms
and five equally spaced values of Vgs Pulse are shown in Figure 3.6. The points at the
top of the graph (negative slope) correspond to positive gate voltage pulses, while the
points at the bottom (positive slope) correspond to negative gate voltage pulses.
42
Figure 3.6 Plots of discharge data over 100 s for Ton = 1 ms for several values of Vgs Pulse on
logarithmic time scales. The initial offset current values and the slopes of the discharging
currents can be seen to depend on Vgs Pulse.
The curvature of the discharge data on a log-log plot slightly deviates from the
logarithmic trend even for Vgs Pulse = 10 V which is not near equilibrium at the end
of the 100 s Toff period. For Ton = 1 ms the traps filled in 2 and 4 V gate pulses mostly
empty in 100 s. For the larger gate voltage magnitudes, 100 s is not long enough for
the current to reach equilibrium, even with the minimum Ton duration we can apply.
The data from these measurements were fit to logarithmic equations, and the
extracted parameters are plotted vs. Vgs Pulse in Figure 3.7.
43
Figure 3.7 Plots of a) the initial offset current values and b) the initial discharge rates vs. Vgs
Pulse for Ton = 1 ms. The initial offset current is defined as the current 1 ms after the gate
transition from ? Vgs Pulse to ground. The initial discharge rate is derived from a logarithmic
fit to the discharge data during the first 100 ms.
The initial offset drain current is approximately linearly dependent on Vgs
Pulse for small voltage magnitudes, with larger than linear offsets induced for large
values of Vgs Pulse. The initial discharge rates follow a linear relationship with Vgs
Pulse. This direct dependence of trap occupation on gate voltage magnitude indicates
that the gate voltage could be used to fill and empty trap states in a predictable way.
The initial offset drain current is seen to depend directly on both the gate
voltage pulse duration (Ton) and magnitude (Vgs Pulse). The logarithmic dependence
on duration indicates that using short pulses will minimize the number of traps
occupied during a non-zero gate voltage pulse. However, even when using 1 ms
pulses, 100 s of Toff is not enough time for all the traps filled at large voltage
magnitudes to empty, as seen in Figure 3.6. Projecting the trend forward, it would
take orders of magnitude more time for some of the measured currents to reach
44
equilibrium. Rather than waiting for them to empty, we use the trap occupation
dependence on gate voltage to force the occupied traps to empty in a controlled
fashion. The process will be modeled in the next section.
3.3 Preisach Model of Hysteresis
The presence of charge traps in our encapsulated CNT FETs is a consequence
of the materials and deposition methods we have available. We would therefore like
to minimize their effects during measurements, most notably by reducing the
hysteresis in transfer curves. Since hysteresis is caused by having different charge
trap occupations when different measurements are taken, one solution is to set a
default trap occupation in between measurements. We have seen that both time and
gate voltage affect trap occupation. One solution is to set the gate voltage to ground
after each measurement pulse and wait for the filled traps to empty, but as we have
seen, for large gate voltages the length of time until the drain current reaches
equilibrium can be orders of magnitude larger than 100 seconds, an impractical time-
period between measurements. Instead we focus on using the gate effect to set a
default trap occupation.
Ignoring the time dependence allows us to use the Preisach hysteresis model50,
51 to predict what kind of gate voltage waveform might set a default trap occupation.
This model is regularly used with ferromagnetic systems but has not been widely
applied to electronic hysteresis. The Preisach model approximates hysteresis by
representing data as a collection of hysterons. A hysteron is a simple relay whose
45
output is 0 if the input is less than ?, 1 if the input is greater than ?, and maintains its
previous value if the input is between ? and ?. A diagram of a hysteron is shown in
Figure 3.8 and its output is given by EQ 3.6.
R?,?
? ?
1
x
Figure 3.8 Diagram of a hysteron relay: x is the input, y is the output, and ? and ? are the
transition x values.
?
?
?
?
?
??
?
?
?
??
?
?
??
xfork
xfor
xfor
xR 1
0
)(,
k = previous y value EQ 3.6
Figure 3.9 shows how a hysteron can be related to a single trap state in a CNT
FET. At point a, the positive field from the gate draws an electron into the trap. A
trapped electron produces a negative field that encourages p-type CNT channel
conduction, so the output from that trap is one and the relay is considered on. At point
b the gate field has been removed, but the electron remains trapped, so the relay is
still on. At point c the negative gate voltage emptied the trap, leaving behind a hole
that produces a positive field, diminishing CNT channel conduction, so the output
from the trap is 0 and the relay is considered off. At point d the gate field has been
removed, but the hole is still trapped, so the relay is still off.
46
Vgs > 0Vgs = 0
Vgs < 0 Vgs = 0
Id
Vgs
Gate
Al2O3
Quartz
CNT
ab
c d
Figure 3.9 Diagrams relating a single trap state in a CNT FET to a hysteron relay in the
Preisach model. Four points on the relay are matched with device diagrams showing the
dielectric polarization and the trap state occupation.
By filling the ?-? parameter space with hysterons and weighting them
properly, any hysteresis loop data can be represented. An example of six hysterons
representing a filled system is shown in Figure 3.10.
?
?
(-1,-1)
(1,1)(-1,1)
1
R1
2
3
4
5
6 R2
R3
R4
R5
R6
a) b)
Figure 3.10 Diagrams of a) ?-? parameter space with six hysterons and b) the associated
relays plotted with equal weights.
47
A Matlab script was written to execute sweeps of the input variable and see
the resulting Preisach output and hysteron states. The script is included in Appendix
E. Dividing the ?-? plane into 100 vertical and horizontal sections produced 5050
equally spaced hysterons. As the input (x) is swept, the output is the sum of the
equally weighted hysterons, computed for each input value. Sweeps from x = ?1 to 0
by steps of 0.01 are shown in Figure 3.11. The red dot in the x-y plane represents the
output at x = 0 for the path taken. The ?-? plane shows which hysterons are on at the
end of the sweep. At x = -1, all the hysterons are off, so when x is swept to 0, most
remain off and the output is low. At x = 1, all the hysterons are on, so when x is swept
to 0 most remain on and the output is high. Our goal is to achieve a neutral hysteron
arrangement, which would correspond to half of the hysterons on and half off. Since
these hysterons are equally weighted, that would relate to the center of the x-y plot (0,
0.5).
48
Figure 3.11 Plots of x-y planes from the Preisach hysteresis model using 5050 equally
weighted hysterons to sweep from a) x = -1 to 0 and c) x = 1 to 0. b) and d) are the ?-? planes
corresponding to x = 0 after each sweep. Hysterons that have been switched ?on? are shown
in red; those in the off state are plotted in black.
In this model, there is no difference in the final output state between sweeping
to a value and jumping directly to that value, since each hysteron only depends on x
being less than ? or greater than ?. The final x value will switch all the hysterons that
would have been triggered in the sweep up to that final value. Figure 3.11 clearly
49
shows that sweeping directly to x = 0 from a large x value cannot reach the center of
the plot. Our method is to approach x = 0 by alternating between positive and
negative inputs that decrease in magnitude. It is similar to a degaussing a cathode ray
tube to reduce magnetic hysteresis. The particular values used are most easily
understood in the for-next loop labeled EQ 3.7.
For i = 0 to N
?
?
?
?
?
?
??
N
i
xx 1max
?
?
?
?
?
?
???
N
i
xx 1max
EQ 3.7
Next i
N is the number of divisions as the magnitude of the input approaches zero.
For N = 4, the inputs would be x = 1, -1, 0.75, -0.75, 0.5, -0.5, 0.25, -0.25, 0, 0.
Preisach output plots for N = 4 and N = 100 are shown in Figure 3.12. Plot a) shows
the output spiraling in toward the center of the plane. Plot b) shows a nearly even
distribution of hysterons on and off, but more are off than on, so the final output is
less than 0.5. When N is increased to 100 the hysterons appear evenly divided
between off and on, and the final output is 0.5.
50
Figure 3.12 Plots of x-y planes from the Preisach hysteresis model using 5050 equally
weighted hysterons to perform an alternating input sweep down to x = 0 a) for N = 4 and c) N
= 100. b) and d) are the ?-? planes corresponding to x = 0 after each sweep. Hysterons that
have been switched ?on? are shown in red; those in the off state are plotted in black.
This simple model indicates that if charge traps behave like input-dependent
relays, this alternating input on the gate should bring the current to its equilibrium
value, setting a default trap state occupation. In the next section this alternating gate
input will be tested on our devices to see if the charge trap effects are reduced.
51
3.4 Hysteresis Reducing Gate Waveforms
Multiple research groups have published waveforms intended to reduce
hysteresis in transfer curves for CNT FETs.37, 42, 52-55 Four will be tested in this
section, including the one derived from the Preisach model in the previous section.
They are shown in Figure 3.13 and labeled simply waveform A, B, C, and D. The rest
of the data in this section will correspond to the colors of each waveform.
Figure 3.13 Waveforms of gate voltage vs. time used to reduce hysteresis in transfer curves.
a) Waveform A sources each gate voltage sequentially. b) Waveform B returns to ground in
between pulses. c) Waveform C returns to ground in between alternating pulses. d)
Waveform D executes an alternative decreasing waveform in between measurement pulses.
Waveform A is a continuous sweep, the standard transfer curve measurement.
It acts as a control and should produce the most hysteresis. Waveform B comes from
a 2010 publication by Pop.42 It sources a gate voltage for Ton = 1 ms before returning
to ground for Toff = 10 s to allow the trap states to empty. Waveform C comes from a
2010 publication by Mattman.54 It is very similar to Waveform B except that it
alternates gate polarity between measurements and Toff = 1 ms. Waveform D is our
method derived from the Preisach model to set a default trap configuration between
measurement pulses. A series of alternating pulses (N = 100) that decrease in
magnitude are executed after each measurement pulse to empty the trap states filled
during each measurement pulse.
52
The same CNT FET (W = 100 ?m, L = 10 ?m, G = 8 ?m) was measured using
all four waveforms. For the pulsed waveforms (B through D) Ton = 1 ms. Gate
voltages were swept between ?10 V by steps of 100 mV. Keithley?s Test Script
Builder was used to load each sweep waveform into the buffer of a Keithley 2612A
sourcemeter which carried out the measurements. The resulting transfer curves are
shown in Figure 3.14 and hysteresis width (?VT), hysteresis height (?Id), and
measurement duration are listed in Table 3.1.
Figure 3.14 Transfer curves of a top-gated aligned CNT FET (W = 100 ?m, L = 10 ?m, G = 8
?m). Each curve corresponds to a different measurement sweep. To observe each curve
clearly, they are offset by increments on 10 V, with the vertical dashed lines corresponding
Vgs = 0 for each sweep. Each method swept between +-10 V by steps of 100 mV.
Waveform ?Id ?VT Time
A 265 uA 2.81 V 0.925 s
B 136 uA 0.90 V 4013 s
C 36.3 uA 0.19 V 1.90 s
D 1.8 uA 0.01 V 29.4 s
Table 3.1 Hysteresis data from the transfer curves in Figure 3.15. ?Id is the difference in
drain current between sweep directions at Vgs = 0. ?VT is the difference in threshold voltage
between sweep directions, extracted from the linear extrapolation method at the point of
maximum transconductance. Time is the measurement duration.
53
All three pulsed methods produce transfer curves with reduced hysteresis from
the sequential sweep, Waveform A. Waveform B cut the hysteresis width by three but
took over an hour to execute. Waveform C was the fastest, completing a full sweep in
less than two seconds. Waveform D showed the least amount of hysteresis and took
30 seconds to execute, which is a reasonable time frame to acquire a transfer curve.
As discussed in section 3.2, hysteresis is not the only indicator of charge trap
behavior. Waveforms B through D all return to Vgs = 0 in between measurement
pulses. The drain current immediately before a measurement pulse should provide an
indication of the device trap state. If a waveform successfully sets a default trap state
occupation in between measurements, this current value should be independent of Vgs
Pulse. The Presiach model was used to predict the drain current at Vgs = 0 in between
measurement pulses for waveforms B ? D. Those results were scaled to match our
output current range and are plotted next to the measured data in Figure 3.15.
Since all three methods successfully reduce hysteresis, observing this drain
current in between measurements better distinguishes the waveform effects than
comparing transfer curves. Considering that the Preisach model does not consider the
time dependence of trap states and that all the hysterons were evenly weighted, the
correspondence between the model and the measurement is impressive. Waveforms B
and C produce drain currents with a strong dependence on Vgs Pulse, while the drain
current in between pulses from Waveform D is close to a single value, independent of
Vgs Pulse. This demonstrates that our proposed method successfully sets a default
charge trap occupation in between each pulse measurement.
54
Figure 3.15 Drain current offsets from a) the Preisach model and b) measurements at Vgs = 0
in between the measurement gate pulses for gate voltage waveforms B, C and D. The
predictions used the Preisach model and were scaled to match the measured data.
To evaluate the effectiveness of this measurement waveform, multiple transfer
curves were taken of the same device with Waveform D for different values of N. The
resulting hysteresis width (?VT) and hysteresis height (?Id) are plotted in Figure 3.16.
Surprisingly, even a single alternating pulse on the way to Vgs = 0 (10 V, -10 V, 0 V)
reduced hysteresis more than the other waveforms. The trap state that input would
induce is not the equilibrium state, but since it is executed in between each
measurement, the dependence on the measurement pulse magnitudes is small.
55
Figure 3.16 Plot of the resulting hysteresis height (?VT) and width (?Id) for transfer curves
taken with Waveform D vs. the number of oscillations (N) in the waveform. Eight transfer
curves between +-10V at a step of 0.1V were taken of the same device.
3.5 Summary
In this chapter we have introduced the phenomenon of defects in dielectric
material that act as traps for free charges. The effects of these charge traps are visible
as hysteresis in transfer curves. By sourcing pulses on the gate electrode of our CNT
FETs and observing how the drain current responds over time, we find that the initial
drain current offset (the distance the drain current is from its equilibrium value 1 ms
after a gate pulse) is linearly dependent on the pulsed gate voltage and has a
logarithmic dependence on pulse duration. This information was used to devise a
measurement scheme that would minimize the effects of trapped charges.
56
By using short pulses on the gate, charges have less time to become trapped
by defects, and therefore the number of charges trapped will be minimized when the
gate voltage is changed. The strong linear dependence of the offset current on the gate
voltage allows us to use a gate voltage waveform to force a default charge trap
occupation in between measurements, eliminating the production of hysteresis in
transfer curves. The time-independent Preisach model is used to derive this
waveform, which consists of alternating polarities as the voltage magnitude collapses,
similar to the method used to degauss a cathode ray tube. Transfer curves of
encapsulated CNT FETs were taken with this gate voltage waveform and with other
published gate voltage waveforms intended to reduce hysteresis in transfer curves.
Our Preisach-derived gate voltage waveform demonstrated the greatest reduction in
hysteresis, with the hysteresis width and height reduced to the level of measurement
noise. With a source-meter capable of 1 ms pulses, a typical transfer curve can be
taken in less than one minute by using this method.
57
Chapter 4: Removal of Metallic CNTs from Mixed Arrays: Burning
In this chapter we discuss the role of the on/off current ratio in carbon
nanotube (CNT) field-effect transistors (FETs). The CNTs arrays produced by
chemical vapor deposition (CVD) contain a mixture of metallic and semiconducting
CNTs, which produce an as-grown on/off ratio near 3. Some applications require a
larger on/off ratio to properly function, so the on/off ratio is increased in CNT devices
on SiO2 by using the standard technique of selective electrical breakdown from Joule
heating, referred to as burning. Improvements are made to the method to preserve
more on-state drain current after burning.
The standard method of burning cannot be directly applied to CNTs on high
frequency substrates, like quartz, because they lack a conductive backplane to be used
as a gate. A method is devised and employed to adapt the standard burning procedure
to devices on quartz. That process is further adapted to simultaneously burn many
CNT channels in parallel. The presence of metallic CNTs in FETs used for high
frequency applications produces a DC current offset when biased. If CNTs are ever to
be used in RF handheld electronics, power consumption will be a concern, so
increasing the on/off ratio of as-grown CNT arrays on HF-compatible substrates is a
necessary step toward that goal.
4.1 On/Off Current Ratio
As mentioned in Chapter 1, carbon nanotubes? (CNTs) structure exhibits a
variety of chiralities which give rise to both metallic and semiconducting properties.
58
The three most common techniques for synthesizing CNTs (arc discharge1, laser
ablation13, and chemical vapor deposition14, 15, 56-59) each produce a mixture of both
types of tubes which typically contain twice as many semiconducting tubes as
metallic tubes.60 However, not all chiralities will be equally likely to grow in all
circumstances. The CNT diameter is related to the catalyst particle size and the
growth temperature and pressure24, 61-64; therefore the ratio of semiconductor to
metallic tubes can be somewhat variable.
The presence of metallic CNTs is non-ideal in most semiconductor
applications. One of the most significant problems is the inability of a CNT device
containing metallic tubes to be gated off. When a positive gate voltage is applied to a
p-type CNT FET, the associated electric field depletes the semiconducting CNTs of
charge carriers but leaves the metallic CNTs unaffected. Therefore the presence of
metallic tubes limits the degree to which a CNT FET can be turned off. For a given
source-drain voltage, the ratio of the on-state current to the off-state current is
referred to as the on/off ratio and is a key metric for the FETs.
Experiments have shown that typical CNT FETs fabricated from CVD grown
nanotubes will have an on/off ratio of around 3. One simple explanation for this is
that the on/off ratio is directly related to the ratio of semiconducting to metallic
CNTs. When a device is gated on, the total current is due to both N metallic CNTs
and 2N semiconducting CNTs; thus 3N total CNTs contribute to the on-state current.
When the device is gated off, only the N metallic CNTs should be conducting, so the
on/off ratio should be close to 3N/N = 3. The on/off ratios of our as-grown CNT
FETs range from 1.5 to 10.
59
Whether or not an on/off ratio near 3 is acceptable depends on the application.
Many DC circuits require an on/off ratio of 1000 or greater to function properly. For
high frequency (HF) purposes, a low on/off ratio merely gives rise to a DC offset
when the channel is biased. With regard to power consumption and/or heat
dissipation, the presence of a constant channel current could be detrimental.
Therefore, the performance of both DC and AC circuits benefit from improved on/off
ratios. Increasing the on/off ratio requires the selective removal of the metallic CNTs.
This is typically accomplished by joule heating in air, leading to electrical
breakdown, and is commonly referred to as burning.
4.2.1 Standard Burning Procedure on SiO2
The standard method for burning metallic CNTs was first published in 2001.65
It is typically performed on thermally oxidized silicon wafers (Si/SiO2) substrate with
CNT contact electrodes in place. The conductive silicon wafer provides a global back
gate to turn off semiconducting tubes while a voltage is held between the CNT
contact electrodes. As the voltage between the CNT contact electrodes is increased,
the amount of current and power dissipated in each CNT increases, causing
significant self-heating. Once the CNT temperature reaches approximately 600 ?C in
air, it will oxidize and physically break at the point of maximum temperature, usually
in the middle of the CNT.66 Though current will be flowing through all the CNTs
bridging the channel, by using the back gate to increase the resistance of the
60
semiconducting CNTs, the metallic CNTs should pass more current and therefore
heat up and burn before the semiconducting CNTs.
