waves, particles, and interactions in reduced dimensions

204
Waves, Particles, and Interactions in Reduced Dimensions A dissertation presented by Yiming Zhang to The Department of Physics in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the subject of Physics Harvard University Cambridge, Massachusetts 2009

Upload: others

Post on 19-Jun-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Waves, Particles, and Interactions in Reduced Dimensions

Waves, Particles, and Interactions in Reduced Dimensions

A dissertation presentedby

Yiming Zhang

toThe Department of Physics

in partial fulfillment of the requirementsfor the degree of

Doctor of Philosophyin the subject of

Physics

Harvard UniversityCambridge, Massachusetts

2009

Page 2: Waves, Particles, and Interactions in Reduced Dimensions

c© 2009 by Yiming ZhangAll rights reserved.

Page 3: Waves, Particles, and Interactions in Reduced Dimensions

Dissertation Advisor: Professor Charles M. Marcus Author: Yiming Zhang

Waves, Particles, and Interactions in Reduced Dimensions

Abstract

This thesis presents a set of experiments that study the interplay between the wave-

particle duality of electrons and the interaction effects in systems of reduced dimensions.

Both dc transport and measurements of current noise have been employed in the studies; in

particular, techniques for efficiently measuring current noise have been developed specifically

for these experiments.

The first four experiments study current noise auto- and cross correlations in various

mesoscopic devices, including quantum point contacts, single and double quantum dots,

and graphene devices.

In quantum point contacts, shot noise at zero magnetic field exhibits an asymmetry

related to the 0.7 structure in conductance. The asymmetry in noise evolves smoothly into

the symmetric signature of spin-resolved electron transmission at high field. Comparison to

a phenomenological model with density-dependent level splitting yields good quantitative

agreement. Additionally, a device-specific contribution to the finite-bias noise, particularly

visible on conductance plateaus where shot noise vanishes, agrees with a model of bias-

dependent electron heating.

In a three-lead single quantum dot and a capacitively coupled double quantum dot, sign

reversal of noise cross correlations have been observed in the Coulomb blockade regime, and

found to be tunable by gate voltages and source-drain bias. In the limit of weak output

tunneling, cross correlations in the three-lead dot are found to be proportional to the two-

lead noise in excess of the Poissonian value. These results can be reproduced with master

iii

Page 4: Waves, Particles, and Interactions in Reduced Dimensions

equation calculations that include multi-level transport in the single dot, and inter-dot

charging energy in the double dot.

Shot noise measurements in single-layer graphene devices reveal a Fano factor indepen-

dent of carrier type and density, device geometry, and the presence of a p-n junction. This

result contrasts with theory for ballistic graphene sheets and junctions, suggesting that the

transport is disorder dominated.

The next two experiments study magnetoresistance oscillations in electronic Fabry-

Perot interferometers in the integer quantum Hall regime. Two types of resistance oscilla-

tions, as a function of perpendicular magnetic field and gate voltages, in two interferometers

of different sizes can be distinguished by three experimental signatures. The oscillations

observed in the small (2.0 µm2) device are understood to arise from Coulomb blockade, and

those observed in the big (18 µm2) device from Aharonov-Bohm interference. Nonlinear

transport in the big device reveals a checkerboard-like pattern of conductance oscillations as

a function of dc bias and magnetic field. Edge-state velocities extracted from the checker-

board data are compared to model calculations and found to be consistent with a crossover

from skipping orbits at low fields to ~E× ~B drift at high fields. Suppression of visibility as a

function of bias and magnetic field is accounted for by including energy- and field-dependent

dephasing of edge electrons.

iv

Page 5: Waves, Particles, and Interactions in Reduced Dimensions

Contents

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iiiTable of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vList of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viiiAcknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

1 Introduction 11.1 Organization of this thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2 Dc transport and current noise . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.2.1 Dc transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41.2.2 Current noise auto- and cross correlations . . . . . . . . . . . . . . . 5

1.3 Material systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91.3.1 GaAs/AlGaAs heterostructures . . . . . . . . . . . . . . . . . . . . . 9

1.4 Basic properties of mesoscopic devices . . . . . . . . . . . . . . . . . . . . . 121.4.1 Quantum point contacts . . . . . . . . . . . . . . . . . . . . . . . . . 121.4.2 Quantum dots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161.4.3 Quantum Hall effects . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2 System for measuring auto- and cross correlation of current noise at lowtemperatures 232.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242.2 Overview of the system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252.3 Amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2.3.1 Design objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262.3.2 Overview of the circuit . . . . . . . . . . . . . . . . . . . . . . . . . . 262.3.3 Operating point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282.3.4 Passive components . . . . . . . . . . . . . . . . . . . . . . . . . . . 292.3.5 Thermalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

2.4 Digitization and FFT processing . . . . . . . . . . . . . . . . . . . . . . . . 312.5 Measurement example: quantum point contact . . . . . . . . . . . . . . . . 32

2.5.1 Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322.5.2 Measuring dc transport . . . . . . . . . . . . . . . . . . . . . . . . . 332.5.3 Measuring noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342.5.4 System calibration using Johnson noise . . . . . . . . . . . . . . . . 35

2.6 System performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372.7 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402.8 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

3 Current noise in quantum point contacts 41

v

Page 6: Waves, Particles, and Interactions in Reduced Dimensions

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423.2 QPC characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 433.3 Current noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.3.1 0.7 structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 483.3.2 Bias-dependent electron heating . . . . . . . . . . . . . . . . . . . . 51

3.4 Conclusion and acknowledgements . . . . . . . . . . . . . . . . . . . . . . . 52

4 Tunable noise cross-correlations in a double quantum dot 534.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 544.2 Device . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 564.3 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 564.4 Double-dot characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . 574.5 Sign-reversal of noise cross correlation . . . . . . . . . . . . . . . . . . . . . 584.6 Master equation simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 594.7 Intuitive explanation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 604.8 Some additional checks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624.9 Conclusion and acknowledgements . . . . . . . . . . . . . . . . . . . . . . . 63

5 Noise correlations in a Coulomb blockaded quantum dot 645.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 655.2 Device . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665.3 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665.4 Noise in the two-lead configuration . . . . . . . . . . . . . . . . . . . . . . . 685.5 Noise in the three-lead configuration . . . . . . . . . . . . . . . . . . . . . . 725.6 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

6 Shot noise in graphene 766.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 776.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 786.3 Shot noise in single-layer devices . . . . . . . . . . . . . . . . . . . . . . . . 796.4 Shot noise in a p-n junction . . . . . . . . . . . . . . . . . . . . . . . . . . . 826.5 Shot noise in a multi-layer device . . . . . . . . . . . . . . . . . . . . . . . . 836.6 Summary and acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . 85

7 Distinct Signatures For Coulomb Blockade and Aharonov-Bohm Interfer-ence in Electronic Fabry-Perot Interferometers 877.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 887.2 Device and measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 897.3 Resistance oscillations in the 2.0 µm2 device . . . . . . . . . . . . . . . . . . 907.4 Resistance oscillations in the 18 µm2 device . . . . . . . . . . . . . . . . . . 937.5 One more signature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 957.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 977.7 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

8 Edge-State Velocity and Coherence in a Quantum Hall Fabry-Perot In-terferometer 998.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1008.2 Device and measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

vi

Page 7: Waves, Particles, and Interactions in Reduced Dimensions

8.3 Checkerboard pattern and interpretation . . . . . . . . . . . . . . . . . . . . 1038.4 Edge-state velocity and energy-dependent dephasing . . . . . . . . . . . . . 1068.5 Nonlinear magnetoconductance in a 2 µm2 device . . . . . . . . . . . . . . . 1088.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1098.7 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

9 Unpublished results 1119.1 Current noise modulated by charge noise . . . . . . . . . . . . . . . . . . . . 1129.2 Quasi-particle tunneling between filling factor 2 and 3 in a constriction . . . 1189.3 The 3/2 quantized plateau in quantum point contacts . . . . . . . . . . . . 1249.4 Non-linear transport in N ≥ 2 Landau levels . . . . . . . . . . . . . . . . . 128

A Fridge Wiring: Thermal Anchoring and Filtering 131A.1 Simple RC filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132A.2 Sapphire heat sinks and circuit boards . . . . . . . . . . . . . . . . . . . . . 133A.3 Mini-circuit VLFX filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135A.4 Thermocoax cables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

B Igor implementation of virtual DACs 140B.1 Igor implementation of virtual DACs . . . . . . . . . . . . . . . . . . . . . . 141

C Effects of external impedance on conductance and noise 152C.1 Effects of external impedance on conductance . . . . . . . . . . . . . . . . . 152C.2 Effects of external impedance on current noise . . . . . . . . . . . . . . . . . 153

D Conductance matrix measurement and multi-channel digital lock-in 156D.1 Simultaneous conductance matrix and current noise measurement . . . . . . 156D.2 Multi-channel digital lock-in . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

E The master equation calculation of current and noise in a multi-lead,multi-level quantum dot 161

vii

Page 8: Waves, Particles, and Interactions in Reduced Dimensions

List of Figures

1.1 Different types of noise, in time and frequency domains . . . . . . . . . . . 71.2 GaAs/AlGaAs heterostructure and conduction band diagram . . . . . . . . 101.3 Conductance as a function of gate voltage in a quantum point contact . . . 131.4 Illustration of a source-drain voltage applied across a barrier in a 1d system 131.5 Conductance as a function of gate voltage in a quantum dot . . . . . . . . . 171.6 Differential conductance as a function of source-drain bias and gate voltage

in a quantum dot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181.7 Simulations of differential conductance as a function of source-drain bias and

gate voltage of a quantum dot . . . . . . . . . . . . . . . . . . . . . . . . . . 191.8 Bulk 2DEG transport in the quantum Hall regime . . . . . . . . . . . . . . 21

2.1 Block diagram of the two-channel noise detection system . . . . . . . . . . . 262.2 Schematic diagram of the amplification lines . . . . . . . . . . . . . . . . . . 272.3 Equivalent circuits valid near dc and at low megahertz . . . . . . . . . . . . 282.4 Biasing the cryogenic transistor . . . . . . . . . . . . . . . . . . . . . . . . . 302.5 Setup for QPC noise measurement by cross-correlation technique . . . . . . 332.6 Power and cross-spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342.7 Circuit model for QPC noise measurements extraction . . . . . . . . . . . . 362.8 Johnson-noise thermometry . . . . . . . . . . . . . . . . . . . . . . . . . . . 372.9 Noise measurement resolution as a function of integration time . . . . . . . 38

3.1 QPC characterization by dc transport . . . . . . . . . . . . . . . . . . . . . 433.2 Noise measurement setup and micrograph of QPC . . . . . . . . . . . . . . 453.3 Demonstration measurements of bias-dependent QPC noise . . . . . . . . . 463.4 The experimental noise factor . . . . . . . . . . . . . . . . . . . . . . . . . . 483.5 Comparison of QPC noise data to the phenomenological Reilly model . . . 503.6 Bias-dependent electron heating in a second QPC . . . . . . . . . . . . . . . 51

4.1 Double-dot device and setup . . . . . . . . . . . . . . . . . . . . . . . . . . . 554.2 Measured and simulated cross-spectral density near a honeycomb vertex . . 584.3 Energy level diagrams in the vicinity of a honeycomb vertex . . . . . . . . . 614.4 Measured cross-spectral density at other bias configurations . . . . . . . . . 63

5.1 Micrograph of three-lead quantum dot, noise measurement setup, and mea-surements in the two-lead configuration . . . . . . . . . . . . . . . . . . . . 67

5.2 Excess Poissonian noise in the two-lead configuration . . . . . . . . . . . . . 715.3 Cross-spectral density in the three-lead configuration . . . . . . . . . . . . . 735.4 Relation between total excess Poissonian noise and cross-spectral density in

the three-lead configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

viii

Page 9: Waves, Particles, and Interactions in Reduced Dimensions

6.1 Characterization of graphene devices using dc transport at B⊥ = 0 and inquantum Hall regime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

6.2 Shot noise in single-layer devices . . . . . . . . . . . . . . . . . . . . . . . . 806.3 Shot noise in a p-n junction . . . . . . . . . . . . . . . . . . . . . . . . . . . 836.4 Shot noise in a multi-layer device . . . . . . . . . . . . . . . . . . . . . . . . 84

7.1 Measurement setup and the electronic Fabry-Perot devices . . . . . . . . . . 907.2 Resistance oscillations as a function of magnetic field for the 2.0 µm2 device 917.3 Magnetic field and gate voltage periods for the 2.0 µm2 device . . . . . . . 927.4 Magnetic field and gate voltage periods for the 18 µm2 device . . . . . . . . 947.5 Resistance oscillations measured in a plane of magnetic field and gate voltage 96

8.1 Measurement setup and the electronic Fabry-Perot device . . . . . . . . . . 1028.2 Nonlinear magnetoconductance in an 18 µm2 interferometer . . . . . . . . . 1048.3 Magnetic field dependence of extracted velocity and damping factor . . . . 1068.4 Nonlinear magnetoconductance in a 2 µm2 device . . . . . . . . . . . . . . . 109

