105
Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating by Yao Ma A thesis submitted in conformity with the requirements for the degree of Master of Applied Science Graduate Department of Electrical and Computer Engineering University of Toronto © Copyright by Yao Ma 2012

Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating

by

Yao Ma

A thesis submitted in conformity with the requirements for the degree of Master of Applied Science

Graduate Department of Electrical and Computer Engineering University of Toronto

© Copyright by Yao Ma 2012

Page 2: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

ii

Abstract

Quantum Transport in InAs Nanowires with Etched Constrictions and

Local Side-gating

Yao Ma

Master of Applied Science

Graduate Department of Electrical and Computer Engineering University of Toronto

2012

To study transport properties in single InAs nanowires (NW) with etched constrictions, a bunch

of back-gated single InAs NW devices were made. The standard device contained a NW section

with an etched constriction, placed between two pre-patterned side-gates. For comparison,

devices either without etched constriction or without side-gates were also fabricated.

Transport measurement results of three devices were presented and discussed. The device

without side-gates exhibited Coulomb blockade due to electron tunneling through double

quantum dots (QDs). The device without the etched constriction displayed conductance

quantization. The standard device showed both Coulomb blockade (due to electron tunneling

through either multiple QDs or single QD) and Fabry-Perot conductance oscillation at different

gate bias regime.

A 3-D electrostatic and 2-D eigenvalue coupled simulation was conducted to explain the

observed conductance quantization. This model suggests that the nonuniform potential

distribution in a thick NW dramatically modifies the confinement energies in the NW.

Page 3: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

iii

Acknowledgments

I begin by sincerely thanking my thesis supervisor Prof. Harry Ruda, Director of the Center for

Advanced Nanotechnology (CAN), first of all for his support and encouragement throughout

completion of the research activities. Prof. Ruda’s exceptional scientific creativity and

knowledge shine through in the feedback he gives his graduate students. He guided me through

all kinds of difficulties and always made sure I was on the right track. I cannot express more of

my gratitude to him.

I reserve special and most emphatic thanks to Dr. Joe Salfi, a former PhD student of CAN, for

teaching me all the necessary experimental skills I needed and inspiring my research curiosity in

every way he could.

I thank Dr. Selva Nair, a senior scientist of CAN, for fruitful discussions on the theoretical

modeling which is presented in Chapter 4. He is always kind, patient and ready to help me. I

learned so much from him.

I am very grateful to Dr. Fangmin Zhang, one of my best friends, for her help in some of the data

processing as well as her encouragement and company during the past three years.

I also want to thank Michael Chen, a master student of CAN. He helped me finish fabricating

and measuring the sample discussed in Chapter 4 during my absence.

To my committee consisting of Prof. Ruda, Prof. Kherani, Prof. Helmy and Prof. Ng, I reserve

special thanks for them to spend time reading my thesis and examine the thesis defence.

Finally, to my parents and my wife, I express my sincere thanks for their support and

encouragement.

Page 4: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

iv

Contents

Chapter 1 Introduction ............................................................................................................. 1

1.1 Semiconductor Nanowires as Building Blocks for Future Electronic Devices .................... 1

1.2 Background Theory .............................................................................................................. 3

1.2.1 Single Electron Tunneling in Semiconductor Quantum Dots........................................ 3

1.2.2 Ballistic Transport in Quantum Wires ......................................................................... 10

1.2.3 Fabry-Perot Electron Interference in Electron Waveguides ........................................ 19

1.2.4 Surfaces and Interfaces in Semiconductor Nanowire Devices .................................... 20

1.3 Motivations, Objectives and Thesis Outline ....................................................................... 22

Chapter 2 Methods .................................................................................................................. 24

2.1 Single InAs Nanowire Device Fabrication ......................................................................... 24

2.2 Experimental Setup for Transport Measurements .............................................................. 29

Chapter 3 Quantum Transport Phenomena in Single InAs Nanowires with Etched

Constrictions ................................................................................................................................ 33

3.1 Introduction ......................................................................................................................... 33

3.2 Side-gating Effect: Hysteresis ............................................................................................. 34

3.3 Coulomb Blockade in an Etched InAs Nanowire without Side-gating .............................. 36

3.4 Transport Phenomena in an InAs Nanowire with Single Etched Constriction and Side-

gating ......................................................................................................................................... 40

3.4.1 Interaction between Side-gate and Back-gate .............................................................. 41

3.4.2 Transport Phenomena in the Short-etched Section — Low Back-gate Bias ............... 43

3.4.3 Transport Phenomena in the Short-etched Section — High Back-gate Bias ............... 48

3.4.4 Transport Phenomena in the Long-etched Section ...................................................... 51

Page 5: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

v

3.5 Conclusions ......................................................................................................................... 52

Chapter 4 Conductance Quantization in a Thick InAs Nanowire with Side-gating ......... 53

4.1 Introduction ......................................................................................................................... 53

4.2 Data Processing ................................................................................................................... 55

4.3 Evidence of Subband Quantization ..................................................................................... 58

4.3.1 Linear and Nonlinear Transport at Low Temperature ................................................. 58

4.3.2 Linear Transport in a Magnetic Field .......................................................................... 60

4.3.3 Linear Transport at Higher Temperatures .................................................................... 62

4.4 Three-dimensional Electrostatic and Two-dimensional Eigenvalue Coupled Modeling of

the Nanowire System ................................................................................................................ 64

4.4.1 Model Description ....................................................................................................... 65

4.4.2 Results of the Model .................................................................................................... 68

4.5 Validation of the Model ...................................................................................................... 75

4.6 Conclusions ......................................................................................................................... 79

Chapter 5 Conclusions and Future Outlook ......................................................................... 81

Bibliography ................................................................................................................................ 83

Page 6: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

vi

List of Tables

Table 3.1: Measured gate leakage current at different gate biases. 1st and 2nd number in

each cell represent Ibg and Isg, respectively, in unit of nA. ............................................... 42

Table 3.2: Key parameters of the QD extracted from the labeled Coulomb diamonds in

Figure 3.6b. .......................................................................................................................... 45

Table 3.3: Extracted lever arms for the three Coulomb peaks marked with arrows in Figure

3.6a, using Eq. (1.5) and assuming the electron temperature to be around 15 K. ............... 46

Table 4.1: Comparison between the calculated and observed subband spacing. The 1st and

2nd column list calculated subband spacing between two successive subbands located

above and below the Fermi level at the onset and end of each plateau, respectively,

based on our coupled simulation. The 3rd column is the extracted average subband

spacing for each plateau from the transconductance diamonds in the stability diagram.

The 4th column is the calculated subband spacing of the “particle in a cylinder”

model. .................................................................................................................................. 73

Page 7: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

vii

List of Figures

Figure 1.1: Schematic diagram of the QD system based on CI model. The QD is coupled to

source and drain through two tunnel barriers and its energy structure is tuned by a

nearby gate. Note that tunnel barriers are characterized by a tunnel resistor and a

capacitor connected in parallel, as indicated in the inset. ...................................................... 4

Figure 1.2: QD in the linear transport regime. (a) If no level in the dot lines up with the

chemical potential in the leads, the electron number is fixed at N-1 due to Coulomb

blockade. (b) The level is in resonance with the chemical potential in the leads, so

the number of electrons can alternate between N-1 and N, resulting in a single

electron current. (c) Calculated Coulomb peak lineshape in the liner response regime

with experimental parameters. ............................................................................................... 6

Figure 1.3: QD in the nonlinear transport regime. (a) Stability diagram of a few-electron

QD with fixed number of electrons on the dot and physics (Coulomb blockade (CB),

single electron tunneling (SET) and double electron tunneling (DET)) labeled in the

corresponding regions. The red line represents the linear transport regime ( 0).

The dashed lines indicate where the current changes due to the onset of transitions

involving excited states. (b) – (e): Energy level alignment diagrams for four points

labeled as gray square, black circle, black square and gray circle, respectively. Top

and bottom black lines represent chemical potentials for the 2nd GS(2) and 3rd

GS(3)electron, thick and thin blue lines represent exited states for the 1st ES(1)and

2nd ES(2) electron. ................................................................................................................. 8

Page 8: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

viii

Figure 1.4: Schematic diagram of two QDs connected in series coupled to each other

through a tunnel barrier. A common gate bias is used to tune the energy levels of the

two dots at the same time. ..................................................................................................... 9

Figure 1.5: 1-D subbands and electron transmission. (a) Dispersion relation in the 1-D

channel. States with negative propagate from right to left and are fed from the

right reservoir with chemical potential and vice versa. The energy interval is

given by the bias V applied between the two reservoirs. Here V < 0, so a net current

flows from left to right. (b) Schematic diagram of the adiabatic limit and the abrupt

limit for the interfaces between the reservoirs and the 1-D channel with the

dimension of the channel and the illustrative energy spectrum of the system labeled.

(c) The single-mode transmission probabilities and total transmission probabilities as

a function of 0 for = 3 based on the saddle-point model, showing T

as a step-like function of the energies in the channel. ......................................................... 13

Figure 1.6: (a) Classical motion of a particle in a 1-D conductor with a magnetic field

perpendicular to the transport direction for different energies and center coordinates,

showing the formation of pure cyclotron orbit (red circle) and skipping orbit (solid

blue curve). While the impurity may momentarily disrupt the forward propagation of

the electron, it is ultimately restored as a consequence of the strong Lorentz force.

Thus the skipping orbit in a conductor at high magnetic fields can suppress

backscattering. (b) Schematic Zeeman splitting of spin-degenerate magneto-subbands

in a magnetic field. .............................................................................................................. 16

Figure 1.7: (a) Stability diagram of 1-D system. Diamond patterns represent conductance

plateaus with labeled conductance values in unit of 2e2/h. Blue arrow denotes a type

of intersection whose corresponding bias levels equal the subband spacing. Inset: The

Page 9: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

ix

spatial distribution of the chemical potential along the 1-D system at finite source and

drain bias. (b) - (d) Energy landscapes for points labeled as black, gray and empty

square, respectively, illustrating the meaning of left/right-tilted stripes and the

evolution of the integer-half-integer plateau behavior expected as the bias is

increased. The half plateau appear when the number of conducting subbands for the

two directions of transport differs by 1 and the integer plateau is retrieved when the

number differs by 2. ............................................................................................................ 17

Figure 2.1: Main steps of device fabrication. (a) 1st exposure and metallization transferred

patterns of the source (S)/drain (D) contacts, side-gate (SG) electrodes and local

markers onto the substrate with respect to the NW. (b) 2nd exposure opened an

etching window for the etchant, which was aligned to previous metal patterns via

local markers. Using resist as the etching mask, the NW section exposed to the

etchant was etched to a desired depth. (c) 3rd exposure and 2nd metallization

transferred patterns of bonding pads and traces (yellow) connecting previous metal

patterns (orange) onto the substrate. Previous metal patterns are barely seen at this

scale. (a) and (b) are actually enlarge views of the dark blue square region. ...................... 28

Figure 2.2: Experimental setup. (a) Custom built sample holder. A top cooper plate presses

down the LCC towards the spring loaded probes mounted inside the G10 fixture,

providing both thermal anchor and electrical contacts for the sample. (b) Cryostat and

electromagnetic system. The sample holder is located in the extender of the cryostat

which can be inserted into the gap of the two magnets and rotated by 90 °. The

magnetic field is controlled by a programmed power supply (Agilent E3631A). (c)

Circuit layout for differential conductance measurement. AC and DC source/drain

biases are coupled by a transformer, attenuated by a divider and then applied onto the

Page 10: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

x

drain contact of the device. Current from the source contact is amplified by a current

preamp and measured by a lock-in. Back/side-gate biases are applied through

protecting resistors to the rear of the sample and side-gate electrodes, respectively. ......... 31

Figure 3.1: (a) Transfer characteristics of a side-gated InAs NWFET, showing distinct

hysteresis at room temperature (black curve) and reduced hysteresis at lower

temperatures (red and blue curves). Green arrows indicate the sweep directions of

side-gate voltage. (b) Transfer characteristics of the same device gated by the back-

gate at room temperature showing little hysteresis. (c) Schematic of the proposed

oxide trap charging mechanism for side-gating. Here the NW is grounded through

source and drain contacts and a positive bias is applied to the side-gate. When the

bias reaches a certain value, the electric field between the gate electrode and the NW

may be high enough to induce electron tunneling from the oxide traps to gate

electrode. Corresponding energy diagram is also shown to illustrate the tunneling

situation. .............................................................................................................................. 35

Figure 3.2: (a) Linear conductance in the pinch-off region of the transfer characteristics of

the device. Three peaks are visible. The second peak is superimposed with RTS. The

third peak denotes the onset of the channel. Inset (I): SEM image of the device with

the source and drain contacts labeled. Inset (II): Transfer characteristics of the device.

The pinch-off region is highlighted with red rectangle. (b) Stability diagram

corresponding to the same back-gate voltage range in (a). Large diamonds reflect the

charge state of the smaller dot while the fast oscillations superimposed on the

boundaries of large diamonds are due to the charging of the larger dot. ............................ 38

Figure 3.3: (a) SEM image of the device studied in this section. Two etched constrictions

were fabricated on the same NW to examine the effect of etching length on transport

Page 11: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

xi

properties. (b) Linear conductance as a function of side-gate and back-gate biases for

the short-etched constriction. The slope of stripes in region I and III is different from

the one in region II. Detailed bias spectroscopy was performed along marked arrows

except for the light green one. ............................................................................................. 41

Figure 3.4: (a) Schematic of the proposed model to explain observed change of relative

coupling between back/side-gate and the channel. Leakage paths from source and/or

drain contacts and side-gate electrodes to the back-gate were probably created during

the wire bonding which can be characterized by variable resistors. (b) Measured gate

leakage current at different gate voltages with side-gate connected to back-gate. ............. 43

Figure 3.5: (a) Linear conductance response along the orange arrow in Figure 3.3b. The

pattern looks very “noisy”, with random switching superimposed onto the

conductance trace. (b) Corresponding bias spectroscopy, showing irregular diamond

patterns that are not closed in the pinch-off region. ............................................................ 44

Figure 3.6: (a) Linear conductance response along the yellow arrow in Figure 3.3b. (b)

Corresponding bias spectroscopy. Three Coulomb diamonds labeled with numbers

can be clearly identified with well defined boundaries. ...................................................... 46

Figure 3.7: Enlarged view of the three Coulomb peaks marked with arrows in Figure 3.6a

(gray curves) and the best fit to theses peaks (black curves). ............................................. 47

Figure 3.8: Linear conductance response of the device at Vbg =2.7 V, corresponding to the

light green arrow in Figure 3.3b. Inset: Linear conductance response of the device at

Vbg =3 V, corresponding to the purple arrow in Figure 3.3b. ............................................. 49

Figure 3.9: Linear conductance in the pinch-off region (a) and the following overshooting

region (b) at Vbg = 3 V (along the purple arrow in Figure 3.3b). (c) Linear

conductance on the first plateau at Vbg = 2.7 V (along the red arrow in Figure 3.3b).

Page 12: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

xii

(d) - (f) Bias spectroscopy corresponding to the linear conductance in (a) - (c). The

grayscale bars are in the unit of 2e2/h. ................................................................................. 50

Figure 3.10: A representative measurement result taken at Vbg = 2V for the long-etched

section. (a) Linear conductance in the pinch-off region. (b) Linear conductance in the

following sub-threshold region. (c) and (d) Bias spectroscopy corresponding to linear

conductance in (a) and (b). Scale bars are in the unit of 2e2/h. ........................................... 51

Figure 4.1: (a) SEM image of the device. This chapter focuses on the measurements on the

section without etching. (b) Gray scale profile along the yellow dashed line providing

a more accurate measure for the geometry dimensions (c) Linear conductance as a

function of side-gate and back-gate biases. Stripes with almost constant slopes

indicate independent coupling to the channel from the two gates. Experiment results

along the two arrows are presented in this chapter. ............................................................. 54

Figure 4.2: Raw data of the finite bias spectroscopy and associated data processing. (a)

Raw data of G-VDS traces (only the positive scan of VDS, i.e., from -30 mV to 30 mV

is plotted) for different Vg. Vg was increased from -8 V to 8 V stepped by 0.05 V

between successive VDS sweeps. (b) G-VDS traces after the subtraction of “self-gating”

effect and compensation for the dip near zero source and drain bias based on the raw

data in (a). (c) G-VDS traces of testing resistors, demonstrating the irrelevance

between the dip feature and the sample conductance. (d) Correction for the dip

feature. By multiplying the conductance in the dip region with the calculated average

peak-to valley ratio, the dip is compensated to an average conductance level outside

the dip. ................................................................................................................................. 57

Figure 4.3: (a) Linear conductance as a function of gate voltage. 11 conductance plateaus

can be identified with the middle 5 ones marked by arrows. Inset: A zoom-in of the

Page 13: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

xiii

2nd plateau, showing the small oscillations superimposed on the plateau. (b) Gray

scale plot of the normalized absolute value of transconductance as a function of gate

voltage and source/drain bias. Diamond patterns around zero bias corresponding to

linear conductance plateaus in (a) are visible. (c) Gate voltage sweeps of G for fixed

values of VDS from 0 to 30 mV, stepped by 3 mV. Black circles highlight half

plateaus. ............................................................................................................................... 59

Figure 4.4: Magnetotransport characters of the device. Differential conductance was

measured as function of side-gate bias without magnetic field (black curve) and with

a magnetic field of 0.6 T which is perpendicular to the sample plane (red curve). Four

plateaus labeled with numbers can be identified. Compared to the case where B = 0 T,

at B = 0.6 T, conductance on each plateau increases and there is a half plateau

appearing between the 3rd and the 4th plateaus possibly due to the Zeeman splitting of

a spin-degenerate level, as indicated by the blue ellipse. The distance in terms of Vsg

between the split plateaus and the original plateau is marked by the green dashed

lines. ..................................................................................................................................... 61

Figure 4.5: Liner transport at higher temperatures. G-Vg curves corresponding to T = 20 K

and 30 K are shifted for clarity. Most conductance plateaus are washed out at T = 30

K, providing a rough estimation for the subband spacing. Inset: G-Vg curve at room

temperature, from which the field-effect mobility can be extracted. .................................. 63

Figure 4.6: Screenshot of the 3-D electrostatic simulation result, showing the simulated

geometry and the potential distribution with source and drain grounded and Vbg = Vsg

= -7 V. .................................................................................................................................. 67

Figure 4.7: Simulated number of occupied subbands as a function of gate bias. Red curve:

result from the coupled simulation carried out by the author. Labeled numbers

Page 14: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

xiv

represent the resembled conductance plateaus. Black curve: result from the approach

adopted by Fort et al [Fort 2012]. ....................................................................................... 70

Figure 4.8: Calculated probability density profiles at Vg = -6 V for the five lowest lateral

modes with corresponding energy levels with respect to conduction band edge

labeled. No degeneracy is observed and the wave functions are pushed up and

sideways by back and side-gate bias, indicating the effect of the nonuniform potential

distribution inside a thick NW at large negative gate bias. ................................................. 71

Figure 4.9: (a) Calculated gate capacitance per unit length as a function of gate bias

from the 3-D electrostatic simulation. Also shown is the back-gate capacitance

calculated using Eq. (4.2) for the same device. (b) Calculated quantum capacitance

per unit length as a function of gate bias using a combination of Eq. (4.8) and results

from the 2-D eigenvalue simulation. ................................................................................... 74

Figure 4.10: Calculated potential distribution profiles inside the NW at the middle slice (x

= 0 nm) and the slice 100 nm left to it (x = -100 nm), at two gate biases (Vg = -2 and -

6 V). ..................................................................................................................................... 76

Figure 4.11: (a) Average electrostatic energy distribution along the NW. From top to

bottom: distributions at Vg = -8 ~ 8 V stepped by 2 V. (b) Lowest subband level

distribution along the NW with fitted Fermi level labeled. From top to bottom:

distributions at Vg = -8 ~ 0 V stepped by 2 V. .................................................................... 77

Figure 4.12: Calculated 1-D carrier concentrations as a function of gate bias based on the

3-D Thomas-Fermi scenario (squares) and the 2-D Schrodinger equation at T = 10 K

(circles) and T = 0 K (triangles). ......................................................................................... 79

Page 15: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

1

Chapter 1

Introduction

1.1 Semiconductor Nanowires as Building Blocks for Future

Electronic Devices

Electronic devices form the basic constituents of modern integrated circuits. They come in many

varieties and are used to process analog signals, perform logic operations and store information

in digital memories. Apart from these purely electronic functions, other devices interact with

light, or sense changes in the environment. Semiconductor nanowires (NWs) offer many

opportunities for the assembly of nanoscale devices by the “bottom-up” paradigm [1] and have

been proposed and intensively studied as a realistic add-on to mainstream Si technology [2]. One

key factor motivating the boom of NW research is that, compared with device structure

fabricated by the “top-down” approach, “bottom-up” synthesized NW materials offer well-

controlled size in at least on critical device dimension, channel width, that is at or beyond the

limits of top-down lithography [1]. In addition, the crystalline structure, smooth surfaces and the

ability to produce radial and axial heterostructures can reduce scattering and result in higher

carrier mobility, meanwhile, offer exciting new possibilities for device architecture [3].