This method is straightforward when dealing with single CNT channels or
aligned arrays where each CNT spans the channel, but is more difficult to execute on
CNT random networks where some metallic CNTs may be necessary for channel
conduction. Each tube-to-tube crossing can provide a defect where local heating may
destroy a CNT prematurely.67 An analysis of the on/off ratio dependence on CNT
density, channel width, and channel length for random network CNT FETs was
performed by a group member, Vinod Sangwan, and is detailed in his publications.27,
68 To avoid the complications associated with random networks, aligned arrays of
CNTs were transferred from quartz to SiO2/Si for devices. Without CNT crossings,
each CNT could be thought of as independent, so the burning process can
theoretically be carried to completion resulting in the removal of all the metallic
CNTs connecting the channel.
Field-effect transistors on SiO2/Si substrates utilizing aligned arrays of CNTs
were prepared following the procedure described in detail in Appendix A. Briefly,
aligned CNTs were grown on a quartz substrate via CVD, covered with a 100 nm
layer of gold in an e-beam evaporator and then covered with a layer of thermal tape.
The gold-covered CNTs were peeled off of the quartz growth substrate and
transferred to a SiO2/Si substrate with the thermal tape. The SiO2/Si sample was then
heated on a hot plate above 120 ?C to melt the tape adhesive and remove the tape,
leaving behind the gold/CNT layer on the SiO2 surface. The gold layer is then etched
away with a potassium chloride chemical etchant. With the aligned arrays of CNTs
61
now on an SiO2/Si substrate, 50 nm gold (with a 5 nm titanium adhesion layer)
contact electrodes were fabricated using standard photolithography, e-beam
deposition and lift-off. CNTs not in the channel region or covered by the electrodes
were removed using standard photolithography followed by O2 plasma etching (100
Watts, 16 SCCM and 200 mTorr for 10 sec) and photoresist removal in acetone. The
resulting structures, shown in Figure 4.1, can also be used as FET devices, with the
silicon substrate acting as the gate.
Si
300nm SiO2
Aligned CNTs
Ti/Au Electrodes
a) b) c)
Figure 4.1 a) Diagram of FET geometry on Si/SiO2 with aligned CNTs in the device channel.
b) Optical image of device geometry and c) SEM image of aligned CNTs in one channel.
Before burning, standard IV curves were taken. The removal of the metallic
CNTs was accomplished by setting a constant gate voltage of 40 V using the silicon
substrate as a global back gate. Then the source-drain bias was swept from 0 to 35 V
and back to 0 V as a rate of 1 volt per second. Standard IV curves were then
measured again. The on/off ratio was calculated with a source-drain bias of -2 V, with
Vgs = -10 V and +10 V for the on and off states, respectively. To increase the on/off
ratio, the burning process was repeated with a higher maximum bias of 40 V for a
second sweep and 45 V for a third sweep.
62
An example of the Id vs. Vds measurements associated with this burning
process is show in Figure 4.2. The drain current is plotted against the source-drain
voltage measured during three sequential voltage sweeps of burning a device with a
channel length of 10 ?m and a channel width of 500 ?m. The noise in the curves at
high source-drain voltages correspond to individual CNTs breaking.
Figure 4.2 Drain current vs. source-drain voltage plot of three burning sweeps of a CNT
channel in SiO2. The gate voltage used during burning was +40V. The device had a channel
width of 250 ?m, a channel length of 10 ?m, and used the silicon global backgate.
Analysis of CNT breakdowns from multiple publications have shown an
approximate breakdown voltage dependence on channel length of ~ 5 V/?m.66, 69
With a 10 ?m long channel, one would expect all the CNTs to experience electrical
breakdown around 50 V source-drain bias. We observe initial breakdown occurring
below half that bias. Output and transfer curves for both CNT FETs containing as-
grown and burned CNT are shown in Figure 4.3.
63
Figure 4.3 IV curves of a CNT FET on SiO2. a) and b) correspond to the as-grown device,
and c) and d) were taken after a burning procedure was executed. The device had a channel
width of 250 ?m, a channel length of 10 ?m, and used the silicon global backgate. The
transfer curves are plotted with drain current on a log scale.
A comparison of the transfer curves indicate that the on/off ratio has been
increased by approximately three orders of magnitude. After burning, however, the
on-state current has been reduced to 6 to 11% of the original on-state current
measured for the as-grown CNT FET before burning, depending on source-drain bias.
64
This is much lower than the 1/3 reduction that would be expected from removing the
metallic CNTs only and leaving all the semiconducting CNTs intact. This implies that
either a large number of semiconducting CNTs have been destroyed along with the
removal of the metallic tubes as a result of this burning process, or the metallic
nanotubes carry higher current on average.
Experiments on highly dense aligned arrays of CNTs have demonstrated that a
single CNT electrical breakdown can trigger breakdowns in adjacent CNTs that
cascade across an entire channel.67, 70 This suggests that semiconducting tubes less
than a critical distance from a metallic CNT will also be destroyed during burning. It
has been shown that a transition from destroying single CNTs to a correlated
breakdown of neighboring CNTs occurs in the density range of 1-30 CNTs per ?m of
channel width. Since the aligned arrays of CNTs used in these measurements are in
the density range of 1-3 CNTs/?m, it can be expected that some of the
semiconducting CNTs will be damaged by this process. As a result, there is a trade-
off between on/off ratio and on-state current magnitude. To preserve current, the
burning process can be stopped once an adequate on/off ratio has been reached.
To establish some statistics for the relationship between improved on/off ratio
and the reduction in on-state current, 36 separate CNT FETs with channel lengths of
10 ?m and channel widths of 200, 500, and 1000 ?m were fabricated on a single chip
and measured before burning, after one burning cycle, and after multiple burning
cycles. A plot of the normalized on-state current (normalized to the as-grown on-state
current) vs. on/off ratio for these devices is shown in Figure 4.4.
65
Figure 4.4 Plot of normalized on-state drain current vs. on/off ratio for 36 different CNT
FETs on SiO2 at multiple stages of burning.
Regardless of the number of burning cycles, a large variation in both on-state
current and on/off ratio is evident. The results range from on-state current reduction
of 74% with an on/off ratio > 103 to an on-state current reduction of 94% with an
on/off ratio near 10. Order of magnitude on-state current losses have been reported
elsewhere.71 It is clear from these results that better burning methods are needed if
high performance CNT FETs with burned channels are to be routinely fabricated. The
next section of this chapter will introduce and develop concepts of pulsed burning
methods as a way of preserving more on-state current and achieving more consistent
final devices.
66
4.2.2 Pulsed Burning on SiO2
One cause of the large distribution in results shown in Figure 4.3 may involve
charge trapping. As discussed in Chapter 3, the field reaching each CNT is a
complicated combination of the source-drain bias, gate-source bias, adjacent CNTs,
and charge traps in the dielectric or on the surface from water molecules. By holding
the gate at a constant 40 V for an extended time period, charges are drawn into traps
and diminish the effect of gating. To reduce this effect, the burning procedure is
adjusted to only source the gate voltage shortly before the burning source-drain bias
is applied.
A plot of a voltage waveform used for employing pulses instead of fixed
voltages during burning is shown in Figure 4.4. Initially, both the gate and source-
drain are grounded. The gate bias is then increased to a large positive voltage (20 V)
to gate off the semiconducting CNTs 100 ms before the source-drain voltage is
sourced for 100 ms to burn out the metallic CNTs. Then both voltages are returned to
ground. An on/off measurement is taken at Vds = 1 V with Vgs = -10 V and +10 V for
the on and off states, respectively. The process is then repeated with a higher pulsed
source-drain bias. Only one period is shown in Figure 4.5.
67
Figure 4.5 Gate voltage and source-drain voltage waveforms used to perform pulsed burning,
then perform an on/off ratio measurement of CNT FETs.
The results of one CNT FET burned with the pulsed gate voltage method are
plotted in Figure 4.6 with the previous data from Figure 4.3 of the 36 devices burned
with a constant gate voltage. The one device that underwent pulsed burning does not
demonstrate any statistics, but it shows the approximate on-state current vs. on/off
ratio path that devices typically exhibit when burned in this way. Though some
devices from the previous method did produce on/off ratios and on-current
magnitudes higher than this sample using pulsed burning, on average this method
preserves more current. Other research groups have published similar pulse-style
burning procedures with shorter time scales and showed improvements over fixed
gate voltages and source-drain sweeps as well.72
68
Figure 4.6 Plot of normalized on-state drain current vs. on/off ratio for a CNT FET on SiO2
that underwent pulsed burning (black) for 36 CNT FETs on SiO2 at multiple stages of
burning (gray).
4.3 Burning on Quartz
Aligned CNTs grown on quartz can be transferred to SiO2/Si substrates for
burning, as was done for the devices in the previous section. Once burned,
encapsulating dielectrics and local top-gates can be added to gate each FET
separately. However, for AC applications the conductive silicon substrate will add
parasitic capacitance. To enable proper high frequency operation, an insulating
substrate like quartz must be used. Transferring CNTs from SiO2 back to quartz after
burning would subject them to the additional processing of etching away the CNT
contact electrodes, covering them in a transfer material like gold, transferring that
layer to a new substrate, etching away the transfer material, and re-depositing contact
69
electrodes and an encapsulating dielectric. Additionally, a very difficult degree of
alignment would be needed to re-use burned channels, which our thermal tape
transfer method cannot accommodate. During the burning process, metallic CNTs are
broken, not completely destroyed, so a slight offset in redefining new contact
electrodes could produce channels with previously broken metallic CNTs now
bridging the new channel. It would be more efficient to burn away metallic CNTs on
the growth quartz substrate.
Since a quartz substrate does not contain a back gate for burning, another way
of controlling the electric field in the channel is necessary. Burning after
encapsulation is impractical since the breakdown temperatures for encapsulated
CNTs have been found to be higher than 600 ?C.73-75 Also, without air to provide
oxygen and dissipate heat, the breakdown of CNTs can damage electrodes. The
method that we employed to burn on quartz before encapsulation was to bring a
separate SiO2/Si piece to rest on top of the CNT contact electrodes to act as a
temporary global top-gate. The contact electrodes were made thicker than previously
(100 nm) to provide an air gap. With the added air gap, the field reaching the CNTs
will be weaker than it would be from a native global back gate, so the gate voltage
used for burning was increased. The surfaces of the quartz sample and the SiO2/Si
sample were carefully cleaned with acetone, methanol, and isopropyl alcohol and
then blown dry with nitrogen. The cleaner the two surfaces, the better the contact will
be, producing a more effective gate field. The SiO2/Si sample was held in place by a
DC probe. Proper conduction was assured by scratching the back of the silicon piece
70
with a diamond scribe and adding silver paint. A diagram of this procedure is shown
in Figure 4.7
Figure 4.7 Diagram of the external Si/SiO2 top-gate burning method on a quartz CNT FET
before dielectric encapsulation.
To access the CNT source-drain electrodes with probes while the SiO2/Si
external top-gate was in place, the contact electrodes were lithographically extended
several millimeters away from the center of the sample. In addition to extending the
contact electrodes, several devices shared the extended contact electrodes to enable
burning multiple devices at once, as shown in Figure 4.7. With these electrodes in
place, the combination of several devices connected in parallel resemble one device
with multiple channels. The number of devices able to be burned simultaneously is
limited by the magnitude of the collective drain currents associated with so many
CNTs in parallel under the source-drain biases necessary to achieve burning. The
largest number of CNT FETs we have burned in parallel is 15, which produced a
drain current greater than 60 mA during burning. The IV curves for 15 CNT FETs on
quartz measured in parallel before and after employing the external top-gate burning
method are shown in Figure 4.9. Before burning, the on-state current at Vds = 1 V was
9.3 mA, and the on/off ratio for Vgs = ? 100 V was 1.3. After burning, the on-state
current at Vds = 1 V is 100 ?A (~ 1% of the as-grown current) and the on/off ratio for
71
Vgs = ? 100 V is 52. These on/off ratios are measured using the external top-gate. The
on/off ratios of the completed devices should improve with a local top-gate as shown
below.
Figure 4.8 Optical image of joined and extended CNT contact electrodes on quartz. These
electrodes link together up to 15 FET channels so that they may be burned simultaneously.
72
Figure 4.9 a), c) Output and b), d) transfer curves for 16 CNT FETs measured in parallel on
quartz using an external Si/SiO2 piece as a top-gate measured a), b) before and c), d) after
burning.
These plots show a reduction in the on-state current to near 10% of the
original as-grown CNTs, which is similar to the reduction observed with devices on
SiO2. After burning is complete, the electrodes connecting the FETs are
lithographically removed. First, a lithographic pattern is defined that is open where
73
the electrodes are to be etched away. Then a potassium iodide gold etchant is used to
remove the 100 nm gold top electrode layer, and after that a fluoroform plasma
(40mTorr, 175W, 18sccm CHF3, 2sccm O2) is used for one minute to etch away the
remaining 5 nm titanium adhesion layer. After the source-drain electrodes are
isolated, a 150 nm Al2O3 dielectric layer and 100 nm gold (with a 5 nm titanium
adhesion layer) local top-gate are added. Optical images of a sample showing
connected devices with a) photoresist in place before etching, b) separated electrodes
after etching, and c) completed devices are shown in Figure 4.10.
Figure 4.10 Optical images of the lithographic steps of isolating and completing CNT FETs
on quartz after electrical breakdown. a) Patterned CNTs and source-drain electrodes are in
place with the connections between FETs uncovered by photoresist. b) The electrodes
previously connecting devices have been removed. c) Patterned ALD alumina layers and
local top-gates were added to complete each device.
74
Output and transfer curves are shown in Figure 4.11 for one of the 15 devices.
A local gate functions much more efficiently than the external silicon top-gate with
an air gap, so the resulting device on/off ratio is greater than 105. All of the devices
burned in parallel demonstrated on/off ratios above 103 after they were completed
with local top-gates. Ambipolarity at large positive gate voltages is visible, as is
typical of encapsulated CNT FETs. These results demonstrate the effectiveness of this
external top-gate burning method on quartz.
Figure 4.11 a) Output and b) transfer curves for a CNT FET on quartz encapsulated in Al2O3
with a local top-gate after external top-gate burning was executed and devices were
separated.
4.4 Summary
CVD growth of CNTs produces a mixture of semiconducting and metallic
tubes. The presence of metallic CNTs prevent a CNT FET from being effectively
turned off, providing a constant drain current when biased that is undesirable for
75
applications that require low power operation. The standard method of removing
metallic CNTs is typically performed on a Si/SiO2 substrate, utilizing the silicon as a
global back gate to turn off semiconducting tubes while a voltage is held between the
CNT contact electrodes.
Burning experiments on SiO2 produce a large variation of resulting on/off
ratios among samples due to the effect of adjacent CNTs, contact resistance, defects,
and variations in breakdown temperature due to chirality. To reduce the effects of
trapped charges due to defects, the standard burning procedure was adjusted to use
pulses instead of constant biases, which resulted in preserving more on-state current.
To remove the metallic CNTs from aligned arrays on quartz substrates, a
method was developed to use an external Si/SiO2 substrate to rest on top of the
source-drain electrodes and act as a temporary global top-gate. Additionally, several
devices were connected in parallel to allow burning up to 15 channels at once. This
method was shown to be successful, consistently increasing the on/off current ratio of
the CNTs from around 3 to greater than 103. If active CVD-grown CNT devices are to
be used for high frequency applications where power consumption is important, this
is a necessary step toward that goal.
76
Chapter 5: HF CNT Measurement Techniques and Limitations
Many research groups have investigated the high frequency (HF) properties of
carbon nanotubes. The standard figures of merit for high frequency devices are cutoff
frequencies, which correspond either to frequencies where operation ceases (hard
cutoffs) or frequencies where the output signal attenuation becomes significant. At
frequencies above attenuation cutoffs, a device may continue to operate, but the
magnitude of the output signal decreases with frequency, eventually reaching a level
where it cannot be detected.
In this chapter, the published HF results from other research groups will be
presented along with the expected cutoff frequencies from their measurements. If
these cutoffs were applied to our devices, significant attenuation would occur at GHz
input frequencies, and device operation would cease near 13 GHz. Our measurements
are performed in a way that overcomes these cutoff frequencies and demonstrates
continued device operation up to input frequencies of 40 GHz. By using a spectrum
analyzer with a low noise floor (near 90 dBm), high frequency output signals are
detectable in spite of attenuation. Also, by using an external mixer in combination
with the spectrum analyzer, output frequencies can be measured up to 110 GHz,
higher than any other directly observed output frequencies from CNT devices to date.
77
5.1 Active Devices: Current Gain Cutoff Frequency
The most commonly used figure of merit for an active device is the current
gain cutoff. It is the frequency at which the current gain drops below 1 (0 dB). Since
our mixing was performed passively, our devices cannot operate at gain, so this cutoff
does not apply to our experiments. It is discussed here because it has received the
most attention in recent publications.
To measure the current gain cutoff frequency, a source-drain and gate bias is
applied to the device as it would be when integrated into a circuit. The most common
method of extracting this frequency is with a vector network analyzer (VNA). VNAs
can be used to extract the parameter H21, which is a hybrid parameter equal to output
current over input current. The input and output currents are compared at the same
frequency, so Fin = Fout. When |H21| is plotted vs. frequency, the point at which it
crosses 0 dB is the current gain cutoff frequency. When extracted this way, this
frequency is sometimes called the extrinsic current gain cutoff frequency because it
corresponds to the current gain of the whole device.
Non-ideal device properties, such as parasitic capacitances from electrodes,
can lower this cutoff from its possible intrinsic value. The intrinsic current gain cutoff
frequency can be extracted by subtracting the parasitics from H21 by using
measurements from control devices, a process called de-embedding. In the event that
the de-embedded |H21| plot does not cross 0 dB in the frequency regime of the VNA,
it can be projected to follow a slope of -20 dB per decade. This method was used to
extract an intrinsic current gain cutoff frequency of 100 GHz for a graphene FET76 in
2010, even though the VNA used was limited to 25 GHz.
78
Another way to extract the intrinsic current gain cutoff frequency is with the
equation 5.1.
GS
m
T
C
g
F
?2int
?
EQ 5.1
Using the largest measured transconductance of CNTs (20 ?S) measured in
2004 and an ideal geometrical capacitance (44 aF/?m), this intrinsic cutoff was
calculated by Burke to scale with channel length as:
? ?mL
GHz
F
gate
T ?
80
int ?
[16] EQ 5.2
Like the de-embedding VNA process, this equation ignores device parasitics.
The largest extrinsic current gain cutoff frequency for a CNT device published so far
is 15 GHz.[77] That particular device used hundreds of CNTs and had a channel length
of 300 nm. Its intrinsic current gain cutoff frequency extracted from the VNA de-
embedding procedure was 80 GHz. Some selected published current gain cutoff
frequencies are plotted vs. publication year in Figure 5.1.
79
76-82
Figure 5.1 Plot of published current gain cutoff frequencies on a log scale for CNT and
graphene devices with sub-micron channel lengths vs. the year of publication.