9.1 Device, noise measurement setup, and charge sensing in conductance . . . . 1149.2 Conductance and current noise near a honey-comb vertex . . . . . . . . . . 1159.3 Maximum current noise as a function of barrier transparency of a nearby dot 1179.4 Device and measurement setup . . . . . . . . . . . . . . . . . . . . . . . . . 1199.5 Diagonal resistance as a function of dc current and magnetic field, at various

temperatures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1209.6 An example of the best fit to bias- and temperature-dependent tunneling

data with the weak tunneling formula . . . . . . . . . . . . . . . . . . . . . 1219.7 Fit error as a function of prefixed e∗ and g . . . . . . . . . . . . . . . . . . . 1229.8 Best-fit e∗ and g as a function of R0

D . . . . . . . . . . . . . . . . . . . . . . 1229.9 Bulk Hall resistance and diagonal resistance as a function of magnetic field,

showing 3/2 quantized plateaus in two different QPCs . . . . . . . . . . . . 1259.10 Temperature and bias dependence of the 3/2 plateaus . . . . . . . . . . . . 1269.11 Bulk Hall and longitudinal resistances as a function of magnetic field, at two

different temperatures and with two different crystal directions . . . . . . . 1299.12 Bulk longitudinal and Hall resistances as a function of dc current and mag-

netic field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

A.1 Photograph of a bank of simple RC filters . . . . . . . . . . . . . . . . . . . 132A.2 Photographs of sapphire circuit boards and heat sinks . . . . . . . . . . . . 135A.3 18 Mini-circuits VLFX-80 filters assembled with the cold finger . . . . . . . 136A.4 Quantum Hall bulk transport . . . . . . . . . . . . . . . . . . . . . . . . . . 136A.5 Thermalcoax cables assembled on the Microsoft fridge . . . . . . . . . . . . 139

B.1 Channel definition table . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141B.2 Virtual DACs parameters table . . . . . . . . . . . . . . . . . . . . . . . . . 143

C.1 Circuit schematics for calculating intrinsic conductance . . . . . . . . . . . 153C.2 Circuit schematics for calculating intrinsic current noise . . . . . . . . . . . 154

D.1 Circuit schematics for measuring conductance matrix and two-channel cur-rent noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

ix

Page 10: Waves, Particles, and Interactions in Reduced Dimensions

D.2 Multi-channel digital lock-in control panel . . . . . . . . . . . . . . . . . . . 158

x

Page 11: Waves, Particles, and Interactions in Reduced Dimensions

Acknowledgements

Writing the acknowledgement section of my thesis signals the end of a long journey,

and gives me a rare chance to look back at my six-year-long Ph.D. life doing experimental

condensed matter physics research at the Marcus lab, Harvard University. There are so

many people who have contributed to make this thesis possible, and I am deeply grateful

to all.

I would like to first thank my advisor Charlie Marcus, for taking me into his research

group, for his superb guidance and support over the years, for his deep insight into experi-

mental physics and unending steam of creative ideas, for educating me the fine art of being

an experimentalist, for the encouragement to open discussions, for teaching me to avoid

using the word “problems” unless precedented by the the word “solved”, and to replace

the use of “I” by “we”, for providing such a well-organized environment unmatched among

physics labs, and finally for teaching me hand-in-hand how to make a perfect latte. I will

cherish all of them as treasuries for life.

I would like to next extend special thanks to Leo DiCarlo and Doug McClure. Leo

had been my day-to-day mentor since I joined the lab the first day, teaching me almost

everything such as making a purchase order, wire-bonding, cryogenic techniques and low-

noise electronics. Doug, joining the lab as an undergraduate a few month before I did,

has been my collaborator on every single project. I feel especially fortunate to be able

to form the “noise team” with Leo and Doug. Our perseverance finally led us out of the

1/f noise dominated regime. By uniting everyone’s strengths, we were able to develop a

noise measurement setup of the state-of-the-art performance, with which we had achieved

xi

Page 12: Waves, Particles, and Interactions in Reduced Dimensions

fruitful results. After Leo’s departure to Yale as a post-doc for Rob Schoelkopf, Doug

and I continued our collaboration on the 5/2 project, and shared some initial success on

the quantum-Hall interferometry experiments. Over the years, I have been really glad to

observe that Doug has matured from a programming genius to a careful experimentalist

now taking the leading role in a very rewarding project.

I should thank all other lab mates, who have defined the unique culture of the Marcus

group with his or her talents and personalities. Reilly, as we call him, not only introduced to

us the “Reilly model” of the QPC 0.7 structure, but he has also been a great source of low-

temperature and high-frequency knowledge—in fact, he owns a vast breath of knowledge,

as he has been known as Reilli-pedia. Jeff Miller is both a fine scientist and a fine artist,

taking amazing pictures out of almost anything. In addition, he and Reilly both possess

the talent of mimicking lab mates, or professors, bringing unlimited supply of humor to the

lab. I am thankful to be able to work with Michi Yamamoto visiting from University of

Tokyo and Reinier Heeres visiting from Delft. Michi had been a very quick learner and very

helpful collaborator, helping me with the single-dot noise measurement, and after returning

to Japan, he was able to install and make improvements on the noise measurement setup in

the Tarucha group. Talented and hard-working, Reinier has helped me with both developing

the “Virtual DACs” routines, and getting the electrons cold (below 20 mK) in the Microsoft

fridge, after dozens of thermal cycles, fixing a newly-emerged problem during each one. I

must thank all other fellow students, post-docs, and visitors, with whom I have had the

privilege to work and have fun with, including Nadya Mason, Jason Petta, Slaven Garaj,

Ferdinand Kuemmeth, Max Lemme, Alex Johnson, Dominik Zumbuhl, Michael Biercuk,

Edward Laird, Jimmy Williams, Sang Chu (who taught me fab), Christian Barthel, Angela

Kou, Vivek Venkatachalam (from Amir Yacoby’s group), Eli Levenson-Falk, Hugh Churchill,

xii

Page 13: Waves, Particles, and Interactions in Reduced Dimensions

Yongjie Hu (from Charlie Lieber’s group), Maja Cassidy, Patrick Herring, Nathaniel Craig,

Jennifer Harlow, Jacob Aptekar, Shu Nakaharai, Susan Watson, and more.

Many thanks to my collaborators on the various projects that I have been involved,

including David Reilly, Michi Yamamoto, Hansres Engel, Bernd Rosenow, Jimmy Williams

and Eli Levenson-Falk, and to the material providers: Micah Hanson, Art Gossard at

UCSB, and Loren Pfeiffer, Ken West at Bell labs, without whom these projects will never be

possible. I would like to thank many fellow students, post-docs and leading scientists in the

field for helpful discussions, including Iuliana Radu, Jeff Miller, Izhar Neder, Nissim Ofek,

Bert Halperin, Marc Kastner, Amir Yacoby, Mike Stopa, Yigal Meir, David Goldhaber-

Gordon, Leonid Levitov, Moty Heiblum, Ady Stern, Claudio Chamon, Jim Eisenstein, Mike

Freedman, Xiao-Gang Wen, Alexei Kitaev, and many more with whom I have communicated

with. I also acknowledge the funding agencies that have supported my research during these

years, including NSF, ARO/ARDA/DTO, Harvard NSEC, Harvard CNS, Microsoft Project

Q and Harvard University.

Additionally, I would like to express my thanks to the lab administrators: James Got-

fredson, Danielle Reuter, and Jess Martin, for keeping the lab running smoothly. I would

like to thank Jim MacArthur at the electronics shop for building amazing DACs for us,

Mark Jackson, Nick Wilson, Nick Dent from Oxford Instruments for their assistance in fix-

ing the fridge problems, and Yuan Lu, JD Deng, Steve Shepard, Noah Clay, Ed Macomber,

John Tsakirgis at CNS for their help with device fabrications.

Moreover, I am thankful that I could be admitted into the Harvard physics program

six years ago. I am thankful to my teachers at Harvard, including Charlie Marcus, Bert

Halperin, Donheen Ham, David Nelson, Eugene Demler, Jene Golovchenko, Mark Irwin

and Yoonjung Lee. I am thankful to be able to gain teaching experience working with

xiii

Page 14: Waves, Particles, and Interactions in Reduced Dimensions

Paul Horowitz, Tom Hayes, and the other TF Anne Goodsell on the demanding course

Physics 123. I am thankful to Sheila Ferguson and other staff of the physics department for

their administrative and personal assistance. I am thankful that Prof. Bert Halperin, Prof.

Donhee Ham, and my advisor Charlie Marcus have agreed to be on my Ph.D. committee,

evaluating my progress and giving me support in every possible way.

I thank Benjamin N. Levy for being my host at Harvard, picking me up when I first

landed in the U.S., sharing meals and watching movies together. I thank all the friends I

met ever since the English Language Program in the summer of 2003, and I really appreciate

your company and friendship in a country so far away from my family.

Last but not least, I am deeply indebted to my parents for their love, trust, encourage-

ment, and never-ending support from around the globe in Hangzhou, China.

xiv

Page 15: Waves, Particles, and Interactions in Reduced Dimensions

Chapter 1

Introduction

In recent years, the advent of semiconductor technology, including the ability to grow

extremely pure and crystalline semiconductor materials, engineering of band-structures,

and advanced lithography methods, has enabled researchers to study numerous intriguing

quantum mechanical phenomena in systems of reduced dimensions [1, 2, 3, 4]. These studies

have been developed into a new branch of physics called mesoscopic physics, which aims

at understanding the world at length scales between the macroscopic and the microscopic

worlds.

When temperatures are lowered to within a few degrees or less from the absolute zero,

electrons in the mesoscopic world can behave sometimes like waves [4, 5, 6, 7, 8, 9, 10, 11],

and sometimes like particles [12, 13]; sometimes both properties coexist [14, 15, 16, 17,

18, 19, 20] and sometimes they compete [21, 22]. The presence of many-body interactions

introduces a whole new level of complications, producing phenomena such as the now-well-

understood Coulomb blockade [12, 13, 3], cotunneling [14, 18] and Kondo [17] physics in

quantum dots, the two-decade-old open question of the “0.7 structure” [23] in quantum

wires, and fractional quantum Hall effects [24] in two-dimensional electron systems under

a strong perpendicular magnetic field, with the possible existence of exotic non-Abelian

quasiparticles [25, 26]. Indeed, the wave-particle duality of electrons and the many-body

1

Page 16: Waves, Particles, and Interactions in Reduced Dimensions

interaction effects, in systems of reduced dimensions, have been the main themes of research

for mesoscopic physics.

1.1 Organization of this thesis

Along these lines, this thesis will present several experimental projects that I have accom-

plished during the six years of my Ph.D. life at Harvard, exploring the interplay between the

wave-particle duality of electrons and the interaction effects, in various mesoscopic devices.

In the rest of this introductory chapter, I will first introduce electronic transport mea-

surements, including dc transport and current noise auto- and cross correlations. Then I

will give a brief description of the material systems, based on which mesoscopic devices are

made, in particular GaAs/AlGaAs heterostructures. Finally in this chapter, I will describe

the basic properties of the various types of mesoscopic devices to be used in subsequent

chapters.

Chapters 2 through 8 form the main body of this thesis, presenting seven projects that

have been published or submitted for publication. Chapter 2 describes the detailed construc-

tion and operation of the two-channel current noise auto- and cross correlation detection

system. Chapter 3 studies shot-noise signatures of the 0.7 structure and spin in quantum

point contacts, as well as bias-dependent electron heating. Chapter 4 realizes sign reversal

of noise cross correlations in a capacitively-couple double quantum dot in a fully controlled

way. Chapter 5 reports the observation of super-Poissonian auto-correlation and positive

cross correlation in a multi-lead quantum dot, and establishes a proportionality between

auto- and cross correlations. Chapter 6 studies shot noise in graphene devices, revealing an

almost constant Fano factor in single-layer devices, and a density-dependent Fano factor in

a multi-layer device. Chapter 7 describes the observation of three distinct signatures for

2

Page 17: Waves, Particles, and Interactions in Reduced Dimensions

Coulomb blockade and Aharonov-Bohm interference in electronic Fabry-Perot interferome-

ters in the integer quantum Hall regime. In the large interferometer that Aharonov-Bohm

interference dominates, Chapter 8 studies edge-state velocity and bias-dependent dephasing

using non-linear transport. Then in the last chapter, I will describe four interesting, yet for

one reason or another unpublished experimental results.

The five appendices that follow provide more technical details related to the experiments

and theoretical calculations. Appendix A describes several methods of filtering and thermal

anchoring for fridge wiring, which are essential for achieving low electron temperatures.

Appendix B provides detailed explanation and source codes for implementing in Igor Pro

the set of tools called “Virtual DACs”, which allows easy definition and simultaneous control

of a linear combination of multiple independent parameters, and are especially useful when

working with devices of many gates. Appendix C gives derivations for the expressions used

to extract the intrinsic conductance matrix and multi-terminal current noise of the device

in the presence of finite-impedance external circuits. Appendix D describes the circuit

for simultaneous measurements of conductance matrix and multi-terminal current noise,

and also describes the operations of a multi-channel digital lock-in developed in-house and

capable of measuring 16 conductance matrix elements simultaneously. Finally, Appendix E

provides the Igor routines for the master equation calculation of current and noise in a

multi-lead, multi-level quantum dot.