Thus far, all the above mentioned functions of electronic devices have been demonstrated using

semiconductor NWs [1, 4, 5, 6]. To keep the scope of this introduction manageable, in the

following part of this section, we only focus on NW devices and applications that are relevant to

this thesis and restrict ourselves within the content of electronics.

Semiconductor NWs are quasi-one-dimensional (quasi-1-D) systems by nature. In

semiconducting 1-D structures (also referred to as quantum wires), carriers are confined in two

Page 16: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

2

dimensions. One of the representative 1-D systems is quantum point contact (QPC), a narrow

constriction between two wide electrically conducting regions, of a width comparable to the

electronic wavelength [7]. The hallmark of QPC is its quantized conductance in unit of e2/h per

electronic subband. Since the conductance of QPC strongly depends on the confinement

potential of the constriction, any potential fluctuation in the vicinity will influence the current

through the QPC. Therefore, QPCs can be used as extremely sensitive charge detectors. In view

of quantum computation in solid-state systems, QPCs may be used as readout devices for the

state of a quantum bit (qubit) [8]. In Section 1.2.2 we will see that conductance quantization is a

universal characteristic of 1-D systems, provided that the transport is ballistic (scattering-free) or

quasi-ballistic. Thus we expect semiconductor NWs to have similar behavior [9, 10] and

applications.

An emerging frontier in condensed matter research is the coherent, optics-like manipulation of

electrons for information processing and storage. In high quality 1-D or quasi-1-D systems, such

phase coherent transport over hundreds of nanometers has been observed [11, 12]. Interferometer,

a staple of coherent optics, has also been realized in these systems. For instance, Fabry-Perot (F-

P) electron/hole interference has been observed in carbon nanotubes [13, 14, 15, 16] and more

recently, in semiconductor NWs [17, 18], which underscores the potential of NWs for on-chip

realization of electron optics [19].

Quantum dots (QDs) are zero-dimensional (0-D) systems in which the motion of carriers is

confined to very small “box-like” volumes. QDs are considered as promising systems for

implementation of spin “qubits” in proposed schemes for quantum computation [20]. The

cylindrical morphology makes NWs natural platform for fabricating QDs: the cylindrical

geometry has already provided lateral confinement for carriers, thus one only needs to create two

longitudinal barriers to form a QD. What makes NWs more attractive to host QDs is that linear

arrays of coupled QDs can be easily made along the axis of a NW if multiple longitudinal

barriers are created. At the ends of the arrays, the QDs can be coupled to 1-D, quasi 1-D or bulk-

like reservoirs, depending on the lateral dimensions of the hosting NWs [21].

Page 17: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

3

1.2 Background Theory

1.2.1 Single Electron Tunneling in Semiconductor Quantum Dots

Among the different QD systems that have been studied, the type that is relevant to this research

involves a semiconducting island coupled to source and drain contacts by two tunnelling barriers.

When the electron tunneling to and from the dot is sufficiently weak, the number of electrons on

the dot will be a well defined integer N. To increase this number by 1, electrons in the reservoirs

must have sufficient energy to overcome the mutual electrostatic Coulomb repulsion among

electrons in the QD plus the quantization energy due to the confinement potential within the QD.

If this condition is satisfied, current flow will occur through sequential tunnelling events of

single electrons through the dot, causing N to fluctuate by one; otherwise the QD is said to

operate in Coulomb blockade regime and no current can flow through the QD. To gain better

control over the QD system, a gate electrode is normally added to electrostatically tune the

energy in the QD. In the following section, a widely used model based on the above device setup

will be introduced and its implications for the lineshape of the Coulomb peaks and the stability

diagram will be discussed. Finally the model will be extended to the case of DQD.

Constant Interaction Model

The constant interaction (CI) model [22] provides an approximate description of the electronic

structure of the QD system. It is based on two assumptions. First, the Coulomb interactions

among electrons in the dot, and between electrons in the dot and those in the environment, are

parameterized by a single, constant capacitance C, which is the sum of the capacitances between

the dot and the source , the drain and the gate . Second, the single-particle energy levels

are independent of the electron-electron interactions. A circuit representation of the QD

system based on above descriptions is depicted in Figure 1.1.

Page 18: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

4

Figure 1.1: Schematic diagram of the QD system based on CI model. The QD is coupled to source and drain through

two tunnel barriers and its energy structure is tuned by a nearby gate. Note that tunnel barriers are characterized by a

tunnel resistor and a capacitor connected in parallel, as indicated in the inset.

Under these assumptions, the total energy of a dot with N electrons in ground state, with

voltages , and applied to the source, drain and gate, respectively, is given by

| | 2⁄ ∑ (1.1)

where | | represents both the unintentional background polarization charge due to

workfunction differences and random charges trapped near the junctions.

The chemical potential of the dot is defined as

N 1 | | (1.2)

where ⁄ is the charging energy. When N is fixed, a change in gate voltage ∆ induces

a change in the chemical potential by | |

∆ | | ⁄ ∆ . The ratio ⁄ is

called lever arm.

The energy required to realize the transition between successive ground states is the so called

addition energy:

1 ∆ (1.3)

QD

RD,CS

Vbg

Drain Source

RD,CD

CG

VDS

R

C

Page 19: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

5

where ∆ is the adjacent single-particle level spacing. To resolve well defined

Coulomb blockade effects, the Coulomb gap, , must greatly exceed the thermal energy:

. A further constrain is that the quantum fluctuations in the electron number, N,

are so small that the electron is well localized on the dot. A hand-waving argument is to consider

the energy uncertainty relationship ∆ ∆ /2, where ∆ ~ and the time to transfer

electron into and out of the island given by ∆ where is the smaller of the two tunnel

resistances [23]. Combining these two expressions together gives . Neglecting

∆ for relatively large QDs, we have , so that = 25.8 kΩ, i.e., the tunnel

resistance should be greater than the quantum resistance of 25.8 kΩ in order to see clear

Coulomb charging effects.

Linear Transport Regime

When the source and drain bias is small ( , ∆ , ), the QD system is operated

in the linear transport regime, where current can only flow when the chemical potential of the dot

lines up within a few of the chemical potential in the leads by tuning the gate bias. This

gives rise to a series of conductance oscillations as a function of gate bias, as illustrated in Figure

1.2a-b. Equating the induced chemical potential change in the dot by gate bias with Eq. (1.3), we

obtain the gate period corresponding to two successive conductance peaks:

∆ (1.4)

As long as , the width of these peaks will essentially be limited by thermal

broadening and therefore smear out at temperatures greater than the energy width of the

Coulomb blockade regime. The conductance linewidth has been calculated in the classical limit

of ∆ by Beenakker as [24]

. (1.5)

Page 20: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

6

where G is the height of the Nth peak. Results using Eq. (1.5) for = 1meV, = 0.25 and

= 3.2aF, which are similar to the extracted device parameters from experiment, are shown in

Figure 1.2c, illustrating the expected lineshape.

Figure 1.2: QD in the linear transport regime. (a) If no level in the dot lines up with the chemical potential in the

leads, the electron number is fixed at N-1 due to Coulomb blockade. (b) The level is in resonance with the

chemical potential in the leads, so the number of electrons can alternate between N-1 and N, resulting in a single

electron current. (c) Calculated Coulomb peak lineshape in the liner response regime with experimental parameters.

Nonlinear Transport Regime

By increasing , multiple dot levels can participate in electron tunneling, which is the situation

of nonlinear transport. Typically the chemical potential of only one reservoir is changed in

experiments and the other one is kept fixed [8]. Here we set the source to be grounded, i.e., = 0.

As changes, the levels of the dot also changes due to the capacitive coupling between the dot

and the drain. A powerful tool, called the stability diagram, which is a plot of differential

0 50 100 150 200 250 3000.0

0.2

0.4

0.6

0.8

1.0

1.2

G/G

max

VG(mV)

N N+1N-1

(a)

(c)

(b)

Page 21: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

7

conductance ⁄ as a function of and , is widely used to investigate the nonlinear

transport characteristics of the QD system. A schematic example of such a diagram for a few-

electron QD is shown in Figure 1.3a. The red line denotes the linear response discussed earlier,

where cross points represent the Coulomb peaks. For each transition point, at either sign of ,

a V-shaped region is outlined. For instance, the point labeled by empty square represents the

alignment of 2 with and . Starting from this point, when 0, two lines labeled

with L and R form edges of the V shape, which denote the alignment of 2 with drain and

source, respectively, as depicted in Figure 1.3b and d. To obtain the slopes of these lines,

consider the change of chemical potential of the dot induced by the change of and based

on Eq. (1.2). On the line labeled with R, ∆ 2 induced by ∆ equals to ∆ 2 induced by ∆ ,

thus we have | | ∆ | | ∆ ; On L, ∆ equals to the summation of ∆ 2 induced by

∆ and ∆ , which gives | |∆ | | ∆ | | ∆ . Solving for these two equations, and

note that ∆ ∆ since = 0, we obtain

∆∆

, ∆∆

(1.6)

Within the V shapes, chemical potentials and associated excited states of the QD fall into the bias

window and single electron tunneling happens. The overlapping area constituted by the outside

of all V shapes forms diamond-like Coulomb blockade regions where the number of electrons on

the QD is fixed at the labelled value. When is further increased such that the bias window is

equal or greater than the addition energy (see Figure 1.3e), the electron number can alternate

between N-1, N and N+1, leading to double electron tunneling. We can imagine that increasing

even more would eventually result in triple electron tunneling, quadruple electron tunneling,

etc.

We can also use the stability diagram to find the energies of the excited states. Where a line of a

transition involving one excited state touches the Coulomb blockade region, the bias window

exactly equals the energy level spacing. Figure 1.3b-d shows the level diagrams for 1

2 , 1 2 and 2 2 , where and represent the ground

state and the excited state for N electrons in the QD.

Page 22: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

8

Figure 1.3: QD in the nonlinear transport regime. (a) Stability diagram of a few-electron QD with fixed number of

electrons on the dot and physics (Coulomb blockade (CB), single electron tunneling (SET) and double electron

tunneling (DET)) labeled in the corresponding regions. The red line represents the linear transport regime ( 0).

The dashed lines indicate where the current changes due to the onset of transitions involving excited states. (b) – (e):

Energy level alignment diagrams for four points labeled as gray square, black circle, black square and gray circle,

respectively. Top and bottom black lines represent chemical potentials for the 2nd GS(2) and 3rd GS(3)electron, thick

and thin blue lines represent exited states for the 1st ES(1)and 2nd ES(2) electron.

Classical Theory of Electrostatically Coupled Double Quantum Dots

Here we consider two QDs connected in series, coupled to each other by a tunnel barrier

represented by a tunnel resistor and a capacitor in parallel. Each dot is capacitively

coupled to a gate bias through a capacitor , and to the source or drain through a tunnel

barrier represented by a tunnel resistor and a capacitor in parallel. The source and

drain bias is applied to the drain, with the source grounded. The cross coupling between

(b) (c) (d) (e)

(a)

V DS

VG

01 2

3 4 5CB CBCB CB CB

SET SETSETSETSET

DET DET DET

DET DET

TET TET

V DS <

0

LR

Page 23: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

9

source and dot 2 ( ) and between drain and dot 1 ( ) are also included. A schematic diagram

of the described DQD system is shown in Figure 1.4.

Figure 1.4: Schematic diagram of two QDs connected in series coupled to each other through a tunnel barrier. A

common gate bias is used to tune the energy levels of the two dots at the same time.

We ignore the influence of discrete quantum states. If stray capacitances are negligible, the

double dot electrostatic energy in the linear transport regime reads [25]

, + (1.7a)

| |

(1.7b)

where is the number of electrons on dot 1(2), is the charging energy of the

individual dot 1(2), is the electrostatic coupling energy, which is the change in the energy of

one dot when an electron is added to the other dot. These energies are expressed in terms of the

capacitances as follows:

EC CC C

C C CM, EC C

C CC C CM

, ECM CM

C CC C CM

(1.8)

where is the total capacitance of dot 1(2):

Similar as the single QD case described in CI model, the chemical potential , of dot

1(2) is defined as

QD1 QD2

RM, CM

VG

source drain

RS, CS1 RD, CD2

CG1 CG2

VDS

CS2 CD1

Page 24: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

10

, , 1,

| |

(1.9a)

, , , 1

| |

(1.9b)

In the case of nonlinear transport, assume the bias voltage is applied to the drain and the source

is grounded, then the bias voltage is coupled to the DQD through the capacitance of the drain and

hence also affects the electrostatic energy of the system. The bias dependence can be accounted

for by replacing CG1(2)VG with CG1(2)VG +CD1(2)VDS in Eq. (1.7) and Eq. (1.9).

1.2.2 Ballistic Transport in Quantum Wires

In this section we consider a simple case where a 1-D conductor is connected to the left and right

electron reservoirs with chemical potentials and . The current carried by N occupied 1-D

modes is the sum of right and left-moving current due to the electron flows from the left and

right reservoirs, and is given by [23]

| | ∑ ∑ , , (1.10)

where is the transmission probability of an incoming wave in mode i on one side, to be

transmitted into mode j on the other side. , is the quasi-equilibrium distribution

function of the left (right) contact. When , a net current can flow through the system

because the distribution functions for left and right-moving electrons will differ, as is depicted in

Figure 1.5a.

In the linear transport regime where , , ,

⁄ , combined with Eq. (1.10) we obtain the linear conductance:

∑ ∑ (1.11)

Page 25: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

11

We now decouple the effect of temperature and transmission probabilities to exam these two

effects separately.

Linear Transport in an Ideal 1-D Conductor

When T 0 K, the distribution function can be approximated as a step function, so Eq. (1.10)

becomes | | ∑ ∑ , which gives ∑ , where

∑ .

We first assume the 1-D conductor is a perfect channel, i.e., there is no backscattering due to

impurities in the channel and surface states/roughness. This assumption allows us to focus on the

effect of the shape of the conductor (see Figure 1.5b) on . We classify the 1-D system into

two types, long and short, based on the aspect ratio, L/W, where L and W are the channel length

and width, respectively. By examining how smooth the transition is from the 1-D channel to the

reservoirs, we can also distinguish the 1-D system with sharp transitions and smooth transitions.

To be more precise, a 1-D system with smooth transition is described by the “adiabatic model”

[26], which assumes the spatial variation of the potential in the transport direction is much

slower than in the transverse direction. The case of sharp transition limit was studied by Szafer

and Stone [27]. Both limits give similar . To appreciate how depends on , we adopt

the saddle point model [28] which assumes the electrostatic potential near the bottleneck of the

constriction has the form of , with the corresponding

subband levels , where i = 1, 2, 3, …, is the potential at the saddle,

and represent the curvatures of the potential. Thus

1 1 exp 1 1 exp 2 . Based on this

expression, transmission probabilities as a function of for = 3 are reproduced in Figure

1.5c, from which we can see that the transmission rises quickly from 0 as the subband level

moves toward the Fermi level of the channel, and approaches unity when the subband level is

below the Fermi level. The total transmission probability T, and also the conductance, is

therefore a step-like function of the energies of the channel. Experimentally, the energy levels of

Page 26: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

12

the channel can be tuned by a gate bias, thus we expect to observe conductance quantization with

the unit of as the gate bias is swept for a 1-D ballistic conductor. Note that even though the

electron transport in the channel is truly ballistic without experiencing any scattering, there still

exists a minimum resistance , which was interpreted by Imry [29] as a contact resistance

arising from the rearrangement of the current from being carried by a large number of modes in

the reservoirs to being carried by very few modes in the channel, as illustrated in Figure 1.5b.

Another important point here is that the conductance is controlled by the mode spacing at the

narrowest point of the channel. Thus either a QPC or a longer wire may exhibit conductance

quantization, provided that the electron transport remains ballistic.

Page 27: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

13

Figure 1.5: 1-D subbands and electron transmission. (a) Dispersion relation in the 1-D channel. States with negative

propagate from right to left and are fed from the right reservoir with chemical potential and vice versa. The

energy interval is given by the bias V applied between the two reservoirs. Here V < 0, so a net current flows from

left to right. (b) Schematic diagram of the adiabatic limit and the abrupt limit for the interfaces between the

reservoirs and the 1-D channel with the dimension of the channel and the illustrative energy spectrum of the system

labeled. (c) The single-mode transmission probabilities and total transmission probabilities as a function of for

= 3 based on the saddle-point model, showing T as a step-like function of the energies in the channel.

One difference between the two limits is that under the adiabatic approximation the off-diagonal

elements of the transmission matrix can be neglected, i.e., = 0 for , whereas an

abrupt transition may give rise to intersubband scattering by itself and disrupt the conductance

N+1

N

N‐1

EF

E1

E2

E3 LμRμ

eV

E

kZ

(a) (b)

(c)

zL

W

0 1 2 3 4 5 6 7 8 9 100

1

2

3

h

T3(E)T2(E)

T(E)

(E-V0)/ ωz

T1(E)

h hh

E3/ ωzE2/ ωzE1/ ωz

Page 28: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

14

quantization. Another difference lies in the response to the length of the channel. When the

channel is short the conductance steps are rounded due to tunneling through the channel for a

Fermi energy just below the next unoccupied level. This behavior has been observed in both

limits. However, as the channel becomes longer, sharp steps are found for a 1-D system in

adiabatic limit whereas resonances as a result of alternatively constructive and destructive

internal reflection within the channel appear as soon as the channel is long enough to damp out

the evanescent modes.

Effects of Disorder and Finite Temperature

Now we relax the previous assumption about a perfect channel and briefly discuss the influence

of disorder on the transmission properties of 1-D system. Backscattering by the potential

fluctuation due to impurities in the channel, surface roughness and surface states is normally

inevitable in real devices, which degrades the conductance below . It is important to realize

that in small nanostructures such as quantum wires the total scattering matrix through the

structure depends on the exact location of the impurities and boundary fluctuations. If such

inhomogeneous effects are important, the conductance in a sample with a slightly different

impurity configuration will be completely different. This behavior contracts from that of

diffusive transport over very long length scales, in which the contribution of many impurities

averages to a constant contribution to the conductance that depends only on their density but not

their exact location.

After we examine the role of transmission probabilities in determining the conductance at

ultralow temperatures, we now consider the influence of temperature on the conductance by

assuming the transmission coefficient of the ith mode is unity for energies greater than the

subband minimum and zero below. Then Eq. (1.11) becomes ∑

∑ ∑ ⁄ . This expression shares the identical feature

with the temperature dependence of Fermi-Dirac distribution function: the increase of

Page 29: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

15

temperature causes the increased rounding of the plateaus until they eventually disappear when

the separation between mode energies is comparable to, or smaller than a few .

Edge States and Zeeman Splitting in 1-D Magnetotransport

A magnetic field perpendicular to the transport plane may also have its effects on the quantized

conductance. Van Wees et al. [30] showed that the plateaus are flattened with increasing

accuracy to about as a function of increasing magnetic field. Additional plateaus are seen to

arise at high fields, but at half steps of .

The flattening of the plateaus with increasing field is taken as a sign of reduced backscattering

and increased transmission due to an increasing amount of current being carried by edge states

[23]. Edge states in a quantum wire with the presence of a magnetic field can be understood in

the following qualitative picture: Assume electrons can only propagate along z directions in a

quantum wire with a hard wall confinement. The magnetic field is applied along –y direction, as

depicted in Figure 1.6. From the quantum mechanical solution of the electron motion in the plane

perpendicular to the magnetic field, the quantized cyclotron radius is given by

2 , corresponding to the associated Landau levels

, n = 0, 1, 2,…where ⁄ is the magnetic length and ⁄ is the cyclotron

frequency. The solution also defines a parameter called center coordinate: / , which

represents the average position of the wavefunction and the center position of the circular

cyclotron orbit of radius . From the viewpoint of the classical motion of the particle in the

waveguide, if is less than the distance from the center coordinate to one wall or the other so

that no scattering of the particle off the wall occurs, pure cyclotron-type motion is possible, as

shown by the red circular orbits in Figure 1.6; if is greater than distance between the center

coordinate to either wall, boundary scattering occurs, giving rise to skipping orbits, as shown by

solid blue orbits in Figure 1.6. If boundary scattering is assumed to be specular, the reflected

momentum is reflected momentum is such that the electron has a net momentum in the plus or

minus z direction, depending whether (and thus ) is positive or negative. The states

Page 30: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

16

associated with these skipping orbits are referred to as edge states. As point out by Buttiker [28],

electrons propagating in edge states cannot be easily scattered over distances larger than the

magnetic length , because even if the electron is backscattered by an impurity, the resulting

Lorenz force exerted on the electron tends to push it back towards the boundary and so maintain

its skipping motion (Figure 1.6).

Figure 1.6: (a) Classical motion of a particle in a 1-D conductor with a magnetic field perpendicular to the transport

direction for different energies and center coordinates, showing the formation of pure cyclotron orbit (red circle) and

skipping orbit (solid blue curve). While the impurity may momentarily disrupt the forward propagation of the

electron, it is ultimately restored as a consequence of the strong Lorentz force. Thus the skipping orbit in a

conductor at high magnetic fields can suppress backscattering. (b) Schematic Zeeman splitting of spin-degenerate

magneto-subbands in a magnetic field.