By using EQ 5.1 for our standard 10 um channel length device, with the
transconductance extracted from IV curves (120 ?S at Vds = 5V) and the measured
gate-source capacitance (30 fF), gives FTint = 640 MHz, demonstrating that if used
actively, our devices would not be expected to perform as well as most recently
published devices. All the devices referenced in Figure 5.1 had sub-micron channel
lengths.
80
5.2 Direct Observation of HF CNT Nonlinear Behavior in Literature
Another way to evaluate the high frequency performance of carbon nanotube
devices is to use their nonlinear properties to convert input frequencies to different
output frequencies, typically called mixing. Mixing will be described in detail in the
next chapter. Direct observation of high frequency nonlinear responses in CNT
devices has ranged from 250 MHz83 to ~100 GHz84 over the past few years. The
results of five such publications are summarized in Table 5.1.
Year Author
Maximum Input
Frequency (GHz)
Output
Frequency Type of Detection
2004 Appenzeller85 0.58 DC DC mixing current
2005 Rosenblatt86 50 DC DC mixing current
2006 Pesetski87 23 10 kHz 10 kHz mixing product
2008 Cobas88 18 DC Diode rectification
2008 Zhong84 106-160 DC DC mixing current
Table 5.1 Summary of high frequency CNT device experiments that directly observed CNT
nonlinear behavior. The input frequencies, output frequencies, and the type of experiment are
listed.
All of these publications take high frequency input signals and convert them
to low frequency or DC signals. There are two reasons for this. One is that the high
output impedance of individual CNTs causes the output signals to be of low
amplitudes. Output DC mixing currents typically require lock-in amplifiers to be
accurately measured. The other barrier to the detection of high frequency output
signals is the high cost of measurement equipment necessary to access high frequency
regimes. Until very recently, 50 GHz was near the limit of commercially available
spectrum analyzers, network analyzers, and oscilloscopes. Signal detection above 50
GHz has been prohibitively expensive. In the next section, we will describe our
method of detecting high frequency output signals.
81
The publication by Pesetski in 2006 was of an active device, where input
signals were placed on the gate. The center frequency between two input signals
separated by 10 kHz was swept up to 23 GHz. The 10 kHz mixing product, due to the
difference between the two input frequencies, was detected for the full range of the
input frequency sweep, indicating that the nonlinear behavior of the device continued
up to 23 GHz. For the rest of publications in Table 5.1, the input signals were placed
on the source and output was measured at the drain (or vice versa).
With this type of measurement setup, the device can be approximated as a
low-pass filter (LPF), shown in Figure 5.2. In a low-pass filter, the capacitor acts as a
complex voltage divider. For low frequency signals, the impedance of the capacitor is
high, and the output voltage is close to the input voltage. As frequency is increased,
the impedance of the capacitor decreases until it acts as a short to ground, and the
voltage output is close to zero. There is a cutoff frequency associated with low-pass
filters, called the corner frequency. It is defined as the frequency at which the output
power drops to one-half of its low frequency value, which corresponds to a drop of
3.01 dB. It occurs when the magnitude of the reactance of the capacitor is equal to the
resistance of the resistor, as shown in EQ 5.3 ? 5.5.
Vin Vout
Rchannel
Cgate
CNTSource Drain
Gate
Vin Vout
a)
b)
Figure 5.2 a) Diagram of CNT FET with the input signal on the source and the output signal
on the drain. b) Circuit diagram of a low-pass filter.
R
CF
Z
C
c ?? ?2
1
?
RC
FC ?2
1
?
EQ 5.3
82
2
1
2 2
??
?
?
?
?
?
?? R
R
jRR
jR
RZ
Z
V
V
CC FF
C
C
FFin
out
EQ 5.4
2
1
2
?
?
?
?
?
?
?
?
?
?
?? CC FF
in
out
FFin
out
P
P
P
P EQ 5.5
A corner frequency indicates output signal attenuation. It does not mean that a
device no longer responds to frequencies above FC, but that the power available is
reduced. Beyond the corner frequency, the output power from a LPF decreases with a
slope of 20 dB per decade. Many publications on the HF performance of CNT
devices take R to be the channel resistance, and C to be the gate capacitance, to
extract a corner frequency. If that were done with one of our typical CNT FETs (W =
100 ?m, L = 10 ?m, G = 8 ?m), the channel resistance would be 7 k?, and the gate-
CNT capacitance would be 100 fF, resulting in FC = 230 MHz.
There is one more cutoff frequency to consider in such a measurement setup.
For an output signal to reach the drain electrode, charges originating at the source
electrode must span the nanotube device channel. If the period of an input frequency
is greater the transit time of charges in the channel, the field in the nanotube will
change direction before charges have reached the drain, sending them back toward
the source. At frequencies well above this transit time cutoff frequency, no signal will
be detected at the drain. This cutoff frequency is given by EQ 5.6, where L is the
channel length.83, 89, 90
L
v
f
drift
tr ?2
?
EQ 5.6
The drift velocity is directly proportional to the carrier mobility (vdrift = E?),
so using high mobility semiconductors, like carbon nanotubes, increases this cutoff
83
frequency. An approximation for this cutoff can be made using the Fermi velocity in
the ballistic limit for metallic or highly-doped semiconducting CNTs, which is 8 x 105
m/s.[16] The input field would need to be strong enough to accelerate charges to that
velocity, and the channels would need to be short enough to allow ballistic transport,
so this Fermi velocity provides an upper limit. For a 10 ?m channel length device, the
upper limit of the transit time cutoff frequency is 13 GHz. This cutoff is one of the
reasons that short device channels are being pursued.
5.3 Tomonaga-Luttinger Liquid Theory
The one-dimensional nature of carbon nanotubes may enable an extension of
this transit time cutoff frequency. For three-dimensional semiconductors, the Fermi
gas model is used and the repulsive Coulomb interactions between charges are
ignored as typical electron densities make them improbable. However, in the one-
dimensional channels of carbon nanotubes, Coulomb interactions between charges are
unavoidable and result in wave-like excitations called plasmons, which behave more
like bosons than fermions. The theoretical analysis of transport in one-dimensional
systems is the Tomonaga-Luttinger Liquid theory.91-93
The velocity of plasmon waves is expected to be greater than the velocity of
non-interacting free charges. The ratio of these velocities is given by g, a
dimensionless parameter that indicates the strength of the electron-electron
interactions. For non-interacting charges in the Fermi model, g = 1. For repulsive
electron-electron interactions, g < 1.
g
v
v Fp ?
EQ 5.7
84
This same effect can be seen when considering a carbon nanotube as a
transmission line with a distributed kinetic inductance.94, 95 The velocity of charges
through a transmission line is known to follow EQ 5.8.
tottotCL
v
1
?
EQ 5.8
For nanotubes on a high frequency insolating substrate, the only inductance
will be the kinetic inductance (Lk). The total capacitance will be a combination of the
quantum capacitance (CQ) and the electrostatic capacitance (CES) from the nanotube
to the gate electrode. These quantities are shown in EQ 5.9, where h is the dielectric
thickness and d is the nanotube diameter.
F
k
ve
h
L
22
?
F
Q
hv
e
C
22
?
? ?dh
CES
/ln
2 ?
?
?
EQ 5.9
Using this transmission line analysis gives the same wave velocity
dependence on g as expected from Tomonaga-Luttinger Liquid Theory. Considered
the four quantum channels in a single CNT, the wave (plasmon) velocity would
follow EQ 5.10.
g
v
C
C
v
CLCLCL
v F
ES
Q
F
ESkQktotk
p ??????
4
1
411 EQ 5.10
The g-parameter for metallic CNTs (or highly doped semiconducting CNTs)
has been measured in tunneling experiments to be between 0.25 and 0.33.[96-99] This
would mean that the velocity of plasmons should be three to four times the Fermi
velocity, allowing plasmons to span longer channels than would possible with non-
interacting charge transport. The publication by Zhong et al.84 in Table 5.1 attempted
to detect this plasmon velocity by using a THz laser to excite broadband 1 ? 1.5 ps
85
signals in a transmission line attached to a CNT FET. By controlling the delay time
between two pulses, resonances can be induced in the channel corresponding to the
overlap of the reflected first pulse with the incident second pulse. These resonances
were detected in devices of several channel lengths as changes in output currents, and
from the resonance time delays, the charge carrier velocities were extracted.
Surprisingly, the velocities were found to match the Fermi velocity of non-interacting
charges and not the expected plasmon velocity. Even at input frequencies near 100
GHz, non-interacting charge motion appears to dominate transport in CNTs.
5.4 Our High Frequency Experimental Method
As seen in the recent publications investigating the nonlinear high frequency
behavior of CNT devices (Table 5.1), high frequency input signals are typically
converted to low frequency or DC output signals because of large output impedance
of individual CNTs and the high cost of high frequency measurement equipment. We
use our CNT FETs as mixers to convert high frequency input signals to even higher
frequency output signals (between 75 and 110 GHz) that are directly observed. This is
accomplished first by decreasing the device output impedance by using hundreds of
CNTs in parallel. Though this increases the output amplitude of the generated signals,
they are still much smaller than the input amplitudes. Detection is accomplished with
a spectrum analyzer which has a low noise floor, near 90 dBm, allowing the precise
measurement of very small signals. Lastly, the financial obstacle is overcome by
combining an external harmonic mixer with our spectrum analyzer to shift the
detection range up to 110 GHz. Our devices mix input frequencies up into the 75 to
86
110 GHz regime, then the external mixer down-converts them to frequencies that the
spectrum analyzer can detect. This allows for the direct observation of high frequency
output signals, something that has not been done with nanotube devices before.
Three cutoff frequencies were mentioned in the previous section. Largely
because of our relatively long device channel lengths, if applied to our devices, these
cutoff frequencies would be below the input frequencies where our devices are
operated. If our devices were used actively, they would have an expected intrinsic
current gain cutoff frequency near 640 MHz. If they were measured passively, with
an input signal on the source and an output signal taken at the drain, they would have
an expected corner frequency of 230 MHz and an expected maximum transit time
cutoff frequency of 13 GHz, beyond which no output signal should be detectable. To
overcome these relatively low frequency restrictions, our devices are used in a new
measurement setup that allows proper operation beyond these cutoff frequencies.
Our CNT FETs are used as passive mixers. Passive operation cannot achieve
gain, so the current gain cutoff does not apply to our experiments. Instead of placing
our input signals on the source, as was done for passive operation in the publications
previously mentioned, our input signals are placed on the gate. This measurement
setup changes the equivalent circuit geometry so that it no longer closely resembles a
low-pass filter. A first order circuit model of our measurement technique is shown in
Figure 5.3. The capacitive coupling from the gate to the CNT is treated as linked to
the middle of the tube. There is a still a capacitance from the input signal to ground
(CGS), but it is in parallel with capacitors from the gate to the CNT (CGCNT) and the
gate to the drain (CGD). As the input frequency is increased, the impedance of all three
87
capacitors decreases. It is only because of the extra resistances from the nanotubes
and the drain to ground that the power lost to the grounded source becomes
significant.
CNTSource Drain
Gate
Vin
Vout
a)
b)
Vin
R/2
CGCNTCGS CGD
R/2
100?
Vout
Figure 5.3 a) Diagram of CNT FET with the input signal on the gate, the output signal on the
drain, and the source grounded. b) Circuit diagram of CNT FET shown in a).
The resistances of our nanotube channels are typically on the order of k?s,
much larger than the 100 ? impedance of our measurement equipment. Therefore, the
equivalent resistance from the drain (where the output signal is taken) to ground will
be close to 100 ?. The output voltage will thus be divided by the impedance of CGD
and the 100 ? path from the drain to ground, given by EQ 5.11.
100
100
?
?
GDin
out
ZV
V
EQ 5.11
DG
GD
fC
j
Z
?2
?
?
EQ 5.12
Using CGD = 30 fF (extracted in Appendix D), at f = 1 GHz, Vout/Vin ? 0.019.
At f = 10 GHz, Vout/Vin ? 0.19. At f = 40 GHz, our maximum possible input
frequency, Vout/Vin ? 0.60. So at low frequencies, small voltages are induced at drain
electrode from the input voltage at the gate. As the input frequency increases, larger
voltages are induced at the drain. Since the source electrode is grounded, this voltage
88
at the drain is equal to the source-drain bias. This demonstrates how an AC signal on
the gate induced an AC signal at the drain. Even in a perfectly symmetric device,
where Cgs = Cgd, the extra impedance through the measurement equipment at the
drain will allow a source-drain bias to be induced. However, attenuation could be
decreased my minimizing the gate-source capacitance, so a slightly asymmetric
device may be preferable.
Note that it is power that is measured by the spectrum analyzer, so a low
output voltage does not necessarily mean the output power is low. Opposite charges
accumulate on the plates of a capacitor, so when the AC voltage on the gate is
positive, a negative voltage is induced at the drain. From the transistor equation in the
linear region presented in Chapter 2 (EQ 2.3), both a positive gate voltage and
negative source-drain voltage decrease the channel conductance. When the AC
voltage on the gate is negative, the induced voltage on the drain is positive, both of
which increase the channel conductance. This difference in conductance maintains
the nonlinear behavior of the CNT FET.
This analysis ignored the current passing through the CNTs, since the
resistance through the measurement equipment is much smaller than the resistance
through the channel. It is only the current passing through the carbon nanotubes that
will produce the mixing signals that we measure. To estimate the attenuation of
mixing output signals, the ratio of the power passing through the CNTs over the total
power needs to be found.
The equations for the three current paths leaving the gate are shown in EQ
5.13 to 5.15. The sum of these currents will be the total current.
89
FCVI GSinGS ?2?
EQ 5.13
R
FC
V
I
GD
in
GD
||100
2
1
?
?
?
EQ 5.14
?
?
?
?
?
?
??
?
100
2
||
22
1 RR
FC
V
I
GCNT
in
GCNT
?
EQ 5.15
There are two current paths responsible for mixing. One passes through the
gate-CNT capacitor, then through half of the nanotube channel into the drain. The
other mixing current path passes through gate-drain capacitor, then through the full
nanotube channel. These currents are given in EQ 5.16. Power goes as the square of
current, so the power ratio is EQ 5.17.
R
FC
V
R
FC
V
I
GD
in
GCNT
in
mix
?
?
??
?
?? 2
1
100
22
1
EQ 5.16
2
??
?
?
??
?
?
?
TOT
CNT
TOT
CNT
I
I
P
P EQ 5.17
The device values used are CGD = 30 fF, CGS = 30 fF, CGCNT = 100 fF, and R =
7 k? as extracted in Chapter 2 and Appendix D. The mixing output power using this
circuit analysis is plotted, along with the output power from a low-pass filter with a
corner frequency of 230 MHz, in Figure 5.4.
90
Figure 5.4 Output power drop as a function of frequency for a low-pass filter with a corner
frequency of 230 MHz and for the nanotube circuit model shown in Figure 5.3.
Figure 5.4 shows that the output attenuation when the input signal is placed on
the gate is less than when it is placed on the source. For the input frequencies that will
be used during our high frequency measurements (25 to 37 GHz), attenuation
between 20 and 30 dB is expected. However, since we are using a spectrum analyzer
with a noise floor near 90 dBm, these output signals will still be detectable.
The other cutoff frequency when devices were used passively came from the
transit time of charges in the channel. With an input signal on the source, charges had
to span the channel to reach the drain, where the output was measured. Since our
input signal is placed on the gate, spanning the device channel is no longer required.
Charges entering the CNT will come from the drain as well as the source. At high
frequencies, these charges will not span our long channel lengths, but the current
coming from the drain will still be detected. The drain electrode charging time will be
given by the resistance of our measurement equipment (100 ?) and the gate-drain
91
capacitance (30 fF). That corresponds to a frequency of 53 GHz, which is greater than
our maximum input frequency of 40 GHz. This means that for all of our
measurements, charges should be able to move from the drain electrode in the CNTs,
producing a measurable output, regardless of channel length.
So, by placing our input signal on our gate electrode and measuring the output
signal on the drain, the output attenuation is reduced and the transit time cutoff is
overcome. By using an external mixer in combination with a spectrum analyzer with
a low noise floor, up to 110 GHz output signals are easily detectable. In the next
chapter, the mechanisms of frequency mixing will be described, and an
approximation of the device drain current under mixing will be derived.
92
Chapter 6: Carbon Nanotube Mixing Theory
In this chapter we will introduce a way of analyzing frequency mixing by
using a simple Taylor expansion. The expected terms from single-tone and two-tone
input signals will be derived, and the nomenclature used to distinguish two-tone
mixing products is presented. The Taylor expansion is then modified for our passive
mixing experiments, and its validity is experimentally tested by comparing the DC
current from a 5 GHz input signal with the expected output. These models will then
be used in the next chapter to understand the results from high frequency mixing
experiments with carbon nanotube field-effect transistors.
6.1.1 Taylor Expansion of Input to Output Nonlinearity
Before discussing the high frequency (HF) mixing experiments done on the
CNT FET devices, a brief review of frequency conversion is given. Frequency
conversion is usually referred to as mixing because two or more input signals are
effectively multiplied to generate new output frequencies. The multiplication comes
from the non-linearity of the device under test.
Conventional analysis of mixing is done by using a Taylor expansion of a
function that is non-linear in one input variable. The classic case is the description of
a transistor with the output drain current Ids expanded in terms of a small AC voltage
?V about the input bias Vds
? ?
? ? ? ?
...
!3!2
)()(
3
3
32
2
2
??
?
??
?
??
?
???
V
V
IV
V
I
V
V
I
VIVVI ddddd
???? EQ 6.1
93
If the relationship between drain current and source-drain voltage in the CNT
FET devices were linear (Ohmic), all the higher order derivatives of the drain current
would be zero. However, as shown in Chapter 2, non-linearities in the output curve
(Id vs. Vds) exist due to Schottky contacts and the gate-induced field, so higher order
terms are necessary for the description of our devices.
For the expansion to be valid, ?V must be small relative to the features of the
function. The magnitude of the oscillating voltage reaching the device depends on the
experimental setup, the frequency used, and the input power. We estimate that the
magnitude of the oscillating voltage reaching the device is significantly less than 1 V,
the typical magnitude of the DC bias. The largest signal output from our signal
generators in the high frequency mixing (75-110 GHz) experiments is +20 dBm (100
mW). Assuming we have an ideal transmission line, the power at the device input
will be the same as the power at the port of the signal generator.
? ? ? ??? ????? IVIVDUTPportP rmsrms Re21Re)()(
EQ 6.2
The maximum power is given by :
0
2
maxmax
2
1
2
1
)()(
Z
V
VIDUTPportP ???
EQ 6.3
where Z0 is the characteristic impedance of the transmission line (i.e. 50 ?).
Thus, Vmax is given by:
? ? VmWZDUTPV 16.35010022 0maxmax ??????
EQ 6.4
However, measurements have shown that the transmission line is lossy in the
HF measurements described here. We typically observe 30 dB loss or more in the
94
signal, so the power reaching the device is at most -10 dBm (100 ?W), and the
magnitude of the AC signal at the device is on the order of:
mVWVdevice 100501002 ????? ?
EQ 6.5
Since the amplitudes of the AC signal are 100 mV or less, we conclude that
the Taylor expansion can be safely applied in the analyses of our devices.
6.1.2 Harmonic Generation
Now consider a single-tone input signal of frequency ? with amplitude A:
? ?tAV ?? cos?