3

Page 18: Waves, Particles, and Interactions in Reduced Dimensions

1.2 Dc transport and current noise

This section will describe electronic transport measurements, including the conventional di-

rect current (dc) transport, and measurements of current noise auto- and cross correlations.

All chapters will study dc transport, while studies of current noise will be the focus for

Chs. 3 through 6.

1.2.1 Dc transport

Since the inception of mesoscopic physics in the 1980’s, dc transport has been the ubiqui-

tous tool, and has played a key role in many discoveries of the field, including integer [5]

and fractional [27] quantum Hall effects, conductance quantization [6, 28] in quantum point

contacts, Coulomb blockade [12, 13] in closed quantum dots, universal conductance fluctua-

tion [4] in open quantum dots, and Aharonov-Bohm interference in ring structures [7], and

in Mach-Zehnder interferometers [11], etc.

Dc transport measures current I through the device in response to a source-drain volt-

age excitation V , or vice versa, extracting the device resistance R = V/I or conductance

G = I/V . Although termed dc transport, the resistances and conductances are often mea-

sured with lock-ins at near dc frequencies (3− 1000 Hz), because the measurement is often

much more quiet away from dc due to the ubiquitous 1/f noise present in devices and in

instruments.

To avoid smearing out features, the lock-in excitation should be kept below the tem-

perature or some intrinsic energy scale of the device. The lock-in excitation can also be

superimposed with another dc bias, to measure differential resistance r = δV/δI or differ-

ential conductance g = δI/δV in the non-linear regime.

There are two ways of biasing for dc transport: voltage bias, applying V and measuring

4

Page 19: Waves, Particles, and Interactions in Reduced Dimensions

I, is more suitable for devices with higher resistances (26 kΩ or higher); current bias, apply-

ing I and measuring V , on the other hand, is more suitable for devices of lower resistances.

There are also two types of measurement configurations: two-wire measurements, where

the voltage is either applied or probed at the same contacts as the source and the drain,

are sensitive to dc wire or ohmic contact resistances; four-wire measurements, where the

voltage probes are different from the source and drain contacts, on the other hand, can

eliminate the resistances from dc wires and ohmic contacts.

1.2.2 Current noise auto- and cross correlations

Current noise, the temporal fluctuation of currents, can yield complimentary information

to dc transport, such as electron temperature, transmission, quantum statistics, and many-

body interaction effects [19, 29, 30]. Chapters 3 through 6 provide studies of current noise

in various mesoscopic devices, including quantum point contacts [31, 32], single [33] and

double [34] quantum dots, as well as graphene devices [35].

We first need to understand two important concepts about current noise: power and

cross spectral densities, which are Fourier transforms of the auto- and cross correlation

functions of current fluctuations. Also, they have the physical meanings of power per unit

frequency, therefore they have the units of A2/Hz. In addition, they can be calculated by

taking the product of Fourier transforms of current fluctuations. The derivations of these

results are provided as follows.

Define the current Iα(t) at terminal α, and its Fourier transform Iα(f) as:

Iα(fn) =1T

∫ T

0Iα(t)e−i2πfntdt,

Iα(t) =+∞∑

n=−∞Iα(fn)ei2πfnt,

where T is the measurement time and fn = n/T . The time-averaged current is 〈Iα〉 =

5

Page 20: Waves, Particles, and Interactions in Reduced Dimensions

1T

∫ T0 Iα(t)dt = Iα(0), and the total cross-correlated power for δIα(t) and δIβ(t) is1:

〈δIαδIβ〉 = 〈IαIβ〉 − 〈Iα〉〈Iβ〉

=+∞∑

n,m=−∞〈Iα(fn)Iβ(fm)ei2πfn+mt〉 − Iα(0)Iβ(0)

=+∞∑

n=−∞Iα(fn)Iβ(−fn)− Iα(0)Iβ(0)

=+∞∑n=1

Iα(fn)I∗β(fn) + I∗α(fn)Iβ(fn)

=∫ +∞

0Sαβ(f)df, (1.1)

and Sαβ(f) = T · [Iα(f)I∗β(f) + I∗α(f)Iβ(f)], (1.2)

where Sαβ(f) is the power (for α = β) or cross (for α 6= β) spectral density.

We may also write:

Sαβ(f) = T · [Iα(f)I∗β(f) + I∗α(f)Iβ(f)]

=1T

∫ T

0dt1

∫ T

0dt2〈δIα(t1)δIβ(t2) + δIα(t2)δIβ(t1)〉e−i2πf(t1−t2)

=2T

∫ T

0dτ

∫ +∞

−∞dt ·Kαβ(t)e−i2πft

= 2Kαβ(f), (1.3)

where τ = (t1 + t2)/2, t = t1 − t2, Kαβ(t) = 12〈δIα(t)δIβ(0) + δIα(0)δIβ(t)〉 is the auto- or

cross correlation function, and Kαβ(f) =∫ +∞−∞ dt ·Kαβ(t)e−i2πft is its Fourier transform.

To summarize, Eq. (1.1) provides literal meanings to the power and cross spectral den-

sities, Eq. (1.2) gives the relationship between them and the Fourier transforms of currents,

and Eq. (1.3) gives the relationship between them and the Fourier transforms of auto- and

cross correlation functions of currents.

Current noise can be classified into different categories according to their frequency

1If α 6= β, it is the total cross-correlated power; otherwise, it is the total power of δIα(t).

6

Page 21: Waves, Particles, and Interactions in Reduced Dimensions

I

Time

(a)

SI

Frequency

(b)

I

Time

(c)

Log

SI

Log Frequency

(d)

1 / fI

Time

(e)

Log

S I

Log Frequency

(f)

1 / f 2

I

Time

(g)

Log

S I

Log Frequency

(h)

1 / f 2

I

Time

(i)

SI

Frequency

(j)

Figure 1.1: Different types of noise: (a,b) white noise, (c,d) 1/f noise, (e,f) Brownianmotion noise, (g,h) random telegraph noise, and (i,j) pick-ups or continuous-wave signals.They are presented in the time domain (a,c,e,g,i), and in the frequency domain as theirpower spectral densities (b,d,f,h,j).

dependence. Figure 1.1 shows different types of noise as a function of time, as well as

their power spectral densities, SI , as a function of frequency. White noise, implying the

auto-correlation to be a delta function in time, has a flat frequency dependence of SI , as

shown in Figs. 1.1(a) and (b). Flicker noise, also called 1/f noise, is the least understood

type; its power spectral density, as its name implies and shown in Figs. 1.1(c) and (d), has

a 1/f dependence on frequency. Brownian motion can be viewed as white noise integrated

over time, and its SI , as shown in Figs. 1.1(e) and (f), has a frequency dependence of 1/f2.

7

Page 22: Waves, Particles, and Interactions in Reduced Dimensions

Random telegraph noise occurs when the current jumps between two meta-stable states; as

shown in Figs. 1.1(g) and (h), its power spectral density is flat up to a corner frequency

given by the transition rates2, before rolling off as 1/f2. Pick-ups or continuous-wave signals

concentrate all the power at a certain frequency, and their power spectral densities are delta

functions in frequency, as shown in Figs. 1.1(i) and (j).

Among these types, white noise is what we are most interested in, because most of the

electron dynamics in mesoscopic systems happen at much shorter time scales than we can

measure, thus they appear white to the measurement setup. Indeed, no types of noise can

be strictly white; otherwise the total power would be infinite. Random telegraph noise,

for example, is white up to a corner frequency set by the rates of dynamics [36]. For this

reason, the noise we and most other people measure is often referred to as zero-frequency

noise in literatures [19, 29, 30].

In particular, two types of white noise are of interest to us. At equilibrium without bias,

only thermal noise is present, and has been used for gain calibration and measurement of

electron temperature, as will be described in detail in the next chapter. At a finite source-

drain bias greater than temperature, shot noise, arising from the quantization of charge

and partial transmission, dominates. It is the shot noise that makes noise measurement

interesting, allowing researches such as shot-noise thermometry [37], detection of fractional

charge [38, 39, 40, 41], observation of two-particle interference [42], and investigation of

interaction effects [43, 44, 45, 46, 34, 33], etc. Our studies of shot noise in different systems

will be presented in Chs. 3 to 6, following Ch. 2, which describes in detail how we measure

noise power and cross spectral densities.

2The power spectral density for random telegraph noise has a Lorentzian shape. Pleasesee Ref. [36] and Sec. 9.1 in Ch. 9 for the explicit expression.

8

Page 23: Waves, Particles, and Interactions in Reduced Dimensions

1.3 Material systems

The material systems of choice for mesoscopic physics research, including Si inversion

layers, GaAs/AlGaAs heterostructures, graphene, carbon nanotubes, and semiconductor

nanowires, often have one or more dimensions restricted, forming one- or two-dimensional

systems. These reduced dimensional systems are realized by the so-called size quantiza-

tion [1]: when the electron motion is restricted to within a certain size in one direction,

the kinetic energy associated with the motion in that direction becomes quantized and sub-

bands are formed; if the density is low enough such that the Fermi energy lies between

the lowest and the first excited sub-bands, and the temperature is much lower than the

sub-band energy spacing, the electron motion in that direction is frozen, effectively reduc-

ing the dimension by one and creating a two-dimensional system. In carbon nanotubes or

nanowires, two of the three dimensions are restricted by size quantization, thus they form

strictly one-dimensional systems.

In addition to reduced dimensions, these material systems often have several other

desired properties. First, they need to be contactable by electrodes, therefore can be probed

by transport measurements. Second, their carrier densities are usually quite low, giving large

Fermi wavelength (∼ 50nm) and allowing easy control by electrostatic gates, which can

further reduce the device dimensions down to zero. Finally, the quality of these materials

are quite high and their mean-free path quite long; in particular, the highest achieved

mobility in GaAs/AlGaAs heterostructures has exceeded 30, 000, 000 cm2/Vs [47, 41].

1.3.1 GaAs/AlGaAs heterostructures

Among these material systems, the GaAs/AlGaAs heterostructures, grown by molecular

beam epitaxy [48] under ultra high vacuum, are the most mature, most flexible and clean-

9

Page 24: Waves, Particles, and Interactions in Reduced Dimensions

Figure 1.2: Schematic of the GaAs/AlGaAs heterostructure for the 010219B wafer grown byMicah Hanson and Arthur Gossard at UCSB. On the right is its conduction band diagram.(Figure adapted from Ref. [49]).

est. They have been used throughout this thesis except Ch. 6. Figure 1.2 shows a schematic

of the GaAs/AlGaAs heterostructure used in Chs. 3 to 5 and grown by Micah Hanson and

Arthur Gossard at UCSB. This heterostructure starts with bulk GaAs substrate, then a

layer of AlxGa1−xAs (x ∼ 0.3), and finally a thin (∼ 10 nm) GaAs cap layer to prevent

oxidation. Inside the AlGaAs layer closer to the lower GaAs/AlGaAs interface than the

upper one, there is a thin δ-doping layer of Si. As shown on the right in Fig. 1.2, the Si

δ-doping layer lowers the conductance band edge towards the Fermi energy, but because Al-

GaAs has a much wider bandgap than GaAs, the global conduction band minimum is at the

GaAs/AlGaAs interface below the doping layer, crossing the Fermi energy. Mobile electrons

are trapped inside the triangular confining potential formed at this GaAs/AlGaAs interface.

Due to size quantization mentioned above, only the lowest sub-band of this confining poten-

tial is populated, creating a two-dimensional electron gas (2DEG). The separation between

the doping layer and the 2DEG is called modulation doping; together with a near perfect

match of lattice constants between GaAs and AlGaAs, it gives GaAs/AlGaAs 2DEG the

10

Page 25: Waves, Particles, and Interactions in Reduced Dimensions

highest mobility among solid state systems.

The GaAs/AlGaAs heterostructures used in Chs. 2, 3, 7, and 8 are grown by Ken West

and Loren Pfeiffer at Bell Labs. The heterostructure used in Chs. 2 and 3 is similar to

the UCSB wafer I just described; the heterostructure used in Chs. 7 and 8, however, is

different in several ways. First, the 2DEG is 200 nm away from the chip surface, and

located in a 30 nm wide GaAs quantum well sandwiched between AlGaAs layers, instead

of at the interface between GaAs and AlGaAs. Second, the sample is doubly doped, with

Si δ-doping layers 100 nm below and above the quantum well. Finally, the two Si layers

are each placed in separate narrow AlGaAs/GaAs/AlGaAs doping wells. These features

have made this heterostructure the highest mobility wafer that we have measured, with a

mobility of ∼ 20, 000, 000 cm2/Vs, two orders of magnitude higher than the UCSB wafer.

The 2DEG can be electrically contacted by Au/Ge ohmic contacts (see Fig. 1.2), en-

abling transport measurements. These contacts are made by depositing Ni/Au/Ge or

Pt/Au/Ge, followed by annealing at around 500 C for tens of seconds. The optimal

annealing recipe will depend on the specific 2DEG material.