The appearance of half integer plateaus at high fields can be straightforwardly explained by the

split of the two-fold spin-degenerate magneto-subbands by an amount of in a magnetic

field, forming distinct spin-up and spin-down states as depicted in Figure 1.6b. However, they

contribute only to the conductance, since the density of states is now reduced by a factor of

two.

Finite Bias Spectroscopy

Similar to the nonlinear transport in QD introduced in Section 1.2.1, employing a DC source and

drain bias as an energy reference gives access to the subband spacing, thus providing a powerful

tool (termed as finite bias spectroscopy) to further characterizes the 1-D system.

z

xB

impurity

(a) (b)EN

EN-1

Page 31: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

17

Picciotto et al. explained how the chemical potential in the channel shifts with the applied finite

bias as a result of satisfying charge neutrality condition [31, 32]. An important point here is that

the charge density in the wire is nearly independent of the bias. This arises because with a small

wire capacitance, a large potential energy penalty is associated with charge imbalance. Charge

neutrality is thus facilitated by a uniform shift of the potential in the wire, such that

/2 (see the inset of Figure 1.7a). This potential shift reflects the screening of the applied bias

by the free charges and is accomplished by the two potential steps at the inlet and outlet of the

wire, respectively.

Figure 1.7: (a) Stability diagram of 1-D system. Diamond patterns represent conductance plateaus with labeled

conductance values in unit of 2e2/h. Blue arrow denotes a type of intersection whose corresponding bias levels equal

the subband spacing. Inset: The spatial distribution of the chemical potential along the 1-D system at finite source

and drain bias. (b) - (d) Energy landscapes for points labeled as black, gray and empty square, respectively,

illustrating the meaning of left/right-tilted stripes and the evolution of the integer-half-integer plateau behavior

expected as the bias is increased. The half plateau appear when the number of conducting subbands for the two

directions of transport differs by 1 and the integer plateau is retrieved when the number differs by 2.

EN

EN+1EN+2(b)

(a)

EN

EN+1EN+2(c)

EN

EN+1EN+2

(d)

VD

S

VG

VD

S>0

N N+1N-1N-2

N

N

N+1/2

N+1/2

N-1/2

N-1/2

N+3/2N-3/2

N-3/2

N-1 N+1

N-1

Page 32: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

18

Glazman and Khaetskii [33] predicted that when | | , additional plateaus will appear in

the differential conductance. If the number of occupied subbands N is large, the additional

plateaus (half plateaus) are predicted to be midway between the plateaus given by . For

small N some deviations of the conductance from half-integer values could be expected. Patel et

al. [34] presented the first evidence for the appearance of the half plateaus. The concept of half

plateaus can be illustrated by sketches of energy landscapes shown in Figure 1.7b-d. When is

small such that same number of subbands (N) lies below and , the total number of

conducting subbands is 2N, and the associated conductance is . When is increased

to a level such that one subband bottom lies in between and , the total number of

conducting subbands becomes 2N+1, generating a conductance of / . When is

further increased such that two subband bottoms reside in between and , the total number

of conducting subbands returns to 2N so the integer plateau with is retrieved.

If we plot G as a function of gate voltage and source and drain bias, just as in Section 1.2.1, we

will obtain similar stability diagram for nonlinear transport in 1-D system. A schematic drawing

is shown in Figure 1.7a. For simplicity we assume all subbands are equally spaced and the

spacing does change with Vg and VDS. To be consistent with the experimental conditions (to be

introduced in Chapter 2), the chemical potential of the source contact is always set to ground.

Similar to Coulomb diamonds for 0-D system, here we see a series of diamond patterns as well,

but with different physical meaning: they represent conductance plateaus with labeled

conductance values in unit of 2e2/h. Similar to the 0-D case, left/right-tilted lines denote the

alignment of subband bottoms with chemical potentials of drain/source, as depicted in Figure

1.7c-d. The intersections (such as the one pointed by the blue arrow in Figure 1.7a) therefore

represent the situation where two successive subband edges both line up with contact potentials,

and the corresponding bias levels provide a direct measure of the subband spacing.

Page 33: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

19

1.2.3 Fabry-Perot Electron Interference in Electron Waveguides

In analogy with the interference due to multiple reflections of light between two reflecting

surfaces of an optical F-P cavity, which gives rise to the constructive and destructive interference

patterns, the wave-nature of electrons allows quantum interference between propagating electron

waves in coherent electron waveguides. Such systems can also be regarded as open quantum dots

with the resonances corresponding to the broad energy levels of the quantum dot [35].

Now we focus the F-P interference in InAs NW system. Assume the propagating electron wave

in the cavity is a plane wave. Constructive interference is observed for 2 2 where N is

an integer and L is the length of the cavity. For a parabolic electron dispersion relation, the

chemical potential in the channel can be expressed as 2⁄ , where is

the Fermi level, is the bottom of the corresponding 1 D subband and is the effective mass.

Meanwhile, the chemical potential is proportional to the gate voltage: . Thus, the system

can be tuned in and out of the constructive interference which modulates carrier transmission

through the channel, leading to periodic conductance oscillations as a function of gate voltage.

The period in Vg is directly related to L. The change of the Fermi wave vector over one period of

the Fabry-Perot oscillations is / and the corresponding change in the carrier density is

2 ⁄ ∆ /| |, where is the gate capacitance per unit length, ∆ is the

oscillation period and D is the orbital degeneracy of the subband. Therefore, the effective length

of the cavity is [17]

2 | |/ ∆ (1.12)

As mentioned in Section 1.2.2, when the source and drain bias satisfies | | , the

total charge in the channel will be unaffected by . In this case, we can view the effect of

as raising the chemical potential of one contact by | | /2 while lowering the other by

| | /2. Now we follow the approach proposed by Liang et al [13] to briefly examine the

nonlinear response upon the application of bias voltage. If the electron transport through the

channel is ballistic, the total energy of the electrons involved in transport is conserved. The total

energy of electrons at the Fermi surface can be expressed as a sum of their kinetic energy and the

local self-consistent potential energy: , where z is the spatial position along the

Page 34: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

20

channel. Using the parabolic electron dispersion relation, the wave number of electrons is given

by 2 2 ⁄ . The resulting phase shifts are given by

2 (1.13)

According to Eq. (1.13), changing the total energy of electrons at the Fermi surface by a source

and drain bias will shift the phase of electron waves. When the bias voltage increases to a critical

value VC such that 2 | | 2 , the condition of constructive interference

is satisfied and conductance resonance will occur. So we can imagine that in a stability diagram

where the differential conductance is plotted as a function of gate and source/drain bias,

conductance peaks/dips corresponding to constructive/destructive interferences will be periodic

in Vg and VDS, and F-P electron interference will manifest itself as chessboard-like patterns which

have been observed by all the above mentioned literatures.

1.2.4 Surfaces and Interfaces in Semiconductor Nanowire Devices

Surface states exist on both clean and adsorbate-covered semiconductor surfaces, originating

from dangling bonds and bonds between adsorbate and semiconductor surface atoms. At abrupt

metal-semiconductor interfaces, the wavefunctions of metal electrons decay exponentially into

the semiconductor. These tails represent metal-induced interface states. This concept also applies

to semiconductor-dielectric interfaces.

The net charge per unit surface area in surface states is given by

| | (1.14)

where and are the donor and acceptor surface state densities, and

are the donor and acceptor distribution functions, respectively [36]. The surface

state distributions can be characterized by a charge neutrality level , above which there are

predominantly acceptors and below which there are mainly donors. Eq. (1.14), together with the

concept of the charge neutrality level, will be adopted in Section 4.4.1 to model the potential

distribution in one of our NW devices.

Page 35: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

21

Surface/interface states have pronounced effects on the nature and efficacy of carrier transport in

field-effect transistors (FETs). Due to the large surface area to volume ratio of NWs, these

effects play more significant roles in NW FETs. The high density of surface states can lead to

Fermi-level pinning, thereby depleting the carriers from near the surface [37] and acting as

scattering centers which reduce carrier mobility. We will return to this point in Chapter 4 where

we will evaluate this surface scattering effect in terms of diameter-dependency of NWs. Another

important consequence of surface states is the degradation of gate control efficacy, which leads

to surface recombination of carriers and also causes hysteresis due to charge-discharge transients

from the surface states [38].

InAs is known to have donor-type surface states which lead to the formation of an accumulation

layer at the surface. This mechanism was explained by Noguchi et al. [39]. Electrons usually

occupy the surface states below the charge neutrality level where the states derive their weight

equally from valence and conduction bands. The conduction band edge Ec at the Γ point in InAs

is much lower than the Χ minima so that is located above Ec. Therefore the surface states

between Ec and can emit electrons into the conduction band, forming the surface

accumulation layers. This important property facilitates the formation of Ohmic contacts

between InAs and most metals [40]. In Section 4.4 we will see that according to the fitting

between a proposed model and the experimental data, the donor-like surface states in one of our

InAs NW devices provide a significant fraction of free carriers in the NW channel and

locates approximately 100 meV above the “effective” .

For the Si/SiO2 interface of MOSFET devices, one can distinguish between charges located deep

inside the oxide (oxide-trapped charges and mobile ionic charges) and charges located at the

interface (fixed oxide charges and interface trap charges) [41]. Some forms of interface trap

charges may also be present at the bare surface of SiO2 in our samples. For example, non-

bridging oxygen (oxygen dangling bond) may be formed when Si-O bond is broken, which may

act either as acceptors or donors. Together with charges in the bulk of the oxide, they can

exchange charges with NWs or gate electrodes and affect the transfer characteristic of NW

devices. This issue will be further discussed in Section 3.2.

Page 36: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

22

It is know from work done on MOSFETs that the presence of large number of capture and

emission processes with a broad distribution of timescales produces a 1/f-type noise spectra,

while in submicrometer devices the individual traps assume importance and produce discrete

switching events and random telegraph signals (RTS) [42]. This phenomenon has been carefully

studied in NW systems such as ZnO [43] and InAs [44, 45] NWFETs. Similar switching events

were observed in our etched NWs and manifest themselves in stability diagrams as abrupt

dislocations of conductance patterns in a similar fashion reported by Klein et al. [46], which

blurs the regular patterns and prevents us from accurately extracting some physical parameters of

the devices. This will be covered with more details in Chapter 3.

1.3 Motivations, Objectives and Thesis Outline

One common method to create tunneling barrier is to narrow the carrier transport channel by dry

or wet etching. QDs defined by such etched constrictions have been realized using top-down [46,

47, 48] and bottom-up [49] approach. Meanwhile, these constrictions can also be regarded as

QPCs [47]. Very recently, transport through single etched constriction fabricated by top-down

approach has been reported [50]. Yet there has not been a systematic study on such a constriction

in single semiconductor NW. This forms the initial motivation of this thesis.

Among the III-V semiconductor NWs, InAs is distinguished by its native electron accumulation

layer, relatively low effective mass, and correspondingly large electronic mean free path and

subband quantization. Combined with its large g-factor and strong spin-orbit coupling, InAs

NWs are extremely suitable for nano-electronic and nano-spintronic devices [51].

Tunneling barrier may be formed between the metallic contact electrode and the NW [52, 53, 54].

For back-gated InAs NWs, the formation of these barriers is attributed to the nonuniform

electrical potential produced by the back-gate and the grounded contacts [17, 54]. Therefore, if

part of the NW section between the two contacts is narrowed, depending on the relative height

between the etched barrier and the contact barriers, we may expect electron transport through

single QD, DQD or a QPC. This expectation is based on the assumption that there is no

Page 37: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

23

additional potential barrier presented in the channel, which is hardly true for real devices due to

the existence of disorder. Hence the actual phenomena observed from the measurements can

provide useful information about disorder. We can then use this information to improve the

design of the device.

For back-gated devices, changing the back-gate voltage will affect the height of all the barriers at

the same time. To gain separate control for the etched barrier, side-gate electrode can be placed

in the vicinity of the constriction. Single etched constriction between a pair of side-gate

electrodes constitutes the standard device architecture studied in this thesis. For comparison,

devices either without etched constriction or without side-gates should also be fabricated.

The above motivations naturally set objectives of the thesis project: successfully fabricate the

three types of devices, perform transport measurements on these devices and understand the

underlying physics from the observed phenomena.

The thesis is organized into five chapters. Chapter 2 describes the pertinent nanowire device

fabrication and experimental methodology. Chapter 3 discusses measurement results of a

standard device and a device without side-gates. Various quantum transport phenomena observed

in the former device are discussed according different gate bias regime. For the latter device, we

only consider the physics in the pinch-off region. Other than quantum transport phenomena, we

also characterize the hysteresis effect of side-gating and the interaction between side-gates and

the back-gate. Chapter 4 focus on the measurement results of a device without etched

constriction. The necessity of performing some preliminary processing on the raw data is first

described. This is followed by presenting the observed phenomena under different measurement

conditions. In the remaining part of this chapter, we show the logic and results of a simulation

conducted to account for the observed phenomena in this device. The validity of this simulation

is also discussed. Finally, we summarize main conclusions of this thesis and propose some future

works of this project.

Page 38: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

24

Chapter 2

Methods

2.1 Single InAs Nanowire Device Fabrication

Transfer Nanowires onto the Substrate

Device fabrication starts with cut pieces (approximately 12 mm 12 mm in size) of heavily-

doped p+ Silicon wafer (resistivity 0.002-0.005 Ω - cm) with a 100 nm thick SiO2 layer

grown by thermal oxidization. These substrates were first patterned with global metallic markers

to position the NWs. After that, they were dipped in N-Methyl-Pyrrolidinone organic stripper at

60 for 15 minutes to remove the residual resist from metallization of markers. This was

followed by a standard solvent cleaning step consisting of immersion in heated (120 ) acetone

and isopropanol (IPA) for 2 minutes and de-ionized (DI) water rinse. These procedures ensure a

clean substrate surface, which is crucial for facilitating good NW adhesion according to our

experience. InAs NWs studied in this thesis were grown by the metal-catalyzed vapour-liquid-

solid (VLS) method [55] in a modified commercial molecular beam epitaxy (MBE) system

(ATC-EP3). They were mechanically transferred to the receiving substrates by putting the

growth substrate upside down on the SiO2 surface of the receiving substrates and gently touching

the rear of the growth substrate with a tweezers. A successful NW transfer yields a desired

density of NWs on the receiving substrates with few broken pieces and undamaged SiO2 surface.

The receiving substrates were then heated at 120 for 15 minutes to get a good adhesion

between NWs and the SiO2 surface so that NWs would not move during the subsequent spin-

coating steps.

Page 39: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

25

Locate Nanowires and Form Ohmic Contacts/Side-gates

To find proper NWs and locate them, samples were loaded into a commercial electron beam

lithography (EBL) system (RAITH Elphy Plus) and the coordinates of the selected NWs were

recorded with respect to the predefined markers. After taking samples out of the EBL system,

they were spin-coated by a Poly-Methyl-Methacrylate (PMMA)-based resist and reloaded into

the EBL chamber for exposure. Note that any spin-coating step is associated with a preheating

(at 120 for a few minutes) before the coating to dehydrate the surface of the substrate and a

soft-baking (at 170 for 90 seconds) after the coating to reduce the remaining solvent content in

the resist. The patterns to be transferred onto the samples include the source/drain electrodes on

top of the NWs, the side-gate electrodes by the side of the NWs, and local markers to align

patterns of different exposures. After the exposure, samples were developed in methyl-isobutyl-

ketone (MIBK) and IPA (1:3) for 40 seconds and were rinsed in IPA to terminate the

development. The residual resist on the exposed regions must be removed to lower the contact

resistance and improve the liftoff. This was done by loading samples into a homemade oxygen

plasma system with the plasma ignited for 8-10 seconds at 100mbar of oxygen and 70 Watts of

total power (the plasma was generated by a commercial McCarroll cavity (Opthos Instruments)

operating at 2.45 GHz.

Prior to metal evaporation, samples were dipped in a solution of 0.3% by weight ammonium

polusulfide in DI water for 8 minutes to get rid of the native oxide at the exposed regions of the

NWs and terminate their surface with Sulfur atoms. This step was found to be essential for

formation of low contact resistance to InAs NWs [56]. After the treatment, samples were quickly

rinsed in DI water, blown dry and immediately loaded into a custom-built evaporation system.

This system is equipped with thermal sources (MDC Vacuum model Re-vap 900) for Au

evaporation and electron-beam sources (Oxford Applied Instruments model EGN4) for Ti

evaporation. The pressure inside the chamber could be brought down to 3 10 mbar within 40

minutes before the metallization, which minimized the re-oxidize of the exposed NW surfaces.

Normally, a Ti (around 10 nm) / Au (around 100 nm) bilayer is evaporated with Ti providing

good adhesion to the SiO2 surface. Liftoff was performed by immersing samples in acetone

Page 40: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

26

overnight followed by the standard solvent cleaning step. A schematic view of samples at this

fabrication stage is shown in Figure 2.1a.

Etch the Nanowires

After the liftoff, samples were loaded in a scanning electron microscope (SEM) system (Hitachi

S-4500) housed in the Department of Material Science and Engineering, University of Toronto to

pick up NWs that have good alignment with side-gate electrodes. Afterwards, samples were

spin-coated with a thinner PMMA resist for finer etching patterns and loaded into the EBL

system. Local metallic markers near centers of NWs patterned during the first exposure were

used to align the etching patterns with respect to side-gates. After the exposure, samples were

developed and treated with oxygen plasma ashing to obtain clean etching windows.

An etchant consisting of citric acid (C6H8O7) and hydrogen peroxide (H2O2) was chosen to etch

the NWs because it gives a high etching selectivity of InAs over GaAs (about 13:1 at room

temperature) at a volume ratio 1:5 [57] for potential usage in engineering our InAs-GaAs core-

shell NW system [58]. The etching mechanism is attributed to an oxidation-reduction reaction at

the III-V semiconductor surface by the hydrogen peroxide, with the dissolution of the oxide

material by the acid [59]. The PMMA resist serves naturally as the etching mask because

C6H8O7/ H2O2 mixture with volume ratios between 0.1 and 100 does not erode photoresist [60].

To optimize parameters for the etching of InAs NW, a bunch of testing samples were made.

These samples contained selected InAs NWs patterned with metal markers. Etching results under

different conditions were compared by taking SEM images of these testing samples. Optimized

etching procedures are given below. First, anhydrous citric acid crystals were dissolved in DI

water at the ratio of 1g C6H8O7: 1ml DI water at least one day prior to the etching to ensure

complete dissolution and room temperature stability. Approximately 15 minutes before

conducting the etching, the liquid citric acid/water mixture was mixed with fresh 30% hydrogen

peroxide and diluted with DI water. A volume ratio C6H8O7: H2O2: DI water of 1:5:25 was

identified to give the most proper etching rate (estimated to be 2 nm/second at 22 ). The 15

minutes delay was used to allow the etchant to return to room temperature, if any temperature

change occurs due to mixing. The etching rate turned out to be very sensitive to the temperature

Page 41: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

27

and the concentration of the solution. To minimize the variation of the two factors, fresh etchant

was always made each time before conducting the etching in our clean room where a relatively

constant room temperature and humidity is maintained. The 30% hydrogen peroxide was

frequently replaced to avoid using the degraded solution. Taking these precautions, a robust

etching rate could be expected. Diameters of the NWs were measured from the SEM images so

the desired etching depth was known. Therefore, the corresponding etching time could be

calculated, and samples were dipped in the etchant for that amount of time. The etching was

ceased by rinsing samples in DI water. The resist was then removed by dipping samples in

acetone for a few minutes followed by the standard cleaning step. A schematic view of samples

after the resist was removed is shown in Figure 2.1b.

Wire-bond the Samples

After the etching step, samples were loaded into the SEM again and only NWs with desired

etching profiles were chosen for the subsequent processing. To connect the electrodes of selected

NWs with the measurement system via wire-bonding, bonding pads with traces connected to the

electrodes need to be made. A copolymer/PMMA bilayer resist, which provides a retrograde

resist profile, was used at this stage to enhance yield for liftoff of thicker metal deposition (Ti (10

nm) / Au (180 nm)). A schematic view of samples after the liftoff is shown in Figure 2.1c.

Page 42: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

28

Figure 2.1: Main steps of device fabrication. (a) 1st exposure and metallization transferred patterns of the source

(S)/drain (D) contacts, side-gate (SG) electrodes and local markers onto the substrate with respect to the NW. (b) 2nd

exposure opened an etching window for the etchant, which was aligned to previous metal patterns via local markers.

Using resist as the etching mask, the NW section exposed to the etchant was etched to a desired depth. (c) 3rd

exposure and 2nd metallization transferred patterns of bonding pads and traces (yellow) connecting previous metal

patterns (orange) onto the substrate. Previous metal patterns are barely seen at this scale. (a) and (b) are actually

enlarge views of the dark blue square region.

To make the surface of NWs as clean as possible before the measurement (thus reduce the

number of surface states), samples were immersed in the stripper again followed by the standard

cleaning step. They were then cut again into a smaller piece (smaller than 6 mm 6 mm) to fit

into the ceramic 28-pin leadless-chip-carriers (LCCs). The anchor of samples onto LCCs was

achieved by melting a small indium piece in between the rear of samples and surface of LCCs’

cavity which serves as the conductive glue. A commercial wedge-bonder (Kulicke and Soffa

model 4526) was employed for the wire-bonding.