. EQ 6.6
Obviously, the higher order terms in EQ 6.1 produce increasingly higher
powers of sinusoidal input, which in turn produces outputs at higher frequencies. For
example, using the trig identity for cosine squared:
? ? ? ?tt ?? 2cos1
2
1
cos
2
??
EQ 6.7
the second order term becomes:
? ? t
AA
V ?? 2cos
22
22
2
??
,
and thus the output frequency is doubled. EQ 6.8
Similarly, the third and fourth-order products are:
? ? t
A
t
A
V ??? 3cos
4
cos
4
3 333
??
EQ 6.9
? ? t
A
t
AA
V ??? 4cos
8
2cos
28
3 4444
???
. EQ 6.10
95
So even with only one input frequency, a non-linear device will produce
harmonics of the input frequency. Also, note the presence of frequency-independent
terms in the even-numbered products. This means that when an AC voltage is applied
to a non-linear device, a DC current is generated. This fact will be examined more
thoroughly later in this chapter.
6.1.3 Two-Tone Mixing Products and Corresponding Nomenclature
If the input signal consists of two tones ?1 and ?2, then the cross product
terms in the expansion produce new frequencies at the sum and difference of ?1 and
?2. The first three products for two-tone mixing are shown below in EQ 6.11.
?
?
? ?
?
tV
tV
V tone
22
11
2
cos
cos
?
?
?
? ?
? ?
? ?
? ?
? ?
?
?
?
?
?
?
?
??
?
?
?
?
?
?
??
??
??
??
??
?
t
V
tVV
t
V
tVV
VV
V tone
21
2
1
2121
21
2
2
2121
2
2
2
1
2
2
02cos
2
11cos
20cos
2
11cos
22
??
??
??
??
?
? ?
? ?
? ?
? ?
? ?
? ?
? ?
? ?
? ?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
???
?
?
?
?
?
???
?
?
?
?
?
???
?
?
?
?
?
???
?
?
?
?
?
???
?
?
?
?
?
???
?
?
?
?
?
?
???
?
?
?
?
?
?
????
?
?
?
?
?
?
tV
tVV
tVV
tV
tVV
tVVV
tVVV
tVV
V tone
21
3
1
212
2
1
21
2
21
21
3
2
212
2
1
21
3
1
2
21
21
3
22
2
1
21
2
21
3
2
03cos
4
1
12cos
4
3
21cos
4
3
30cos
4
1
12cos
4
3
01cos
4
3
3
2
10cos
4
3
3
2
21cos
4
3
??
??
??
??
??
??
??
??
?
EQ 6.11
96
To identify peaks in the output spectrum, we introduce the standard naming
convention used for identifying the various frequencies that arise. By convention the
ordered pair (n, m), which represents n?1 + m?2 where n and m are integers, is used
to identify the particular output frequency of interest. Note that these integers are
NOT the same as those in CNT chiral vectors. For the rest of this dissertation, the
ordered pair (n, m) will always be used to refer to a particular mixing output
frequency.
Furthermore, the sum |n| + |m| is used to identify the mixing order of a
particular measurement, which also corresponds to the lowest order expansion that
would produce an output at that frequency. However, higher order products can also
fall into lower order frequencies. For example a 3rd order expansion will produce a
signal at the 3rd harmonic of F1, (3, 0) = 3F1 + 0F2. A 5
th order expansion will also
produce a signal at the same frequency from the 4th harmonic mixing with the
fundamental, 4F1 ? F1 = 3F1, which is still (3, 0). Since the Taylor expansion carries
the coefficient of 1/(order factorial), each output frequency is typically dominated by
the contribution from the lowest order term.
As an illustration, consider Figure 6.1, where the output from the spectrum
analyzer centered at 75.5 GHz is shown. The input frequencies are 25.1667 GHz ? 50
kHz. Since the acquisition window is centered near 3F1,2, higher ordered odd terms up
to the 11th order can be seen before the spectral output drops into the noise floor at
-90 dB. The labeling procedure illustrated below is useful for identifying particular
output frequencies of interest that are present in the acquisition window of the
97
spectrum analyzer. We note here that the mixing products shown below will be
evaluated in detail in Chapter 6.
Figure 6.1 Output power spectrum read by an Agilent MXG analog signal generators when
using a CNT FET (W = 100 ?m, L = 10 ?m, G = 8 ?m) as a passive mixer. The input signals
on the gate were -10 dBm at frequencies = 25.1667 GHz ? 50 kHz. The peaks are labeled
with their corresponding (n, m) notation.
6.2 Modified Taylor Expansion for Passive CNT Mixing
Now we modify the Taylor expansion to describe how the CNT FET devices
were measured. Our AC input signal is placed on the gate electrode. Because of the
large number of CNTs in the channel, a signal on the gate will also induce a signal in
the channel (as shown in the previous chapter), resulting in two AC voltages in our
expansion.
dsdsgsgs VVVVVV ??? ???? ,
EQ 6.12
98
? ?
? ?
? ?
?
?
?
?
?
?
?
?
?
?
?
?
?
?
??
?
?
??
?
?
??
?
?
??
?
?
????
...
!2
)(
...
!2
)(
),(,
2
2
2
2
2
2
ds
ds
d
ds
ds
d
gs
gs
d
gs
gs
d
dsgsddsdsgsgsd
V
V
I
V
V
I
V
V
I
V
V
I
VVIVVVVI
??
??
?? EQ 6.13
Substituting in the definitions of transconductance and dynamic conductance
(EQ 2.3 And EQ 2.4) produces EQ 6.14.
? ?
? ?
? ?
?
?
?
?
?
?
?
?
?
?
?
?
??
?
?
??
??
?
?
??
????
...
!2
)(
...
!2
)(
),(,
2
2
ds
ds
D
dsD
gs
gs
m
gsm
dsgsddsdsgsgsd
V
V
G
VG
V
V
g
Vg
VVIVVVVI
??
??
?? EQ 6.14
Our high frequency mixing experiments are done in the absence of a constant
source-drain bias, which means the mixing is being done passively. More specifically,
this means that our devices do not have a source-drain bias because the only power
reaching the device is applied to the gate electrode. The primary reason for this
experimental configuration is that we were unable to find bias-tees rated in the W-
frequency regime, and our electrodes were not designed to be large enough to allow
an extra DC probe to add that bias.
Passive mixing cannot produce gain, but apart from a reduced output signal, it
can perform the same functions as active mixing, and if necessary, the output signals
can be amplified post-mixing. A significant advantage of passive mixing is that it
eliminates the power losses associated with the source-drain bias-induced current
through metallic CNTs in the as-grown networks, as mentioned in Chapter 4. Recall
that we demonstrated that the metallic CNTs can be successfully burned out, which
reduces the total number of CNTs and effectively increases the channel impedance.
Here, the mixing signals being measured are related to the device output impedance,
99
so samples with more CNTs will produce larger output signals, making them easier to
detect. Thus, passive mixing allows us to use samples with the largest number of
CNTs while avoiding the unwanted constant drain current from the presence of
metallic CNTs.
Without a DC source-drain bias, the first term in the expansion, the DC drain
current, will be zero. The transconductance and all its derivatives will also be zero for
Vds = 0, as shown in Chapter 2. Our expansion then reduces to EQ 6.15.
? ? ? ? ...
!2
)(,
2
??
?
?
???? ds
ds
D
dsDdsgsgsd
V
V
G
VGVVVI
???? EQ 6.15
This shows that even though our input signal is placed on the gate, it is the
induced signal across the channel that generates all the mixing products. This probing
arrangement allows us to ground the source with one probe, apply a signal on the gate
with another, and measure the output on the drain. Our last step is to take the
derivatives of the dynamic conductance with respect to the gate voltage instead of the
source drain voltage, a substitution derived in Chapter 2 (EQ 2.5). This form is useful
in the DC mixing current measurements in the next section. The final form of our
expansion is EQ 6.16.
? ?
? ?
...
!2
)(
2
??
?
?
??? ds
gs
D
dsDdsd
V
V
G
VGVI
???
EQ 6.16
We briefly mention a subtle point about EQ 6.16. It only applies to the bulk
of the nanotube, assuming that the contact resistances are negligible. However, when
the contacts resistances are comparable with the resistance of the CNT, an adjustment
to include the contact effects will have to be made to EQ 6.16. When unbiased, our
CNT FETs are in an off-state, and EQ 6.16 is expected to apply.
100
6.3 Measurement of CNT DC Drain Current
As mentioned in the single-tone analysis of section 6.1.2, the even power
products in the expansion contain a DC term, so an AC input voltage should produce
a DC current. Previous groups have reported measuring this DC current, referred to as
the mixing current, in the absence of an external source-drain bias in CNT FETs and
diodes.85, 86, 88 The effect was described by Appenzeller as ?the increase in current for
one half cycle of an AC oscillation is different from the decrease in current for the
other half of the cycle,?85 so over time, more charges are pushed in one direction than
in the other, resulting in a net DC current. Considering only the second-order mixing
product, the DC mixing current due to the CNT should be:
? ?
gs
mixCNT
V
GA
VI
?
?
??
4
2
?
EQ 6.17
To test this relation, the DC mixing current was measured as a function of
gate voltage for our CNT FETs. At the same time, measurements of conductance
were taken so that both sides of EQ 6.17 could be compared. The following
experimental setup was used: Gate
Bias T
S/D
Bias T
Spectrum
Analyzer
Source
Meter 2
Source
Meter 1
Signal
Generator
100 MHz
LPF
Figure 6.2 Diagram of the experimental setup used to measure the DC mixing current
produced from a single frequency input. The black, red, and blue lines correspond to SMA,
SMC, and BNC cables respectively.
101
The components used were:
Agilent MXG Analog Signal Generators (N5183A): 100 kHz ? 40 GHz
Anritsu K250 Bias Tees: DC ? 40 GHz
Keithley 2400 Source-Meters: DC
Picoprobe 40A-GSG-150?m-LP probes: DC ? 40 GHz
Crystek 100 MHz Low Pass Filter (CLPFL ? 0100)
Agilent ESA-E Series Spectrum Analyzer (E4407B): 9kHz ? 26.5 GHz
SMA coaxial cables: DC ? 18 GHz (shown in black)
SMC coaxial cables: DC ? 10 GHz (shown in red)
BNC coaxial cables: DC ? 3 GHz (shown in blue)
The bias tees allow the setting of a DC bias but also the measurement of the
DC current. Typically, measurements of DC mixing current require the use of a lock-
in amplifier. However, measuring this current was possible with a source-meter
because of the large number of CNTs in the device channel.
To ensure a large enough signal to accurately measure, a large input power of
18 dBm was used. This corresponds to a maximum voltage of 3.55 V. Not all of the
voltage drop occurs across the CNTs, so this estimate provides an upper limit for the
input voltage, A. We note that this input amplitude is 25 times larger than the
maximum amplitude used in the high frequency mixing experiments.
The microwave signal generator sourced a constant 5 GHz 18 dBm AC signal.
The measurement procedure was to set a gate voltage with the first source-meter, then
the source-drain bias was swept with the second source-meter between ? 50 mV by
10 mV increments. The resulting Id vs. Vds data for each gate voltage was fit to a line,
where the extracted slope was the conductance, and the intercept was the DC mixing
current. The gate voltage was swept between ?10 V in steps of 100 mV. The
measured conductance is plotted in Figure 6.3, clearly showing the non-linear
features of the device.
102
Figure 6.3 Conductance vs. gate voltage for a top-gated CNT FET on quartz (W = 100 ?m, L
= 10 ?m, and G = 8 ?m). Measurements were made with an 18 dBm 5 GHz signal input on
the gate. Conductances were computed at each gate voltage by sweeping the source-drain
bias between ?50 mV by 10 mV steps and equal to the extracted slope of each sweep.
The derivative of this conductance (with respect to gate voltage) was taken to
compare it with the measured DC mixing current. To calculate the derivative, a line
was fitted over five points (400 mV span) and the slope extracted. Using A = 3.55 V,
EQ 6.17 becomes:
gs
mixCNT
V
G
I
?
?
?? 15.3
. Both sides of that equation are plotted in
Figure 6.4.
103
Figure 6.4 Plot of the measured drain current and the derivative of the measured conductance
with respect to gate voltage scaled by 3.15 while a 5 GHz signal is input on the gate. The 3.15
comes from A = 3.55 V in EQ 6.17.
There appears to a qualitative agreement between the two quantities, with the
least amount of agreement for large negative gate voltages. This is expected, since the
expansion used should only apply when the voltage drop is primarily across the
nanotube, which is when the semiconducting CNTs are gated off. When the
semiconducting CNTs are gated on, the gate-independent mixing due to the Schottky
barrier will compete with the mixing due to CNT non-linearities. In addition, when
the semiconducting CNTs are gated on, the resistance of the CNT network is now
roughly the same as the contact resistance between the CNT and metal electrodes.
Previously, the contact resistance was estimated to be about 4 k?. Correcting for
these voltage drops produces EQ 6.18 and Figure 6.5.
104
tot
CNT
mixCNTmix
R
R
II ?
EQ 6.18
Figure 6.5 Plot of the measured drain current and the derivative of the measured conductance
with respect to gate voltage, times the nanotube resistance, divided by the total channel
resistance, scaled by 3.15 while a 5 GHz signal is input on the gate. The 3.15 comes from A =
3.55 V and EQ 6.17.
The correspondence between the measured mixing current and the calculated
change in conductance is much better when considering the effects of contacts. The
observed deviations could be due to several causes. First, we have only considered
the second order DC contribution, while all higher order even terms will add to the
DC current and have their own gate voltage dependences. Moreover, the conductance
that we used for our derivative is the DC conductance. Nanotube conductance may be
frequency-dependent, though this measurement indicates that conductance at 5 GHz
is close to the DC conductance. Lastly, the amplitude used for our input signal was an
105
upper limit which assumed that all the power fell across the CNTs. Considering all of
these effects, the agreement between the curves is reasonable and suggests that the
Taylor expansion we use to represent the drain current is appropriate.
We note that no source-drain biases are used in our high frequency mixing
measurements (to be discussed in Chapter 6), and our input signals are centered at Vgs
= 0. This is a regime where the semiconducting CNTs are gated off and the
agreement between the DC mixing current and the change in conductance is
reasonably high.
6.4 Summary
We have presented an approximation of the drain current of a CNT FET as a
Taylor expansion of a small input voltage. A single tone input will generate
harmonics of the input frequency, while a two-tone input will produce many mixing
products. The expected amplitudes of up to third-order products are listed, and the
nomenclature used to specific mixing products is presented. The Taylor expansion
was then modified to represent passive mixing (Vds = 0) and expressed in terms of the
gate voltage dependence of conductance. A single tone mixing experiment was
performed to test the validity of this expansion. The DC mixing current was measured
and compared with the expected output derived from the derivative of conductance
with respect to gate voltage. The two were shown to match well in the regions where
the approximation was valid. This verifies that our Taylor expansion is a valid
representation of our devices under small AC inputs.
106
Chapter 7: CNT Mixing Data
In this chapter, the results from two-tone mixing experiments are presented
with output frequencies in two different regimes. This division is due to the
availability of experimental components needed to perform the measurements.
Frequency regimes will be referred to by the type of coaxial connectors used in each
regime. The three regimes involved in our experiments are SMA (up to 26.5 GHz), K
(up to 40 GHz), and W (up to 110 GHz). We first discuss experiments with output
frequencies from 250 MHz to 26.5 GHz (SMA regime). In this regime, CNT FETs
are compared with control devices to determine if any mixing signals come from the
experimental setup itself. Then we present the measurements performed with output
frequencies from 75 to 110 GHz (W regime). In this regime, the mixing output
amplitudes are compared with their expected values from our Taylor expansion of
drain current. Lastly, measurement sweeps of the output frequency in the W regime
are made with one input signal fixed and the other swept to compare different types
of devices and understand their frequency dependent outputs.
7.1 250 MHz to 25 GHz (SMA Regime) Measurements
Mixing experiments in the 100 kHz to 26.5 GHz are discussed in this section.
The measurements are limited to 26.5 GHz by the spectrum analyzer. The
experimental setup used in the SMA regime is shown in Figure 7.1.
107
Power
Divider
Spectrum
Analyzer
Signal
Generator (F1)
Signal
Generator (F2)
DUT
10 MHz reference signal
Figure 7.1 Diagram of the experimental setup used to perform mixing experiments in the
SMA regime. The black, red, and blue lines correspond to BNC, SMA, and K coaxial cables
respectively. A 10 MHz reference signal is generated by the spectrum analyzer and input to
both signal generators. The outer wings of the probes are grounded through the coaxial
cables.
The components used and the frequencies over which they are rated are:
Agilent MXG Analog Signal Generators (N5183A): 100 kHz ? 40 GHz
Anritsu Power Divider (K240C): DC ? 40 GHz
Picoprobe 40A-GSG-150?m-LP probes: DC ? 40 GHz
Agilent ESA-E Series Spectrum Analyzer (E4407B): 9kHz ? 26.5 GHz
BCN coaxial cables: DC ? 3 GHz (shown in black)
SMA (3.5 mm) coaxial cables: DC ? 26.5 GHz (shown in red)
K (2.92 mm) coaxial cables: DC ? 46 GHz (shown in blue)
The first signal generator is connected to the power divider through a K
coaxial cable. The second signal generator is connected to the power divider through
a male-male (M-M) K connector. The attenuation through the coaxial cable is about 1
dB greater than through the M-M connector, so to have roughly equal input powers
from both sources, the output amplitude of the first signal generator is 1 dB higher
than the second signal generator. The power divider is not a directional coupler, but
merely divides the power at each terminal to the other two. All the CNT devices
measured contain outer electrodes separated by 300 ?m that are grounded by the
outer electrodes of the GSG probes. A 10 MHz reference signal from the spectrum
analyzer is used to synchronize both signal generators. The signal generators and
spectrum analyzer were controlled with Labview with GPIB communication. Unless
108
otherwise specified, all the data was acquired with the spectrum analyzer window set
to the following: a center frequency of 1.2 MHz, a resolution bandwidth of 3 kHz,
and the number of points per sweep is 600. The two tones from the signal generators
were separated by 100 kHz .
kHzFF in 501 ??
kHzFF in 502 ??
EQ 7.1
The center frequency between the two tones (Fin) was swept from 0.25 to 25
GHz in steps of 250 MHz. The output spectra were captured at acquisition windows
centered at the input frequency
Measurements were performed on four different types of devices. Later, we
refer to them as devices under test (DUTs). Three of the four devices are controls
used to assess the parasitic attenuation in the measurement setup, shown in Figure
7.2. The first two controls, through electrodes and a parallel plate capacitor, allow for
direct measurement of the system parasitics. For the parallel plate capacitor, the
center electrodes are vertically separated by an ALD Al2O3 dielectric layer (65 nm
thick). This structure has a DC capacitance near 44 pF. The third control consisted of
the same electrode geometry as the CNT FETs but with all the CNTs etched out of
the channel (essentially, a channel width of zero). This control (DC capacitance ~ 30
fF) allows for a direct comparison with the CNT device to clearly show the
contribution of the tubes to the output signals. The CNT FET structure was described
previously in Chapter 2. The CNT FET device measured had a channel width (W) of
100 ?m, a channel length (L) of 10 ?m, and a gate length (G) of 8 ?m with a DC
capacitance of ~100 fF. The capacitances are extracted in Appendix D.