Devices of arbitrary shape and size are created by applying negative voltages on sur-

face gates, usually made with Cr/Au or Ti/Au using electron-beam lithography. Negative

voltages raise the conduction band relative to the Fermi energy underneath the gates, and

if the conduction band minimum becomes higher than the Fermi energy, all the electrons

under the gates are depleted. This approach is especially flexible since each gate voltage

can be precisely controlled with independent digital-to-analog converters (DACs), tailoring

the confining potential of the device.

11

Page 26: Waves, Particles, and Interactions in Reduced Dimensions

1.4 Basic properties of mesoscopic devices

In this section, I will be introducing the basic properties of the various mesoscopic devices

to be studied in later chapters. Quantum point contacts (QPCs) are short one-dimensional

(1d) wires, and show quantized conductance as a function of gate voltage. Quantum dots

and double quantum dots are zero-dimensional (0d) systems; when their sizes are sufficiently

small, and their leads are sufficient opaque, Coulomb charging energy dominates, and they

exhibit single-electron tunneling behaviors. Subject to a strong perpendicular magnetic

field, two-dimensional (2d) electronic systems can exhibit integer and fractional quantum

Hall effects, when transport occurs along the edges, making the system effectively one-

dimensional. The subsequent three subsections will describe these properties in detail.

1.4.1 Quantum point contacts

Quantum point contacts, formed by depleting two facing gates as shown in the inset of

Fig. 1.3, are the simplest gated structure. The gates restrict the electron motion in one

direction, making the QPCs one-dimensional due to size quantization. As shown in Fig. 1.3,

the conductance g through a QPC, measured as a function of the gate voltage Vg, shows

steps of size 2e2/h. Indeed, the quantized conductance suggests the realization of an electron

waveguide, and is the hallmark of 1d ballistic transport [6, 1, 2].

The derivation of quantized conductance in 1d systems is pretty straightforward, as

given below. Consider that in a 1d system, the density of states with positive wave vector k

is ρ+(ε) = (dk/dε)/2π, and the velocity is v(ε) = (dε/dk)/~, where ε is the kinetic energy.

The product of these two quantities cancels the dispersion terms, giving a constant incident

rate per unit energy: ρ+(ε) · v(ε) = 1/h. When a source-drain voltage Vsd is applied across

a barrier in a 1d system, as shown in Fig. 1.4, the transmitted current is simply given by

12

Page 27: Waves, Particles, and Interactions in Reduced Dimensions

12

10

8

6

4

2

0g

[ e2 /

h ]

-1600 -1200 -800 -400 0Vg [ mV ]

Quantum Point Contact

500 nm

Figure 1.3: Conductance as a function of gate voltage in a quantum point contact. Inset:Scanning electron micrograph of a device with identical design to the one measured.

μdμs

k(ε)τ(ε)

ds

eVsd

Figure 1.4: Illustration of a source-drain voltage applied across a barrier in a 1d system,with transmission probability τ , in the zero-temperature limit.

the total indicant rate eVsd/h multiplied by the transmission probability τ and the electron

charge e: I = (e2/h)τ · Vsd. Therefore, the conductance is just g = (e2/h)τ . Note that

even in the absence of any barriers, with τ = 1, the conductance e2/h is still finite and only

related to fundamental constants, regardless of the length and width of the 1d system.

A QPC is not a single-channel 1d system though. First, the size of quantized conduc-

tance has been doubled to 2e2/h at zero magnetic field due to spin degeneracy, which can

be lifted at high magnetic fields, as in Ch. 3. Second, size quantization in the constricted

direction leads to sub-bands with sub-band energy spacing given by the confining poten-

13

Page 28: Waves, Particles, and Interactions in Reduced Dimensions

tial curvature. As a function of gate voltage, the sub-bands can be populated one by one,

leading to conductance steps, as observed in Fig. 1.3.

Between adjacent plateaus, the conductance risers are not infinitely sharp, but have

some width. A more realistic model for QPC is to assume the confining potential has the

form of a saddle point [50, 51]: V (x, y) = V0−mω2xx

2/2 +mω2yy

2/2, where x is the current

flowing direction, and y is the confinement direction. In this model, the sub-band energy

spacing is given by ~ωy. For the sub-band indexed n with sub-band edge at εn, the energy-

dependent transmission has the form τn(ε) = 1/(1 + e2π(εn−ε)/~ωx), therefore, the widths of

risers are given by ~ωx when kBT ~ωx.

Taking both spin and sub-bands into account, explicit expressions for transport, both

conductance and noise, at arbitrary bias can be calculated using the Landauer-Buttiker

formalism [52, 53, 54, 55]. The dc current is:

I =e

h

∫ ∑n,σ

τn,σ(ε)[fs(ε)− fd(ε)]dε, (1.4)

where σ denotes the spin, and fs (fd) is the Fermi distribution in the source (drain). The

differential conductance can be calculated by taking the derivative with respect to the

source-drain voltage, and assuming the bias the applied symmetrically:

g =e2

h

∫ ∑n,σ

τn,σ(ε)12

[−fs(ε)

ε− fd(ε)

ε

]dε

=e2

h

∫ ∑n,σ

τn,σ(ε)1

2kBTfs(ε)[1− fs(ε)] + fd(ε)[1− fd(ε)]dε. (1.5)

And finally, the current noise is given by:

SI =2e2

h

∫ ∑n,σ

τn,σ(ε)fs(ε)[1− fs(ε)] + fd(ε)[1− fd(ε)]dε

+2e2

h

∫ ∑n,σ

τn,σ(ε)[1− τn,σ(ε)][fs(ε)− fd(ε)]2dε. (1.6)

14

Page 29: Waves, Particles, and Interactions in Reduced Dimensions

Note that the first term in SI is exactly 4kBTg, which can be viewed as thermal noise,

even in the nonlinear regime3. This is why we define the partition noise as SPI ≡ SI−4kBTg

in Ch. 3, and its expression given in Eq. (3.2) is just the second term in Eq. (1.6).

Understanding most features in a QPC, we need to note one more subtle feature that

has remained an open problem until now: the shoulder-like feature near 0.7 × 2e2/h, as

can be seen in Fig. 1.3. This is termed 0.7 structure [23], and found to be related to

spin and many-body interactions, yet its exact origin is still under debate. Since the 0.7

structure happens below the 2e2/h plateau, there are only two spin channels, and the

natural question to ask is whether the transmission coefficients are different for the two spin

channels. Conductance would give the sum of them; the additional information brought

by the shot-noise measurements [56, 31], which is related to∑

σ τσ(1 − τσ), can give the

second equation needed. Indeed, the shot-noise study of the 0.7 structure, presented in

Ch. 3 show that they are different, and the spin is almost fully-polarized for conductance

above 0.7× 2e2/h even at zero magnetic field.

Quantum point contacts, as simple as they are, can already demonstrate the wave-

particle duality of electrons and many-body interaction effects. On the quantized con-

ductance plateaus, the electrons flow in channels of reflectionless waveguides, and show

vanishing shot noise. On the risers, on the other hand, shot noise arises in the presence of

partial transmission, demonstrating the particle nature of electrons. Finally, the 0.7 struc-

ture in QPCs, a many-body interaction effect, makes the single-particle picture to break

down, and has remained to be understood.

3This result assumes symmetrically applied bias. Although in experiments, bias is oftenapplied only on one side, interaction effects tend to effectively symmetrize the bias.

15

Page 30: Waves, Particles, and Interactions in Reduced Dimensions

1.4.2 Quantum dots

While the QPCs have left the electrons to flow freely in one direction, quantum dots confine

the electrons on 0d islands, which are connected to outside reservoirs by QPCs. Quantum

dots can be classified as open or closed dots, depending on the coupling between the dots

and the reservoirs. In open quantum dots, each QPC passes at least one channel, thus

the charge on the island is not quantized, and transport shows wave-like behaviors such as

universal conductance fluctuation and weak localization [4]. In closed quantum dots, when

the transmission through each QPC is less than one, Coulomb charging becomes dominant

and quantizes the charge on the island; transport occurs when electrons tunnel on and off

the island one by one, explicitly demonstrating the particle nature of electrons [3, 57]. Since

the open quantum dots are out of the scope of this thesis, I will only describe the closed dots

in the sequential tunneling limit below. Readers who wish to learn systematically about

quantum dots should refer to several excellent reviews [3, 4, 57, 58] available on these topics.

Shown in the inset of Fig. 1.5 is an example of a quantum dot design, which consists

of two plunger gates and two QPC leads. Conductance as a function of a plunger gate

voltage in the closed dot regime, as shown in Fig. 1.5, is zero almost everywhere, except

at some regularly spaced gate voltage points, drastically different from that in QPCs [see

Fig. 1.3]. This phenomenon, termed Coulomb blockade (CB) [3, 57], occurs because in

contrast to QPCs, where the non-interacting picture can explain most of the observations,

closed quantum dots are in the interaction-dominated limit, and electrons are quantized on

the island.

To understand the observed CB behavior, we first consider the simplest model of

16

Page 31: Waves, Particles, and Interactions in Reduced Dimensions

0.04

0.03

0.02

0.01

0.00g

[ e2 /

h ]

-860 -840 -820Vg [ mV ]

Quantum Dot 500 nm

Figure 1.5: Conductance as a function of gate voltage in a quantum dot. Inset: Scanningelectron micrograph of a device with identical design to the one measured.

Coulomb energy with N electrons on the dot [57]:

E =1

2C(CgVg − eN)2, (1.7)

where C is the dot total capacitance, and Cg is the capacitance between the dot and the

gate Vg. This energy is purely classical, and the only quantum ingredient needed is that the

QPCs are in the tunneling regime, quantizing the number of electrons on the island. The

number of electrons N is chosen to minimize this Coulomb energy, but since N is restricted

to be an integer, transport is blocked most of the time, except when CgVg/e equals half

integers, in which case N can fluctuate between two values without energy cost, giving finite

conductance. This model explains the CB behavior observed in Fig. 1.5, and gives the gate

voltage period between the CB peaks as e/Cg, expected to be roughly constant.

Figure 1.6 shows the differential conductance g = dI/dVsd as a function of Vg and Vsd,

revealing much richer structures. First, the conductance is zero in diamond-shaped regions,

which is the well-known Coulomb diamonds. Inside these diamonds, the electron number

is well defined, and transport is blocked. The boundaries of the diamonds show sharp

conductance peaks, and they converge to a single point at zero bias, corresponding to a

zero-bias CB peak seen in Fig. 1.5. In addition, outside the diamonds, there are lines of

17

Page 32: Waves, Particles, and Interactions in Reduced Dimensions

-1224 -1220 -1216Vg [ mV ]

2

1

0

-1

-2

Vsd

[ m

V ]

0.03

0.02

0.01

0.00

g [ e2 / h ]

Figure 1.6: Differential conductance as a function of source-drain bias and gate voltage ina quantum dot.

finite conductance parallel to the diamond boundaries, and ending at the boundaries with

the opposite slope. Moreover, the positive-slope lines generally have higher conductance

than the negative-slope lines.

These features can be understood from energy level diagrams and the electrostatic

considerations of the dot, and are captured by CB simulations in the sequential tunneling

limit4, as shown in Fig. 1.7. Near a diamond vertex at zero-bias, transport occurs when

electrons tunnel on and off the discrete energy levels of the dot, and the electron number

fluctuates between two integer values N and N + 1. Because differential conductance is

measured, the lines correspond to either the source (for positive-slope lines) or drain (for

negative-slope lines) chemical potential aligning with the dot levels. Therefore, the height

of the diamonds from zero-bias gives the charging energy e2/C (assuming charging energy is

much greater than level spacings), and the positive-slope lines have a slope of Cg/(C −Cs),

4The same simulations are also used in Ch. 5, and the source procedures are provided inAp. E.

18

Page 33: Waves, Particles, and Interactions in Reduced Dimensions

-1.0

-0.5

0.0

0.5

1.0

Vsd

[ m

V ]

-10 -5 0 5 10Vg [ mV ]

0.060.040.020.00

-10 -5 0 5 10Vg [ mV ]

g [ e2 / h ]

(a) Single-level simulation (b) Multi-level simulation

s d s d s ds d

A

A

B

B

C

C

D

D

Figure 1.7: (a) Single-level and (b) multi-level simulations of differential conductance asa function of source-drain bias and gate voltage of a quantum dot. At the four pointsindicated by A, B, C, and D, the corresponding energy diagrams are shown at the bottom.The parameters of the simulation are chosen rather arbitrarily.

while the negative-slope ones have a slope of −Cg/Cs, where Cs denotes the capacitance

between the source and the dot. Furthermore, the conductance for positive-slope lines

is proportional to (C − Cs)/C, and that for negative-slope lines is proportional to Cs/C;

since Cs/C is usually less than 0.5 (equals 0.3 for the simulations shown in Fig. 1.7), the

positive-slope lines would have higher conductance than the negative-slope lines.

All the observed features can be captured by the single-level CB picture described above,

as its simulation shows in Fig. 1.7(a), except the presence of lines outside the diamonds.

Indeed, they are related to transport through multiple excited states [59, 60, 61], and can

be easily taken into account in a multi-level simulation, as shown in Fig. 1.7(b).