S D

S D

SG

SG

SG

SG

SG

SG

S D

SiO2

p++ Si

local marker

(b)(a)

(c) SG

bonding pad

trace

Page 43: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

29

Once the electrodes on top of NWs are bonded to the pins of LCCs, the single NW devices

become subject to the threat of electro-static discharge (ESD) which according to our experience

can vaporize or melt the NWs. Thus, all necessary precautions have to be taken to avoid the

ESD damage to the devices. A custom LCC chip holder was assembled onto the stage of the

bonder providing a resistance of 50 kΩ to ground for each pin. Our measurement system also

contains a homemade electrical breakout box that can ground LCC pins when they are not being

connected to instrument terminals, which will be further introduced in the next section.

Therefore, the devices are safe when samples are either in the wire-bonder chip carrier or inside

the measurement system. Samples were temporarily kept in a LCC chip holder inserted in a

conductive mat after the wire-bonding and were ready to be loaded into the measurement system.

When they have to be carried by hand, one must properly ground him/herself to minimize the

ESD damage.

2.2 Experimental Setup for Transport Measurements

Electrical measurement was performed in a closed cycle He cryostat (Advanced Research

Systems model ARS DE 202) with a temperature range of 7-350 K. The temperature could be

adjusted and monitored by a heater and a sensor attached to the copper coldfinger of the cryostat,

under the control of a temperature controller (LakeShore model 330). The LCC was placed

inside a custom built copper hip holder screwed onto the coldfinger. A top copper plate was used

to press the pins at the back of LCC against the bottom spring loaded probes. These probes were

mounted inside a plastic (G10) fixture attached to the copper frame. This design ensures both

electrical contact and thermal anchoring for the LCC. The back ends of 24 probes were soldered

to 12 low frequency twisted pairs with bandwidth practically limited to approximately 50 kHz by

wiring capacitance. An image of the sample holder is shown in Figure 2.2a with the above

components marked in the image. The twisted pairs were connected to the measurement circuit

by a vacuum electrical feedthrough (24 pin Fisher). The circuit was placed inside a homemade

RF-shielded electrical breakout box (see Figure 2.2b) with toggle switches which provide choice

between translating the feedthrough to coaxial connectors demanded by electrical instruments

and grounding the electrodes of devices through 1 MΩ resistors. The latter protects the devices

Page 44: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

30

against ESD damage when the electrodes are not connected to the terminals of electrical

instruments.

The DC voltage biases (i.e., the gate biases Vbg and Vsg and source/drain DC bias VDS) were

provided by the source-measurement unit (SMU) of a semiconductor characterization system

(Keithley model 4200), with the current being measured at the same time. A pair of side-gates

was always connected together. Since gate bias is normally in the range of several volts, a

protecting resistor (1 MΩ) was always placed between the SMU and the pin that provided the

bias in case of a serious gate leakage to protect the SMU. The source/drain AC bias

2 came from the AC output of a low-frequency (100 kHz) lock-in (Stanford Research

Instruments SR830). The amplitude of A was chosen such that the root mean square (RMS)

of | |√

~ to avoid electron heating by . VDS and Vds were inductively coupled to each

other using a magnetically shielded signal transformer (Triad Magnetics), attenuated by 1/100

via a resistor-divider network, and fed to the drain electrode of the NW. Current from the source

electrode Ids was first amplified by a low-noise current preamplifier (Ithaco Instruments model

DL1211) and then obtained using the lock-in. The differential conductance is therefore obtained

as ⁄ , which we will refer to as “conductance” for convenience throughout this thesis.

The circuit layout of the differential conductance measurement is shown in Figure 2.2c. The

input resistance of the current preamp depends on the chosen sensitivity and is either 20 Ω or 200

Ω in all the measurements done for the thesis. In either case, it is trivial compared to the NW

resistance (at least a few KΩ), so the chemical potential of the source electrode can be viewed as

pinned to the ground potential all the time.

The applied magnetic field was generated by an adjustable gap electromagnet (Alpha model

4600) controlled by a power supply (Alpha model 3002), which was in turn controlled by a

programmable power supply (Agilent model E3631A) to set the step and invert the direction of

the magnetic field. The magnetic field intensity could be varied from -0.6 T to 0.6 T. The metal

frame that carries the extender of the cryostat was modified so that the extender could rotate 90 °

within the gap between the two magnets, providing a range for the direction of the magnetic field

from being parallel to being perpendicular to the sample plane. Figure 2.2b shows an image of

the electromagnetic system. The intensity of the magnetic field could be verified by inserting a

probe of a gaussmeter (LakeShore model 450) to the gap region between the two magnets.

Page 45: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

31

Figure 2.2: Experimental setup. (a) Custom built sample holder. A top cooper plate presses down the LCC towards

the spring loaded probes mounted inside the G10 fixture, providing both thermal anchor and electrical contacts for

the sample. (b) Cryostat and electromagnetic system. The sample holder is located in the extender of the cryostat

which can be inserted into the gap of the two magnets and rotated by 90 °. The magnetic field is controlled by a

programmed power supply (Agilent E3631A). (c) Circuit layout for differential conductance measurement. AC and

DC source/drain biases are coupled by a transformer, attenuated by a divider and then applied onto the drain contact

of the device. Current from the source contact is amplified by a current preamp and measured by a lock-in.

Back/side-gate biases are applied through protecting resistors to the rear of the sample and side-gate electrodes,

respectively.

top copper plate

G10 fixture

copper frame

LCC with sample inside

(a) (b)

cold finger

temperature sensors

twisted pairs

AgilentE3631A

electromagnets

vacuumfeedthrough

power supply(Alpha 3002)

electrical breakout box

metal frame

extender of the cryostat

cryostat

Vbg

Vsg

A/V Lock-in

VDS Vds

(c)

transformer

voltage divider current preamp

Page 46: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

32

Ground loops occur when the grounds of apparatus and circuits are actually at different

potentials. They could couple stray AC signals (especially the 60 Hz noise and its harmonics)

from other equipments into the measurement circuit which degrade the signal-to noise ratio.

Thus, efforts were put to eliminate them by troubleshooting while monitoring the change of the

noise spectrum using a digital sampling oscilloscope (Agilent model Infiniium 54845a).

After the sample was loaded into the cryostat, it was always pumped overnight to a pressure less

than 10-5 mbar before any measurement was performed. This was proved to be necessary to

stabilize the transistor characteristics and obtain minimal hysteresis of the devices, which might

be a result of the detachment of species on the surface of NWs due to the pumping, thus reduced

the number of surface/interface states.

Page 47: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

33

Chapter 3

Quantum Transport Phenomena in Single

InAs Nanowires with Etched Constrictions

3.1 Introduction

As mentioned in Section 1.3, if we selectively etch part of an ideal NW, we expect to see a

transition of transport from 1-D to 0-D on the same device: conductance quantization when the

channel is turned on by the gate bias, and Coulomb blockade due to electron tunneling through

either DQD or a single dot when the channel is turned off. In reality, disorder exists in NWs and

the etching might induce further surface roughness or surface states. This may lead to the

formation of multiple tunneling junctions (MTJ) in the pinch-off region and smear the

conductance plateaus when the 1-D transport takes place. Adding a pair of side-gate at the two

sides of the etched constriction allows locally tuning the potential in the constriction. Side-gating

may exhibit hysteresis [61]. Although this effect can be explored for memory applications [62], it

is detrimental for transport properties of NWFETs [63].

In this chapter we present typical transport measurement results in etched single InAs NWs. Two

types of device were studied: Section of NW with an etched constriction controlled by the back-

gate, and section of NW with an etched constriction controlled by both the back-gate and the

side-gates. We first discuss the hysteresis of side-gating in the following section since it does not

belong to the category of quantum transport phenomena. This is followed by discussion of

results on the first type of device shown in Section 3.3. Then we focus on different transport

phenomena observed in the second type of device in Section 3.4. This section also includes a

proposed electrostatic model to account for the observed interaction between the back-gate and

side-gates.

Page 48: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

34

3.2 Side-gating Effect: Hysteresis

In this section we describe a universal side-gating effect observed in our side-gated devices.

Figure 3.1 shows the typical transfer characteristic of one such device in terms of side-gating.

For the G-Vsg curve at each temperature, side-gate voltage was swept toward more positive

values and then swept back toward more negative values, with the back-gate grounded. At room

temperature, large hysteresis was clearly observed. The conductance increased after the positive

sweep of Vsg and decreased after the negative sweep, as indicated by the green arrows. The

hysteresis was reduced at lower temperature (red curve), and eventually disappeared at 200 K

(blue curve). Figure 3.1b displays the transfer characteristic of the same device in terms of back

gating (with side-gate grounded) from which we could barely see the hysteresis. The

measurement was taken at room temperature and at lower temperatures hysteresis was even

harder to resolve. All the measurements shown in Figure 3.1 were performed at the same gate-

voltage sweep rate.

The transfer characteristics of top-gated InAs NWFETs have been studied by Dayeh et al. [63,

64] where the observed hysteresis was attributed to the capacitive effects of donor-like interface

trap states. They found that slower gate-voltage sweep lead to a charge neutral interface with

reduced capacitance and smaller hysteresis. It is this reason that caused the hysteresis in our

devices, we should also observe distinct hysteresis from the back-gate voltage sweeps because

donor-type interface states exist all over the surface of InAs NWs, as introduced in Section 1.2.4.

The absence of significant hysteresis in back-gate voltage sweeps therefore indicates that this

scenario alone is not enough to account for the different magnitude of hysteresis between side-

gating and back gating.

Page 49: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

35

Figure 3.1: (a) Transfer characteristics of a side-gated InAs NWFET, showing distinct hysteresis at room

temperature (black curve) and reduced hysteresis at lower temperatures (red and blue curves). Green arrows indicate

the sweep directions of side-gate voltage. (b) Transfer characteristics of the same device gated by the back-gate at

room temperature showing little hysteresis. (c) Schematic of the proposed oxide trap charging mechanism for side-

gating. Here the NW is grounded through source and drain contacts and a positive bias is applied to the side-gate.

When the bias reaches a certain value, the electric field between the gate electrode and the NW may be high enough

to induce electron tunneling from the oxide traps to gate electrode. Corresponding energy diagram is also shown to

illustrate the tunneling situation.

The side-gating effects on Si NWs fabricated by top-down approach were studied by Matsukawa

et al. [65]. They observed the same sign hysteresis in the n-type NWs and attributed it to the

effect of positively charged oxide traps due to both impact ionization and injection of holes from

side-gates as the gate voltage was swept to large positive values. This viewpoint provides a clue

-6 -4 -2 0 2 4 6

0.02

0.04

0.06

0.08

0.10G

(2e2 /h

)

300K 250K 200K

Vsg(V)

-3 -2 -1 0 1 2

0.00

0.04

0.08

0.12

0.16

0.20

G(2

e2/h

)

Vbg(V)

Ti/Au

SiO2

NW

(a)

(b) (c)

E field

EC

Page 50: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

36

to explain our data. As mentioned in Section 1.2.4, SiO2 has interface and bulk traps whose

charge state changes with gate voltage. Interface traps are populated continuously as the gate

voltage is tuned, regardless of gate type, thus they cannot be used to account for the different

hysteresis behaviors between the back-gate and side-gates. Oxide traps, on the other hand, are

charged only with carrier injection at gate fields above 3×105 V/cm [66] and may response

differently to back-gate and side-gate biases. Experimentally, if a sweep to large positive values

of gate voltage creates a negative shift in the threshold voltage, this means holes are injected into

oxide trap states (or electrons leave the trap states), enhancing the accumulation of electrons at

the surface of InAs NW. As can be seen from SEM images of devices with side-gates (Figure

3.3a and Figure 4.1a), the end edge of all side-gates are sharp tips. These sharp terminals of side-

gates were not intentionally made (all side-gate patterns were drawn as rectangles), but results of

the proximity effect during the exposure [67]. Similar to the “point discharge” principle, the

electric field near the sharp tips may be strong enough (such strong field may be absent in back

gating) to induce the tunneling of trapped electrons from the trap states to side-gate electrodes

when side-gate voltage is increased above a certain positive value. This scenario is depicted in

Figure 3.1c. The trap states will remain positively charged until the side-gate voltage is

decreased below a certain negative value which favors the reverse tunneling from gate electrodes

to oxide traps. This tunneling mechanism should be suppressed in lower temperature [66], which

is consistent with our observation that hysteresis at side-gating decreased as temperature was

lowered.

3.3 Coulomb Blockade in an Etched InAs Nanowire without

Side-gating

We begin the study of etched InAs NWs by fist investigating the simplest case: an etched InAs

NW section tuned by a global back-gate. Inset I of Figure 3.2a is the SEM image of the device.

The diameter of the NW and the etched constriction are 50 ± 2 nm and 25 ± 2 nm, respectively.

The etched constriction has a length of 110 ± 2 nm and the distance between the two contacts is

610 ± 2 nm. The measurement was performed at T= 10 K. The AC source and drain bias was set

to a RMS of 1 mV with a frequency of 1450 Hz.

Page 51: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

37

The transfer characteristics of the device are shown in inset II of Figure 3.2a. The channel is

turned on at Vbg > 0.95 V. The remarkable conductance dip at Vbg > 1.5 V is most likely due to

the universal conductance fluctuations (UCF) with a magnitude of e2/h, originating from

electrons being multiply scattered by randomly distributed disorder in the channel [23]. UCF has

been observed in InAs NWs and appear to be random in Vg [54] with no regular pattern in the

Vg-VDS plane [68].

Figure 3.2a is the enlarged view of the pinch off region marked by the red rectangle in the inset

II, from which three linear conductance peaks around Vbg = 0.71 V, 0.8 V and 0.93 V can be

identified. The third peak is the mark of the end of pinch off region and onset of the channel.

Most part of the second peak is superimposed with high conductance state of RTS so the

conductance value is higher than its normal value. The stability diagram of the same back-gate

voltage range in Figure 3.2a is shown in Figure 3.2b. Two large irregular Coulomb diamonds can

be identified, whose boundaries are superimposed with small “kinks”. The two diamonds

correspond to the two Coulomb blockade regions separated by three conductance peaks in Figure

3.2a. Such behavior is characteristic for an unequal double quantum dot structure controlled by a

common gate, which has been realized based on some 1-D systems [69, 70, 71]. The large

diamonds reflect the charge state of the smaller dot while the “kinks” are due to charging of the

larger dot. Coulomb blockade is lifted whenever an energy level of the smaller dot is aligned

with an energy level of the larger dot, or they fall into the bias window, so that an electron can

tunnel sequentially through the two dots.

Page 52: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

38

Figure 3.2: (a) Linear conductance in the pinch-off region of the transfer characteristics of the device. Three peaks

are visible. The second peak is superimposed with RTS. The third peak denotes the onset of the channel. Inset (I):

SEM image of the device with the source and drain contacts labeled. Inset (II): Transfer characteristics of the device.

The pinch-off region is highlighted with red rectangle. (b) Stability diagram corresponding to the same back-gate

voltage range in (a). Large diamonds reflect the charge state of the smaller dot while the fast oscillations

superimposed on the boundaries of large diamonds are due to the charging of the larger dot.

We can adopt the model of DQD introduced in Section 1.2.1 to do some preliminary analysis.

Applying a voltage to the back-gate tunes both dots simultaneously. In a qualitative picture, the

smaller dot is expected to contain fewer electrons than the larger dot and the capacitive coupling

of the gate to dot 1 is expected to be smaller than that of dot 2 (recall Figure 1.4). Tuning the

back-gate voltage will therefore fill dot 1 more slowly than dot 2. Following this logic, it is

straightforward to understand that “kinks” superimposed on the boundaries of large diamonds are

the result of electrons leaving/entering the larger dot and consequently lowering/raising the

chemical potential of the smaller dot due to interdot coupling.

(a)

(b)

(I) (II)

0.000

0.003

0.006

0.009

0.012

0.015G

(2e2 /h

)

0.650 0.675 0.700 0.725 0.750 0.775 0.800 0.825 0.850 0.875 0.900 0.925-30

-20

-10

0

10

20

30

Vbg(V)

V DS(m

V)

-0.0060

0.0011

0.0048

0.025

0.130.23

0.8 1.0 1.2 1.4 1.6 1.80.0

0.1

0.2

0.3

0.4

0.5

0.6

G(2

e2 /h)

Vbg(V)

G(2e2/h)

S

D

Page 53: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

39

We focus on one edge of the large diamond to see how we can extract information from the

“kinks”. Depending on how the double dots are connected to the contacts, there are two possible

scenarios. (A): The smaller dot is directly coupled to the source. In this case, edges represented

by the blue line denote the alignment of the energy level of the smaller dot µ to the chemical

potential of the source µS and edges represented by the orange line denote the alignment of the

energy level of the larger dot µ to the chemical potential of the drain . (B): The smaller dot is

directly coupled to the drain. In this case edges represented by the blue line denote the alignment

of µ to µS and edges represented by the orange line denote the alignment of to the . Take

the alignment of µ to µS as an example. According to Eq. 1.9, this condition means the following

relation has to be satisfied: ∆ ∆| |

∆| | .

Similarly, we can build up such relations for other alignment conditions. This leads to the

following expressions for the slopes of the blue and orange border lines in two scenarios:

Scenario A:

Blue line ( = ): ∆ ∆⁄ ⁄

Orange line ( = ): ∆ ∆⁄ ⁄

Scenario B:

Blue line ( = ): ∆ ∆⁄ ⁄

Orange line ( = ): ∆ ∆⁄ ⁄

Since the origin of the DQD is not clear, it is hard to determine which scenario it would be for

the configuration of the DQD system. The source and drain contacts used during the experiment

are labeled in the inset (I) of Figure 3.2a. Intuitively, if one believes it is the contact barriers plus

the etched barrier that defines the DQD, scenario B is probably the case because the length of the

unetched section close to the source (270 nm) is a bit longer than that of the unetched section

close to the drain (240 nm). However, more reliable speculation on the origin of the DQD should

be based on the estimated dot length, which in turn requires the knowledge of gate-to-dot

capacitances Cbg1 and Cbg2.

Page 54: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

40

The shift of right-tilted borderlines, e.g., the blue arrow, is indicative of the interdot coupling

strength . The separation of the left-tilted borderlines, e.g., the orange arrow, is indicative of

the addition energies of the larger dot ∆ where ∆ is the level spacing of the

larger dot. The separation of Coulomb peaks in linear transport regime measures the addition

energies of the smaller dot ∆ . Nonetheless, as can be seen from Figure 3.2b,

severe RTS shift differential conductance patterns at certain back-gate voltage ranges, which

prevents accurate extraction of capacitances from above relations.

3.4 Transport Phenomena in an InAs Nanowire with Single

Etched Constriction and Side-gating

In the previous section we presented the measurement result of an etched InAs NW controlled by

a global back-gate. As mentioned in Section 1.3, one drawback of back-gated NW devices is that

tuning the back-gate voltage not only changes the energy in the channel but also modifies the

coupling between the channel and contacts. To gain separate control over the channel, we added

a pair of side-gates to the two sides of each etched constriction. In this section we focus on

transport properties for such a device. Figure 3.3a is the SEM image of this device. The diameter

of the NW and etched constrictions are 56 ± 2 nm and 36 ± 2 nm, respectively. Two etched NW

sections with lengths 780 nm and 740 nm, are separated by a common source electrode. The

purpose of having two etched sections on the same NW is to compare transport properties at

different length of etched constrictions. The length of the top left constriction is 170 ± 2 nm, to

which we will refer as “short-etched constriction”. The length of the bottom right constrain is

260 ± 2 nm and we will refer to it as “long-etched constriction”. All the measurements shown in

this section were performed at T= 10 K. The AC source and drain bias was set to a RMS of 1 mV

with a frequency of 210 Hz.

Page 55: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

41

Figure 3.3: (a) SEM image of the device studied in this section. Two etched constrictions were fabricated on the

same NW to examine the effect of etching length on transport properties. (b) Linear conductance as a function of

side-gate and back-gate biases for the short-etched constriction. The slope of stripes in region I and III is different

from the one in region II. Detailed bias spectroscopy was performed along marked arrows except for the light green

one.

3.4.1 Interaction between Side-gate and Back-gate

Figure 3.3b shows the linear conductance as a function of side-gate and back-gate biases for the

short-etched constriction. A series of dark stripes running from top left to bottom right with

conductance less than 0.15 conductance quantum can be identified but the slopes of these stripes

vary from place to place. In general, slopes in region I and III are similar and they are steeper

than slopes in region II. These stripes are linear conductance peaks and there slopes represent the

relative coupling strength between the side-gates to the channel and the back-gate to the channel.

The variation of the slopes therefore indicates the change of relative coupling strength at

different gate biases.

To understand the origin of this phenomenon, we notice that the leakage current of both side-gate

and back-gate are significant in this device and they keep changing as gate bias varies. The

measured gate leakage current is summarized in Table 3.1 with the first/second number in each

cell represent Ibg and Isg, respectively, in unit of nA. From this table we can conclude that Ibg is

large as Vbg increases, regardless of the value of Vsg. The absolute value of Isg, on the other hand,

(a) (b)

-0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.01.0

1.5

2.0

2.5

3.0

3.5

Vsg(V)V

bg(V

)

-1.00E-03

0.0930

0.187

0.281

0.375

G(2e2/h)

Ι

ΙΙ

ΙΙΙ

Page 56: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

42

becomes large only when the voltage difference between Vsg and Vbg is large. This is a strong

indication that the back-gate has a leakage to the source or the drain, while side-gate does not but

has a leakage to the back-gate.