109
Figure 7.2 Optical images and side profile views of the different control DUTs used in the
SMA two-tone mixing experiments. a) Through electrodes, b) parallel plate capacitors, c)
FET geometry. The spacing between the center of the horizontal electrodes in each device is
150 ?m.
The spectrum analyzer is used to measure the mixing output power levels.
The output power attenuation is defined as the signal generator input power (17 dBm
in these measurements) subtracted from the measured output power. Since the two-
tone inputs are similar in magnitude, the calculated output power attenuation is the
average of both peaks.
The output power attenuation for the three controls and the CNT device is
plotted vs. input frequency in Figure 7.3. Obviously, the through electrodes and
parallel plate capacitor devices attenuate the input signals nearly identically. The
output from these two controls demonstrates the parasitic attenuation in the
experimental setup due primarily to attenuation in the probes, but also to the cables
and connectors. This attenuation runs from 8 dB at low frequencies to 30 dB at 25
GHz. We note that this output amplitude is not necessarily the power that drops
across the channel of the CNT FET device because for the FET device the signal
passes directly from the gate to the drain, and thus the signal across the FET channel
is probably smaller. The attenuation for both of the FET structures (with and without
a) b)
c)
110
CNTs) is qualitatively similar. At low frequencies the attenuation is larger for the
control device (-55 dB) than the CNT FET device. At 20 GHz, the attenuation is
nearly identically for the two devices. By 25 GHz, the attenuation of all four device
structures is similar (~ -30 dB).
Figure 7.3 Plot of output power attenuation for four different device types vs. input frequency
used in two-tone mixing. The input frequency was swept from 0.25 to 25 GHz by steps of
250 MHz, and the outputs were taken at the input frequency. The parallel plate capacitor used
had a DC capacitance near 44 pF. The CNT FET measured had W = 100 ?m, L = 10 ?m, and
G = 8 ?m.
Output mixing spectra for the CNT FET and the capacitor control are plotted
in Figure 7.4 at input frequencies of 1 GHz and 25 GHz, near the minimum and
maximum of our spectrum analyzer. The output window is centered about the input
frequency. The spectra of the two devices are separated in each plot to more easily
see their features. At 1 GHz, both the CNT FET and the capacitor?s spectra contain
111
many mixing products. At 25 GHz, the two spectra now have fewer mixing products,
but are similar. Since the capacitor itself should not contain any non-linearities, we
attribute the observed mixing signals to the experimental setup. Furthermore, the
devices are not properly impedance-matched so significant power is reflected back
from the device to the input of the signal generators, where mixing can occur. Mixing
can also take place at the input of the spectrum analyzer, but the powers reaching the
input of the spectrum analyzers are significantly lower than the rated input limit of
20 dBm. Thus, undesired mixing at the spectrum analyzer should be a negligible
contribution to the parasitic mixing observed in the capacitor control device.
Figure 7.4 Output spectra captured from a spectrum analyzer from two-tone mixing
experiments performs on a parallel plate capacitor and a CNT FET. The vertical axis is input
power subtracted from the output power. The input frequencies were a) 1 GHz and b) 25
GHz. The two spectra in each plot are separated by 1 MHz to more clearly see their features.
The similarities between the mixing products of the CNT FET and the control
around the input frequencies are problematic for quantitative analysis of the data.
There is no clear way to determine if the mixing products are coming from the CNTs
or the equipment.
112
To address this problem, output spectra were taken at three times the input
frequency, centered about the third-order harmonics. This arrangement minimizes the
contributions from the input signals passing directly through the devices. The output
spectra for the CNT FET and the three controls at three different input frequencies are
plotted in Figure 7.5. As the input frequency increases, the mixing products from the
controls decrease, falling near or below the noise floor by input frequencies of 3 GHz.
The mixing products from the CNT FET remain strong for all three input frequencies.
The two highest amplitude mixing products remain in the through-electrode control
spectra even at an output frequency of 24.75 GHz. However, their amplitudes are
20 dB (100 times) below the CNT FET. Since this mixing is performed passively, the
CNTs cannot be amplifying products from the system, so any CNT FET mixing
products of higher output amplitudes than the controls must be due to the CNTs and
not the equipment. Comparing the CNT FET to the control FET device without CNTs
clearly shows that above output frequencies of 9 GHz the CNTs are responsible for
the mixing products.
113
Figure 7.5 Output spectra captured at three times the input frequency from a spectrum
analyzer from two-tone mixing experiments performs on a parallel plate capacitor and a CNT
FET. The vertical axis is input power subtracted from the output power. The input
frequencies were a) 1 GHz, b) 3 GHz, and c) 8.25 GHz. The two spectra in each plot are
separated by 100 kHz to more clearly see their features.
114
Figure 7.5 represents only a small portion of the data taken. To compare data
over the complete frequency range examined, the peak amplitudes at specific mixing
product frequencies are plotted in Figure 7.6 for the four types of devices. The spectra
used were captured at three times the input frequency. Only the third and fifth-order
mixing products are plotted, as the higher order products are near the noise floor for
the CNT FET. Each point in the graphs represents an average of the two
corresponding peaks, labeled in the (n, m) notation. The amplitudes of each peak
were computed as the maximum value in a 50 kHz span centered at each peak. The
noise floor was computed as the minimum value in the 50 kHz span between the two
center peaks. Spectra were captured with a resolution bandwidth of 3 kHz.
115
Figure 7.6 Mixing product amplitudes vs. output frequency for a) a CNT FET, b) an FET
without CNTs in the channel, c) a parallel plate capacitor, and d) through electrodes used as
devices in two-tone mixing. The vertical axis is input power subtracted from the output
power. The output spectra were taken at three times the input frequencies.
The CNT FET produces output mixing peak amplitudes spaced similarly over
the measured frequency regime, with a gradual decrease in magnitude of around 10
dB from 7 GHz to 25 GHz. The mixing products from the FET electrode geometry
without CNTs fall completely below the noise floor by an output frequency of 3 GHz.
Both the capacitor and through electrodes show strong mixing products for output
frequencies below 3 GHz, which fall quickly near the noise floor at higher
frequencies. This indicates that either the mixing due to the equipment cannot
116
respond at these higher frequencies, or the attenuation of the input signals becomes
large enough that they no longer meet the power threshold where equipment mixing
can occur. Note also that at lower frequencies the attenuation of the input signal
through the CNT FET is much higher (Figure 7.3) than the capacitor and through
electrodes, so the amplitudes resulting in the mixing observed in these controls are
not present in the CNT FET device. These plots show that the mixing output
produced by the CNT FET device above 3 GHz is due to the presence of nanotubes
and not other components. This is well below the frequencies that will be used in the
W regime.
7.2 75 to 110 GHz (W Regime) Measurements with External Mixer
Carbon nanotubes are being considered for high frequency applications
because other electronic materials are either very expensive or cannot operate in high
frequency regimes. Because of this, equipment using conventional electronic
materials to source and analyze high frequency signals substantially increase in cost
with increasing maximum frequency. The signal generators we are using, which are
limited to 40 GHz, cost around 25,000 USD and the spectrum analyzer we are using,
which is limited to 26.5 GHz, costs around 44,000 USD. Moving up a frequency
regime can increase the cost to hundreds of thousands of USD. For this reason, direct
observation of AC signals above the K regime is prohibitively expensive.
To access higher frequency regimes, Agilent produces external harmonic
mixers that are compatible with their spectrum analyzers. These function by receiving
117
high frequency input signals that are then mixed down with a local oscillator (LO)
produced by the spectrum analyzer to frequencies in the SMA regime that are passed
on to the spectrum analyzer where they can be accurately measured. These inputs are
then corrected for attenuation during the external down mixing and plotted at the
original mixing input frequency. By having a stand-alone external mixer responsible
for a limited frequency regime, the range of a spectrum analyzer can be economically
extended. Each external mixer costs about $4,000. We chose the highest frequency
external mixer available from Agilent, which operates from 75 ? 110 GHz, which is
the maximum of the W-band (1 mm). This corresponds to the highest output
frequencies directly observed with CNT FETs.
7.2.1 Experimental Setup
The experimental setup we use in the W regime is shown in Figure 7.7.
Power
Divider
Spectrum
Analyzer
Signal
Generator (F1)
Signal
Generator (F2)
DUT
10 MHz reference signal
External
Mixer
Figure 7.7 Diagram of the experimental setup used to perform mixing experiments in the W
regime. The black, red, blue, and green lines correspond to BNC, SMA, K, and W coaxial
cables respectively. A 10 MHz reference signal is generated by the spectrum analyzer and
input to both signal generators. The outer wings of the probes are grounded through the
coaxial cables.
The components used and the frequencies over which they are rated are:
Agilent MXG Analog Signal Generators (N5183A): 100 kHz ? 40 GHz
Anritsu Power Divider (K240C): DC ? 40 GHz
Picoprobe 40A-GSG-150?m-LP probe: DC ? 40 GHz (shown in blue)
Cascade Mictrotech i110-A-GSG-150?m probe: DC ? 110 GHz (shown in green)
118
Agilent 11970 Series Harmonic Mixer (11970W): 75-110 GHz
Agilent ESA-E Series Spectrum Analyzer (E4407B): 9kHz ? 26.5 GHz
BCN coaxial cables: DC ? 3 GHz (shown in black)
SMA (3.5 mm) coaxial cables: DC ? 26.5 GHz (shown in red)
K (2.92 mm) coaxial cables: DC ? 46 GHz (shown in blue)
W (1 mm) coaxial cables: DC ? 110 GHz (shown in green)
The input of the experimental setup is the same as in the SMA regime in the
previous section. No bias is applied to the gate or the source-drain, so this is a passive
mixing experimental setup. The output probe has been replaced with a W probe,
which connects to a W coaxial cable and then a waveguide adaptor on the external
mixer. The external mixer converts the W regime frequencies down to the SMA
regime. The two SMA cables connecting the external mixer and the spectrum
analyzer correspond to the local oscillator (LO) signal from the spectrum analyzer
(4.15 to 6.09 GHz) and the intermediate frequency (IF) signal from the mixer
carrying the down-converted signal to be measured. The nominal uncertainty of the
spectrum analyzer with this external mixer is ?5.6 dB for the input powers we use.
7.2.2 Two-Tone Mixing in W Regime with 100 kHz Separation
By using a W-band external mixer, we are able to measure output signals in
the 75 ? 110 GHz regime. However our signal generators are limited to a maximum
output frequency of 40 GHz, leaving a gap from 40 to 75 GHz in our acquisition
window. Therefore, we are unable to observe the fundamentals of our input
frequencies with this experimental setup. It is possible to observe mixing products at
two times our input frequencies, but that would limit the maximum output frequency
to 80 GHz (2 ? 40 GHz = 80 GHz). To utilize the entire frequency range of our
119
external mixer with our equipment and a two-tone input separated by 100 kHz, our
acquisition window will be three times our input signals. This is why the mixing
products centered at three times the input frequency were measured in the SMA
regime in the previous section. The frequency input and output ranges are: 25 ? Fin ?
36.67 GHz and 75 ? Fout ? 110 GHz.
First, we discuss measurements where the input frequencies span the complete
output range of the external mixer in steps of 500 MHz. Then we observe the mixing
product peak amplitudes as a function of input power at a fixed frequency to see if
they correspond to the expected dependence from our Taylor expansion. Lastly, we
use the CNT mixers in a conventional setup, with a fixed local oscillator frequency
and a variable input frequency. The output range of the external mixer is varied with
the output mixing products. That output is used to characterize CNT FETs with
different gate lengths and channel lengths.
7.2.2.1 Varying the Output Frequency from 75 to 110 GHz
Mixing measurements were performed on a CNT FET (W = 100 ?m, L = 10
?m, G = 8 ?m) as described above. Output spectra near the minimum and maximum
frequencies of the external mixer (Fout = 75.5 GHz and 109.5 GHz) are shown in
Figure 7.8. They clearly demonstrate mixing in both windows, indicating that the
carbon nanotubes continue to act as mixers up to 109.5 GHz. The output powers of
the signal generators were fixed at 16 and 15 dBm to provide symmetric input powers
after the 1 dB loss through the extra cable connected to the first signal generator.
120
There is a reduction in output peak amplitude from 75.5 to 109.5 GHz. This is not a
trend across this frequency regime, as will be seen next, but an unusually large
difference between these two particular input frequencies. These two spectra were
chosen because they represent the maximum and minimum output frequency
measured in this sweep.
Figure 7.8 Output spectra at a) 75.5 GHz and b) 109.5 GHz of a CNT FET used to mix two
input frequencies separated by 100 kHz. The output frequencies are three times the input
frequencies. The vertical axis is the input power (15 dBm) subtracted from the output power.
This measurement was repeated by sweeping through the input frequencies
that span the output range of the external mixer by steps of 500 MHz. This
corresponds to F1,2 = 25 to 36.67 GHz ? 50 kHz. The signal generators and spectrum
analyzer were controlled with Labview over GPIB. The spectrum analyzer window
was set to 1.2 MHz, the resolution bandwidth was 3 kHz, and the number of points in
each sweep was 600. The amplitudes of each peak were computed as the maximum
value in a 50 kHz span centered at each peak. The noise floor was computed as the
minimum value in the 50 kHz span between the two center peaks.
121
Figure 7.9 Mixing product amplitudes vs. output frequency for a CNT FET (W = 100 ?m, L =
10 ?m, G = 8 ?m) used as devices in two-tone mixing. The vertical axis is input power
subtracted from the output power. The output frequency is three times the input frequency.
In Figure 7.9 we show three output peaks corresponding to third and fifth-
order mixing. There is no appreciable decrease in output power over the frequency
range measured. Point-to-point variations can change by as much as 7 dB. While this
falls within the uncertainty of external mixer and spectrum analyzer, it could also be
due to standing waves forming at specific frequencies in components of the
experimental setup, resulting in higher reflection that reduces the output signal.
To compare these output mixing amplitudes with the attenuation expected
from the circuit model in Chapter 5, data from both the SMA and W regimes are
combined. Since all six of the measured peaks show similar trends, only the highest
amplitude peaks, (1,2) and (2,1), are compared. This data, along with the expected
attenuation from the circuit model and a low-pass filter with a corner frequency of
122
230 MHz is shown in Figure 7.10. Both models are scaled to the same low frequency
output power of -40 dBm.
Figure 7.10 Mixing produce amplitudes plotted against two input frequency regimes for a
CNT FET FET (W = 100 ?m, L = 10 ?m, G = 8 ?m) used as a device in two-tone mixing.
The FET circuit model frequency-dependent output power and a low-pass filter with a corner
frequency of 230 MHz are also plotted, matched to the low frequency output.
As expected, the low-pass filter model undershoots the mixing product output
power from our measurements. The circuit analysis from Chapter 5 provides a much
better approximation. This result is evidence that by placing our input signal on the
gate instead of the drain, the signal attenuation is reduced. The corner frequency from
the resistance of the channel and gate capacitance does provide a rough estimate of
where significant attenuation will begin, but predicting output amplitudes requires a
more careful circuit analysis. Any loss in output power is undesirable, but as was
shown with our experiments, input frequencies beyond FC do not prevent CNT
123
response. We already lose 30 dB at Fin = 25 GHz from our experimental setup, using
components rated up to that frequency. Preventing high power losses at frequencies
an order of magnitude above that will be nontrivial.
Two quick comments are made about the output signals we detect. The
external harmonic mixers can produce their own mixing signals when the local
oscillator frequency (LO) from the spectrum analyzer and input radio frequency (RF)
from the device are separated by the intermediate frequency (IF) of the down-mixed
signal. As a result, false signals can appear in the spectrum analyzer window. To
confirm that the spectra seen are real outputs from the device and not false signals
from the external mixer, the spectrum analyzer can be put into signal identification
mode100 which eliminates false signals and only displays input signals with
uncalibrated amplitudes. Signal identification mode was used at both 75.5 and 109.5
GHz for CNT FET devices, and the output peaks in Figure 7.8 remained unchanged.
We also briefly mention that a false mixing output was once observed on the
through electrodes control when the W probe was lowered too far, broke through the
electrode pad and penetrated into the quartz substrate. As a result a mixing output
signal appeared and faded sporadically over time. The probe tip material is listed as
?special? by Cascade and may contain a semiconductor, thus forming a Schottky
diode which could explain the production of a mixing signal. When the controls were
probed properly, no such anomalous mixing signals were observed.
124
7.2.2.2 Varying the Input Power
The mixing product coefficients in the Taylor expansion we use to represent
our drain current under mixing (EQ 5.16) contain no direct frequency dependence,
but the attenuation of input signals from our experimental setup can vary with input
frequency, so comparisons of measured data with that approximation are best
accomplished at fixed frequencies as a function of input signal amplitudes. Mixing
experiments were performed at an output frequency of 75.5 GHz by sweeping the
first signal generator?s output power (labeled A1) from -10 to 20 dBm in steps of 0.5
dBm. The second signal generator was held at a fixed 15 dBm output power (labeled
A2). Each captured spectrum was an average of 25 sweeps in the spectrum analyzer.
Note that A1 and A2 are the powers reported by the signal generators, not the powers
reaching the device. From measurements in the SMA regime, a drop of 30 dB is
expected for input frequencies near 25 GHz. The spectra at A1 = 0, 10, and 20 dBm
for A2 = 15 dBm are shown in Figure 7.11. Vertical lines mark the center of each
spectrum.
The asymmetry of the spectra is seen to shift from left to right as the input
power of the higher-frequency signal generator is increased. Also, the peaks seem to
increase in amplitude as expected. To better analyze the changes with input power,
Figure 7.12 shows a plot of the six mixing output powers vs. the first signal
generator?s output power over the full range of powers used.
125
Figure 7.11 Output spectra at 75.5 GHz of a CNT FET used to mix two input frequencies
separated by 100 kHz. The output frequencies are three times the input frequencies. One
signal generator power was kept fixed (A2 = 15 dBm) while the other was 0 dBm (black), 10
dBm (red), and 20 dBm (blue).
Figure 7.12 Plot of the output power of six mixing products vs. the output power of one
signal generator in a two-tone mixing setup performed on a CNT FET (W = 100 ?m, L = 10
?m, G = 8 ?m). The other signal generator was held at a fixed output power of 15 dBm. The
126
output frequency was 75.5 GHz, which is three times the input frequencies, which were
separated by 100 kHz.
Each peak intersects its complementary peak at the same A1 value of 16 dBm
(green dashed vertical line). This is where the output spectra would be symmetric and
is expected since it is 1 dB above where the second signal generator is held at 15
dBm. However, the drop in output power for some of the peaks was not expected.
Consider for example the (0, 3) peak. Using only one term in the expansion, this
output amplitude should only depend on the output power of the second signal
generator, which is fixed, so it should continue as a straight line. By including a
second term in the expansion, the fifth-order contribution will have an A1 dependence
as well, so it should increase slightly with A1, but certainly not decrease. This
decrease in (0, 3) indicates that the power reaching the device from the second signal
generator is not held fixed but depends on the power from the first signal generator.