At zero-bias, transport occurs only through one level: the lowest unoccupied level in

the N -electron ground state, which is also the highest occupied level in the (N +1)-electron

19

Page 34: Waves, Particles, and Interactions in Reduced Dimensions

ground state. We call the orbital levels above it electron excited levels, because they are

empty in the ground state and allow electrons to be excited on to them; similarly, we call

the orbital levels below hole excited levels, because they are full in the ground state, thus

allow holes to be excited on to them. The alignment of the chemical potential in the leads

to these excited levels can produce the lines outside the diamonds. The electron excited

levels can only be aligned with higher chemical potential of the two leads, while the hole

excited levels can only be aligned with the lower one. At negative bias, when the source

chemical potential is higher, an electron (hole) excited level produces a positive- (negative-)

slope line, as illustrated in the diagram C (D) of Fig. 1.7. At positive bias, on the other

hand, the drain chemical potential is higher, thus an electron (hole) excited level would

produce a negative- (positive-) slope line. As we have seen, transport through excited state

explains the presence of lines outside the diamonds; since there are ∼ 100 electrons in the

quantum dot studied in Fig. 1.6, there should be many orbital levels, producing many lines

observed outside the diamond regions.

1.4.3 Quantum Hall effects

A clean two-dimensional electron system, placed in a strong perpendicular magnetic field,

can exhibit integer [5] and fractional [27] quantum Hall (QH) effects, where the Hall conduc-

tance is precisely quantized at integer or fractional multiples of the conductance quantum

(e2/h). Quantum Hall effects are yet another example that wave and particle phenom-

ena as well as interaction effects show up in one way or another. While the integer QH

effects can be understood in a non-interacting picture, as will be described shortly, the

fractional QH effects will not be present without interactions. While the currents are

being carried in chiral waveguides with very little backscattering, measurements of shot

20

Page 35: Waves, Particles, and Interactions in Reduced Dimensions

0.10

0.08

0.06

0.04

0.02

0.00

RX

X [

h / e

2 ]

1086420 B [ T ]

1

1/2

1/3

1/41/51/61/81/10

3/4

3/5

0

RX

Y [ h / e2 ]

18

16

14

12

10

8

6

4

2

0

GX

Y [ e2 / h ]

1612840Filling Factor

Figure 1.8: Bulk longitudinal resistance RXX (red) and Hall resistance RXY (black) as afunction of magnetic field. Inset: Hall conductance GXY ≡ 1/RXY as a function of fillingfactor, converted from magnetic field using a sheet density of 2.6× 1015 m−2.

noise [38, 39, 40] have revealed fractional quasi-particle charge in the fractional QH regime.

While the current-carrying edges can be brought together to interfere in a Fabry-Perot in-

terferometer, Coulomb charging may dominate the transport and quantize the charge.

An example of transport measurement in the QH regime is shown in Fig. 1.8: bulk longi-

tudinal resistance RXX and Hall resistance RXY are measured as a function of perpendicular

magnetic field B, showing both integer and fractional QH plateaus in RXY with vanishing

RXX. The quantized conductance, regardless of sample sizes, reminds us of transport in

quantum point contacts. Indeed, we can plot in the inset of Fig. 1.8 Hall conductance

GXY ≡ 1/RXY as a function of the filling factor ν ≡ n/nφ, showing remarkable resemblance

to the QPC conductance versus gate voltage data in Fig. 1.3. Here, n is the electron sheet

density, nφ = B/φ0 is the flux density, and φ0 = h/e is the flux quantum.

21

Page 36: Waves, Particles, and Interactions in Reduced Dimensions

As mentioned above in Sec. 1.4.3, quantized conductance is the hallmark of ballistic

1d transport. Although the Hall bar sample is two-dimensional, a perpendicular magnetic

field discretizes the 2d Fermi sea into Landau levels (LL’s) at EN = (N + 1/2) · ~ωc, where

N (= 0, 1, 2, . . .) is the LL index, ωc = eB/m∗ is the cyclotron frequency, and m∗ is the

effective mass for electrons. Chiral transport occurs along the 1d channels where LL’s

cross the Fermi energy near the edges, in opposite directions along opposite edges. The

degeneracy of each LL, for each spin, is the same as the flux density nφ = B/φ0, therefore,

the number of filled LL’s is just the filling factor ν ≡ n/nφ. Since each LL contributes to

one 1d channel, the conductance is thus quantized to ν · e2/h for ν = integers. Due to

the chirality of edge transport, backscattering is strongly suppressed, and resistance can be

quantized to an unprecedented level such that it becomes the new standard for resistance.

Although integer QH effects are well understood in the non-interacting picture described

above, interactions are essential for fractional QH effects, as energy gaps need to open up

within the otherwise degenerate LL’s. Compared to integer QH effects, fractional QH

effects show much more interesting physics, yet they are also much less understood. In

addition to the fractional charge for quasi-particles in the fractional QH regime, as observed

in shot-noise measurements [38, 39, 40], these quasi-particles are expected to be neither

bosons nor fermions, but anyons that obey fractional or non-Abelian statistics. Interference

experiments in Fabry-Perot interferometers are proposed [62, 63, 64] to reveal their effects,

but several experiments [65, 66, 67, 68] on Fabry-Perot interferometers trying to observe

the anyonic statistics have remained inconclusive, possibly dominated by Coulomb blockade

rather than interference [69, 22]. Many other outstanding problems in the QH regime have

also remained to be answered, such as spin polarization, edge reconstruction and neutral

modes, anisotropic and reentrant states, etc.

22

Page 37: Waves, Particles, and Interactions in Reduced Dimensions

Chapter 2

System for measuring auto- andcross correlation of current noise atlow temperatures

L. DiCarlo1, Yiming Zhang1, D. T. McClure1, C. M. MarcusDepartment of Physics, Harvard University, Cambridge, Massachusetts 02138

L. N. Pfeiffer, K. W. WestAlcatel-Lucent, Murray Hill, New Jersey 07974

We describe the construction and operation of a two-channel noise detection system

for measuring power and cross spectral densities of current fluctuations near 2 MHz in

electronic devices at low temperatures. The system employs cryogenic amplification and

fast Fourier transform based spectral measurement. The gain and electron temperature are

calibrated using Johnson noise thermometry. Full shot noise of 100 pA can be resolved with

an integration time of 10 s.2

1These authors contributed equally to this work.

2This chapter is adapted with permission from Rev. Sci. Instrum. 77, 073906 (2006). c©(2006) by the American Institute of Physics.

23

Page 38: Waves, Particles, and Interactions in Reduced Dimensions

2.1 Introduction

Over the last decade, measurement of electronic noise in mesoscopic conductors has suc-

cessfully probed quantum statistics, chaotic scattering and many-body effects [19, 29, 30].

Suppression of shot noise below the Poissonian limit has been observed in a wide range of

devices, including quantum point contacts [70, 71, 72], diffusive wires [73, 74], and quan-

tum dots [75], with good agreement between experiment and theory. Shot noise has been

used to measure quasiparticle charge in strongly correlated systems, including the fractional

quantum Hall regime [38, 39, 40] and normal-superconductor interfaces [76], and to inves-

tigate regimes where Coulomb interactions are strong, including coupled localized states in

mesoscopic tunnel junctions [43] and quantum dots in the sequential tunneling [77, 46] and

cotunneling [45] regimes. Two-particle interference not evident in dc transport has been

investigated using noise in an electronic beam splitter [72].

Recent theoretical work [78, 79, 80, 81] proposes the detection of electron entanglement

via violations of Bell-type inequalities using cross-correlations of current noise between

different leads. Most noise measurements have investigated either noise autocorrelation

[70, 73, 38, 82, 72, 37, 45] or cross correlation of noise in a common current [71, 83, 74, 84,

75, 43], with only a few experiments [85, 86, 87, 88] investigating cross correlation between

two distinct currents. Henny et al. [85, 86] and Oberholzer et al. [87] measured noise

cross correlation in the acoustic frequency range (low kilohertz) using room temperature

amplification and a commercial fast Fourier transform (FFT)-based spectrum analyzer.

Oliver et al. [88] measured cross correlation in the low megahertz using cryogenic amplifiers

and analog power detection with hybrid mixers and envelope detectors.

In this chapter, we describe a two-channel noise detection system for simultaneously

measuring power spectral densities and cross spectral density of current fluctuations in

24

Page 39: Waves, Particles, and Interactions in Reduced Dimensions

electronic devices at low temperatures. Our approach combines elements of the two meth-

ods described above: cryogenic amplification at low megahertz frequencies and FFT-based

spectral measurement.

Several factors make low-megahertz frequencies a practical range for low-temperature

current noise measurement. This frequency range is high compared to the 1/f noise corner

in typical mesoscopic devices. Yet, it is low enough that FFT-based spectral measurement

can be performed efficiently with a personal computer (PC) equipped with a commercial

digitizer. Key features of this FFT-based spectral measurement are near real-time opera-

tion and sufficient frequency resolution to detect spectral features of interest. Specifically,

the fine frequency resolution provides information about the measurement circuit and am-

plifier noise at megahertz, and enables extraneous interference pickup to be identified and

eliminated. These two features constitute a significant advantage over both wideband ana-

log detection of total noise power, which sacrifices resolution for speed, and swept-sine

measurement, which sacrifices speed for resolution.

2.2 Overview of the system

Figure 2.1 shows a block diagram of the two-channel noise detection system, which is inte-

grated with a commercial 3He cryostat (Oxford Intruments Heliox 2VL). The system takes

two input currents and amplifies their fluctuations in several stages. First, a parallel resistor-

inductor-capacitor (RLC) circuit performs current-to-voltage conversion at frequencies close

to its resonance at fo = (2π√LC)−1 ≈ 2 MHz. Through its transconductance, a high elec-

tron mobility transistor (HEMT) operating at 4.2 K converts these voltage fluctuations into

current fluctuations in a 50 Ω coaxial line extending from 4.2 K to room temperature. A

50 Ω amplifier with 60 dB of gain completes the amplification chain. The resulting signals

25

Page 40: Waves, Particles, and Interactions in Reduced Dimensions

0.3 K 4.2 K

Digitize& FFT

Cross Spectrum

PowerSpectrum 1

V1

V2

R L C 60 dB

R L C 60 dB

Device

PowerSpectrum 2

Figure 2.1: Block diagram of the two-channel noise detection system, configured to measurethe power spectral densities and cross spectral density of current fluctuations in a multi-terminal electronic device.

V1 and V2 are simultaneously sampled at 10 MS/s by a two-channel digitizer (National

Instruments PCI-5122) in a 3.4 GHz PC (Dell Optiplex GX280). The computer takes the

FFT of each signal and computes the power spectral density of each channel and the cross

spectral density.

2.3 Amplifier

2.3.1 Design objectives

A number of objectives have guided the design of the amplification lines. These include (1)

low amplifier input-referred voltage noise and current noise. (2) simultaneous measurement

of both noise at megahertz and transport near dc, (3) low thermal load, (4) small size,

allowing two amplification lines within the 52 mm bore cryostat, (5) maximum use of

commercial components, and (6) compatibility with high magnetic fields.

2.3.2 Overview of the circuit

Each amplification line consists of four circuit boards interconnected by coaxial cables,

as shown in the circuit schematic in Fig. 2.2(a). Three of the boards are located inside

26

Page 41: Waves, Particles, and Interactions in Reduced Dimensions

4.2 K

Sapphire

C2

R3C3

Q1

UT-85

1.6 KVdac

300 K

FR-4UT-85

R5

R6 R7

C6

C5

C4

LOCK IN

Ih

AU1447

SS/SS

SR560R2 R4

SS/SS

CRYOAMP

DIG.SPLITTER

SINK

C1

R1L1

UT-34CRES

1 cm

0.3 K

0.3 K

R1R2R3R4R5R6R7C1C2C3C4C5C6L1

51050

15011

102x515222.22.222

2x33

(a)

(b) (c)

Figure 2.2: (a) Schematic diagram of each amplification line. Values of all passive compo-nents are listed in the accompanying table. Transistor Q1 is an Agilent ATF-34143 HEMT.(b) Layout of the CRYOAMP circuit board. Metal (black regions) is patterned by etchingof thermally evaporated Cr/Au on sapphire substrate. (c) Photograph of a CRYOAMPboard. The scale bar applies to both (b) and (c).

the 3He cryostat. The resonant circuit board [labeled RES in Fig. 2.2(a)] is mounted on

the sample holder at the end of the 30 cm long coldfinger that extends from the 3He

pot to the center of the superconducting solenoid. The heat-sink board (SINK) anchored

to the 3He pot is a meandering line that thermalizes the inner conductor of the coaxial

cable. The CRYOAMP board at the 4.2 K plate contains the only active element operating

cryogenically, an Agilent ATF-34143 HEMT. The four-way SPLITTER board operating

at room temperature separates low- and high- frequency signals and biases the HEMT.

Each line amplifies in two frequency ranges, a low-frequency range below ∼ 3 kHz and a

high-frequency range around 2 MHz.