Vbg\Vsg (V) 0 1 2 3

1 270/-70 220/0 205/25 18/310

2 880/-470 600/-2 600/1 545/90

3 1600/-1000 1100/-180 1000/0 1000/12 Table 3.1: Measured gate leakage current at different gate biases. 1st and 2nd number in each cell represent Ibg and

Isg, respectively, in unit of nA.

Based on the analysis of gate leakage current, we propose a simple model to account for the

observed gate coupling variation. During the wire bonding, the oxide underneath the bonding

pads may be destroyed by the wedge of the wire bonder, creating leakage paths to the back-gate.

These leakage paths can be characterized by variable resistors, as depicted in Figure 3.4a. In the

case of linear transport measurement, the NW is grounded through source and drain contacts, so

a leakage from source and/or drain to the back-gate would form a gate leakage current loop.

Similarly, a leakage from the side-gate electrode to the back-gate allows current to flow

whenever there is a voltage difference between the two gates. The coupling between the two

gates is weak when the voltage difference between the two gates is small, and becomes stronger

when the voltage difference is larger. Therefore, we see in Figure 3.3b that in region II where the

voltage difference between the two gates is small, the two gates are relatively independent from

each other, rendering constant slope of the stripes. When the voltage difference of the two gates

is large (region I and III), the mutual coupling between the two gates becomes stronger, leading

to the steeper bend of stripes, which means the coupling between the back-gate and the channel

becomes relatively weaker compared to the coupling between the side-gates and the channel.

Page 57: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

43

Figure 3.4: (a) Schematic of the proposed model to explain observed change of relative coupling between back/side-

gate and the channel. Leakage paths from source and/or drain contacts and side-gate electrodes to the back-gate

were probably created during the wire bonding which can be characterized by variable resistors. (b) Measured gate

leakage current at different gate voltages with side-gate connected to back-gate.

If we connect the side-gate and the back-gate together, there will be no voltage difference

between the two gates and thus no current flow in between. The gate leakage current in case

should reflect the leakage from the back-gate to the contacts. This result is plotted in Figure 3.4b,

from which we can conclude that the back-gate coupling to the channel keeps decreasing and

become constant at Vg > 1.5 V. The decrease of back-gate coupling to the channel as back-gate

voltage increases is responsible for the steeper slope in region I than the one in region III.

3.4.2 Transport Phenomena in the Short-etched Section — Low

Back-gate Bias

At low back-gate bias, the contacts are expected to be more opaque, and the Fermi level in the

channel may be below some potential fluctuations, leading to tunneling through MTJ. This

scenario was confirmed in this device by performing bias spectroscopy near the pinch off regime

of back gating. Figure 3.5 shows such a measurement along the orange arrow in Figure 3.3b,

where the back-gate voltage was kept at Vg = 1 V and side-gate bias was swept. At Vsg < 2.75 V,

S/D

SiO2

Vds

Vsg

Vbg

Ids

Ibg

Isg

0.0 0.5 1.0 1.5 2.0 2.5 3.00

200

400

600

800

1000

I g(n

A)

Vg(V)

side gate

(a) (b)

Page 58: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

44

the linear transport is completely pinched off (see the linear response in Figure 3.5a), but from

the bias spectroscopy in Figure 3.5b, several irregular diamond-shaped regimes of different

height can be seen. The diamonds do not close, and the slopes are not constant, which is a strong

sign of charge transport through multiple QDs defined by a number of defects along the 1-D

channel [11, 71]. At 3.06 V< Vsg < 3.175 V and Vsg > 3.175 V, the first and second conductance

plateaus show up with average conductance values of 0.06 and 0.12 conductance quantum,

respectively. There are several quasi-periodical oscillations superimposed on the conductance

plateaus. This feature will be discussed in the next section. However, the conductance trace is not

smooth. Rather, it is corroded by random conductance switching, which is a strong indication of

the presence of RTS.

Figure 3.5: (a) Linear conductance response along the orange arrow in Figure 3.3b. The pattern looks very “noisy”,

with random switching superimposed onto the conductance trace. (b) Corresponding bias spectroscopy, showing

irregular diamond patterns that are not closed in the pinch-off region.

Changing the gate voltage shifts the Fermi level relative to the barriers in the structure. As the

Fermi level is raised above some of the barriers in the MTJ, the dots merge to form larger dots

with increased capacitances. At certain gate bias ranges, we find sign of formation of single QD.

0.00

0.02

0.04

0.06

0.08

0.10

0.12

0.14

G(2

e2 /h)

2.6 2.7 2.8 2.9 3.0 3.1 3.2 3.3 3.4-30

-20

-10

0

10

20

30

Vsg(V)

V DS

(mV)

-1.00E-03

0.0402

0.101

0.190

0.267

G(2e2/h)

(b)

(a)

Page 59: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

45

Figure 3.6b is the bias spectroscopy along the dark green arrow in Figure 3.3b, where the side-

gate was connected to the back-gate. In the gate voltage range of 1.15 V < Vg < 1.3 V, three

Coulomb diamonds (labeled with numbers) with well defined boundaries are clearly seen.

Applying the CI model introduced in Section 1.2.1, we obtain the key parameters of the QD as

gate bias varies, which are summarized in Table 3.2 as follows:

Diamond EC (meV) C (aF) Cg (aF) α=Cg/C CS (aF) CD (aF)

1 11 14.6 3.7 0.25 10.4 0.5

2 10 16 3.7 0.23 11.4 0.9

3 7.5 21.3 4 0.19 16 1.3 Table 3.2: Key parameters of the QD extracted from the labeled Coulomb diamonds in Figure 3.6b.

From Table 3.2 we see that the total capacitance increases as gate voltage increases, a

phenomenon in electronic QD system that has been reported as a result of increased QD size at

higher gate bias [22, 72]. It is also obvious from Table 3.2 that the coupling between the QD and

source is much stronger than the coupling between the QD and the drain, which explains the

asymmetry of slopes for left-tilted and right tilted diamond edges in Figure 3.6b. Although due to

the gate leakage, we may not directly use the extracted gate capacitance to estimate the length of

the QD, the change of C and Cg of QD is a strong indication that the single QD is formed

between the contact barriers.

Page 60: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

46

Figure 3.6: (a) Linear conductance response along the yellow arrow in Figure 3.3b. (b) Corresponding bias

spectroscopy. Three Coulomb diamonds labeled with numbers can be clearly identified with well defined boundaries.

In the classical regime, the line shape of an individual Coulomb peak can be described by Eq.

(1.5). From the best fit (using least square method for the nonlinear curve fitting) to the three

peaks marked with arrows in Figure 3.6a, we obtain the estimated lever arms for each peak,

assuming the electron temperature is around 15 K. These lever arms are summarized in Table 3.3.

These values represent lever arms at the three intersections of Coulomb diamonds 1 and 2 in

Figure 3.6b, and they are very close to the lever arms of Coulomb diamonds 1 and 2 estimated in

Table 3.2 using the CI model. This confirms that our approaches to estimate parameters of QD

are reliable.

α1 α2 α3

0.24 0.22 0.2 Table 3.3: Extracted lever arms for the three Coulomb peaks marked with arrows in Figure 3.6a, using Eq. (1.5) and

assuming the electron temperature to be around 15 K.

0.000

0.008

0.016

0.024

0.032

0.040

G(2

e2 /h)

G(2e2/h)

1.00 1.05 1.10 1.15 1.20 1.25 1.30 1.35 1.40 1.45 1.50 1.55 1.60 1.65 1.70-30

-20

-10

0

10

20

30(b)

Vg(V)

VD

S(m

V)

-1.0E-030.0120.0390.0690.0980.130.160.21

1 2 3

(a)

Page 61: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

47

Figure 3.7: Enlarged view of the three Coulomb peaks marked with arrows in Figure 3.6a (gray curves) and the best

fit to theses peaks (black curves).

Around the labeled Coulomb diamonds in Figure 3.6b, we do not observe conductance stripes

running in parallel with edges of diamonds, which means excited states of QD is blurred by the

thermal energy , i.e., ΔE < . This situation is further confirmed by the quasi-periodicity

of Coulomb oscillations for the marked Coulomb peaks and diamonds in Figure 3.6. Eq. (1.4)

can be simplified to ∆ ⁄ under the condition that ΔE EC. From Figure 3.6b we can see

that for the labeled diamonds EC is around 10 meV, so the above condition is satisfied. Using a

general expression for the quantum confinement energy [73]: ∆ ~ ⁄ , and assuming the

electron temperature to be 15 K (which gives a satisfactory consistency of lever arms estimated

by Eq. (1.5) compared with values obtained by CI model, as described above), we have

⁄ < 1.3 meV, which gives L > 320 nm. Since the smallest energy quantization of a QD

is determined by its longest dimension, the above result means the length of the QD is at least

320 nm, which is most likely formed between the two contact barriers.

0

0.002

0.004

0.006

0.008

0.01

1.15 1.17 1.19 1.21 1.23 1.25 1.27

G (2

e2/h)

Vg (V)

Original peaks

Fitted peaks

Page 62: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

48

3.4.3 Transport Phenomena in the Short-etched Section — High

Back-gate Bias

At high back-gate bias, the contacts become more transparent and the Fermi energy overcomes

most of the potential barriers in the channel, so the carrier transport is more likely to take place

in a clean 1-D system. This intuition is supported by the linear conductance response taken at

=2.7 V, which is shown in Figure 3.8. Quasi-periodic oscillations can be seen superimposed

on the first conductance plateau and the pinch-off regime. Such superimposition of oscillations

on the conductance plateaus has been attributed to F-P electron interference [15, 16] as

introduced in Section 1.2.3, and suggests that the transport in our NW is likely to be ballistic. If

not, the carrier’s path length would have been randomized by scattering leading to the

randomization of the phase and smearing of F-P oscillations [17]. Compared with the “noisy”

pattern in low back-gate bias region (Figure 3.5a), random conductance switching is absent here.

Similar switching events observed in InAs NWs have been modeled by the scattering of

electrons from a gate-voltage-dependent potential barrier produced when a defect captures an

electron [44]. As the back-gate bias is increased, the Fermi level in the channel may be well

above the trap level and therefore may overcome the scattering potential generated by the

electron trapping events. This viewpoint explains the suppressed random conductance switching

in the high back-gate bias regime. One interesting feature is the overshooting at the onset of each

plateau, which manifests itself as two thick stripes in Figure 3.3b. This phenomenon has been

attributed to the tunneling via discrete impurities [74], back scattering from defects [75] and

enhanced many-body interactions [76]. Another striking feature is the occurrence of F-P

oscillations in the pinch-off region, which, to our knowledge, has never been reported before.

This feature differs significantly from those in the low back-gate bias case where Coulomb

blockade due to electrons tunnel through either MTJ or a single dot is observed. Electron

interference requires that electron waves are phase coherent. In the pinch-off region electrons in

the contacts face a potential barrier in the channel. If the observed conductance is due to the

thermionic emission of electrons over the barrier, the phase information of electron waves cannot

be maintained. Electron transport by tunneling through the barrier is phase coherent, but a more

quantitative investigation needs to be done to better understand this mechanism in our device.

Page 63: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

49

Figure 3.8: Linear conductance response of the device at Vbg =2.7 V, corresponding to the light green arrow in

Figure 3.3b. Inset: Linear conductance response of the device at Vbg =3 V, corresponding to the purple arrow in

Figure 3.3b.

To further demonstrate the existence of electron interference, bias spectroscopy was performed

in the pinch-off and overshooting region at Vbg = 3 V (the linear conductance response at Vbg = 3

V is shown in the inset of Figure 3.8) and the first conductance plateau at Vbg = 2.7 V (indicated

by the rectangle in Figure 3.8). The result is shown in Figure 3.9d-f, with corresponding linear

conductance shown in Figure 3.9a-c. Clear chess-board pattern in the grayscale nonlinear

conductance plot, which is a signature of F-P interference as introduced in Section 1.2.3, can be

seen in Figure 3.9d-e. The interference pattern diminishes as |VDS| is increased, suggesting the

occurrence of heating or dephasing when the electron energy distribution deviates from

equilibrium [13]. For the nonlinear transport on the first conductance plateau (Figure 3.9f), the

conductance peaks around zero source and drain bias due to constructive interference are barely

seen. This may indicate a strong role of the above mentioned mechanism.

0.0 0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.20.00

0.05

0.10

0.15

0.20

0.25

G(2

e2 /h)

Vsg(V)

-0.2 0.0 0.2 0.4 0.60.00

0.02

0.04

0.06

0.08

0.10

0.12

G(2

e2/h

)Vsg(V)

Page 64: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

50

Figure 3.9: Linear conductance in the pinch-off region (a) and the following overshooting region (b) at Vbg = 3 V

(along the purple arrow in Figure 3.3b). (c) Linear conductance on the first plateau at Vbg = 2.7 V (along the red

arrow in Figure 3.3b). (d) - (f) Bias spectroscopy corresponding to the linear conductance in (a) - (c). The grayscale

bars are in the unit of 2e2/h.

In Figure 3.9d, a decrease of the critical value VC mentioned in Section 1.2.3 with increased side-

gate voltage can be clearly identified. The condition which determines VC described in Section

1.2.3 implies that the length of the cavity should be inversely proportional to VC. Therefore, the

decrease of VC reflects the increase of the cavity length as side-gate bias is increased. In principle,

we can also use Eq. (1.12) to estimate the cavity length. However, in this device side-gates are

coupled to the back-gate and this coupling changes with the voltage differences between the two

gates. The resulting gate capacitance per unit length required for Eq. (1.12) may not be constant.

A more reliable way to estimate the cavity length is to use the condition that determines VC. In

the numerical integral of Eq. (1.13), can be obtained by performing a 3-D electrostatic

simulation using the same device geometry. Another remarkable feature of Figure 3.9d-e is the

different slopes of right and left sloping differential conductance lines, which may arise from

different electron reflection probabilities at the two interfaces [13].

0.000

0.003

0.006

0.009

0.012

0.00

0.03

0.06

0.09

0.12

0.04

0.05

0.06

0.07(a)

(b)

(c)

Vsg(V)

-0.3 -0.2 -0.1 0.0 0.1 0.2-30

-20

-10

0

10

20

30

(d)2.0E-04

0.0010

0.0055

0.029

0.15

0.30 0.36 0.42 0.48 0.54 0.60

(e)0.0040

0.063

0.12

0.18

0.22

0.80 0.88 0.96 1.04 1.12 1.20

(f)0.040

0.065

0.11

0.17

0.28VD

S(m

V)G

(2e2 /h

)

Page 65: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

51

3.4.4 Transport Phenomena in the Long-etched Section

A series of measurement was also performed for the long-etched section, in a similar way as

what has been done for the short-etched section. The main difference here is that Coulomb

blockade dominates the physics in the pinch-off region even at high back-gate biases, unlike the

case for the short-etched section where F-P oscillations extend all the way to the complete pinch-

off at high back-gate biases. Figure 3.10 is a representative measurement result taken at Vbg = 2

V. The result is separated into two parts to magnify the feature near the complete pinch-off

region (Figure 3.10a and c). Similar to Figure 3.9, here Figure 3.10a and b are the linear

conductance response and Figure 3.10c and d are the corresponding bias spectroscopy. Familiar

F-P oscillations and interference pattern can be identified in Figure 3.10b and d, but in Figure

3.10a and c we see Coulomb peaks and irregular diamonds, which suggests electron tunneling

through MTJ. Since the only difference between the two sections is the length of the etched

constrictions, the comparison indicates that transport through longer constriction may experience

stronger potential fluctuation, in consistent with the result of dry-etched constrictions [48].

Figure 3.10: A representative measurement result taken at Vbg = 2V for the long-etched section. (a) Linear

conductance in the pinch-off region. (b) Linear conductance in the following sub-threshold region. (c) and (d) Bias

spectroscopy corresponding to linear conductance in (a) and (b). Scale bars are in the unit of 2e2/h.

0.0000

0.0005

0.0010

0.0015

0.0020

0.0025

(d)(c)

G(2

e2 /h)

(a) (b)

-0.60 -0.55 -0.50 -0.45 -0.40 -0.35-30

-20

-10

0

10

20

30

(a)

Vsg(V)

VD

S(m

V)

-8.0E-04

0.0017

0.0057

0.00970.026

-0.35 -0.30 -0.25 -0.20 -0.15 -0.10 -0.05 0.00

0.0

0.023

0.045

0.068

0.090

0.000

0.005

0.010

0.015

0.020

0.025

Page 66: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

52

3.5 Conclusions

Transport measurement was performed on InAs NWs with etched constrictions. Some etched

NW devices have pairs of side-gate in the vicinity of etched constrictions. A universal side-

gating effect observed in all side-gated devices was first presented and the observed hysteresis

was attributed to carrier injection at high gate fields due to the sharp tip of side-gates. For the

measured device without side-gates, we did not observe conductance quantization. When the

channel was turned on, the conductance trace fluctuated with a magnitude of e2/h as a function of

back-gate voltage, a possible indication of UCF in a weakly disordered 1-D channel. When the

channel was turned off, i.e., in the pinch-off region, Coulomb blockade arise from electron

tunneling through DQD was observed. For the measured device with side-gates, non-constant

coupling from side-gates and back-gate to the channel was discovered from the linear

conductance map as a function of side-gate and back-gate bias. A simple electrostatic model was

proposed to account for this phenomenon. Conductance plateaus were observed in this device. F-

P oscillations were found in the sub-threshold regime and on the first conductance plateau. At

low back-gate bias regime, close to the pinch-off region, signs of electron tunneling through

either multiple QD or single QD were observed at different biases. RTS was found being

superimposed onto the conductance trace for the whole side-gate voltage range. At high back-

gate bias regime, RTS disappeared as a result of Fermi level overcoming the trap level. Near the

pinch-off region, transport in the long-etched section still exhibited the sign of tunneling through

MTJ, whereas in the short-etched section, F-P oscillations extended all the way to complete

pinch-off. The latter phenomenon has never been reported before in the literature and requires

further investigation.

Page 67: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

53

Chapter 4

Conductance Quantization in a Thick InAs

Nanowire with Side-gating

4.1 Introduction

In Section 1.2.2, we discussed the conductance quantization in a 1-D conductor as a result of the

formation of 1-D subbands. At low temperature where subband spacing significantly excesses

the thermal energy , as long as the transport is ballistic or quasi-ballistic, the conductance

quantization should be resolved. As a 1-D/quasi-1-D system, NW provides a natural platform to

explore the 1-D transport phenomenon. Nonetheless, very few related result have been reported

in NW system [9, 10], presumably due to the blur of conductance quantization by

impurity/surface scattering. From this perspective, our MBE-grown InAs NWs possess the

advantage of producing relatively cleaner channels compared with those grown by other growth

techniques due to the unique principle MBE technology [77].

In designing the experiment to study conductance quantization in single NW devices, the

consideration involves a trade-off between the two prerequisites for observing this phenomenon.

Apparently, thinner NWs have larger subband spacing owing to stronger lateral confinement, so

it is easier to satisfy the requirement of ∆ . Meanwhile, thinner NWs also suffer stronger

surface scattering which dramatically reduce the transmission probabilities of conductance

plateaus [78, 79].

In this chapter we focus on an InAs NW section without etching. A thick NW was picked mainly

for the sake of obtaining a deep etching profile on one section of the NW. Figure 4.1a is the SEM

image of the fabricated device. The diameter of the NW is 95 2 nm. The distance between the

Page 68: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

54

source and drain contact is about 800 nm. The distances between side-gates and the NW are 60

nm and 140 nm, respectively. Figure 4.1b is the gray scale profile along the yellow dashed line in

Figure 4.1a from which we obtain more accurate measure of dimensions.

Figure 4.1: (a) SEM image of the device. This chapter focuses on the measurements on the section without etching.

(b) Gray scale profile along the yellow dashed line providing a more accurate measure for the geometry dimensions

(c) Linear conductance as a function of side-gate and back-gate biases. Stripes with almost constant slopes indicate

independent coupling to the channel from the two gates. Experiment results along the two arrows are presented in

this chapter.

Other than the measurement discussed in Section 4.3.3, all other measurements were performed

at T= 10 K. The AC source and drain bias was set to a RMS of 0.3 mV with a frequency of 210

Hz. The linear conductance as a function of side-gate and back-gate biases (see Figure 4.1c) was

first measured to get guidance for the subsequent experiments. A series of stripes with almost

constant slope can be identified from this plot, which differs significantly from the case

described in Section 3.4.1. Gate leakage currents Isg and Ibg were very small within the bias

ranges, which means the coupling between side-gate and back-gate can be neglected. This

confirms the model proposed in Section 3.4.1, where relative changes of gate-channel coupling

strength in that device are caused by changes of the coupling between the two gates. From the

slope of the stripes we obtain the ratio of gate capacitance ⁄ 5.5. This result is in

0 100 200 300 400 500 600 700 8000

40

80

120

160

200

Gra

y sc

ale

D istance (nm )

95 60140

-8 -6 -4 -2 0 2 4 6 8-8

-6

-4

-2

0

2

4

6

8

Vsg(V)

Vbg

(V)

0.00

0.984

1.90

2.82

3.70

4.50

G(2e2/h)

(a) (c)

(b)

Page 69: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

55

agreement with our expectation since side-gates only have a local effect on the channel. Besides,

as mentioned in Section 3.2, owing to the proximity effect during the exposure, the end of side-

gate forms a sharp tip, which would further reduce the side-gate coupling to the channel.