Since the powers reaching the input of the device are unknown, they are extracted
from the third-order harmonics of the input tones and used to predict the mixing peak
output power dependence on input power. The reason why the signal power from the
second signal generator reaching the device did not stay constant and the method of
recalibration are presented in Appendix F.
The mixing peak amplitudes could be directly analyzed at low input powers
where the device input powers are as expected, but fewer peaks are produced at those
input powers. To consider the six main peaks observed, the powers produced by the
signal generators are converted to the powers incident upon the device. To summarize
the conversion method, when ignoring the fifth-order and higher expansion terms, the
third-order harmonics of the two fundamentals, (0, 3) and (3, 0), do not depend on the
127
input power of the other tone; therefore their outputs should primarily depend only on
one signal generator. The (3, 0) data are used to extract the signal power at the input
of the device from signal generator #1 and the (0, 3) data are used to extract the same
from signal generator #2. After the input powers reaching the CNT FET are known,
they are used with our Taylor expansion of drain current under mixing to produce the
expected output amplitude dependence on input power for six mixing products, which
is compared with the measured data.
Considering only one term in each peak?s expansion, the output power as a
function of the input power for all six peaks are listed in EQ 7.2, where Aout is the
output power in dBm, Ain is the device input power in dBm, and ?n is given by EQ
7.3.
? ? 221
4
3
2010 3)2,1( ininout AALogLogA ???
?
?
?
?
?
?? ?
? ? 212
4
3
2010 3)1,2( ininout AALogLogA ???
?
?
?
?
?
?? ?
? ? 23
4
1
2010 3)3,0( inout ALogLogA ??
?
?
?
?
?
?? ?
? ? 13
4
1
2010 3)0,3( inout ALogLogA ??
?
?
?
?
?
?? ?
? ? 241
16
5
2010 5)4,1( ininout AALogLogA ???
?
?
?
?
?
??
?
?
? ? 214
16
5
2010 5)1,4( ininout AALogLogA ???
?
?
?
?
?
??
?
?
EQ 7.2
42 Rnn ?? ?
!
1
1
1
nV
G
n
gs
D
n
n ?
?
?
?
??
EQ 7.3
By matching the expected output to the (0, 3) and (3, 0) measured data, the
beta parameters were ?3 = 0.537 mW and ?5 = 20.0 mW. These values
128
give:
?
?
?
2
2
gs
D
V
G
55.6 ?S / V2 and
?
?
?
4
4
gs
D
V
G
6.78 ?S / V4. Figure 7.13 displays the
measured data for each mixing product plotted against the expected product behavior
using EQ 7.2. The agreement for the (0, 3) and (3, 0) plots is nearly perfect. This may
seem inevitable since they were used to predict the input amplitudes, but they were
not transposed point-by-point. Simple functions fit their behavior very well. The other
four mixing product projections follow the same trends as the data, falling short at
large signal generator powers. Including more terms in the expansion would raise the
projected output power in each plot. These plots confirm that the mixing behavior we
observe in our high frequency two-tone mixing experiments follow the input signal
amplitudes expected from our Taylor expansion of drain current. The lesson from
these measurements is that at low input powers, the powers reaching the device are
directly related to the powers generated by the signal generators. At larger input
powers, even though unexpected power-dependent attenuation occurs in the system,
the measured mixing product amplitudes follow the device input power dependencies
derived from the Taylor expansion.
129
Figure 7.13 Plots of the output power of six mixing product peaks vs. the output power of one
signal generator in a two-tone mixing setup performed on a CNT FET (W = 100 ?m, L = 10
?m, G = 8 ?m). The black dots are the measurements taken, and the red lines are the expected
mixing product amplitudes based on a single expansion term per peak. The other signal
generator was held at a fixed output power of 15 dBm. The output frequency was 75.5 GHz +
300 kHz, which is three times the input frequency of the first signal generator.
130
7.2.2.3 Examining the (1, 2) to (3, 0) Amplitude Difference
An additional confirmation cross-check for the output amplitudes is to
compare the amplitudes of the (1, 2) and (2, 1) products with the (0, 3) and (3, 0)
peaks for equal inputs. An example of these amplitudes is shown in Figure 7.14.
Since all four peaks are 3rd order products, they have the same coefficients in front of
the first terms in their expansion and should maintain the same ratio independent of
input frequency and power.
Figure 7.14 Output spectra at 75.5 GHz of a CNT FET used to mix two input frequencies
separated by 100 kHz. The two input amplitudes were equal, producing the symmetric
spectrum. The (1,2) and (3,0) mixing peak amplitudes are shown in red. The output
frequencies are three times the input frequencies.
For Pin1 = Pin2,
? ?
? ?
93
1
4
1
1
4
3
2
3
2
3
3
2
3
)0,3(
)1,2(
??
?
?
?
?
?
?
?
?
?
?
?
?
?
in
in
out
out
P
P
P
P
?
?
EQ 7.4
131
A power ratio of 9 corresponds to 9.54 dB, so to first order approximation,
these peaks should always have an amplitude difference of 9.54 dB. If the expansion
includes more terms and the input amplitudes are high enough, this amplitude
difference can decrease, shown in the two-term expansion case of EQ 7.5 ? 8.
? ? ? ? ? ? ? ? ?
?
?
?
?
?
????
2
5
33
5
5
5
3
3)2,1( 8
25
4
3
8
25
4
3
VVVVVI d ?????
? ? ? ? ? ? ? ? ?
?
?
?
?
?
????
2
5
33
5
5
5
3
3)3,0( 16
25
4
1
16
25
4
1
VVVVVI d ?????
EQ 7.5
2
2
5
3
2
2
5
3
)3,0(
)2,1(
16
25
4
1
8
25
4
3
??
?
?
??
?
?
?
??
?
?
??
?
?
?
?
V
V
P
P
out
out
?
?
?
?
EQ 7.6
For
2
5
3 V??
?
?
: 9
4
1
4
3
2
5
3
2
5
3
)3,0(
)2,1(
?
??
?
?
??
?
?
??
?
?
??
?
?
?
?
?
?
?
out
out
P
P
EQ 7.7
For
2
5
3 V??
?
?
: 4
16
25
8
25
2
2
2
2
)3,0(
)2,1(
?
?
?
?
?
?
?
?
?
?
?
?
?
?
V
V
P
P
out
out
EQ 7.8
From the derivatives of GD extracted previously,
?
5
3
?
?
164 for our measured
device. That would require input voltages on the order of 13V or higher to see the (1,
2) and (3, 0) peaks come closer together, which is much larger than we ever approach.
We are therefore in the regime when only the first term in the expansion should
significantly contribute the output product height, and we expect a constant 9.54 dB
amplitude difference in our measurements.
132
Symmetric spectra at 75.5 GHz and 109.5 GHz were captured for twelve
different CNT FET device geometries. The devices measured covered all of the
different FET geometries fabricated on one chip, which are mentioned in Chapter 2.
The particular channel widths, channel lengths, and gate lengths for each device are
given in Table 7.1
Device 1 2 3 4 5 6 7 8 9 10 11 12
W 100 100 100 100 100 100 100 100 100 100 50 25
L 10 10 10 10 5 10 25 50 75 10 10 10
G 5 8 10 12 3 8 21 46 71 8 8 8
Table 7.1 Channel widths, channel lengths, and gate lengths in micrometers for twelve CNT
FETs that were used for two-tone high frequency mixing measurements
The difference between the (1, 2) and (3, 0) peak amplitudes for all 12 devices
are plotted in Figure 7.15. All the points fall within the maximum uncertainty of the
equipment (shown in green) of the expected value. This shows that the essential
nanotube nonlinearity that generates these mixing peaks continues to operate up to
110 GHz.
133
Figure 7.15 Output amplitude difference between the (2,1) and (3,0) mixing products when
the input powers are equal. Twelve different CNT FET devices with different electrode
geometries were measured. The expected amplitude difference of 9.54 dB should be
independent of frequency and device type. The experimental uncertainty bounds are
displayed as green lines.
7.2.3 Heterodyne Detection at 37 GHz LO Operation
The two-tone mixing experiments performed in the last section had the two
input frequencies separated by 100 kHz so that many mixing peaks could be observed
in a small output window (1.2 MHz). For real-world applications, a mixer is typically
used for heterodyne signal detection. This is accomplished by having one input, the
local oscillator (LO) at the fixed frequency and the second input signal varied over a
specified frequency range. The resulting output signal is selected to correspond to one
particular mixing product. This arrangement is now used with our CNT FETs acting
as mixers to take a RF input signal from the K frequency regime (DC ? 40 GHz) and
134
mix it up to the W regime that we can observe with our external mixer experimental
setup (75 ? 110 GHz).
Our first signal generator produces our LO signal at a fixed frequency and
power of 37 GHz and 18 dBm, respectively. Our second signal generator produces
our RF signal from 1 GHz to 36 GHz at a fixed power of 10 dBm. The output mixing
product of interest is (2, 1). The advantage of using this arrangement is that
magnitude of the fixed LO signal is much larger than the varied RF signal, so the
frequency dependent attenuation of the RF signal will have a minimized effect on the
output amplitude. Since the LO power is 7 dB greater than the RF power (including
the 1 dB loss from extra cable), the contribution to the output power of the mixing
product from the LO signal is 14 dB greater than the contribution from RF signal, as
shown in EQ 7.9.
? ? IRALOALogLogA ininout ???
?
?
?
?
?
?? 2
4
3
2010 3)1,2( ?
EQ 7.9
Several different devices were measured in this configuration. The input
frequency of the RF signal was swept from 1.25 GHz to 35.75 GHz by steps of
250 MHz. The spectra were captured in a 120 kHz window, resolution bandwidth of
3 kHz, and the peak amplitude extracted as the maximum value in that window. The
signal generators and spectrum analyzer were controlled with Labview over GPIB.
The results for the three control devices mentioned earlier are plotted in
Figure 7.16. They are through electrodes, a parallel plate capacitor, and a 10 ?m
channel length, 8 ?m gate length FET geometry without CNTs in the channel. No
peaks were visible in the output frequency window, so the output signals for each
form the noise floor, indicating that if any unintended mixing does occur in this
135
frequency regime without a CNT FET as the DUT, it is not detectable with our
equipment. This suggests that any mixing signals observed in the measurement of
CNT FETs is attributed to the presence of CNTs.
Figure 7.16 Output power of the mixing product from the 37 GHz LO mixing measurements
for the three control devices. The LO signal is fixed at 37 GHz and 18 dBm while the RF
signal is swept from 1.25 to 35.75 GHz at a power of 10 dBm.
Also measured were CNT FETs with channel lengths of 10 ?m but with
different gate lengths (5, 10, 15 ?m). The results of those measurements are shown in
Figure 7.17. For devices that have gates that do not entirely span the channel, the
capacitance of the gate to the CNTs is expected to be greater than the capacitance of
the gate to the source and drain. Increasing the gate length increases all three
capacitances similarly, but as the gate begins to overlap the source and drain
electrodes, the electrode capacitance grows considerably while the gate-CNT
capacitance remains constant. This increased gate-source capacitance provides a low
136
resistance path to ground that will reduce the amount of the signal passing through the
channel of the device. Figure 7.17 shows this increasing loss of signal as the gate
length is increased.
Figure 7.17 Output power of the mixing product from the 37 GHz LO mixing measurements
for CNT FETs with channel lengths of 10 ?m but different gate lengths. The LO signal is
fixed at 37 GHz and 18 dBm while the RF signal is swept from 1.25 to 35.75 GHz at a power
of 10 dBm.
By performing the same circuit analysis as in Chapter 5, the extra attenuation
due to the increased gate-source and gate-drain capacitances from gate overlap can be
predicted. Those predictions with no overlap, 1 ?m of overlap, and 5 ?m of overlap
are shown in Figure 7.18. Increased attenuation with increasing gate overlap is
shown. However, the slopes of all three curves are predicted to be similar for the
input frequency range of 25 to 37 GHz, while the data appears to show an increase in
slope with increasing gate overlap. Acquiring a more accurate representation of the
137
output mixing peak amplitudes requires an analysis beyond lumped elements,
representing the CNT as a transmission line.
Figure 7.18 Predicted attenuation from the CNT FET circuit model in Chapter 5 for three
different amounts of gate overlap. a) is on a logarithmic frequency scale over five orders of
magnitude and b) is on a linear scale over the HF input frequency range, 25 to 37.5 GHz.
The last devices compared with this measurement setup are CNT FETs with
different channel lengths. The output data for a device with a 5 ?m channel length
and a 75 ?m channel length are plotted in Figure 7.19. For both devices, the gate
length is shorter than the channel length. The results for both devices are nearly
identical, which is surprising considering that the channel impedance of the L = 75
?m device is 40 times greater than the L = 5 ?m device. These results show that the
transit time cutoff, present when the input signal is placed on the source and the
output is taken at the drain, does not apply to this measurement setup. Using the
ballistic limit of the Fermi velocity, the maximum transit length of charges at our
maximum input frequency of 37 GHz is 3.5 ?m. This only corresponds to 5% of the
138
75 ?m channel length. This suggests that only a few microns of the nanotube length
near the drain are active at such high frequencies; charges in the rest of the nanotube
length do not contribute to the output current. If charges needed to span the entire
device channel, no output would be observed in the long channel length device.
Figure 7.19 Output power of the mixing product from the 37 GHz LO mixing measurements
for CNT FETs with different channel lengths. The LO signal is fixed at 37 GHz and 18 dBm
while the RF signal is swept from 1.25 to 35.75 GHz at a power of 10 dBm.
Output detection beyond the transit time cutoff from the Fermi velocity could
be explained with plasmon excitations, which are expected to travel faster than non-
interacting charges. If the plasmon velocity was four times the Fermi velocity,
plasmons could span the 5 ?m channel at such high input frequencies but not the 75
?m channel. The plasmon velocity would need to be 21 times the Fermi velocity for
plasmons originating at the source to the reach the drain in one period. The more
likely explanation is that by placing the input signal on the gate, charge motion
139
responsible for mixing occurs at both ends of the channel. These devices demonstrate
CNT non-linear operation at the highest frequency directly observed to date and at
channel lengths 100 times larger the submicron channel lengths previously assumed
necessary for GHz operation.77-79, 83, 85
7.3 Summary
In this chapter, we have reviewed our results from two-tone mixing
experiments in two frequency regimes. Measurements from 1 GHz to 26.5 GHz
(SMA regime) showed that at low frequencies output mixing signals are visible from
control devices, demonstrating unwanted mixing from the equipment used in the
experimental setup. These third-order output signals die out around Fout = 5 GHz in
control devices while they remain in the CNT FETs, confirming that the CNTs are
responsible for the mixing behavior seen above this frequency.
In the high frequency regime, the mixing product outputs are found to only
slightly decrease in amplitude from 75 to 110 GHz. The mixing product amplitudes
are compared with their expected values from our Taylor expansion of drain current
by sweeping the input power while holding the input frequency fixed. Some
unexpected output behavior is seen and attributed to input power limitations of the
power divider used. Recalibration of the input powers reaching the device, extracted
from the third harmonic products of the two input tones, allows us to project the
expected output amplitudes of other peaks. Results closely matching the data show
our results to be consistent with the Taylor expansion form of mixing we use.
140
Finally, sweeps of the output frequency range are made in a practical mixing
configuration using a fixed 37 GHz LO input to compare different types of devices.
The controls did not produce any measurable mixing products in these measurements
while the CNT FETs did. In devices with gates overlapping the source and drain
electrodes, the output signal quickly decreases with frequency, as expected with
increased gate-source capacitance and our circuit model. No strong dependence on
channel length was observed, with mixing signals from a 75 ?m channel length
device of similar magnitude to devices with 10 and 5 ?m channel lengths, indicating
that our measurement technique overcomes the transit time cutoff frequency. These
experiments clearly show CNT mixing behavior up to 110 GHz, the highest output
frequency directly observed to date, with no indication of the output dropping below
the noise floor at even higher frequencies.
141
Chapter 8: Summary and Conclusions
In this dissertation, field-effect transistors using aligned arrays of carbon
nanotubes were used as passive mixers to convert input frequencies to output
frequencies as high as 110 GHz, the highest CNT FET output frequencies observed to
date. After describing the design and fabrication of the devices, the role of defects in
the dielectric layer acting as charge traps was investigated by observing how the drain
current approached an equilibrium value in response to non-zero gate pulses when
held at a constant source-drain bias. These currents were observed to follow a power
law with time until nearly reaching equilibrium. By observing the behavior of charge
traps as a function of the gate voltage magnitude and pulse duration, a measurement
method that minimizes the effects of traps was developed. By using short
measurement pulses, the number of traps filled during each pulse was minimized. A
default trap state occupation was set in between each measurement pulse by applying
an alternating decreasing voltage waveform to the gate, similar to a degaussing
method used on cathode ray tubes. This measurement method was shown to reduce
hysteresis in transfer curves beyond other published measurement waveforms. The
time-independent Preisach hysteresis model was used to derive this waveform and to
predict the drain current output for the waveforms considered, matching the measured
data well. For passive high frequency measurements, only small input voltages
centered at Vgs = 0 were used, so the presence of charge traps was not detrimental. For
DC measurements, our method reduced the hysteresis in transfer curves to the level of
measurement noise, with a standard transfer curve taking less than one minutes to
acquire.
142
We then presented a method to successfully remove metallic CNTs from as-
grown aligned arrays on quartz, increasing the on/off current ratio to 103 and higher.
The presence of metallic CNTs in the channel prevented the FET from being gated
off, adding a constant channel current that does not contribute to device operation
whenever the device is biased. Our high frequency mixing experiments were
performed passively, without any bias, so this constant current under bias did not
apply to our measurements. For active devices, this would be a necessary step to
avoid unwanted power losses. Since increasing the on/off ratio also reduces the on-
state current, the devices we used for our high frequency mixing experiments used as-
grown nanotube arrays to minimize the device input impedance and maximize our
output signal magnitudes.
The last three chapters focused on high frequency CNT measurements. First,
the work by previous research groups was presented. The experiments most similar to
ours used CNT FETs as passive mixers by placing input signals on the source and
measuring outputs at the drain. Two frequency cutoffs were expected in such a
measurement setup. First, the device resembled a low-pass filter with an associated
RC corner frequency, above which significant attenuation would occur. The second
cutoff frequency corresponded to the transit time necessary for charges to span the
channel. This transit time cutoff frequency was a hard cutoff, above which no output
signal should be detected. Because of these cutoffs, submicron channel lengths were
used and only DC or low frequency output signals were measured.
Our experimental method placed the input signal on the gate instead of the
drain, even though our CNT FETs were used passively. This changed the circuit
143
model so that it no longer directly resembled a low-pass filter and the output
attenuation should be decreased. This measurement technique also overcame the
transit time cutoff frequency because signals would be induced at the drain electrode
without needing to span the channel. Instead of merely measuring the DC or low
frequency mixing output signals, high frequency outputs were directly measured
using an external mixer and a spectrum analyzer with a low noise floor (-90 dBm).
After introducing our experimental setup, generic frequency mixing was
introduced. We presented an approximation of the drain current of a CNT FET as a
Taylor expansion of a small input voltage. A single-tone mixing experiment was
performed to test the validity of this expansion. The DC mixing current was measured
and compared with the expected output derived from the derivative of conductance
with respect to gate voltage. The two were shown to match well in the regions where
the approximation was valid.