The low-frequency equivalent circuit is shown in Fig. 2.3(a): a resistor (R1 = 5 kΩ)

to ground, shunted by a capacitor (C1 = 10 nF), converts an input current i to a voltage

on the HEMT gate. The HEMT amplifies this gate voltage by ∼ −5 V/V on its drain,

which connects to a room temperature voltage amplifier at the low frequency port of the

27

Page 42: Waves, Particles, and Interactions in Reduced Dimensions

SR560

(a) (b)

150 Ω

50 Ω

5 kΩ 96 pF10 nF

66 µH

IhVdac

i i

Vh,d

AU-1447

DMM

50 ΩVh,ds

+

Figure 2.3: Equivalent circuits characterizing the amplification line in the (a) low-frequencyregime (up to ∼ 3 kHz), where it is used for differential conductance measurements, and inthe (b) high-frequency regime (few megahertz), where it is used for noise measurement.

SPLITTER board. The low-frequency voltage amplifier (Stanford Research Systems model

SR560) is operated in single-ended mode with ac coupling, 100 V/V gain and bandpass

filtering (30 Hz to 10 kHz). The bandwidth in this low-frequency regime is set by the input

time constant.

The high-frequency equivalent circuit is shown in Fig. 2.3(b). The inductor L1 = 66 µH

dominates over C1 and forms a parallel RLC tank with R1 and the capacitance C ∼ 96 pF

of the coaxial line connecting to the CRYOAMP board. Resistor R4 is shunted by C2 to

enhance the transconductance at the CRYOAMP board. The coaxial line extending from

4.2 K to room temperature is terminated on both sides by 50 Ω. At room temperature, the

signal passes through the high-frequency port of the SPLITTER board to a 50 Ω amplifier

(MITEQ AU-1447) with a gain of 60 dB and a noise temperature of 100 K in the range

0.01− 200 MHz.

2.3.3 Operating point

The HEMT must be biased in saturation to provide voltage (transconductance) gain in the

low (high) frequency range. R4, R5 + R6 and supply voltage Vdac determine the HEMT

28

Page 43: Waves, Particles, and Interactions in Reduced Dimensions

operating point (R1 grounds the HEMT gate at dc). A notable difference in this design

compared to similar published ones regards the placement of R4. In previous implementa-

tions of similar circuits [89, 90, 91], R4 is a variable resistor placed outside the refrigerator

and connected to the source lead of Q1 via a second coaxial line or low-frequency wire.

Here, R4 is located on the CRYOAMP board to simplify assembly and save space, at the

expense of having full control of the bias point in Q1 (R4 fixes the saturation value of the

HEMT current Ih). Using the I-V curves in Ref. [91] for a cryogenically cooled ATF-34143,

we choose R4 = 150 Ω to give a saturation current of a few mA. This value of satura-

tion current reflects a compromise between noise performance and power dissipation. As

shown in Fig. 2.4, Q1 is biased by varying the supply voltage Vdac fed at the SPLITTER

board. At the bias point indicated by a cross, the total power dissipation in the HEMT

board is IhVh,ds + I2hR4 = 1.8 mW, and the input-referred voltage noise of the HEMT is

∼ 0.4 nV/√

Hz.

2.3.4 Passive components

Passive components were selected based on temperature stability, size and magnetic field

compatibility. All resistors (Vishay TNPW thin film) are 0805-size surface mount. Their

variation in resistance between room temperature and 300 mK is < 0.5%. Inductor L1 (two

33 µH Coilcraft 1812CS ceramic chip inductors in series) does not have a magnetic core

and is suited for operation at high magnetic fields. The dc resistance of L1 is 26(0.3) Ω at

300(4.2) K. With the exception of C1, all capacitors are 0805-size surface mount (Murata

COG GRM21). C1 (two 5 nF American Technical Ceramics 700B NPO capacitors in

parallel) is certified nonmagnetic.

29

Page 44: Waves, Particles, and Interactions in Reduced Dimensions

Figure 2.4: Drain current Ih as a function of HEMT drain-source voltage Vh,ds, with theHEMT board at temperatures of 300 K (dashed) and 4.2 K (solid). These curves wereobtained by sweeping the supply voltage Vdac and measuring drain voltage Vh,d with anHP34401A digital multimeter (see Fig. 2.3(a)). From Vh,d and Vdac, Ih and Vh,ds were thenextracted. Dotted curves are contours of constant power dissipation in the HEMT board.The HEMT is biased in saturation (cross).

2.3.5 Thermalization

To achieve a low device electron temperature, circuit board substrates must handle the heat

load from the coaxial line. The CRYOAMP board must also handle the power dissipated

by the HEMT and R4. Sapphire, having good thermal conductivity at low temperatures

[92] and excellent electrical insulation, is used for the substrate in the RES, SINK and

CRYOAMP boards. Polished blanks, 0.02 in. thick and 0.25 in. wide, were cut to lengths

of 0.6 in. (RES and CRYOAMP) or 0.8 in. (SINK) using a diamond saw. Both planar

surfaces were metallized with thermally evaporated Cr/Au (30/300 nm). Circuit traces were

then defined on one surface using a Pulsar toner transfer mask and wet etching with Au

and Cr etchants (Transene types TFA and 1020). Surface mount components were directly

soldered.

The RES board is thermally anchored to the sample holder with silver epoxy (Epoxy

30

Page 45: Waves, Particles, and Interactions in Reduced Dimensions

Technology 410E). The CRYOAMP (SINK) board is thermalized to the 4.2 K plate (3He

pot) by a copper braid soldered to the back plane.

Semirigid stainless steel coaxial cable (Uniform Tube UT85-SS/SS) is used between the

SINK and CRYOAMP boards, and between the CRYOAMP board and room temperature.

Between the RES and SINK boards, smaller coaxial cable (Uniform Tube UT34-C) is used

to conserve space.

With this approach to thermalization, the base temperature of the 3He refrigerator is

290 mK with a hold time of ∼ 45 h. As demonstrated further below, the electron base

temperature in the device is also 290 mK.

2.4 Digitization and FFT processing

The amplifier outputs V1 and V2 (see Fig. 2.1) are sampled simultaneously using a commer-

cial digitizer (National Instruments PCI-5122) with 14-bit resolution at a rate fs = 10 MS/s.

To avoid aliasing [93] from the broadband amplifier background, V1 and V2 are frequency

limited to below the Nyquist frequency of 5 MHz using 5-pole Chebyshev low-pass filters,

built in-house from axial inductors and capacitors with values specified by the design recipe

in Ref. [94]. The filters have a measured half power frequency of 3.8 MHz, 39 dB suppression

at 8 MHz and a passband ripple of 0.03 dB.

While the digitizer continuously stores acquired data into its memory buffer (32 MB per

channel), a software program processes the data from the buffer in blocks of M = 10 368

points per channel. M is chosen to yield a resolution bandwidth fs/M ∼ 1 kHz, and to be

factorizable into powers of two and three to maximize the efficiency of the FFT algorithm.

Each block of data is processed as follows. First, V1 and V2 are multiplied by a Hanning

window WH [m] =√

2/3[1 − cos(2πm/M)] to avoid end effects [93]. Second, using the

31

Page 46: Waves, Particles, and Interactions in Reduced Dimensions

FFTW package [95], their FFTs are calculated:

V1(2)[fn] =M−1∑m=0

WH[m]V1(2)(tm)e−i2πfntm , (2.1)

where tm = m/fs, fn = (n/M)fs, and n = 0, 1, ...,M/2. Third, the power spectral densities

P1,2 = 2|V1,2|2/(Mfs) and the cross spectral density X = 2(V ∗1 · V2)/(Mfs) = XR + iXI are

computed.

As blocks are processed, running averages of P1, P2, and X are computed until the

desired integration time τint is reached. With the 3.4 GHz computer and the FFTW algo-

rithm, these computations are carried out in nearly real-time: it takes 10.8 s to acquire and

process 10 s of data.3

2.5 Measurement example: quantum point contact

In this section, we demonstrate the two-channel noise detection system with measurements

of a quantum point contact (QPC). While the investigation of bias-dependent current noise

in QPCs is the main topic of Ch. 3, we here describe the techniques used for measuring

dc transport, as well as the circuit model and the calibration (based on Johnson-noise

thermometry) that are used for extracting QPC noise.

2.5.1 Setup

A gate-defined QPC4 is connected to the system as shown in the inset of Fig. 2.5. The two

amplification lines are connected to the same reservoir of the QPC. In this case, the two input

3We have achieved fully real-time operation with a newly purchased 3.16 GHz Core 2Duo computer.

4This device is named QPC 1 in Ch. 3.

32

Page 47: Waves, Particles, and Interactions in Reduced Dimensions

500nm

VdcVac

Vg1

Vg2

Figure 2.5: Inset: setup for detection of QPC current noise using cross-correlation, andelectron micrograph of a device identical in design to the one used. The QPC is defined bynegative voltages Vg1 and Vg2 applied on two facing gates. All other gates in the device aregrounded. Main: linear conductance g(Vsd = 0) as a function of Vg2 at 290 mK, measuredusing amplification line 1. Vg1 = −3.2 V.

RLC tanks effectively become a single tank with resistance R′ ≈ 2.5 kΩ, inductance L′ ≈

33 µH and capacitance C ′ ≈ 192 pF. The QPC current noise couples to both amplification

lines and thus can be extracted from either the single channel power spectral densities or

the cross spectral density. The latter has the technical advantage of rejecting any noise not

common to both amplification lines.

2.5.2 Measuring dc transport

A 25 µVrms, 430 Hz excitation Vac is applied to the other QPC reservoir and used for lock-in

measurement of g. A dc bias voltage Vdc is also applied to generate a finite Vsd. Vsd deviates

from Vdc due to the resistance in-line with the QPC, which is equal to the sum of R1/2 and

ohmic contact resistance Rs. Vsd could in principle be measured by the traditional four-

wire technique. This would require additional low-frequency wiring, as well as filtering to

prevent extraneous pick-up and room-temperature amplifier noise from coupling to the noise

33

Page 48: Waves, Particles, and Interactions in Reduced Dimensions

!"

!#

!"

Figure 2.6: Power spectral densities P1 and P2, and real and imaginary parts XR and XI

of the cross spectral density, at base temperature and with the QPC pinched off (g = 0),obtained from noise data acquired for τint = 20 s. Inset: expanded view of XR nearresonance, along with a fit using Eq. (2.3) over the range 1.7 to 2.3 MHz.

measurement circuit. For technical simplicity, here Vsd is obtained by numerical integration

of the measured bias-dependent g:

Vsd =∫ Vdc

0

dV

1 + (R1/2 +Rs)g(V )(2.2)

Figure 2.5 shows linear conductance g(Vsd = 0) as a function of gate voltage Vg2, at a fridge

temperature Tfridge = 290 mK (base temperature). Here, g was extracted from lock-in

measurements using amplification line 1. As neither the low frequency gain of amplifier 1

nor Rs were known precisely beforehand, these parameters were calibrated by aligning the

observed conductance plateaus to the expected multiples of 2e2/h. This method yielded a

low-frequency gain −4.6 V/V and Rs = 430 Ω.

2.5.3 Measuring noise

Figure 2.6 shows P1, P2, XR, and XI as a function of frequency f , at base temperature and

with the QPC pinched off (g = 0). P1(2) shows a peak at the resonant frequency of the RLC

tank, on top of a background of approximately 85(78)× 10−15 V2/Hz. The background in

34

Page 49: Waves, Particles, and Interactions in Reduced Dimensions

P1(2) is due to the voltage noise SV,1(2) of amplification line 1(2) (∼ 0.4 nV/√

Hz). The peak

results from thermal noise of the resonator resistance and current noise (SI,1 + SI,2) from

the amplifiers5. XR picks out this peak and rejects the amplifier voltage noise backgrounds.

The inset zooms in on XR near the resonant frequency. The solid curve is a best-fit to the

form

XR(f) =X0R

1 + (f2 − f2o )2/(f∆f3dB)2

, (2.3)

corresponding to the lineshape of white noise band-pass filtered by the RLC tank. The

fit parameters are the peak height X0R, the half-power bandwidth ∆f3dB and the peak

frequency fo. Power spectral densities P1(2) can be fit to a similar form including a fitted

background term:

P1(2)(f) = PB1(2) +

P 01(2)

1 + (f2 − f2o )2/(f∆f3dB)2

. (2.4)

2.5.4 System calibration using Johnson noise

Chapter 3 presents measurements of QPC excess noise, defined as SPI (Vsd) = SI(Vsd) −

4kBTeg(Vsd) (SI is the total QPC current noise spectral density). The extraction of SPI

from measurements of X0R requires that a circuit model for the noise detection system be

defined and that all its parameters be calibrated in situ. The circuit model we use is shown

in Fig. 2.7. Within this model,

SPI =

(X0R

G2X

− 4kBTeReff

)(1 + gRsReff

)2

, (2.5)

where GX =√G1G2 is the cross-correlation gain and Reff = 2πf2

oL′/∆f3dB is the total

effective resistance parallel to the tank6. Calibration requires assigning values for Rs, Te,

5Further below, the contribution from amplifier current noise is shown to be negligible.