Therefore, as long as the back-gate bias is small, the back-gate always has a much stronger

coupling to the channel than the side-gate does.

Most results presented in this chapter are based on measurements along the yellow arrow in

Figure 4.1c, where side-gates and back-gate were connected together. The only exception is the

magnetotransport measurement described in Section 4.3.2, where the back-gate was grounded

and the side-gate bias was scanned, i.e., along the cyan arrow in Figure 4.1c. To explain the main

experimental observations, a 3-D electrostatic and 2-D eigenvalue coupled simulation was

performed and the results are presented in Section 4.4 and 4.5.

4.2 Data Processing

This section describes the necessity of performing some processing to the raw data. Figure 4.3a

shows the traces of differential conductance G versus DC source and drain bias VDS (only the

positive scan of VDS, i.e., from -30 mV to 30 mV is plotted) for different gate voltages Vg. The

gate voltage was increased from -8 V to 8 V stepped by 0.05 V between successive VDS sweeps.

One noticeable feature of this plot is the dip of the traces around zero DC bias. To investigate the

origin of this dip, we connected two successive pins of a LCC with a 10 KΩ surface mount

resistor and loaded this test chip into the measurement system. The only difference between the

test measurement and the sample measurement is the replacement of the sample with the test

resistor, all other features were kept the same. The same G-VDS measurement was performed and

the result is shown in Figure 4.2c (red curve, both positive and negative scan of VDS is plotted).

The same dip appears, indicating that it is a feature of the measurement system rather than a

feature of the sample itself. So effort needs to be made to correct this system error, so that real

sample character can be resolved.

The sample resistance in this plot ranges from a few KΩ to a few MΩ. To check how the dip

feature evolves with the change of resistance, we replaced the 10 KΩ resistor with a 1 MΩ one

Page 70: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

56

and the G-VDS relation is shown as the black curve in Figure 4.2c. G is magnified 100 times for

comparison with the 10 KΩ case. The dip is still visible for the 1 MΩ resistor. To compare the

peak-to valley ratio for the two cases, G is first averaged from the positive and negative scan via

/2 where and represent the G value at a given

VDS from positive and negative scan, respectively. After this symmetric treatment, the peak can

be defined as the average outside the dip region and the valley is defined as the average inside

the dip. The calculated peak-to valley ratio for both cases is 1.03, meaning the dip is independent

of resistance. Thus, we can correct the dip for the whole range of the measured conductance. By

multiplying each conductance in the valley with the peak-to-valley ratio, the dip is compensated

to an average conductance level outside the dip, as shown in Figure 4.2d.

Page 71: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

57

Figure 4.2: Raw data of the finite bias spectroscopy and associated data processing. (a) Raw data of G-VDS traces

(only the positive scan of VDS, i.e., from -30 mV to 30 mV is plotted) for different Vg. Vg was increased from -8 V to

8 V stepped by 0.05 V between successive VDS sweeps. (b) G-VDS traces after the subtraction of “self-gating” effect

and compensation for the dip near zero source and drain bias based on the raw data in (a). (c) G-VDS traces of testing

resistors, demonstrating the irrelevance between the dip feature and the sample conductance. (d) Correction for the

dip feature. By multiplying the conductance in the dip region with the calculated average peak-to valley ratio, the

dip is compensated to an average conductance level outside the dip.

Another feature of Figure 4.2a is the asymmetry of the traces with respect to VDS. Kristensen et al.

have attributed this asymmetry to the effect of VDS influencing the confinement potential and

termed as “self-gating” [80], i.e., the confinement potential is not only defined by the sample

parameters, the geometry and the gate bias, but also, to some extent, by the DC source and drain

(b)(a)

-30 -20 -10 0 10 20 300.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

-30 -20 -10 0 10 20 30

G(2

e2 /h)

VDS(mV)

-30-25-20-15-10 -5 0 5 10 15 20 25 30

0.95

1.00

1.05

1.10

1.15

G(2

e2 /h)

VDS(mV)

10K 1M

-30-25-20-15-10 -5 0 5 10 15 20 25 30VDS(mV)

before correction after correction

G(2e

2/h)

1.11

1.12

1.13

1.14

1.15

1.16

1.17

(c) (d)

Page 72: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

58

bias. Following the same correction method reported in literatures [80, 81], this trivial

electrostatic effect is subtracted from the data by using | | /2,

making the G-VDS traces perfectly symmetric with respect to VDS. Figure 4.2b shows the G-VDS

traces after the subtraction of the “self-gating” effect and the compensation for the dip.

4.3 Evidence of Subband Quantization

4.3.1 Linear and Nonlinear Transport at Low Temperature

The G-VDS traces in Figure 4.2b exhibits a series of “bundles” with high density of traces. Since

all traces are separated in Vg equally, bundles therefore represent situations where G does not

change with Vg, which are exactly the conductance plateaus described in Section 1.2.2. Bundles

around zero bias are linear conductance plateaus whereas those at high bias can be classified into

half plateaus because their conductance values locate roughly in the middle of two successive

linear conductance plateaus’ values.

The linear conductance as a function of Vg is plotted in Figure 4.3a. Eleven plateaus can be

identified superimposed with small oscillations. Similar oscillations on conductance plateaus

were attributed by Biercuk et al. as Fabry-Perot interference [82]. The inset of Figure 4.3a shows

an enlarged view of the 2nd plateau from which we can see that these oscillations are quasi-

periodical in Vg with an average period of ∆ 0.12V. Plugging in the calculated gate-to-wire

capacitance per unit length Cg 30 aF (to be discussed in Section 4.4.2) into Eq. 1.12, we obtain

the effective length of the F-P cavity L 90 nm. This could be explained by the presence of a

scatterer such as a point defect located 90 nm away from a contact. The first nine plateaus

have an average plateau height of about 0.25 times conductance quantum G0 (2e2/h), which

implies that some electron scattering processes prevent full transmission of electrons from the

source to the drain. The heights of the 10th and the 11th plateaus, however, are about 0.5G0.

Another interesting feature is the change of Vg span on each plateau—decreases from about 1.0

V for the first four plateaus to about 0.5 V for the middle five plateaus, and then increase to

Page 73: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

59

about 1.0 V for the last two plateaus. Explaining these two phenomena is the main motivation for

performing a simulation and will be covered in full detail in Section 4.4.

Figure 4.3: (a) Linear conductance as a function of gate voltage. 11 conductance plateaus can be identified with the

middle 5 ones marked by arrows. Inset: A zoom-in of the 2nd plateau, showing the small oscillations superimposed

on the plateau. (b) Gray scale plot of the normalized absolute value of transconductance as a function of gate voltage

and source/drain bias. Diamond patterns around zero bias corresponding to linear conductance plateaus in (a) are

visible. (c) Gate voltage sweeps of G for fixed values of VDS from 0 to 30 mV, stepped by 3 mV. Black circles

highlight half plateaus.

(a)

(b)0.00.51.01.52.02.53.03.54.04.5

111098

765

432

G(2

e2 /h)

1 Vg(V)

G(2

e2 /h)

-6.0 -5.8 -5.6 -5.4 -5.2 -5.00.3

0.4

0.5

-8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8-30-20-10

0102030

Vg(V)

V DS(

mV)

0

1|dG/dVg| (a.u.)

(c)

-8 -6 -4 -2 0 2 4 6 8 10 12 14 16 180.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

G(2

e2 /h)

Vg(V)

Page 74: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

60

Half plateaus can be better resolved by plotting gate voltage sweeps of G for fixed values of VDS

as shown in Figure 4.3c. The left-hand trace was measured at VDS = 0 and the successive traces,

offset to the right for clarity, were incremented in steps of 3 mV up to 30 mV. For finite bias, the

integer plateaus becomes less pronounced as VDS is incremented up to 21 mV while additional

structure develops at conductance values approximately midway between the original integer

values, as indicated by black circles. At higher VDS all the structure becomes smeared out.

Rossler et al. pointed out that half plateaus can only be observed in clean samples, presumably

because scattering events in the NW become more likely when more unoccupied subband states

are energetically available at larger VDS [83]. At even higher bias, the conductance is usually

obscured by noise [50] or increases/decreases due to various self-gating effects [80]. Although

we do not observe the retrieval of integer plateaus, the presence of half plateaus can be

interpreted as the result of the cleanliness of the channel, which reduces the probability for

backscattering.

In Section 1.2.2 we mention that subband spacing can be extracted directly from the stability

diagram of a 1-D system. We can also plot the stability diagram in terms of transconductance

(dG/dVg, which is calculated by numerical differentiation from the measured differential

conductance G = dIds/dVds) rather than the differential conductance. The transconductance is zero

(or small) on conductance plateaus and show peaks in the transition regions between plateaus.

Such a stability diagram therefore emphasizes the boundaries of diamond patterns and makes the

identification of the intersections much easier. Figure 4.3c is the gray scale plot of the

normalized absolute value of transconductance calculated from the data in Figure 4.2b. A series

of diamonds around zero bias corresponding to linear conductance plateaus in Figure 4.3a are

visible. We will leave further discussion of these diamonds for Section 4.4.2 where a comparison

between extracted subband spacing from this diagram and calculated subband spacing from a

theoretical model will be provided.

4.3.2 Linear Transport in a Magnetic Field

To study the magnetotransport properties of the device, a magnetic field perpendicular to the

sample plane was applied and the liner conductance was measured as a function of side-gate bias

Page 75: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

61

(Vbg was fixed at 0 V) at different field intensities. Figure 4.4 shows the linear conductance

response at B = 0 T (black curve) and 0.6 T (red curve). Four plateaus as labeled can be

identified in the scanned side-gate bias range, corresponding to the 5th -8th plateaus in the G-Vg

curve. Two distanced features arise at B = 0.6 T: The conductance on each plateau increases and

there is an additional half plateau appearing between the 3rd and the 4th plateaus (indicated by the

blue dashed ellipse in Figure 4.4).

Figure 4.4: Magnetotransport characters of the device. Differential conductance was measured as function of side-

gate bias without magnetic field (black curve) and with a magnetic field of 0.6 T which is perpendicular to the

sample plane (red curve). Four plateaus labeled with numbers can be identified. Compared to the case where B = 0 T,

at B = 0.6 T, conductance on each plateau increases and there is a half plateau appearing between the 3rd and the 4th

plateaus possibly due to the Zeeman splitting of a spin-degenerate level, as indicated by the blue ellipse. The

distance in terms of Vsg between the split plateaus and the original plateau is marked by the green dashed lines.

The first phenomenon is a possible indication of the reduced backscattering. As introduced in

Section 1.2.2, the presence of a static magnetic field perpendicular to the wire axis can result in

the formation of hybrid magneto-electric subbands which are a mixture of the quantization due to

the lateral confining potential of the wire and the quantizing effect of the magnetic flux in terms

of Landau level formation. As the magnetic field increases, more and more current is carried by

the edge states that can greatly suppress the backscattering. A simple estimation can provide

-8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8

1.4

1.6

1.8

2.0

2.2

2.4

2.6

2.8

G(2

e2 /h)

Vsg(V)

B=0T B=0.6T

1

2

3

4

Page 76: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

62

some insights about the emergence of edge states. At B = 0.6 T, the cyclotron radius of the 1st

Landau level is 0 33 nm. From the perspective of classical motion, up to 66

nm away from the boundary, the injected electrons can still scattered by the boundary and form

skipping orbit. So there should be significant contribution to the conductance from the edge

states.

The emergence of the half integer plateau between the 3rd and 4th plateaus is possibly due to the

Zeeman splitting of a spin-degenerate energy level as introduced in Section 1.2.2. Assume the g

factor of the InAs NW to be 15, which equals its bulk value, then at B = 0.6 T, 0.5 meV.

As will be discussed in Section 4.4.2, the predicted spacing between the 7th and the 8th subbands

by a theoretical model is about 2.5 meV. At B = 0 T, the two plateaus are spaced about 4.5 V

apart in terms of side-gate bias. Accordingly, the two split plateaus should be located equidistant

from the two sides of the original plateau with a side-gate bias range of 0.45 V, assuming the

capability of the side-gate to tune the energy levels in the channel does not change with the

applied magnetic field. From Figure 4.4, the side-gate spacing between the two split plateaus and

the original plateau is marked by the green dashed lines and is about 0.75 V. This discrepancy

could arise from a faulty previous assumption about gate control, and/or a change of subband

spacing due to the effect of Landau quantization.

4.3.3 Linear Transport at Higher Temperatures

The linear conductance as a function of gate bias was also measured at temperatures higher than

10 K and the result is shown in Figure 4.5. The conductance responses at T = 20 K and 30 K are

shifted for clarity. Most plateaus disappear as temperature increases from 10 K to 30 K. As is

introduced in Section 1.2.2, the conductance plateaus eventually disappear when the separation

between mode energies is comparable to, or smaller than a few (use 3 here). The fact

that plateaus can be observed at 10 K and are smeared out at 30 K gives a rough lower and upper

boundary for the subband spacing: ∆ ~ (2.6 meV, 7.8 meV), which is consistent with the

observed subband spacing from the transconductance map.

Page 77: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

63

Figure 4.5: Liner transport at higher temperatures. G-Vg curves corresponding to T = 20 K and 30 K are shifted for

clarity. Most conductance plateaus are washed out at T = 30 K, providing a rough estimation for the subband

spacing. Inset: G-Vg curve at room temperature, from which the field-effect mobility can be extracted.

The inset of Figure 4.5 is the G-Vg relation at room temperature. Curves corresponding to

positive and negative Vg scan are presented, showing little hysteresis. The difference between the

traditional transistor character at room temperature and the conductance quantization at low

temperature is distinct. From this figure we can extract the gate voltage dependent field-effect

mobility using

(4.1)

where L is the channel length and is the summation of back-gate capacitance and side-

gate capacitance and the quantity can be obtained by numerical differentiation of the

measured G-Vg characteristic. We can apply an analytical expression for the back-gate

capacitance of a cylinder separated from a conducting plane by a dielectric slab:

/ (4.2)

-8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 80.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

5.5

G(2

e2 /h)

Vg(V)

30K 20K 10K

-5 -4 -3 -2 -1 0 1 2 3 4 50

1

2

3

4

G(2

e2 /h)

Vg(V)

room temperature

Page 78: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

64

where R is the NW radius, t is the thickness of the dielectric layer and = 2.2 is an effective

gate dielectric constant taking the inhomogeneous dielectric surroundings of the NW into

account [84]. Using in the parameters of our device (L = 800 nm, t = 100 nm and R = 50 nm)

gives 55 aF. Using the fact that ⁄ 5.5, we obtain 10 aF and in turn

= 65 aF. Substituting into Eq. (4.1) gives the peak field-effect mobility

7900 cm2/(V-s). However, Eq. (4.2) assumes a metallic channel with infinite length. It has been

reported that relaxing these assumptions leads to a reduction of about 30% of close to the

depletion regime of a thin Si NW [85]. In addition, the presence of side-gates in our device

would screen the electric fields of the back-gate and further decreases . Therefore,

calculated by applying Eq. (4.2) for will be overestimated and 7900 cm2/(V-s) is actually the

lower boundary of the peak field-effect mobility .

As will be shown in Section 4.4.2, the calculated using a 3-D electrostatic simulation which

accounts for all the above factors gives 24 aF. Using this result, 21000 cm2/(V-s).

Recently Ford et al. studied the diameter-dependent mobility of InAs NWs [79] and found the

mobility increases with increasing diameter, which the authors attributed to the reduction of

surface roughness scattering. Another study [58] using the InAs NWs grown by similar

conditions as the NWs used for this thesis showed similar trend, and the mobility should be

around 5000 cm2/(V-s) for an InAs NW whose diameter is 100 nm, assuming a liner relationship

between the mobility and the diameter. The estimated here, whether based on Eq. (4.2) or

the simulation, is much higher than 5000 cm2/(V-s), which indicates a dramatic reduction of the

impurity scattering for the NW studied in this Chapter, which in turn supports the observed

quasi-ballistic nature of the transport in this device.

4.4 Three-dimensional Electrostatic and Two-dimensional

Eigenvalue Coupled Modeling of the Nanowire System

Although the presence of linear conductance plateaus and finite bias half plateaus provide strong

evidence of subband quantization, and from the temperature-variate measurement we deduct a

Page 79: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

65

rough range of 2.6 ~ 7.8 meV for the subband spacing, detailed band structure remains to be

unclear. If subband spacing is almost constant and does not vary gate bias (such as the subband

quantization due to a saddle point potential introduced in Section 1.2.1), both the gate span and

the height of each conductance plateau should be roughly the same, assuming the capability of

tuning the energy in the channel by the gate bias does not change and the transmission

probabilities for each mode are identical. As mentioned in Section 4.3.1, the linear conductance

response in Figure 4.3a exhibits variable gate span of each plateau and doubled height for the last

two plateaus. We understand these phenomena as signs of deviation from constant equally-

spaced subband structure. Very recently, Ford et al. have used the subband spacing calculated

from the simple “particle in a cylinder” model combined with a derived relation between the gate

bias and the energy in the channel to explain the observed conductance plateaus in InAs NWs

[10]. However, this procedure is only valid for a thin NW (the thickest NW they measured has a

diameter of 35nm) so that the lateral potential difference inside the NW can be neglected. The

diameter of our NW is 95nm, the potential distribution from the bottom to the top of the NW

may play an important role in modifying the solution of 2-D Schrodinger equation and in turn

dramatically deviate the subband spacing from the “particle in a cylinder” case. Our instinct was

further supported by the failure of trying to use the subband spacing of the “particle in a cylinder”

case to explain our data, as will be shown in Section 4.4.2. The above considerations form the

motivation of performing a detailed simulation to gain better insight into the experimental results.

4.4.1 Model Description

The strategy of the modeling is getting the potential distribution in the NW from a 3-D

electrostatic simulation and coupling it to the 2-D Schrodinger to obtain the subband spacing.

The geometry of the 3-D simulation (see Figure 4.6) is exactly the same as the real device with

all dimensions obtained by fabrication parameters and SEM images. The potential distribution

V(x, y, z) is governed by the 3-D Poisson equation

· , , , , (4.3)

where , with each dielectric material volume assigned their respective values of (15.5

for InAs, 3.9 for SiO2 and 1 for the vacuum). The space charge density is set to be

Page 80: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

66

| | in the NW and zero otherwise. Here, is the uniform

background density of ionized donors, and are 3-D electron and hole concentrations

and are given by

2 / | | (4.4a)

2 / | | (4.4b)

where is the effective mass of electrons in InAs, * * 3/2 * 3/2 2/3( )dh lh hhm m m= + is the effective mass

of holes in InAs with contributions from light holes ( ) and heavy holes ( ), and are

conduction band edge and Valance band edge, respectively, is the Fermi level of the system,

1 exp is the complete Fermi-Dirac integral of order j

with Γ being the gamma function.

Charging of the surface states is included using a simple model employing a spatially continuous

surface state density with a uniform energy distribution. The surface charge density is

| | | |

| | (4.5)

where is the surface states, | | , | | , is the surface

potential and is the surface charge neutrality level described in Section 1.2.4.

Page 81: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

67

Figure 4.6: Screenshot of the 3-D electrostatic simulation result, showing the simulated geometry and the potential

distribution with source and drain grounded and Vbg = Vsg = -7 V.

The 3-D simulation was carried out using the electrostatic solver of a commercial finite-element

analysis (FEA) software COMSOL Multiphysics. Dirichlet boundary conditions were used for

the back and side-gates (Vbg = Vsg = Vg) and source/drain electrodes (VS = VD = 0), since the main

goal of the model is to explain linear transport phenomena observed with back-gate and side-

gates connected. Neumann boundary condition, i.e., the electric field perpendicular to the

boundary planes is zero, was applied to the top and side boundaries of the entire simulation

domain. The size of the domain was chosen to be large enough so that a change in simulation

results with further increase in domain size would be negligibly small. The potential distribution

in the simulation domain is shown in Figure 4.6 with Vbg = Vsg = -7 V.

The x axis is assumed to be the transport dimension and is centered at x = 0 nm. The calculated

potential distributions V(x, y, z) at different cross sections, x = xc, of the NW serve as input to the

2-D Schrodinger equation

, , , , (4.6)

Page 82: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

68

This equation was solved at different Vg within the NW domain using Dirichlet boundary

condition ( 0), and the eigenvalue solver of COMSOL Multiphysics. Subband quantization

as a function of gate bias and the position along the NW can therefore obtained once E is

computed. A key assumption of the coupled model is that the linear conductance of the system is

determined by the subband quantization at the middle slice (x = 0) of the NW (which will be

justified in Section 4.5), so Eq. (4.6) was only solved with 0, , for different Vg.