We then performed two-tone mixing experiments on CNT FETs and control
devices to eliminate any mixing contributions from the experimental setup. By using
an external harmonic mixer to convert our device output signals to frequencies that
our spectrum analyzer could detect, our CNT FETs demonstrated generated output
frequencies as high as 110 GHz. Up to 11th order mixing products were observed in
this regime. The difference between two third-order mixing product amplitudes was
found to be within the measurement uncertainty of the difference expected from our
drain current approximation, and all the mixing product amplitudes were found to
depend on input power as expected.
144
The attenuation of our output signals followed the frequency dependence
expected from our circuit model. Power loss is never an intended device feature, but
certain levels of power loss may be acceptable for many applications. In
telecommunication systems, for example, a link budget accounts for all the gains and
losses from a transmitter to a receiver. Each component does not need to provide gain
itself; the total losses need merely fall within a range. Our two-tone W regime mixing
experiments are an example of high frequency function in spite of large power loss.
Input signals of 36.5 GHz were mixed up to 109.5 GHz, experiencing a total power
loss near 70 dB from signal generator output power to device output power. Then the
external mixer down-mixed the produced signals to below 6 GHz where they were at
a power level where they could still be measured. Amplification in that frequency
regime is a much easier task. By mixing passively, no power was wasted in powering
the CNT devices themselves.
The output signals from devices with different channel lengths were
compared. If measured with the input signal on the source, the transit time cutoff of
the long channel length device (75 ?m) should have prevented any signal from being
detected. With our measurement technique, the output signals from both devices were
nearly identical, possibly indicating that only a few microns of the device length near
the drain are active at high frequencies. This shows that high frequency performance
can be achieved in CNT devices without slow and costly submicron fabrication
methods like electron beam lithography. Standard photolithography and even shadow
mask deposition can produce devices of this scale.
145
Further experiments could clarify these results. It would first be beneficial to
fill in the gap (25 to 75 GHz) of our output frequency window with additional
external mixers. This would provide continuous data that could be better compared
with circuit models. We would also like to access even higher output frequency
regimes (> 110 GHz) to see if other limitations begin to appear. To verify the
mechanism of signal propagation through the device, electrode geometries with long
channel lengths like those in Figure 8.1 would be measured. If only a few microns of
channel length near the drain are active at high input frequencies, devices a), b), and
c) should have similar output signals, even though the gate-CNT capacitance is
changing. In devices e) and f), the gate-drain capacitance will be very small and
charge motion should only occur in CNTs themselves under the gate. The transit time
cutoff should apply in such a case, with a transit length equal to the ungated length of
CNTs from the gate to the drain. If HF output signals were detected in such devices,
different mechanisms of mixing/and or signal propagation would be necessary.
CNTSource Drain
a) Gate
CNTSourc Drain
b) Gate
CNTSource Drain
c) Gate
CNTSource Drain
d) Gate
CNTSource Drain
e) Gate
CNTSource Drain
f) Gate
Figure 8.1 Electrode geometries for a long channel length device that could indicate how an
input signal on the gate induces a mixing current in the CNT channel.
We have increased the observable output frequency regime for carbon
nanotube field-effect transistors. Even uncommonly large devices (channel lengths of
146
75 ?m) continued to operate up to 110 GHz without indication that performance was
beginning to fail. By mixing up input frequencies in the SMA regime, we have
demonstrated an economical way of producing clean high frequency output signals in
an expensive frequency regime. Passive operation eliminates the need for a local
power supply, allowing for incorporation into small and portable circuits such as
radio frequency identification (RFID) tags. Our devices could be made wafer scale by
standard photolithography methods, and the attenuation of the generated output
frequencies could be improved with more carefully designed electrode geometries.
Output current and power could also be easily increased by using wider channels
(more CNTs).
147
Appendix A: Fabrication Procedure for Aligned CNT FETs on Si/SiO2
Substrate
These instructions produce a top local gate, bottom global gate aligned CNT
FET on a silicon/silicon dioxide substrate. The typical device geometries are: silicon
substrate with 500 nm thermal oxide, aligned CNT channel, 50 nm Ti/Au source-
drain, 75 nm Al2O3 dielectric, 100 nm Ti/Au gate. If the metallic CNTs are to be
removed by joule heating, that should be done after patterning the CNTs and before
depositing the Al2O3 dielectric. Two substrates are necessary, the quartz CNT growth
substrate and the silicon/silicon dioxide device substrate.
Quartz CNT Growth Substrate
Purchase Wafers:
3? diameter, single crystal quartz wafers, 0.5mm thick, ST-cut 42?45?, single side
polished (SSP), Hoffman Materials
Dicing:
CNTs align themselves perpendicular to the wafer flat, so either mark the back side of
the wafer with a diamond scribe to indicate the alignment direction or dice into
shapes which allow one to identify the alignment direction.
Cover wafers with photoresist to protect surface during dicing. Many photoresists
would do.
spin HMDS at 4000 rpm for 60 sec
spin positive photoresist, OIR 908-35 at 4000 rpm for 60 sec
bake on hotplate at 90 C for 60 sec
dice with diamond saw, both sides of wafers covered by tape
remove tape
soak and sonicate in Acetone for 10 minutes
rinse with acetone, methanol, and isopropanol (IPA)
blow dry with nitrogen
Annealing: necessary for CNT alignment. To prevent cracking, heating and cooling
rates should be < 10 C/min.
heat quartz pieces in open tube furnace (air) to 900 C
hold for 12 hours
cool to room temperature
148
Catalyst Deposition: Ferritin - for medium density and partial alignment over whole
sample surface.
dilute ferritin solution with DI to 1 mg/mL
soak quartz pieces in solution for 20 min in refrigerator, 3 C
rinse with DI
blow dry with nitrogen
keep ferritin refrigerated - do not let freeze
CVD: Hydrogen Anneal - to clean the quartz surface of residual photoresist.101 To
prevent cracking, heating and cooling rates should be < 10 C/min.
We use a three-zone tube furnace from Lindberg Blue, 2 in quartz tube, mass flow
controllers, and quartz boats from Glass Tech. All three zones are set to the same
temperature.
insert samples in boats into the center of the tube
flush with 200 sccm argon for five minutes
flow 1700 sccm argon and 1900 sccm hydrogen
heat to 400 C
hold temperature and flow rates for 1 hour
cool to room temperature
stop hydrogen flow
stop argon flow after 5 minutes
remove samples
CVD: CNT Growth
Many papers report using different gases: methane, ethylene, methanol and ethanol
bubblers, water bubblers, etc. No significant difference in density was found from
these other sources. To prevent cracking, heating and cooling rates should be < 10
C/min.
insert samples in boats into the center of the tube
heat in air to 700 C - to burn away the organics in ferritin
hold for 10 min (oxidizing step)
cool to room temperature
seal tube ends
flush with 1000 sccm argon, hydrogen, and methane for 5 min
stop argon and methane flow
set hydrogen flow to 250 sccm
heat to 925 C
hold for 10 min (reducing step)
introduce methane flow 1000 sccm
hold for 20 min
stop methane flow
cool to room temperature
stop hydrogen flow
flush with 1000 sccm argon for 5 min
remove samples
149
Remove CNTs from Substrate: Uses single sided thermal tape (120 C transistion
temperature, Revalpha 3195MS, Nitto Denko)
e-beam evaporate 100 nm Au over whole quartz piece
cut a piece of thermal tape with scissors
remove plastic coating from adhesive side
carefully press tape against gold coated quartz piece
smooth tape over the surface with fingers
carefully peel tape away from quartz substrate with tweezers, removing gold layer
and CNTs
Silicon Device Substrate
Purchase Wafers:
3? diameter silicon wafers, 0.381mm thick, 1-0-0 orientation, 500 nm thermal
oxide, Silicon Quest Int?l
Clean Wafers and Dice Wafers:
soak full wafers in TCE, acetone, methanol, and IPA for 5 min
blow dry with nitrogen
bake at 120 C in an over for 20 min
dice wafer into quarter pieces with a diamond scribe
rinse with acetone, methanol, IPA
blow dry with nitrogen
Transfer CNTs from Thermal Tape to Silicon Device Substrate
e-beam evaporate 100 nm Au over whole Si/SiO2 piece
carefully press the adhesive side of the thermal tape piece carrying the Au/CNT
layer to the Si/SiO2 piece
wipe a clean razor blade across the tape surface to ensure uniform contact
place Si/SiO2/Au/CNT/Tape assembly on a 150 C hot plate
remove the tape with tweezers as it peels away
gently rinse Si/SiO2 piece with acetone, methanol, and IPA
blow dry with nitrogen
run 45 sec oxygen plasma (100 W, 200 mTorr, 16 sccm O2) in reactive ion etcher
soak sample in diluted gold etchant (2:1 DI to gold etchant) for 2 min
rinse with DI
blow dry with nitrogen
Photolithography: Source-Drain Electrodes
For the high frequency CNT devices, some channel lengths are < 1um, so a stepper
was used instead of a contact aligner. The procedure would be very similar if using a
contact aligner.
spin HMDS at 4000 rpm for 60 sec
spin positive photoresist, OIR 908-35 at 4000 rpm for 60 sec
bake on hotplate at 90 C for 60 sec
150
align source-drain mask pattern in stepper
expose pattern in stepper (400um chuck, 0.24s exposure)
bake on hotplate at 120 C for 60 sec
develop in OPD 4262 for 60 sec
e-beam evaporate 5 nm Ti and 45 nm Au (CHA Mark IV)
liftoff in acetone for 1 hr
use tweezers to remove large flakes of metal weakly sticking to the sample
use strong acetone sprayer to ensure liftoff of small features
rinse in acetone, methanol, IPA
blow dry with nitrogen
Photolithography: Pattern CNTs
This step defines the channel width for each FET and prevents separate devices from
being shorted together.
spin OIR 908-35 at 4000 rpm for 60 sec
bake at 90 C for 60 sec
align the CNT pattern mask to the sample alignment marks or source-drain
electrodes
expose mask pattern in contact aligner for 13 sec
develop in OIR 4262 for 60 sec
rinse in DI for 30 sec
run 10 sec oxygen plasma (100 W, 200 mTorr, 16 sccm O2) in reactive ion etcher
expose full sample in contact aligner for 13 sec
develop in OIR 4262 for 60 sec
soak in acentone for 3 min
rinse with acetone, methanol, IPA
blow dry with nitrogen
CVD: Hydrogen Anneal - best if done immediately before the next ALD step.
Included in ensuring a clean surface, this process encourages p-type devices.101
insert samples in boats into the center of the tube
flush with 200 sccm argon for five minutes
flow 1700 sccm argon and 1900 sccm hydrogen
heat to 400 C
hold temperature and flow rates for 1 hour
cool to room temperature
stop hydrogen flow
stop argon flow after 5 minutes
remove samples
Atomic Layer Deposition: Dielectric Layer
As is discussed in Chapter 3, the deposition temperature will have an effect on the
dielectric quality and the charge traps in the sample. There is no procedural difference
when using different temperatures. We use a Beneq TFS 500 system.
place Si/SiO2 samples on a clean silicon wafer (the wafer will be used to
determine an etch rate later)
151
load growth recipe (Al2O3 baseline, 220 C)
vent system and insert sample
pump system and wait until it reaches temperature
run recipe - set target for 75 nm
vent lines
vent system and remove sample
pump system
log run
Photolithography: Determining Etch Rate:
dice the silicon wafer that held the Si/SiO2 pieces in the ALD into a few pieces
spin OIR 908-35 at 4000 rpm for 60 sec on silicon pieces
bake at 90 C for 60 sec
expose any mask pattern in contact aligner for 13 sec
develop in OIR 4262 for 60 sec
rinse in DI for 30 sec
dilute buffered oxide etch (BOE) with DI to be 1:3, giving a total molarity of:
rinse a silicon piece in dilute BOE for 60 sec
soak in DI for 2 min
soak in acetone for 2 min
rinse with acetone, methanol, IPA
determine etch depth with profilometer
using that rate, predict the length or time needed to etch 90% of the total dielectric
thickness (67.5 nm in our case)
rinse a different silicon piece in the dilute BOE for that length of time
soak in DI for 2 min
soak in acetone for 2 min
rinse with acetone, methanol, IPA
determine etch depth with profilometer
use the new depth and time to predict a more accurate etch rate
continue trial etching in BOE with different silicon pieces if necessary
Photolithography: Dielectric Etching
spin OIR 908-35 at 4000 rpm for 60 sec on Si/SiO2 samples
bake at 90 C for 60 sec
align the dielectric pattern mask to the sample alignment marks or source-drain
electrodes
expose in contact aligner for 13 sec
develop in OIR 4262 for 60 sec
rinse in DI for 30 sec
use the same dilute BOE as when determining the etch rate
add 5% to the total etch time, as it is worse to etch too little than too much
rinse the Si/SiO2 samples in dilute BOE for the new etch time
soak in DI for 2 min
soak in acetone for 2 min
rinse with acetone, methanol, IPA
152
Photolithography: Gate
spin HMDS at 4000 rpm for 60 sec
spin positive photoresist, OIR 908-35 at 4000 rpm for 60 sec
bake on hotplate at 90 C for 60 sec
align gate mask pattern to sample alignment marks in stepper
expose pattern in stepper (400um chuck, 0.24s exposure)
develop in OPD 4262 for 60 sec
rinse with DI for 30 sec
blow dry with nitrogen
e-beam evaporate 5 nm Ti and 95 nm Au (CHA Mark IV)
liftoff in acetone for 1 hr
use tweezers to remove large flakes of metal weakly sticking to the sample
use strong acetone sprayer to ensure liftoff of small features
rinse in acetone, methanol, IPA
blow dry with nitrogen
153
Appendix B: Fabrication Procedure for Top-Gate aligned CNT FET on
PET with Tri-Layer Dielectric and Vertical Integration
These instructions produce an integrated top-gate aligned carbon nanotube
FET circuit on a PET substrate with a tri-layer dielectric. The tri-layer dielectric was
needed because the CNTs in the channel tend to meander through the PMMA
dielectric and short the gate to the source-drain. The Al2O3 layer acts as a physical
barrier, and the two layers of PMMA are necessary with flexible devices to produce a
thick enough dielectric (around 600 nm after printing) to prevent significant gate
leakage. The instructions are only one method of fabricating such a device. Other
lithographic means could achieve a similar result. Three transfer substrates and one
device substrate are necessary to complete device.
Figure B.1 Diagram of completed integrated circuit using CNT FETs with aligned CNT
channels on PET
154
Figure B.2 a) Optical image of completed integrated CNT FET circuit on PET and b) SEM
image of CNTs between fingered electrodes before dielectric layers and top-gate are added.
Quartz Transfer Substrate: Quartz Piece with CNTs
Purchase Wafers:
3? diameter, single crystal quartz wafers, 0.5mm thick, ST-cut 42?45?, single side
polished (SSP), Hoffman Materials
Dicing:
CNTs align themselves perpendicular to the wafer flat, so either mark the back side of
the wafer with a diamond scribe to indicate the alignment direction or dice into
shapes which allow one to identify the alignment direction.
Cover wafers with photoresist to protect surface during dicing. Many photoresists
would do.
spin HMDS at 4000 rpm for 60 sec
spin positive photoresist, OIR 908-35 at 4000 rpm for 60 sec
bake on hotplate at 90 C for 60 sec
dice with diamond saw, both sides of wafers covered by tape
remove tape
soak and sonicate in Acetone for 10 minutes
rinse with acetone, methanol, and isopropanol (IPA)
blow dry with nitrogen
Annealing: necessary for CNT alignment. To prevent cracking, heating and cooling
rates should be < 10 C/min.
heat quartz pieces in open tube furnace (air) to 900 C
hold for 12 hours
cool to room temperature
Catalyst Deposition: Ferritin - for medium density and partial alignment over whole
sample surface.
dilute ferritin solution with DI to 1 mg/mL
soak quartz pieces in solution for 20 min in refrigerator, 3 C
rinse with DI
155
blow dry with nitrogen
keep ferritin refrigerated - do not let freeze
CVD: Hydrogen Anneal - to clean the quartz surface of residual photoresist.101 To
prevent cracking, heating and cooling rates should be < 10 C/min.
We use a three-zone tube furnace from Lindberg Blue, 2 in quartz tube, mass flow
controllers, and quartz boats from Glass Tech. All three zones are set to the same
temperature.
insert samples in boats into the center of the tube
flush with 200 sccm argon for five minutes
flow 1700 sccm argon and 1900 sccm hydrogen
heat to 400 C
hold temperature and flow rates for 1 hour
cool to room temperature
stop hydrogen flow
stop argon flow after 5 minutes
remove samples
CVD: CNT Growth
Many papers report using different gases: methane, ethylene, methanol and ethanol
bubblers, water bubblers, etc. No significant difference in density was found from
these other sources. To prevent cracking, heating and cooling rates should be < 10
C/min.
insert samples in boats into the center of the tube
heat in air to 700 C - to burn away the organics in ferritin
hold for 10 min (oxidizing step)
cool to room temperature
seal tube ends
flush with 1000 sccm argon, hydrogen, and methane for 5 min
stop argon and methane flow
set hydrogen flow to 250 sccm
heat to 925 C
hold for 10 min (reducing step)
introduce methane flow 1000 sccm
hold for 20 min
stop methane flow
cool to room temperature
stop hydrogen flow
flush with 1000 sccm argon for 5 min
remove samples
Release Layer:
2 hour vapor phase deposition of SAM release layer (tridecafluoro-1,1,2,2-
tetrahydrooctyle trichlorosilane)
Silicon Transfer Substrate 1: Silicon Piece with Source-Drain Electrodes
156
Purchase Wafers:
3? diameter silicon wafers, 0.381mm thick, 1-0-0 orientation, Silicon Quest Int?l
Clean Wafers and Release Layer:
soak full wafers in TCE, acetone, methanol, and IPA for 5 min
blow dry with nitrogen
bake at 120 C in an over for 20 min
2 hour vapor phase deposition of SAM release layer (tridecafluoro-1,1,2,2-
tetrahydrooctyle trichlorosilane)
Photolithography: Source-Drain Electrodes
e-beam evaporate 30 nm Au (CHA Mark IV) over full wafer
dice wafer into quarter pieces with diamond scribe
gently blow clean with nitrogen
spin positive photoresist, OIR 908-35 at 4000 rpm for 60 sec
bake on hotplate at 90 C for 60 sec
expose source-drain mask pattern in contact aligner for 13 sec
develop in OPD 4262 for 60 sec
rinse with DI for 30 sec
blow dry with nitrogen
expose whole sample in contact aligner for 13 sec
etch electrode pattern in diluted gold etchant (2:1 DI to gold etchant) for 20 sec
rinse with DI for 30 sec
develop in OPD 4262 for 60 sec
rinse with acetone, methanol, and IPA
blow dry with nitrogen
Silicon Transfer Substrate 2: Silicon Piece with Gate Electrodes, Interconnects,
and Dielectric
Purchase Wafers:
3? diameter silicon wafers, 0.381mm thick, 1-0-0 orientation, Silicon Quest Int?l
Clean Wafers and Release Layer:
soak full wafers in TCE, acetone, methanol, and IPA for 5 min
bake at 120 C in an over for 20 min
2 hour vapor phase deposition of SAM release layer (tridecafluoro-1,1,2,2-
tetrahydrooctyle trichlorosilane)
Photolithography: Gate Electrodes
e-beam evaporate 30 nm Au (CHA Mark IV) over full wafer
dice wafer into quarter pieces with diamond scribe
gently blow clean with nitrogen
spin positive photoresist, OIR 908-35 at 4000 rpm for 60 sec
bake on hotplate at 90 C for 60 sec
expose source-drain mask pattern in contact aligner for 13 sec
157
develop in OPD 4262 for 60 sec
rinse with DI for 30 sec
blow dry with nitrogen
expose whole sample in contact aligner for 13 sec
etch electrode pattern in diluted gold etchant (2:1 DI to gold etchant) for 20 sec
rinse with DI for 30 sec
develop in OPD 4262 for 60 sec
rinse with acetone, methanol, and IPA
blow dry with nitrogen
Shadowmask Deposition: Interconnects and Dielectric Layer
align interconnect shadowmask to the gate electrodes on sample
e-beam evaporate 600 nm Au (CHA Mark IV) through the shadowmask
remove mask
spin e-beam resist PMMA A7 at 4000 rpm for 60 sec
bake at 120C for 2 min
align dielectric layer shadowmask to the gate electrodes on sample
e-beam evaporate 25 nm of Al2O3 (CHA Mark IV) through the shadowmask
remove mask
spin e-beam resist PMMA A7 at 4000 rpm for 60 sec
align dielectric etching shadowmask to interconnects on sample
run 5 minute oxygen plasma (100 W, 200 mTorr, 16 sccm O2) in reactive ion
etcher
Assembling Device
Purchase PET:
Polyethylene terephthalate with one side treated for adhesion and the other coated
with ITO.