6Within the model, Reff = (1/(1/g + Rs) + 1/R′)−1. The best-fit ∆f3dB to the mea-surement shown in Fig. 2.6 with the QPC pinched off (g = 0) gives R′ = 2.4 kΩ. This

35

Page 50: Waves, Particles, and Interactions in Reduced Dimensions

C'L'R'

Rs

4kBTe/Rs

4kBTe/R'4kBTegSIP

g

G1

SV,1

SI,1

G2

SV,2

SI,2

Figure 2.7: Circuit model used for extraction of the QPC partition noise SPI . G1(2) is the

voltage gain of amplification line 1(2) between HEMT gate and digitizer input.

and GX . While the value Rs is obtained from the conductance measurement, GX and Te

are calibrated from thermal noise measurements. The procedure demonstrated in Fig. 2.8

stems from the relation7 X0R = 4kBTeReffG

2X , valid at Vsd = 0.

First, XR(f) is measured over τint = 30 s for various Vg2 settings at each of three elevated

fridge temperatures (Tfridge = 3.1, 4.2, and 5.3 K). X0R and Reff are extracted from fits to

XR(f) using Eq. (2.3) and plotted parametrically [open markers in Fig. 2.8(a)]. A linear fit

(constrained to pass through the origin) to each parametric plot gives the slope dX0R/dReff

at each temperature, equal to 4kBTeG2X . Assuming Te = Tfridge at these temperatures,

GX = 790 V/V is extracted from a linear fit to dX0R/dReff(Tfridge), shown in Fig. 2.8(b).

Next, the base electron temperature is calibrated from a parametric plot of X0R as a

function of Reff obtained from similar measurements at base temperature [solid circles in

Fig. 2.8(a)]. From the fitted slope dX0R/dReff [black marker in Fig. 2.8(b)] and using the

calibrated GX , a value Te = 290 mK is obtained. This suggests that electrons are well

small reduction from 2.5 KΩ reflects small inductor and capacitor losses near the resonantfrequency.

7The full expression within the circuit model is X0R = (4kBTeReff + (SI,1 +SI,2)R2

eff)G2X .

The linear dependence of X0R on Reff observed in Fig. 2.8(a) demonstrates that the quadratic

term from amplifier current noise is negligible.

36

Page 51: Waves, Particles, and Interactions in Reduced Dimensions

Figure 2.8: Calibration by noise thermometry of the electron temperature Te at base fridgetemperature and the cross-correlation gain GX . (a) X0

R as function of Reff (both from fitsto XR(f) using Eq. (2.3)), at base (solid circles) and at three elevated fridge temperatures(open markers). Solid lines are linear fits constrained to the origin. (b) Slope dX0

R/dReff

(from fits in (a)) as a function of Tfridge. Solid line is a linear fit (constrained to the origin)of dX0

R/dReff at the three elevated temperatures (open markers).

thermalized to the fridge.

2.6 System performance

The resolution in the estimation of current noise spectral density from one-channel and

two-channel measurements is determined experimentally in this section. Noise data are

first sampled over a total time τtot = 1 h, with the QPC at base temperature and pinched

37

Page 52: Waves, Particles, and Interactions in Reduced Dimensions

Figure 2.9: (a) X0R as a function of time t, for τint of 10 s (open circles) and 100 s (solid

circles). (b) Standard deviations σ1 and σR as a function of τint. The solid line is a fit toσR of the form Cτ

−1/2int , with best-fit value C = 0.30 × 10−15 s1/2V2/Hz. (c) σR/

√σ1σ2 as

a function of τint. The dashed line is a constant 1/√

2.

off. Dividing the data in segments of time length τint, calculating the power and cross

spectral densities for each segment, and fitting with Eqs. (2.3) and (2.4) gives a sequence of

τtot/τint peak heights for each of P1, P2 and XR. Shown in open (solid) circles in Fig. 2.9(a)

is X0R as a function of time t for τint = 10(100) s. The standard deviation σR of X0

R is

1(0.3) × 10−16 V2/Hz. The resolution δSI in current noise spectral density is given by

σR/(G2XR

2eff) [see Eq. (2.5)]. For τint = 10 s, δSI = 2.8× 10−29 A2/Hz, which corresponds

to full shot noise 2eI of I ∼ 100 pA.

The effect of integration time on the resolution is determined by repeating the analysis

38

Page 53: Waves, Particles, and Interactions in Reduced Dimensions

for different values of τint. Fig. 2.9(b) shows the standard deviation σ1 (σR) of P 01 (X0

R) as a

function of τint. The standard deviation σ2 of P 02 , not shown, overlaps closely with σ1. All

three standard deviations scale as 1/√τint, consistent with the Dicke radiometer formula

[96] which applies when measurement error results only from finite integration time, i.e., it

is purely statistical. This suggests that, even for the longest segment length of τint = 10 min,

the measurement error is dominated by statistical error and not by instrumentation drift

on the scale of 1 h.

Figure 2.9(c) shows σR/√σ1σ2 as a function of τint. This ratio gives the fraction by

which, in the present measurement configuration, the statistical error in current noise spec-

tral density estimation from X0R is lower than the error in the estimation from either P 0

1 or

P 02 alone. The geometric mean in the denominator accounts for any small mismatch in the

gains G1 and G2. In theory, and in the absence of drift, this ratio is independent of τint and

equal to 1/√

2 when the uncorrelated amplifier voltage noise [SV,1(2)] dominates over the

noise common to both amplification lines. The ratio would be unity when the correlated

noise dominates over SV,1(2).

The experimental σR/√σ1σ2 is close to 1/

√2 (dashed line). This is consistent with

the spectral density data in Fig. 2.6, which shows that the backgrounds in P1 and P2 are

approximately three times larger than the cross-correlation peak height. The ratio deviates

slightly below 1/√

2 at the largest τint values. This may result from enhanced sensitivity to

error in the substraction of the P1(2) background at the longest integration times.

A similar improvement relative to estimation from either P 01 or P 0

2 alone would also

result from estimation with a weighted average (P 01 /G

21 + P 0

2 /G22)G2

X/2. The higher res-

olution attainable from two channel measurement relative to single-channel measurement

in this regime has been previously exploited in noise measurements in the kilohertz range

39

Page 54: Waves, Particles, and Interactions in Reduced Dimensions

[71, 83, 84].

2.7 Discussion

We have presented a two-channel noise detection system measuring auto- and cross corre-

lation of current fluctuations near 2 MHz in electronic devices at low temperatures. The

system has been implemented in a 3He refrigerator where the base device electron temper-

ature, measured by noise thermometry, is 290 mK. Similar integration with a 3He -4He

dilution refrigerator would enable noise measurement at temperatures of tens of millikelvin.

2.8 Acknowledgements

We thank N. J. Craig, J. B. Miller, E. Onitskansky, and S. K. Slater for device fabrication.

We also thank H.-A. Engel, D. C. Glattli, P. Horowitz, W. D. Oliver, D. J. Reilly, P. Roche,

A. Yacoby, Y. Yamamoto for valuable discussion, and B. D’Urso, F. Molea and H. Steinberg

for technical assistance. We acknowledge support from NSF-NSEC, ARDA/ARO, and

Harvard University.

40

Page 55: Waves, Particles, and Interactions in Reduced Dimensions

Chapter 3

Current noise in quantum pointcontacts

L. DiCarlo1, Yiming Zhang1, D. T. McClure1, D. J. Reilly, C. M. MarcusDepartment of Physics, Harvard University, Cambridge, Massachusetts 02138

L. N. Pfeiffer, K. W. WestAlcatel-Lucent, Murray Hill, New Jersey 07974

M. P. Hanson, A. C. GossardDepartment of Materials, University of California, Santa Barbara, California 93106

We present measurements of current noise in QPCs as a function of source-drain bias,

gate voltage, and in-plane magnetic field. At zero bias, Johnson noise provides a measure

of the electron temperature. At finite bias, shot noise at zero field exhibits an asymme-

try related to the 0.7 structure in conductance. The asymmetry in noise evolves smoothly

into the symmetric signature of spin-resolved electron transmission at high field. Compari-

son to a phenomenological model with density-dependent level splitting yields quantitative

agreement. Additionally, a device-specific contribution to the finite-bias noise, particularly

visible on conductance plateaus (where shot noise vanishes), agrees quantitatively with a

model of bias-dependent electron heating.2

1These authors contributed equally to this work.

2This chapter is adapted from Ref. [32] and from Phys. Rev. Lett. 97, 036810 (2006)[with permission, c© (2006) by the American Physical Society].

41

Page 56: Waves, Particles, and Interactions in Reduced Dimensions

3.1 Introduction

The experimental discovery nearly two decades ago [6, 28] of quantized conductance in quan-

tum point contacts (QPCs) suggested the realization of an electron waveguide. Pioneering

measurements [70, 71, 72] of noise in QPCs almost a decade later observed suppression of

shot noise below the Poissonian value due to Fermi statistics, as predicted by mesoscopic

scattering theory [97, 54]. Shot noise has since been increasingly recognized as an important

probe of quantum statistics and many-body effects [19, 29, 30], complementing dc transport.

For example, shot-noise measurements have been exploited to directly observe quasiparticle

charge in strongly correlated systems [38, 39, 40, 76], as well as to study interacting local-

ized states in mesoscopic tunnel junctions [43] and cotunneling [45] and dynamical channel

blockade [46, 33] in quantum dots.

Paralleling these developments, a large literature has emerged concerning the appear-

ance of an additional plateau-like feature in transport through a QPC at zero magnetic

field, termed 0.7 structure. Experiments [23, 98, 99, 100, 101, 102, 103] and theories [104,

105, 106, 107, 108, 109] suggest that 0.7 structure is a many-body spin effect. Its underlying

microscopic origin remains an outstanding problem in mesoscopic physics. This persistently

unresolved issue is remarkable given the simplicity of the device.

In this chapter, we review our work [31] on current noise in quantum point contacts—

including shot-noise signatures of 0.7 structure and effects of in-plane field B‖—and present

new results on a device-specific contribution to noise that is well described by a model that

includes bias-dependent heating in the vicinity of the QPC. Notably, we observe suppression

of shot noise relative to that predicted by theory for spin-degenerate transport [97, 54]

near 0.7 × 2e2/h at B‖ = 0, consistent with previous work [110, 56]. The suppression

near 0.7 × 2e2/h evolves smoothly with increasing B‖ into the signature of spin-resolved

42

Page 57: Waves, Particles, and Interactions in Reduced Dimensions

Figure 3.1: (a) Linear conductance g0 as a function of Vg2 (Vg1 = −3.2 V), for B‖ rangingfrom 0 (red) to 7.5 T (purple) in steps of 0.5 T. The series resistance Rs ranging from 430 Ωat B‖ = 0 to 730 Ω at B‖ = 7.5 T has been subtracted to align the plateaus at multiplesof 2e2/h. (b,c) Nonlinear differential conductance g as a function of Vsd, at B‖ = 0 (b) and7.5 T (c), with Vg2 intervals of 7.5 and 5 mV, respectively. Shaded regions indicate the biasrange used for the noise measurements presented in Figs. 3.3(c) and 3.4.

transmission. We find quantitative agreement between noise data and a phenomenological

model for a density-dependent level splitting [109], with model parameters extracted solely

from conductance. In the final section, we investigate a device-specific contribution to

the bias-dependent noise, particularly visible on conductance plateaus (where shot noise

vanishes), which we account for with a model [71] of Wiedemann-Franz thermal conduction

in the reservoirs connecting to the QPC.

3.2 QPC characterization

Measurements are presented for two QPCs defined by split gates on GaAs/Al0.3Ga0.7As

heterostructures grown by molecular beam epitaxy. For QPC 1(2), the two-dimensional

electron gas [2DEG] 190(110) nm below the heterostructure surface has density 1.7(2) ×

1011 cm−2 and mobility 5.6(0.2)×106 cm2/Vs. Except where noted, all data are taken at the

base temperature of a 3He cryostat, with electron temperature Te of 290 mK. A magnetic

field of 125 mT, applied perpendicular to the plane of the 2DEG, was used to reduce bias-

43

Page 58: Waves, Particles, and Interactions in Reduced Dimensions

dependent heating [71] (see section below). Each QPC is first characterized at both zero and

finite B‖ using near-dc transport measurements. The differential conductance g = dI/dVsd

(where I is the current and Vsd is the source-drain bias) is measured by lock-in technique

as discussed in Sec. 2.5.2. The B‖-dependent ohmic contact and reservoir resistance Rs in

series with the QPC is subtracted.

Figure 3.1 shows conductance data for QPC 1 (see micrograph in Fig. 3.2). Linear-

response conductance g0 = g(Vsd ∼ 0) as a function of gate voltage Vg2, for B‖ = 0 to 7.5 T

in steps of 0.5 T, is shown in Fig. 3.1(a). The QPC shows the characteristic quantization

of conductance in units of 2e2/h at B‖ = 0, and the appearance of spin-resolved plateaus

at multiples of 0.5 × 2e2/h at B‖ = 7.5 T. Additionally, at B‖ = 0, a shoulder-like 0.7

structure is evident, which evolves continuously into the 0.5 × 2e2/h spin-resolved plateau

at high B‖ [23].