From the expression of space charge density we see that this model assumes two sources for the

free carriers inside the NW: background impurities inside the NW and surface charges due to

surface states. Their relative contributions to the space charge density provide one more

parameter to play with. The charge neutrality condition is satisfied at Vg = 0, where potential is

zero everywhere and charges due to the free carriers in the NW are compensated for by

background impurities in the NW and surface states at fixed positions. This condition is

equivalent to treating the NW as “floating” in the vacuum. Thus, a simple 2-D electrostatic

simulation that mimics the floating situation could be adopted, from which could be

estimated once the portion of free electrons coming from surface states was determined. To

reduce the number of fitting parameters in this coupled model, was set to be 150 meV above

. Following this choice, and the portion of free electrons coming from surface states at Vg =

0 were adjusted until both the 3-D electrostatic simulation and the 2-D eigenvalue simulation

gave an identical depletion (threshold) voltage of Vg -7.7 V, which is consistent with the

experimental result.

4.4.2 Results of the Model

The best fit gives = 82 meV above and almost 100% of the portion of free electrons

coming from surface states at Vg = 0. The latter reinforces the reason to observe the conductance

quantization in a thick NW at low temperature: there is little backscattering in the channel and

the effect of surface scattering is also reduced for a thick NW. The calculated energy

quantization at the center of the NW as a function of Vg is obtained. Subbands with energy levels

below are occupied, and the number of occupied subbands versus Vg is plotted in Figure 4.7

(red curve). A series of steps (labeled with corresponding number) resemble the observed

Page 83: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

69

conductance plateaus and the two main features of experimental data are basically met here: the

gate span of each step decreases from about 1 V for the first few steps to about 0.5 V for the

following steps, and then increase to above 1 V for the 10th and 11th steps. Starting from the 9th

step, subbands become two-fold degenerate, tuning the gate bias would therefore dragging two

subbands across the Fermi level at the same time, giving rise to a conductance plateau that is

doubled in height. For comparison, the energy levels calculated from the “particle in a cylinder”

model are shifted by the average electrostatic energy of the middle slice of the NW (

| | 0, , ) at different Vg, obtained from the 3-D electrostatic simulation, to

get the occupied number of subbands below . This N-Vg relation is also plotted in Figure 4.7

(black curve). This approach was adopted by Ford et al. [10] and three features can be concluded

from the result: first, the onset of the first step happens at Vg -4.7 V, which is not consistent

with the calculated threshold voltage of the 3-D electrostatic simulation; second, the gate span of

each step does not follow the observed trend; third, most steps are double-heighted due to the

onset of two-fold degenerate subbands. The comparison between the two modelling approaches

is straightforward: if we only carry out the 3-D electrostatic simulation and use the calculated

average potential to directly shift the energy levels corresponding to the well known “particle in

a cylinder” problem, the result is far from being self-consistent and deviates severely from the

observed features. They therefore cannot be used to explain our data. On the other hand, our

coupled model, while showing deviation, captures the main features of the experimental data,

and therefore stands out from its counterpart.

Page 84: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

70

Figure 4.7: Simulated number of occupied subbands as a function of gate bias. Red curve: result from the coupled

simulation carried out by the author. Labeled numbers represent the resembled conductance plateaus. Black curve:

result from the approach adopted by Fort et al [10].

The deviation of subband spacing from the case of “particle in a cylinder” can be understood by

considering the screening effect from free carriers in the NW. When the channel is enhanced at

positive gate bias, the free electron density is high. The channel behaves more like a metallic one

in this case, strongly screening the electric field due to the potential difference between the gate

and the contacts. The potential variation inside the NW is thus negligible. When the channel is

depleted at large negative gate bias, however, the free electron density is low. The screening

effect is strongly suppressed in this case, generating large potential differences across the NW

(can be up to several hundred mV, see Figure 4.10). It is this nonuniform potential distribution

inside the NW that strongly modifies the electronic confinement potential and the subband

spacing. If the NW is thin, the absolute value of threshold voltage is usually small. This means

the nonuniformity of potential distribution inside the NW at depletion is not severe, so we only

need to consider the geometrical confinement, i.e., the “particle in a cylinder” model alone is

enough to describe the subband spacing. This model predicts that certain subbands show two-

fold degeneracy arising from the structural symmetry of cylindrical NW. As mentioned above,

-8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 50

4

8

12

16

20

24

28

32

36

40

44

48

num

ber o

f occ

upie

d su

bban

ds

Vg(V)

3D electrostatic simulation + "particle in a cylinder" model 3D electrostatic simulation + 2D eigenvalue simulation

3 4 5 6 7 89

1011

1 2

Page 85: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

71

when the gate bias is close to 0, starting from the 9th step, the step height doubled owing to the

two-fold degeneracy of subbands. This is a strong indication of the recovery of symmetrical

confinement with the absence of the nonuniform potential distribution. Figure 4.8 shows the

calculated probability density profiles, | , | , for the five lowest lateral modes (subbands) at

Vg = -6 V. No degeneracy is observed here and the wave functions are pushed up and sideways

by back and side-gate bias, which differ significantly from the “particle in a cylinder” case.

These profiles therefore confirm our point that the nonuniform potential distribution inside the

NW at large negative bias breaks the symmetry and strongly modifies the electron confinement,

and in turn, the subband spacing.

Figure 4.8: Calculated probability density profiles at Vg = -6 V for the five lowest lateral modes with corresponding

energy levels with respect to conduction band edge labeled. No degeneracy is observed and the wave functions are

pushed up and sideways by back and side-gate bias, indicating the effect of the nonuniform potential distribution

inside a thick NW at large negative gate bias.

We have learned that the potential distribution across a thick NW plays an important role in

determining the electronic confinement. Since a change in Vg modifies the potential distribution,

we expect the electronic confinement, and in turn, subband spacing, to change with Vg as well,

which has been verified by our model. As mentioned earlier, in the stability diagram of 1-D

transport in terms of transconductance, the intersections of left/right tilted stripes of linear

conductance diamonds give a direct measure of the subband spacing between two successive

subbands, located above and below the Fermi level, respectively. In developing this concept, we

Page 86: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

72

have assumed that the subband spacing is independent of gate bias. In our case, subband spacing

keeps changing with gate bias, and we can imagine that the transconductance diamonds will be

distorted with unequal slopes for left/right tilted boundaries, depending on whether the spacing is

increasing or decreasing. This unique feature, combined with the superimposed Fabry-Perot

oscillations, dramatically complicates the transconductance patterns. Nevertheless, we can still

trace the left/right tilted white stripes (transconductance peaks) in Figure 4.3b and locate the

intersections. From the corresponding source and drain bias, the subband spacing is obtained and

listed in the 4th column of Table 4.1. A prediction of the change of subband spacing based on our

coupled simulation is shown in the first two columns of Table 4.1. The 1st column is the subband

spacing between two successive subbands located above and below the Fermi level at the onset

of each plateau; the 2nd column is the subband spacing between the same two subbands, but at

the end of each plateau. The two spacing, ΔEstart and ΔEend, thus track the evolution of spacing

corresponding to each plateau. The measured spacing ΔEexperiment from transconductance

diamonds should therefore be located between the corresponding ΔEstart and ΔEend, and can be

viewed as an average effect of the change of subband spacing. From the table, we can see that

the predicted spacing for the 9th and 10th plateaus shows some deviation from the observed

values. All other predicted spacing matches the observed spacing with the right trend. For

comparison, subband spacing of the symmetric cylindrical confinement is listed in the last

column of Table 4.1, and it clearly shows that the subband spacing of the “particle in a cylinder”

cannot be used to explain the observed spacing.

Page 87: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

73

Plateau ΔEstart ΔEend ΔEexperiment ΔEsymmetric

(meV) (meV) ( ± 0.5 meV) (meV)

1 9.2 7.8 9 5.9

2 7.5 6.2 7.5 7.8

3 6.7 5.8 6 2.8

4 3.6 3.6 4 6.8

5 2.8 3.1 3 5.6

6 3.7 2.6 4 5.5

7 2.8 2.4 2.5 8.8

8 6.7 7.6 7 2.7

9 2.4 4.5 7.5 1.4

10 6.3 3.6 9 12.1

11 8.1 9.8 9 2.3 Table 4.1: Comparison between the calculated and observed subband spacing. The 1st and 2nd column list calculated

subband spacing between two successive subbands located above and below the Fermi level at the onset and end of

each plateau, respectively, based on our coupled simulation. The 3rd column is the extracted average subband

spacing for each plateau from the transconductance diamonds in the stability diagram. The 4th column is the

calculated subband spacing of the “particle in a cylinder” model.

A byproduct of the model is the capability to perform capacitance calculations. The 3-D

electrostatic simulation includes the screening effect of nonuniform carrier distribution by

treating the channel as a semiconductor, as well as the edge effect due to the distortion of the

electric field near metal contacts [85]. It can therefore provide a more accurate estimate of the

gate capacitance. We define the total gate-to-wire capacitance as the total space charge

induced by the voltage difference between the gate (Vg) and the electrodes (grounded):

| | (4.7)

The volume integration was performed within the entire NW domain and the result is shown as

red dots in Figure 4.9a. Given that ⁄ 5.5, should be a factor of 0.85 of . was

also calculated using Eq. (4.2) with our device geometry and the result is shown as black dots in

Figure 4.9 as a comparison. As mentioned in Section 4.3.3, Khannal et al. [85] assumed a

semiconducting channel with finite length and their 3-D electrostatic simulation gave a 30%

reduction of close to the depletion regime of a thin Si NW compared with the value

Page 88: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

74

calculated from Eq. (4.2). of our device only accounts for about 30% of the value calculated

from Eq. (4.2), indicating a further reduction of due to the screening effect from the side-

gates.

Figure 4.9: (a) Calculated gate capacitance per unit length as a function of gate bias from the 3-D electrostatic

simulation. Also shown is the back-gate capacitance calculated using Eq. (4.2) for the same device. (b)

Calculated quantum capacitance per unit length as a function of gate bias using a combination of Eq. (4.8) and

results from the 2-D eigenvalue simulation.

Quantum capacitance (QC), first introduced by Luryi [86], is an important concept in NW

devices, which accounts for the fact that the gate field penetrates through the wire since it is not

completely screened on the wire surface as is the case in an ideal macroscopic conductor.

Physically, the QC is the energy broadened density of states evaluated at the Fermi Level:

⁄ ∑ ⁄, ∑ 2 ⁄ |, (4.8)

where , are eigenvalues of Eq. (4.6) solved at the middle slice of the NW, i.e.,

energy levels of lateral modes due to the strongest confinement along the NW. Calculated QC as

a function of gate bias is shown in Figure 4.9b. The effect of discrete well-spaced modes on the

QC is clearly seen.

-8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 815

20

25

30

35

40

45

50

55

60

65

70

75

-8.0 -7.6 -7.2 -6.8 -6.4 -6.0 -5.6 -5.2 -4.8 -4.4 -4.0

(b)

Cap

acita

nce

(pF/

m)

Vg(V)

Cbg Cg

(a)

Vg(V)

Capacitance (pF/m

)

0

200

400

600

800

1000

1200

1400

1600

1800

2000

2200

2400

Page 89: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

75

QC affects how the band edge under the gate responds to the gate bias. To see this, we consider

an electrostatic model for a ballistic transistor [87]. The potential at the top of the barrier in the

channel is ⁄ , where is the mobile charge in the

channel, . Assume the source is grounded. In terms of energy, we obtain the

conduction band edge under the gate: / , where /

and / . Taking the derivative of with respect to , and using the fact that

⁄ , we have

| | 1⁄ (4.9)

In fact, the gate-to-channel capacitance is a series combination of the geometric ( ) and

quantum ( ) capacitance [88]. Comparing the capacitance values in Figure 4.9, we can see that,

except for the depletion case where is exponentially reduced, , which indicates that

. Therefore we expect that the value of | |⁄ is small and will keep decreasing

as Vg increases, reflecting the increasing screening effect due to the enlarged charge density in

the channel. This leads to an inefficient gate tuning of the energy levels in the channel. From

Figure 4.3a and b we see that for the left three plateaus about 1 V change in Vg can move two

successive subbands with a spacing of around 7 meV across the Fermi level, giving a

| |⁄ value of about 0.007, which is indeed very small compared with reported value

for InAs single NWFET [45].

4.5 Validation of the Model

As mentioned in Section 4.4.1, a key assumption of the model is that the linear conductance of

the NW is determined by the middle slice of the NW. This assumption originates from the logic

that at negative gate biases, since source and drain are grounded, the middle of the NW should

possess the most negative average potential, and therefore highest conduction band edge. The

presence of side-gates pointing to the middle of the NW further enhances this effect. At positive

gate biases with high carrier concentration, the gate has weaker capability in dragging the

Page 90: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

76

conduction band edge of the middle region below the grounding level. Applying this assumption

will not generate a result that is far from the real case.

To support this assumption qualitatively, we show in Figure 4.10 the calculated potential

distribution profiles inside the NW at two slices: the middle slice (x = 0 nm) and the slice 100 nm

left to the middle one (x = -100 nm), at two gate biases (Vg = -2 and -6 V). The potential

distributions are very similar for the two slices at both gate biases, implying that the middle

region (at least from -100 nm to 100 nm) of the NW has a relatively uniform potential

distribution which can be represented by the distribution in the middle slice. Note that the

potential differences for the two biases are about 90 meV and 300 meV across the NW,

respectively, which supports our statement in Section 4.4.2 that nonuniform potential

distributions up to hundreds of meV exist with lateral cross section of the NW at negative gate

bias and are responsible for the modification of electronic confinement.

Figure 4.10: Calculated potential distribution profiles inside the NW at the middle slice (x = 0 nm) and the slice 100

nm left to it (x = -100 nm), at two gate biases (Vg = -2 and -6 V).

To further support the assumption, we calculated the average electrostatic energies at different

positions along the transport dimension x of the NW using

Page 91: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

77

| | , , (4.10)

The position dependent energies at different gate biases are plotted in Figure 4.11a. From the top

to bottom, curves represent axial potential energy distributions at Vg = -8 ~ 8 V stepped by 2 V.

The middle section of the NW has the highest electrostatic energies at negative gate biases and

the lowest electrostatic energies at positive gate biases but with flatter potential profiles. The plot

also shows that the axial potential varies slowly except near the contact regions. This is an

essential condition to apply Thomas-Fermi picture that carrier concentration is determined by

local potential energy (Eq. (4.4)), which validates our approach of the 3-D electrostatic

simulation.

Figure 4.11: (a) Average electrostatic energy distribution along the NW. From top to bottom: distributions at Vg = -8

~ 8 V stepped by 2 V. (b) Lowest subband level distribution along the NW with fitted Fermi level labeled. From top

to bottom: distributions at Vg = -8 ~ 0 V stepped by 2 V.

The potential energy shifts the quantized energy levels in the channel. To see how this is

achieved, we use the lowest subband energy of the middle slice of the NW 0 shifted by

the potential energy differences along the x axis to represent the lowest subband distribution

020406080

100120140

-400 -360 -320 -280 -240 -200 -160 -120 -80 -40 05055606570758085

Elec

trost

atic

ene

rgy(

meV

)

(b)

E 1(m

eV)

x(nm)

EF

(a)

Page 92: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

78

along the transport dimension of the NW: 0 0 , and plot the

results at different gate biases in Figure 4.11. From the top to bottom, curves represent axial E1

distributions at Vg = -8 ~ 0 V stepped by 2 V. The fitted Fermi level is also shown in this plot.

This plot clearly shows that along the transport dimension of the NW, the lowest subband energy

at the middle of the NW is the highest. This approach can be extended to obtain the axial

distribution of multiple subbands and those at the middle slice of the NW all possess highest

energies. That is to say, the Fermi level needs to be above the subband levels at the middle of the

NW first for them to conduct. So we only need to perform the 2-D eigenvalue simulation at the

middle slice of the NW and use the corresponding eigenvalues to predict conductance

quantization characteristics.

Ideally, one should conduct the 3-D Poisson and Schrodinger self-consistent simulation coupled

through the carrier concentration to get a more accurate modeling of the device. However, this

type of simulation is computationally intense. Although our 3-D and 2-D coupled simulations are

not self-consistent, a comparison of the calculated carrier concentrations based on the two

simulations gives an estimate of how much the two simulations deviate from each other.

From the Thomas-Fermi scenario, the 1-D carrier concentration can be calculated by performing

the integration of the 3-D carrier concentration obtained from Eq. (4.4) over the middle cross

section of the NW:

0, , (4.11)

From the 2-D Schrodinger equation, the 1-D carrier concentrations at finite temperature and zero

temperature are given in the following forms:

∑ | , | 2 2 ⁄, (4.12a)

∑ | , | 4 2 ⁄, (4.12b)

where , and are eigenfunctions and eigenvalues of the Schrodinger equation and

the integrations are performed over the middle slice of the NW.

Page 93: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

79

The calculated 1-D carrier concentrations using Eq. (4.11) and Eq. (4.12) at different gate biases

are shown in Figure 4.12. calculated from Schrodinger at T = 0 K and T = 10 K are very

close to each other. At Vg < -3 V, the two simulations give almost identical 1-D carrier

concentrations. At Vg > -3 V, the two simulations begin to deviate from each other. This

comparison shows that the coupled simulations possess good quality at negative gate biases.

Figure 4.12: Calculated 1-D carrier concentrations as a function of gate bias based on the 3-D Thomas-Fermi

scenario (squares) and the 2-D Schrodinger equation at T = 10 K (circles) and T = 0 K (triangles).

4.6 Conclusions

The transport properties of a thick InAs NW device controlled by a pair of side-gates and back-

gate were studied. The trivial electrostatic “self-gating” effect was first subtracted by making the

conductance versus source and drain bias traces symmetric with respect to the bias. Conductance

compensation was added around the zero bias regions of the traces to cure the system error. The

existence of linear conductance plateaus and half plateaus provides strong evidence of subband

quantization in the NW. The observation of conductance quantization at low temperature is very

likely to be a result of reduced backscattering in the channel. Magnetotransport measurements

indicate the emergence of magneto-electric-hybrid subbands characterized by edge states and

Zeeman splitting of these spin-degenerate states. Most of the linear plateaus disappeared as the

-8 -7 -6 -5 -4 -3 -2 -1 0

123456789

10111213

n 1D(m

-1)

Vg(V)

Thomas-Fermi Schrodinger (10K) Schrodinger (0K)

108

Page 94: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

80

temperature was increased to 30 K, suggesting a rough range of 2.6 ~ 7.8 meV for the subband

spacing. The calculated room temperature electron mobility is higher than reported values of

InAs NWs with similar growth conditions, implying a cleaner channel for this NW, which

supports the quasi-ballistic nature of transport in the NW.

A 3-D electrostatic and 2-D eigenvalue coupled simulation was conducted to explain the

observed linear conductance features. The best fit to the data was achieved for the condition that

the background impurity concentration in the channel is very low and almost 100% of the free

carriers come from the surface sates. This condition reinforces the reason to observe the

conductance quantization in a thick NW at low temperature: there is little backscattering in the

channel and the effect of surface scattering is also reduced for a thick NW. The calculated

subband spacing from the model was compared with the values extracted from the

transconductance stability diagram and shows satisfactory agreement. The spacing differs

significantly from the “particle in a cylinder” case, where a uniform potential distribution inside

the NW is assumed, strongly indicating a modification to the electronic confinement of the NW

from the nonuniform potential distribution inside the NW, as confirmed by the model. The

underlying assumption of the model is that it is the subband quantization in the middle region of

the NW that determines the linear transport so that one only needs to solve the 2-D eigenvalue

problem within the middle slice of the NW to get the representative subband spacing. This

assumption was supported qualitatively by the potential distribution profiles along the NW from

the model and quantitatively by the calculated average potential energy and lowest subband level

distribution along the NW. Finally, a comparison of calculated 1-D carrier concentration based

on the 3-D Thomas-Fermi scenario and the 2-D Schrodinger equation shows the good quality of

the coupled model at negative gate biases.

Page 95: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

81

Chapter 5

Conclusions and Future Outlook

This thesis studied electron transport properties in three types of single InAs NW devices.

Devices with local side-gating exhibited a universal hysteresis in the transfer characteristics,

which was attributed to carrier injection at high gate fields due to the sharp tip of side-gates. Not

all the devices with single etched constriction showed sign of conductance quantization. We

understand this as the influence of disorder which is sample-specific. For the device without

side-gates, Coulomb blockade due to electrons tunneling through DQD was presented in the

pinch-off region. For a side-gated device with an etched constriction, F-P conductance oscillation

was found superimposed onto the first conductance plateau, which is a strong indication of the

quasi-ballistic nature of the transport. In the pinch-off region, as the back-gate voltage increased,

Coulomb blockade due to electrons tunneling through MTJ, single QD and F-P oscillation was

successively observed. F-P oscillation extending all the way to the pinch-off region has never

been reported to our knowledge. Remarkable conductance quantization was observed in a thick

InAs NW without etched constriction, which is very likely to be the result of reduced surface

scattering for a thick NW. The calculated room temperature electron mobility is higher than

reported values of InAs NWs with similar growth conditions, implying a cleaner channel for this

NW, which supports the quasi-ballistic nature of transport in the NW. According to the

conducted simulation, the background impurity concentration in the channel is very low and

almost 100% of the free carriers come from the surface sates. This result reinforces the reason to

observe the conductance quantization in a thick NW at low temperatures. The calculated subband

spacing from the simulation was compared with the values extracted from the transconductance

stability diagram and shows satisfactory agreement. The spacing differs significantly from the

“particle in a cylinder” case, where a uniform potential distribution inside the NW is assumed.