Melinex 453/700 from Dupont Teijin Films
Transfer Printing: Details of the procedure can be found in group publications.22, 23
cut piece of PET with scissors to appropriate size for the device substrate
sandwich PET device substrate between Si transfer substrate 1 and a blank Si
piece with a release layer (Si 1/RL/Au -> PET/ITO -> RL/Si blank)
transfer print source-drain electrode pattern from silicon transfer substrate #1 to
PET device substrate (Nanonex 2500 imprint machine, 500 PSI, 180 C, 3 min)
sandwich PET device substrate between quartz transfer substrate and a blank Si
piece with a release layer (Qu/CNT/RL -> Au/PET/ITO -> RL/ Si blank)
transfer print CNTs from quartz transfer substrate to PET device substrate
(Nanonex 2500 imprint machine, 500 PSI, 100 C, 3 min)
align Si transfer substrate #2 to PET device substrate in Nanonex contact aligner
add blank Si piece with a release layer to the top PET device substrate (Si
2/RL/Au/PMMA/ Al2O3/PMMA -> CNT/Au/PET/ITO -> RL/Si Blank)
transfer print gate electrodes, interconnects, and dielectric layers from Si transfer
substrate #2 to the PET device substrate (Nanonex 2500 imprint machine, 500
PSI, 100 C, 3 min)
158
Appendix C: Fabrication Procedure for Top Local Gated CNT FET on
Quartz
These instructions produce a top-gate aligned CNT FET on a quartz substrate.
The typical device geometries are: 500 um quartz substrate, aligned CNT channel, 50
nm Ti/Au source-drain, 75 nm Al2O3 dielectric, 100 nm Ti/Au gate. If the metallic
CNTs are to be removed by joule heating, that should be done after patterning the
CNTs and before depositing the Al2O3 dielectric.
Figure C.1 a) side-view and b) top-view diagram of a completed CNT FET on quartz. c)
Optical image of completed device.
Purchase Wafers:
3? diameter, single crystal quartz wafers, 0.5mm thick, ST-cut 42?45?, single side
polished (SSP), Hoffman Materials
Dicing:
CNTs align themselves perpendicular to the wafer flat, so either mark the back side of
the wafer with a diamond scribe to indicate the alignment direction or dice into
shapes which allow one to identify the alignment direction.
Cover wafers with photoresist to protect surface during dicing. Many photoresists
would do.
spin HMDS at 4000 rpm for 60 sec
spin positive photoresist, OIR 908-35 at 4000 rpm for 60 sec
bake on hotplate at 90 C for 60 sec
dice with diamond saw, both sides of wafers covered by tape
remove tape
soak and sonicate in Acetone for 10 minutes
rinse with acetone, methanol, and isopropanol (IPA)
blow dry with nitrogen
159
Annealing: necessary for CNT alignment. To prevent cracking, heating and cooling
rates should be < 10 C/min.
heat quartz pieces in open tube furnace (air) to 900 C
hold for 12 hours
cool to room temperature
Alignment Marks: (only needed for iron catalyst)
Many photoresists would do.
spin negative photoresist, NR7-1500 PY at 4000 rpm for 60 sec
bake on hotplate at 90 C for 60 sec
expose alignment mark mask pattern in contact aligner, 16 sec
bake on hotplate at 120 C for 60 sec
develop in RD6 for 15 sec
rinse with DI for 30 sec
e-beam evaporate 30nm Ti (CHA Mark IV)
liftoff photoresist in RR2, 85 C
rinse in acetone, do not let dry
rinse in photostripper (Microstrip 2001) for 3 minutes
rinse in acetone, methanol, IPA
blow dry with nitrogen
Catalyst Deposition: Ferritin - for medium density and partial alignment over whole
sample surface.
dilute ferritin solution with DI to 1 mg/mL
soak quartz pieces in solution for 20 min in refrigerator, 3 C
rinse with DI
blow dry with nitrogen
keep ferritin refrigerated - do not let freeze
Catalyst Deposition: Iron - for patterned high density and alignment. The same
patterning could be done with ferritin, with a small reduction in density.
To evaporate the thin film of iron, use a slow deposition rate (0.1 A/s). Use a new
single iron pellet in a clean crucible and focus the beam on the pellet before
depositing.
160
Figure C.2 Diagram of the CNT FET device geometry with the iron catalyst lithography mask
filled in.
spin positive photoresist OIR 908-35 at 4000 rpm for 60 sec
bake on hotplate at 90 C for 60 sec
align catalyst mask to sample alignment marks in contact aligner
expose in contact aligner, 13 sec
develop in OPD 4262 for 60 sec
rinse with DI for 30 sec
e-beam evaporate 0.5 nm Fe (CHE Mark IV)
expose sample surface without mask in contact aligner, 13 sec
soak in OPD 4262 for 3 min
soak in photostripper, Microstrip 2001 for 3 min
soak in acetone for 3 min
rinse with DI
blow dry with nitrogen
CVD: Hydrogen Anneal - to clean the quartz surface of residual photoresist.101 To
prevent cracking, heating and cooling rates should be < 10 C/min.
We use a three-zone tube furnace from Lindberg Blue, 2 in quartz tube, mass flow
controllers, and quartz boats from Glass Tech. All three zones are set to the same
temperature.
insert samples in boats into the center of the tube
flush with 200 sccm argon for five minutes
flow 1700 sccm argon and 1900 sccm hydrogen
heat to 400 C
hold temperature and flow rates for 1 hour
cool to room temperature
stop hydrogen flow
stop argon flow after 5 minutes
remove samples
CVD: CNT Growth
161
Many papers report using different gases: methane, ethylene, methanol and ethanol
bubblers, water bubblers, etc. No significant difference in density was found from
these other sources. To prevent cracking, heating and cooling rates should be < 10
C/min.
insert samples in boats into the center of the tube
heat in air to 700 C - to burn away the organics in ferritin and make isolated iron
oxide nanoparticles from iron films
hold for 10 min (oxidizing step)
cool to room temperature
seal tube ends
flush with 1000 sccm argon, hydrogen, and methane for 5 min
stop argon and methane flow
set hydrogen flow to 250 sccm
heat to 925 C
hold for 10 min (reducing step)
introduce methane flow 1000 sccm
hold for 20 min
stop methane flow
cool to room temperature
stop hydrogen flow
flush with 1000 sccm argon for 5 min
remove samples
Photolithography: Source-Drain Electrodes
For the high frequency CNT devices, some channel lengths are < 1um, so a stepper
was used instead of a contact aligner. The procedure would be very similar if using a
contact aligner.
Figure C.3 Diagram of the CNT FET device geometry with the source-drain lithography
mask filled in.
spin HMDS at 4000 rpm for 60 sec
spin positive photoresist, OIR 908-35 at 4000 rpm for 60 sec
bake on hotplate at 90 C for 60 sec
align source-drain mask pattern to sample alignment marks in stepper (if used)
expose pattern in stepper (400um chuck, 0.24s exposure)
162
bake on hotplate at 120 C for 60 sec
develop in OPD 4262 for 60 sec
e-beam evaporate 5 nm Ti and 45 nm Au (CHA Mark IV)
liftoff in acetone for 1 hr
use tweezers to remove large flakes of metal weakly sticking to the sample
use strong acetone sprayer to ensure liftoff of small features
rinse in acetone, methanol, IPA
blow dry with nitrogen
Photolithography: Pattern CNTs
This step defines the channel width for each FET and prevents separate devices from
being shorted together.
Figure C.4 Diagram of the CNT FET device geometry with the nanotube channel lithography
mask filled in.
spin OIR 908-35 at 4000 rpm for 60 sec
bake at 90 C for 60 sec
align the CNT pattern mask to the sample alignment marks or source-drain
electrodes
expose mask pattern in contact aligner for 13 sec
develop in OIR 4262 for 60 sec
rinse in DI for 30 sec
run 10 sec oxygen plasma (100 W, 200 mTorr, 16 sccm O2) in reactive ion etcher
expose full sample in contact aligner for 13 sec
develop in OIR 4262 for 60 sec
soak in acentone for 3 min
rinse with acetone, methanol, IPA
blow dry with nitrogen
CVD: Hydrogen Anneal - best if done immediately before the next ALD step.
Included in ensuring a clean surface, this process encourages p-type devices.101 To
prevent cracking, heating and cooling rates should be < 10 C/min.
insert samples in boats into the center of the tube
flush with 200 sccm argon for five minutes
163
flow 1700 sccm argon and 1900 sccm hydrogen
heat to 400 C
hold temperature and flow rates for 1 hour
cool to room temperature
stop hydrogen flow
stop argon flow after 5 minutes
remove samples
Atomic Layer Deposition: Dielectric Layer
As is discussed in Chapter 3, the deposition temperature will have an effect on the
dielectric quality and the charge traps in the sample. There is no procedural difference
when using different temperatures. We use a Beneq TFS 500 system.
place quartz samples on a clean silicon wafer (the wafer will be used to determine
an etch rate later)
load growth recipe (Al2O3 baseline, 220 C)
vent system and insert sample
pump system and wait until it reaches temperature
run recipe - set target for 75 nm
vent lines
vent system and remove sample
pump system
log run
Photolithography: Determining Etch Rate:
dice the silicon wafer that held the quartz pieces in the ALD into a few pieces
spin OIR 908-35 at 4000 rpm for 60 sec on silicon pieces
bake at 90 C for 60 sec
expose any mask pattern in contact aligner for 13 sec
develop in OIR 4262 for 60 sec
rinse in DI for 30 sec
dilute buffered oxide etch (BOE) with DI to be 1:3, giving a total molarity of:
rinse a silicon piece in dilute BOE for 60 sec
soak in DI for 2 min
soak in acetone for 2 min
rinse with acetone, methanol, IPA
determine etch depth with profilometer
using that rate, predict the length or time needed to etch 90% of the total dielectric
thickness (67.5 nm in our case)
rinse a different silicon piece in the dilute BOE for that length of time
soak in DI for 2 min
soak in acetone for 2 min
rinse with acetone, methanol, IPA
determine etch depth with profilometer
use the new depth and time to predict a more accurate etch rate
continue trial etching in BOE with different silicon pieces if necessary
164
Photolithography: Dielectric Etching
Figure C.5 Diagram of the CNT FET device geometry with the dielectric layer lithography
mask filled in.
spin OIR 908-35 at 4000 rpm for 60 sec on quartz samples
bake at 90 C for 60 sec
align the dielectric pattern mask to the sample alignment marks or source-drain
electrodes
expose in contact aligner for 13 sec
develop in OIR 4262 for 60 sec
rinse in DI for 30 sec
use the same dilute BOE as when determining the etch rate
add 5% to the total etch time, as it is worse to etch too little than too much
rinse the quartz samples in dilute BOE for the new etch time
soak in DI for 2 min
soak in acetone for 2 min
rinse with acetone, methanol, IPA
Photolithography: Gate
Figure C.6 Diagram of the CNT FET device geometry with the gate lithography mask filled
in.
spin HMDS at 4000 rpm for 60 sec
spin positive photoresist, OIR 908-35 at 4000 rpm for 60 sec
bake on hotplate at 90 C for 60 sec
align gate mask pattern to sample alignment marks in stepper
165
expose pattern in stepper (400um chuck, 0.24s exposure)
develop in OPD 4262 for 60 sec
rinse with DI for 30 sec
blow dry with nitrogen
e-beam evaporate 5 nm Ti and 95 nm Au (CHA Mark IV)
liftoff in acetone for 1 hr
use tweezers to remove large flakes of metal weakly sticking to the sample
use strong acetone sprayer to ensure liftoff of small features
rinse in acetone, methanol, IPA
blow dry with nitrogen
166
Appendix D: Capacitance Calculations
The capacitances of the parallel plate capacitors introduced in Chapter 2 were
extracted in several ways. First, the capacitance equation for parallel plate
arrangements was used based on the electrode geometries. This equation is given in
EQ D.1, where k is the dielectric constant, ?0 is the permittivity of free space, A is the
area of overlap, and d is the separation of the plates.
d
Ak
C 0
?
?
EQ D.1
The AC capacitance was also measured using an LCR meter at 100 kHz (HP
4275A Multi-Frequency LCR Meter). The third method used scattering (S)
parameters taken from VNA measurements from 100 kHz to 20 GHz. The impedance
of an ideal capacitor should be inversely related to frequency and capacitance, as
shown in EQ D.2. Using the input impedance calculated from S11 (EQ D.3), the
capacitance for each parallel plate capacitor was calculated in this high frequency
regime. The results of each of the three techniques is given in Table D.1. The parallel
plate calculation and LCR meter measurement correspond well. The results from the
VNA measurements at higher frequencies are of the same order of magnitude.
fCj
Z cap ?2
1
?
EQ D.2
1
1
50
11
11
?
?
?
S
S
Z in
EQ D.3
C1 C2 C3 C4 C5
Parallel Plate Calculation 44.0 pF 15.9 pF 2.75 pF 440 fF 27.5 fF
LCR Meter (100 kHz) 43.5 pF 15.5 pF 2.61 pF 381 fF 15.5 fF
S Parameter Extraction 20 pF 8 pF 2 pF 250 pF 30 fF
Table D.1 The calculated capacitances for five fabricated parallel plate capacitors using a
parallel plate calculation, an LCR meter, and S Parameters taken with a vector network
analyzer.
167
It is also useful to know the different capacitances in our CNT FETs, the
capacitance between the gate and the CNTs (CGCNT), the capacitance between the gate
and the sources (CGS), and the capacitance between the gate and the drain (CGD). The
capacitance between the gate and the CNTs can be approximated as the capacitance
between a cylinder and an infinite plane, given in EQ D.4, where h is the distance
between the cylinder and the plane, and d is the cylinder diameter.
? ?dh
k
L
CGCNT
/ln
2 0???
EQ D.4
Using a dielectric constant of 8, a dielectric thickness of 65 nm, a CNT
diameter of 1.5 ?m, and the channel length of 10 ?m, the approximate gate-CNT
capacitance is 1.2 pF per CNT. The capacitance of a CNT FET was also taken from
VNA measurements as described earlier. A plot of CGCNT extracted from S
Parameters is given in Figure D.1. It calculates CGCNT to be between 100 and 250 fF,
which corresponds to 80 to 210 CNTs, which is consistent with our growth density in
a 100 ?m wide channel.
168
Figure D.1 Gate-CNT capacitance (CGCNT) plotted as a function of frequency extracted from
scattering parameters taken on a CNT FET (W = 100 ?m, L = 10 ?m, G = 8 ?m).
To estimate the gate capacitance to the drain and source, S Parameters were
taken on an FET electrode geometry without any CNTs in the channel (W = 0). The
capacitance associated with that measurement should give the gate-drain capacitance.
Since the source electrodes cannot be probed in a GSG configuration, their
capacitance cannot be measured directly. However, the electrode proximity to the
gate is similar for both the sources and the drain, so CGS ? CGD. In addition to device
measurements, a simulation of the FET geometry without CNTs was made with
Microwave Office, which produces theoretical scattering parameters. The resulting
capacitances from both the VNA measurements and the Microwave Office
Simulation are plotted in Figure D.2.
169
Figure D.2 Gate-drain capacitance (CGD) plotted as a function of frequency extracted from
scattering parameters taken on an FET electrode geometry without CNTs in the channel (W =
0, L = 10 ?m, G = 8 ?m). The same capacitance from a Microwave Office simulation using
the same device geometry is plotted in red.
The theory and the data plots overlap, giving a gate-drain capacitance near 30
fF. When referenced in this dissertation, the gate capacitance values used are CGCNT =
100 fF, CGS = CGD = 30 fF.
170
Appendix E: Preisach Hysteresis Model Matlab Script
clear all; close all;
n=100; % n is the of vertical divisions
N=n*(n+1)/2; % N is the total number of hysterons
h=2/(n+1); % h is the separation between
hysterons
a0=-1+h/2; % a0 is the first alpha value
b0=-1+3*h/2; % bo is the first beta value
% The locations of each hysteron (a, b) are filled in.
i=1;
for i1=1:n
for i2=1:i1
a(i,1)=a0 + (i2-1)*h;
b(i,1)=b0;
i=i+1;
end
b0=b0+h;
end
% The initial default hysteron occupation (R0) is set by
sweeping waveform D.
Osc=100;
for i1=1:Osc+1
VgsMax(i1,1)=1-(i1-1)/Osc;
end
for i1=1:Osc
x=VgsMax(i1,1);
for i2=1:N
if x>b(i2,1);
R0(i2,1)=1;
end
if xb(i2,1);
R0(i2,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if 0b(i1);
R(i1,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if 0b(i1);
R(i1,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if 0b(i1);
R(i1,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if 0b(i1);
R(i1,1)=1;
end
if -xb(i1);
R(i1,1)=1;
end
if 0b(i1);
R(i1,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if 0b(i1);
R(i1,1)=1;
end
if -xb(i1);
R(i1,1)=1;
end
if 0b(i1);
R(i1,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if 0b(i1);
R(i1,1)=1;
end
if -xb(i1);
R(i1,1)=1;
end
if 0b(i1);
R(i1,1)=1;
end
if xb(i2,1);
R(i2,1)=1;
end
if xb(i2,1);
R(i2,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if xb(i2,1);
R(i2,1)=1;
end
if xb(i2,1);
R(i2,1)=1;
end
if xb(i1);
R(i1,1)=1;
end
if xb(i2,1);
R(i2,1)=1;
end
if xb(i2,1);
R(i2,1)=1;
end
if x