This deviation strongly indicates a modification to the electronic confinement of the NW from

the nonuniform potential distribution inside the NW. Finally, the underlying assumption of the

Page 96: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

82

simulation was validated. Overall the simulation successfully captured the main features of the

measurement results.

The complicated device fabrication turned out to be extremely difficult. Due to the long

processing period and low yield, only a few devices were measured. Meanwhile, the low

resolution of our EBL also limited the length of etched constrictions. To reach more convincing

conclusions on transport through single etched constriction in NW, a group of devices with

different etching length and aspect ratios need to be fabricated and studied. Nevertheless, this

thesis provides preliminary insight for a novel design of single InAs NW devices. Compared

with existing results reported from literature, at least two striking achievements have been

accomplished: first, Fabry-Preot oscillations in the sub-threshold region of a quasi-1D system

have been observed; second, conductance quantization has been observed in a thick NW, and the

simulation advances the understanding to the role of high gate biases on the confinement energy

of thick NWs. This platform exhibits potential of integrating QD, QPC and interferometer onto a

single NW, which is very promising for quantum computing applications.

Page 97: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

83

Bibliography

[1] W. Lu and C. M. Lieber. Nanoelectronics from the bottom up. Nature Materials 6(11): 841-

850, 2007.

[2] ITRS Organization, International Technology Roadmap for Semiconductors Report, 2011

Edition, 2011.

[3] W. Lu, P. Xie and C. M. Lieber. Nanowire transistor performance limits and applications.

IEEE Transactions on Electron Devices, 55(11): 2859-2876, 2008.

[4] Y. Li, F. Qian, J. Xiang and C. M. Lieber. Nanowire electronic and optoelectronic devices.

Materials Today, 9(10): 18-27, 2006.

[5] B. Z. Tian, X. L. Zheng, T. J. Kempa, Y. Fang, N. F. Yu, G. H. Yu, J. L. Huang and C. M.

Lieber. Coaxial silicon nanowires as solar cells and nanoelectronic power sources. Nature, 449:

885-889, 2007.

[6] Y. Cui, Q. Q. Wei, H. K. Park and C. M. Lieber. Nanowire nanosensors for highly sensitive

and selective detection of biological and chemical species. Science, 293: 1289-1292, 2001.

[7] H. van Houten and C. W. J. Beenakker. Quantum point contacts. Physics Today, 49(7): 22-27,

1996.

[8] R. Hanson, L. P. Kouwenhoven, J. R. Petta, S. Tarucha, L. M. K. Vandersypen. Spins in few-

electron quantum dots. Reviews of Modern Physics, 79(4): 1217-1265, 2007.

[9] W. Lu, J. Xiang, B. P. Timko, Y. Wu and C. M. Lieber. One-dimensional hole gas in

germanium/silicon nanowire heterostructures. Proceedings of the National Academy of Science,

102(29): 10046-10051, 2005.

[10] A. C. Ford, S. B. Kumar, R. Kapadia, J. Guo and A. Javey. Observation of degenerate one-

dimensional sub-bands in cylindrical InAs nanowires. Nano Letters, 12: 1340-1343, 2012.

Page 98: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

84

[11] P. L. McEuen, M. Bockrath, D. H. Cobden, Y-G. Yoon and S. G. Louie. Disorder,

pseudospins and backscattering in carbon nanotubes. Physical Review Letters, 83(24): 5098-

5101, 1999.

[12] P. Roulleau, T. Choi, S. Riedi, T. Heinzel, I. Shorubalko, T. Ihn and K. Ensslin. Suppression

of weak antilocalization in InAs nanowires. Physical Review B, 81: 155449, 2010.

[13] W. Liang, M. Bockrath, D. Bozovic, J. H. Hafner, M. Tinkham and H. Park. Fabry-Perot

interference in a nanotube electron waveguide. Nature, 411: 665-669, 2001.

[14] A. Javey, J. Guo, Q. Wang, M. Lundstrom and H. J. Dai. Ballistic carbon nanotube field-

effect transistors. Nature, 424: 654-657, 2003.

[15] K. Grove-Rasmussen, H. I. Jorgensen and P. E. Lindelof. Fabry-Perot interference, Kondo

effect and Coulomb blockade in carbon nanotubes. Physica E, 40: 92-98, 2007.

[16] T. Kamimura, Y. Ohono and K. Matsumoto. Carbon nanotube Fabry-Perot device for

detection of multiple single charge transitions. Japanese Journal of Applied Physics, 48: 025001,

2009.

[17] A. V. Kretinin, R. Popovitz-Biro, D. Mahalu and H. Shtrikman. Multimode Fabry-Perot

conductance oscillations in suspended stacking-faults-free InAs nanowires. Nano Letters, 10:

3439-3445, 2010.

[18] H. A. Nilsson, M. T. Deng, P. Caroff, C. Thelander, L. Samuelson, L.-E. Wernersson and H.

Q. Xu. InSb nanowire field-effect transistors and quantum-dot devices. IEEE Journal of Selected

Topics in Quantum Electronics, 17(4): 907-914, 2011.

[19] G. Liu, Y. Zhang and C. N. Lau. Electron double slit interferometers based on carbon

nanotubes. Nano Letters, 11: 4043-4046, 2011.

[20] D. Loss and D. P. DiVincenzo. Quantum computation with quantum dots. Physical Review

A, 57(1): 120-126, 1998.

Page 99: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

85

[21] J. Salfi, S. Roddaro, D. Ercolani, L. Sorba, I. Savelyev, M. Blumin, H. E. Ruda and F.

Beltram. Electronic properties of quantum dot systems realized in semiconductor nanowires.

Semiconductor Science and Technology, 25: 024007, 2010.

[22] L. P. Kouwenhoven, D. G. Austing and S. Tarucha. Few-electron quantum dots. Reports of

Progress in Physics, 64: 701-736, 2001.

[23] D. K. Ferry, S. M. Goodnick and J. Bird. Transport in Nanostructures. Cambridge

University Press, Cambridge, 2009.

[24] C. W. J. Beenakker. Theory of Coulomb-blockade oscillations in the conductance of a

quantum dot. Physical Review B, 44(4): 1646-1656, 1991.

[25] W. G. van der Wiel, S. De Franceschi, J. M. Elzerman, T. Fujisawa, S. Tarucha, L. P.

Kouwenhoven. Electron transport through double quantum dots. Reviews of Modern Physics,

75(1): 1-22, 2003.

[26] A. Yacoby and Y. Imry. Quantization of the conductance of ballistic point contacts beyond

the adiabatic approximation. Phyaical Review B, 41(8): 5341-5350, 1990.

[27] A. Szafer and A. D. Stone. Theory of quantum conduction through a constriction. Physical

Review Letters, 62(3): 300-303, 1989.

[28] M. Buttiker. Quantum transmission of a saddle-point constriction. Physical Review B,

41(11): 7906-7909, 1990.

[29] Y. Imry. Physics of Mesoscopic Systems. World Scientific Press, Singapore, 1986.

[30] B. J. van Wees, L. P. Kouwenhoven, H. van Houten, C. W. J. Beenakker, J. E. Mooij, C. T.

Foxon and J. J. Harris. Quantized conductance of magnetoelectric subbands in ballistic point

contacts. Physical Review B, 38(5): 3625-3627, 1988.

M. Buttiker. Absence of backscattering in the quantum hall effect in multiprobe conductors.

Physical Review B, 38(14): 9375-9389, 1988.

Page 100: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

86

[31] R. de Picciotto, H. L. Stormer, L. N. Pfeiffer, K. W. Baldwin and K. W. West. Four-

terminal resistance of a ballistic quantum wire. Nature, 411: 51-54, 2001.

[32] R. de Picciotto, L. N. Pfeiffer, K. W. Baldwin and K. W. West. Nonlinear response of a

clean one-dimensional wire. Physical Review Letters, 92(3): 036805, 2004.

[33] L. I. Glazman and A. V. Khaetskii. Nonlinear quantum conductance of a lateral

microconstraint in a heterostructure. Europhysics Letters, 9(3): 263, 1989.

[34] N. K. Patel, L. Martin-Moreno, M. Pepper, R. Newbury, J. E. F. Frost, D. A. Ritchie, G. A.

C. Jones, J. T. M. B. Janssen, J. Singleton and J. A. A. J. Perenboom. Ballistic transport in one

dimension: additional quantization produced by an electric field. Journal of Physics: Condensed

Matter, 2(34): 7241, 1990.

[35] S. Amasha, I. G. Rau, M. Grobis, R. M. Potok, H. Shtrikman and D. Goldhaber-Gordon.

Coulomb blockade in an open quantum dot. Phyaical Review Letters, 107(21): 216804, 2011.

[36] W. Monch. Semiconductor Surfaces and Interfaces, Third Eidition. Springer-Verlag, Berlin,

2001.

[37] L. Polenta, M. Rossi, A. Cavallini, R. Calarco, M. Marso, R. Meijers, T. Richter, T. Stoica

and H. Luth. Investigation on localized states in GaN nanowires. ACS Nano, 2(2): 287-292, 2008.

[38] R. Agarwal. Heterointerfaces in semiconductor nanowires. Small, 4(11): 1872-1893, 2008.

[39] M. Noguchi, K. Hirakawa and T. ikoma. Intrinsic electron accumulation layers on

reconstructed clean InAs (100) surfaces. Physical Review Letters, 66(17): 2243-2246, 1991.

[40] D. B. Suyatin, C. Thelander, M. T. Bjork, I. Maximov and L. Samuelson. Sulfur passivation

for ohmic contact formation to InAs nanowires. Nanotechnology, 18: 105307, 2007.

[41] S. M. Sze. Physics of Semiconductor Devices. Wiley Interscience, New York, 1981.

[42] P. Dutta and P. M. Horn. Low-frequency fluctuations in solids: 1/f noise. Reviews of

Modern Physics, 53(3): 497-516, 1981.

Page 101: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

87

[43] H. D. Xiong, W. Wang, Q. Li, C. A. Richter, J. S. Suehle, W-K. Hong, T. Lee and D. M.

Fleetwood. Random telegraph signals in n-type ZnO nanowire field effect transistors at low

temperature. Applied Physics Letters, 91: 053107, 2007.

[44] J. Salfi, I. G. Savelyev, M. Blumin, S. V. Nair and H. E. Ruda. Direct observation of single-

charge-detection capability of nanowire field-effect transistors. Nature Nanotechnology, 5: 737-

741, 2010.

[45] J. Salfi, N. Paradiso, S. Roddaro, S. Heun, S. V. Nair, I. G. Savelyev, M. Blumin, F.

Beltram and H. E. Ruda. Probing the gate-voltage-dependent surface potential of individual InAs

nanowires using random telegraph signals. ACS Nano, 5(3): 2191-2199, 2011.

[46] L. J. Klein, K. L.M. Lewis, K. A. Slinker, S. Goswami, D. W. van der Weide, R. H. Blick, P.

M. Mooney, J. O. Chu, S. N. Coppersmith, M. Friesen and M. A. Eriksson. Quantum dots and

etch-induced depletion of a silicon two-dimensional electron gas. Journal of Applied Physics, 99:

023509, 2006.

[47] M. Larsson, D. Wallin and H. Q. Xu. A highly tunable lateral quantum dot realized in

InGaAs/InP by an etching technique. Journal of Applied Physics, 103: 086101, 2008.

[48] M. Manoharan, S. Oda and H. Mizuta. Impact of channel constrictions on the formation of

multiple tunnel junctions in heavily doped silicon single electron transistors. Applied Physics

Letters, 93: 112107, 2008.

[49] I. Shorubalko, R. Leturcq, A. Pfund, D. Tyndall, R. Krischek, S. Schon and K. Ensslin. Self-

aligned charge read-out for InAs nanowire quantum dots. Nano Letters, 8(2): 382-385, 2008.

[50] C. Rossler, M. Herz, M. Bichler and S. Ludwig. Freely suspended quantum point contact.

Solid State Communications, 150: 861-864, 2010.

[51] J. Baugh, J. S. Fung, J. Mracek and R. R. LaPierre. Building a spin quantum bit register

using semiconductor nanowires. Nanotechnology, 21: 134018, 2010.

Page 102: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

88

[52] S. De Franceschi, J. A. van Dam, E. P. A. M. Bakkers, L. F. Feiner, L. Gurevich and L. P.

Kouwenhoven. Single-electron tunneling in InP nanowires. Applied Physics Letters, 83(2): 344-

346, 2003.

[53] Z. H. Zhong, Y. Fang, W. Lu and C. M. Lieber. Choerent single charge transport in

molecular-scale silicon nanowires. Nano Letters, 5(6): 1143-1146, 2005.

[54] T. S. Jespersen, M. Aagesen, C. Sorensen, P. E. Lindelof and J. Nygard. Kondo physics in

tunable semiconductor nanowire quantum dots. Physical Review B, 74(23): 233304, 2006.

[55] R. Wagner and W. Ellis. Vapor-liquid-solid mechanism of single crystal growth. Applied

Physics Letters, 4(5): 89, 1964.

[56] D. Y. Petrovykh, M. J. Yang and L. J. Whitman. Chemical and electronic properties of

sulfur-passivated InAs surfaces. Surface Science, 523(3): 231-240, 2003.

[57] G. C. DeSalvo, R. Kaspi and C. A. Bozada. Citric acid etching of GaAs Sb , Al . Ga . Sb,

and InAs for heterostructure device fabrication. Journal of the Electrochemistry Society, 141(12):

3526-3531, 1994.

[58] K. L. Kavanagh, J. Salfi, I. Savelyev, M. Blumin and H. E. Ruda. Transport and strain

relaxation in wurtzite InAs-GaAs core-shell heterowires. Applied Physics Letters 98(15): 152103,

2011.

[59] G. C. DeSalvo, W. F. Tseng and J. Comas. Etch rates and selectivities of citric

acid/hydrogen peroxide on GaAs, Al0.3Ga0.7 As, In0.2Ga0.8As, In0.53Ga0.47As, In0.52Al0.48As, and

InP. Journal of the Electrochemistry Society, 139(3): 831-835, 1992.

[60] C. Juang, K. J. Kuhn and R. B. Darling. Selective etching of GaAs and Al . Ga . As with

citric acid/hydrogen peroxide solutions. Journal of Vacuum Science & Technology B, 8(5): 1122-

1124, 1990.

[61] R. Jia, S. Kasai, Q. Wang, S. B. Long, J. B. Niu, Z. G. Li and M. Liu. Surface-induced large

side-gating phenomenon in GaAs quantum wire transistors and its removal by surface

passivation using Si interface control layer. Applied Physics Letters, 90: 132124, 2007.

Page 103: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

89

[62] H. A. Nilsson, C. Thelander, L. E. Froberg, J. B. Wagner and L. Samuelson. Nanowire-

based multiple quantum dot memory. Applied Physics Letters, 89: 163101, 2006.

[63] S. A. Dayeh, C. Soci, P. K. L. Yu, E. T. Yu and D. Wang. Influence of surface sates on the

extraction of transport parameters from InAs nanowire field effect transistors. Applied Physics

Letters, 90: 162112, 2007.

[64] S. A. Dayeh, C. Soci, P. K. L. Yu, E. T. Yu and D. Wang. Transport properties of InAs

nanowire field effect transistors: the effects of surface states. Journal of Vacuum Science &

Technology B, 25(4): 1432-1436, 2007

[65] T. Matsukawa, S. Kanemaru, M. Masahara, M. nagao, H. Tanoue and J. Itoh. Programmable

conductivity of silicon nanowires with side-gates by surface charging. Japanese Journal of

Applied Physics, 42: 2422-2425, 2003.

[66] M. Radosavljevic, M. Freitag, K. V. Thadani and A. T. Johnson. Nonvolatile molecular

memory elements based on ambipolar nanotube field effect transistors. Nano Letters, 2(7): 761-

764, 2002.

[67] E. H. Anderson, D.L. Olynick, W. Chao, B. Harteneck and E. Veklerov. Influence of sub-100

nm scattering on high-electron beam lithography. Journal of Vacuum Science, 19(6): 2504-2507,

2001.

[68] T. S. Jespersen, M. L. Polianski, C. B. Sorensen, K. Flensberg and J. Nygard. Mesoscopic

conductance fluctuations in InAs nanowire-based SNS junctions. New Journal of Physics, 11:

113025, 2009.

[69] D. Wallin, A. Fuhrer, L. E. Froberg, L. Samuelson, H. Q. Xu, S. Hofling and A. Forchel.

Detection of charge states in nanowire quantum dots using a quantum point contact. Applied

Physics Letters, 90: 172112, 2007.

[70] A. Fuhrer, L. E. Froberg, J. N. Pedersen, M. W. Larsson, A. Wacker, M-E. Pistol and L.

Samuelson. Few electron double quantum dots in InAs/InP nanowire heterostructures. Nano

Letters, 7(2): 243-246, 2007.

Page 104: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

90

[71] P. Stokes and S. I. Khondaker. Evaluating defects in solution-processed carbon nanotube

devices via low-temperature transport spectroscopy. ACS Nano, 4(5): 2659-2666, 2010.

[72] M. Hofheinz, X. Jehl, M. Sanquer, G. Molas, M. Vinet and S. Deleonibus. Capacitance

enhancement in Coulomb blockade tunnel barriers. Physical Review B, 75(23): 235301, 2007.

[73] F. A. Zwanenburg, A. A. van Loon, G. A. Steele, C. E. W. M. van Rijmenam, T. Balder, Y.

Fang, C. M. Lieber and L. P. Kouwenhoven. Ultrasmall silicon quantum dots. Journal of Applied

Physics, 105: 124314, 2009.

[74] J. Weis, R. J. Haug, K. v. Klitzing and K. Ploog. Transport spectroscopy of a confined

electron system under a gate tip. Physical review B, 46(19): 12837-12840, 1992.

[75] H. T. Chou, S. Luscher, D. Goldhaber-Gordon, M. J. Manfra, A. M. Sergent, K. W. West

and R. J. Molnar. High-quality quantum point contacts in GaN/AlGaN heterostructures. Applied

Physics Letters, 86: 073108, 2005.

[76] T. Morimoto, M. Henmi, R. Naito, K. Tsubaki, N. Aoki, J. P. Bird and Y. Ochiai.

Resonantly enhanced nonlinear conductance in long quantum point contacts near pinch-off.

Physical Review Letters, 97: 096801, 2006.

[77] N. Wang, Y. Cai and R. Q. Zhang. Growth of nanowires. Material Science and Engineering

R, 60: 1-51, 2008.

[78] A. Motayed, M. Vaudin, A. V. Davydov, J. Melngailis, M. He and S. N. Mohammad.

Diameter dependent transport properties of gallium nitride nanowire field effect transistors.

Applied Physics Letters, 90: 043104, 2007.

[79] A. C. Ford, J. C. Ho, Y. L. Chueh, Y. C. Tseng, Z. Fan, J. Guo, J. Bokor and A. Javey.

Diameter-dependent electron mobility of InAs nanowires. Nano Letters, 9(1): 360-365, 2009.

[80] A. Kristensen, H. Bruus, A. E. Hansen, J. B. Jensen, P. E. Lindelof, C. J. Marckmann, J.

Nygard and C. B. Sorensen. Bias and temperature dependence of the 0.7 conductance anomaly in

quantum point contacts. Physical Review B, 62(16): 10950-10957, 2000.

Page 105: Quantum Transport in InAs Nanowires with Etched ... · Quantum Transport in InAs Nanowires with Etched Constrictions and Local Side-gating Yao Ma Master of Applied Science Graduate

91

[81] G. Scappucci, L. D. Gaspare, E. Giovine, A. Notargiacomo, R. Leoni and F. Evangelisti.

Conductance quantization in etched Si/SiGe quantum point contact. Physical Review B, 74(3):

035321, 2006.

[82] M. J. Biercuk, N. Mason, J. Martin, A. Yacoby and C. M. Marcus. Anomalous conductance

quantization in carbon nanotubes. Physical Review Letters, 94(2): 026801, 2005.

[83] C. Rossler, S. Baer, E. de Wiljes, P.- L. Ardelt, T. Ihn, K. Ensslin, C. Reichi and W.

Wegscheider. Transport properties of clean quantum point contacts. New Journal of Physics, 13:

113006, 2011.

[84] O. Wunnicke. Gate capacitance of back-gated nanowire field-effect transistors. Applied

Physics Letters, 89: 083102, 2006.

[85] D. R. Khanal and J. Wu. Gate coupling and charge distribution in nanowire field effect

transistors. Nano Letters, 7(9): 2778-2783, 2007.

[86] S. Luryi. Quantum capacitance devices. Applied Physics Letters, 52(16): 501-503, 1988.

[87] M. A. Khayer and R. K. Lake. The quantum and classical capacitance limits of InSb and

InAs nanowire FETs. IEEE Transactions on Electron Devices, 56(10): 2215-2223, 2009.

[88] P. J. Burke. Carbon nanotube devices for GHz to THz applications. Proceedings of SPIE,

5593: 52-61, 2004.