Figures 3.1(b) and 3.1(c) show g as a function of Vsd for evenly spaced Vg2 settings

at B‖ = 0 and 7.5 T, respectively. In this representation, linear-response plateaus in

Fig. 3.1(a) appear as accumulated traces around Vsd ∼ 0 at multiples of 2e2/h for B‖ = 0,

and at multiples of 0.5×2e2/h for B‖ = 7.5 T. At finite Vsd, additional plateaus occur when

a sub-band edge lies between the source and drain chemical potentials [111]. The features

near 0.8 × 2e2/h (Vsd ∼ ±750 µV) at B‖ = 0 cannot be explained within a single-particle

picture [112]. These features are related to the 0.7 structure around Vsd ∼ 0 and resemble

the spin-resolved finite bias plateaus at ∼ 0.8× 2e2/h for B‖ = 7.5 T [99, 101].

3.3 Current noise

QPC current noise is measured using the cross-correlation technique (see Fig. 3.2) discussed

in Sec. 2.5.3. Johnson-noise thermometry allows in situ calibration of Te and the ampli-

44

Page 59: Waves, Particles, and Interactions in Reduced Dimensions

Digitize& FFT

R L C 60 dB

R L C 60 dB

Vg1

Vg2500 nm

VdcCrossSpectrum

290 mK 4.2 K

Figure 3.2: Equivalent circuit near 2 MHz of the system measuring QPC noise by cross-correlation on two amplification channels. The scanning electron micrograph shows a deviceof identical design to QPC 1. The QPC is formed by negative voltages Vg1 and Vg2 appliedon two facing electrostatic gates. All other gates on the device are grounded.

fication gain in the noise detection system by the procedure previously demonstrated in

Sec. 2.5.4.

We characterize the QPC noise at finite bias by the excess noise, defined as SPI (Vsd) =

SI(Vsd)− 4kBTeg(Vsd), where SI is the total QPC current noise spectral density. Note that

SPI is the noise in excess of 4kBTeg(Vsd) rather than 4kBTeg(0) and thus differs from excess

noise as discussed in Refs. [70] and [56]. In the absence of 1/f and telegraph noise as well as

bias-dependent electron heating, SPI originates from the electron partitioning at the QPC.

Experimental values for SPI are extracted from simultaneous measurements of cross-

spectral density and of g as described in Sec. 2.5.3. With an integration time of 60 s, the

resolution in SPI is 1.4 × 10−29 A2/Hz, corresponding to full shot noise 2eI of I ∼ 40 pA.

SPI as a function of dc current I for QPC 1 with gates set to very low conductance (g0 ∼

0.04× 2e2/h) [Fig. 3.3(b)] exhibits full shot noise, SPI = 2e|I|, demonstrating an absence of

1/f and telegraph noise at the noise measurement frequency [113].

Figure 3.3(c) shows SPI (Vsd) in the Vsd range −150 µV to +150 µV [shaded regions

in Figs. 3.1(b) and 3.1(c)], at B‖ = 0 and Vg2 settings corresponding to open markers in

Fig. 3.3(a). Similar to when the QPC is fully pinched off, SPI vanishes on plateaus of linear

45

Page 60: Waves, Particles, and Interactions in Reduced Dimensions

Figure 3.3: (a) Linear conductance g0 as a function of Vg2 at B‖ = 0. Solid marker and openmarkers indicate Vg2 settings for the noise measurements shown in (b) and (c), respectively.(b) SP

I as a function of dc current I with the QPC near pinch-off. The dotted line indicatesfull shot noise SP

I = 2e|I|. (c) Measured SPI as a function of Vsd, for conductances near

0 (circles), 0.5 (squares), 1 (upward triangles), 1.5 (squares), and 2 ×2e2/h (downwardtriangles). Solid lines are best-fits to Eq. (3.1) using N as the only fitting parameter. Inorder of increasing conductance, best-fit N values are 0.00, 0.20, 0.00, 0.19, and 0.03.

conductance. This demonstrates that bias-dependent electron heating is not significant in

QPC 1. In contrast, for g ∼ 0.5 and 1.5×2e2/h, SPI grows with |Vsd| and shows a transition

from quadratic to linear dependence [70, 71, 72]. The linear dependence of SPI on Vsd at

high bias further demonstrates the absence of noise due to resistance fluctuations. Solid

curves superimposed on the SPI (Vsd) data in Fig. 3.3(c) are best-fits to the form

SPI (Vsd) = 2

2e2

hN[eVsd coth

(eVsd

2kBTe

)− 2kBTe

], (3.1)

with the noise factor N as the only free fitting parameter. Note that N relates SPI to Vsd, in

contrast to the Fano factor [19, 30], which relates SPI to I. This fitting function is motivated

by mesoscopic scattering theory [97, 54, 19, 30], where transport is described by transmission

coefficients τn,σ (n is the transverse mode index and σ denotes spin) and partition noise

46

Page 61: Waves, Particles, and Interactions in Reduced Dimensions

originates from the partial transmission of incident electrons. Within scattering theory, the

full expression for SPI is

SPI (Vsd) =

2e2

h

∫ ∑n,σ

τn,σ(ε)[1− τn,σ(ε)](fs − fd)2dε, (3.2)

where fs(d) is the Fermi function in the source (drain) lead. Equation (3.1) follows from

Eq. (3.2) only for the case of constant transmission across the energy window of transport,

with N = 12

∑τn,σ(1 − τn,σ). Furthermore, for spin-degenerate transmission, N vanishes

at multiples of 2e2/h and reaches the maximal value 0.25 at odd multiples of 0.5 × 2e2/h.

Energy dependence of transmission can reduce the maximal value below 0.25, as discussed

below.

While Eq. (3.1) is motivated by scattering theory, the value of N extracted from fitting

to Eq. (3.1) simply provides a way to quantify SPI (Vsd) experimentally for each Vg2. We

have chosen the bias range e|Vsd| . 5kBTe for fitting N to minimize nonlinear-transport

effects while extending beyond the quadratic-to-linear crossover in noise that occurs on the

scale e|Vsd| ∼ 2kBTe.

The dependence of N on conductance at B‖ = 0 is shown in Fig. 3.4(a), where N is

extracted from measured SPI (Vsd) at 90 values of Vg2. The horizontal axis, gavg, is the average

of the differential conductance over the bias points where noise was measured. N has the

shape of a dome, reaching a maximum near odd multiples of 0.5 × 2e2/h and vanishing

at multiples of 2e2/h. The observed N (gavg) deviates from the spin-degenerate, energy-

independent scattering theory in two ways. First, there is a reduction in the maximum

amplitude of N below 0.25. Second, there is an asymmetry in N with respect to 0.5×2e2/h,

resulting from a noise reduction near the 0.7 feature. A similar but weaker asymmetry is

observed about 1.5× 2e2/h. The reduction in the maximum amplitude can be understood

as resulting from an energy dependence of transmissions τn,σ; the asymmetry is a signature

47

Page 62: Waves, Particles, and Interactions in Reduced Dimensions

!"

#

Figure 3.4: (a) Experimental N as a function of gavg at B‖ = 0 (red circles) along withmodel curves for nonzero (solid) and zero (dashed) proportionality of splitting, γn (see text).(b) Experimental N as a function of gavg in the range 0 − 1 × 2e2/h, at B‖ = 0 T (red),2 T (orange), 3 T (green), 4 T (cyan), 6 T (blue), and 7.5 T (purple). The dashed curveshows the single-particle model (γn = 0) at zero field for comparison.

of 0.7 structure, as we now discuss.

3.3.1 0.7 structure

We investigate further the relation between the asymmetry in N and the 0.7 structure by

measuring the dependence of N (gavg) on B‖. As shown in Fig. 3.4(b), N evolves smoothly

from a single asymmetric dome at B‖ = 0 to a symmetric double dome at 7.5 T. The latter

is a signature of spin-resolved electron transmission. Notably, for gavg between 0.7 and 1 (in

units of 2e2/h), N is insensitive to B‖, in contrast to the dependence of N near 0.3×2e2/h.

We compare these experimental data to the shot-noise prediction of a phenomenolog-

ical model [109] for the 0.7 anomaly. This model, originally motivated by dc transport

data, assumes a lifting of the twofold spin degeneracy of mode n by an energy splitting

∆εn,σ = σ · ρn · γn that grows linearly with 1D density ρn (with proportionality γn) within

that mode. Here, σ = ±1 and ρn =√

2m∗/h∑

σ(√µs − εn,σ +

√µd − εn,σ), where µs(d) is

the source(drain) chemical potential and m∗ is the electron effective mass. Parameters of

48

Page 63: Waves, Particles, and Interactions in Reduced Dimensions

the phenomenological model are extracted solely from conductance. The lever arm convert-

ing Vg2 to energy (and hence ρn) as well as the transverse mode spacing are extracted from

transconductance (dg/dVg2) data [Fig. 3.5(a)] [112]. Using an energy-dependent transmis-

sion τn,σ(ε) = 1/(1 + e2π(εn,σ−ε)/~ωx,n) for a saddle-point potential [50, 51], the value ωx,n

(potential curvature parallel to the current) is found by fitting linear conductance below

0.5 × 2e2/h (below 1.5 × 2e2/h for the second mode), and γn is obtained by fitting above

0.5(1.5)× 2e2/h, where (within the model) the splitting is largest [see Fig. 3.5(b)]. We find

~ωx,1(2) is ∼ 500(300) µeV and γ1(2) ∼ 0.012(0.008) e2/4πε0 for the first (second) mode.

Note that the splitting 2 ·ρn ·γn is two orders of magnitude smaller than the direct Coulomb

energy of electrons spaced by 1/ρn. Using these parameters, SPI (Vsd) is calculated using

Eq. (3.2), and N is then extracted by fitting SPI (Vsd) to Eq. (3.1). The calculated values of

N (gavg) at B‖ = 0 are shown along with the experimental data in Fig. 3.4(a). For compar-

ison we include calculation results accounting for energy-dependent transmission without

splitting (γn = 0). The overall reduction of N arises from a variation in transmission across

the 150 µV bias window (comparable to ~ωx), and is a single-particle effect. On the other

hand, asymmetry of N about 0.5 and 1.5× 2e2/h requires nonzero γn.

Magnetic field is included in the model by assuming a g-factor of 0.44 and adding the

Zeeman splitting to the density-dependent splitting, maintaining the parameters obtained

above. Figure 3.5(c) shows calculated N (gavg) at B‖ corresponding to the experimen-

tal data, reproduced in Fig. 3.5(d). Including the magnetic field in quadrature or as a

thermally weighted mixture with the intrinsic density-dependent splitting gives essentially

indistinguishable results within this model. Model and experiment show comparable evolu-

tion of N with B‖: the asymmetric dome for B‖ = 0 evolves smoothly into a double dome

for 7.5 T, and for conductance & 0.7× 2e2/h, the curves for all fields overlap closely. Some

49

Page 64: Waves, Particles, and Interactions in Reduced Dimensions

!"

#

Figure 3.5: (a) Transconductance dg/dVg2 as a function of Vsd and Vg2. Blue lines tracethe alignment of mode edges with source and drain chemical potentials; their slope andintersection give the conversion from Vg2 to energy and the energy spacing between modes,respectively. As two crossing points are observed between the first and second modes (themodel attributes this to spin-splitting in the first mode), we take the midpoint as thecrossing point for the blue lines. (c) Measured linear conductance (red) as a function ofVg2 at B‖ = 0, and linear conductance calculated with the model (black solid) with best-fitvalues for ωx,n and γn. Single-particle model takes γn = 0 (black dashed). (c) Model N asa function of gavg in the range 0− 1× 2e2/h, at B‖ = 0, 2, 3, 4, 6, and 7.5 T. (d) Same asFig. 3.4(b).

differences are observed between data and model, particularly for B‖ = 7.5 T. While the

experimental double dome is symmetric with respect to the minimum at 0.5 × 2e2/h, the

theory curve remains slightly asymmetric with a less-pronounced minimum. We find that

setting the g-factor to ∼ 0.6 in the model reproduces the measured symmetrical double

dome as well as the minimum value of N at 0.5×2e2/h. This observation is consistent with

reports of an enhanced g-factor in QPCs at low density [23, 101].

Recent theoretical treatments of 0.7 structure have also addressed its shot-noise signa-

ture. Modelling screening of the Coulomb interaction in the QPC, Lassl et al. [114] qualita-

tively reproduce the B‖-dependent N . Jaksch et al. [115] find a density-dependent splitting

in density-functional calculations that include exchange and correlation effects. This theory

justifies the phenomenological model and is consistent with the observed shot-noise suppres-

sion. Using a generalized single-impurity Anderson model motivated by density-functional

calculations that suggest a quasi-bound state [116], Golub et al. [117] find quantitative

50

Page 65: Waves, Particles, and Interactions in Reduced Dimensions

Figure 3.6: Experimental N as a function of gavg at B‖ = 0 (red circles) for QPC 2, alongwith model curves for nonzero (solid) and zero (dashed) proportionality of splitting γn.Model calculations include bias-dependent electron heating.

agreement with the B‖-dependent N .

3.3.2 Bias-dependent electron heating

In contrast to QPC 1, noise data in QPC 2 show evidence of bias-dependent electron

heating. Figure 3.6 shows N (gavg) at B‖ = 0 over the first three conductance steps, ex-

tracted from fits using Eq. (3.1) to SPI (Vsd) data over the range |Vsd| ≤ 400 µV at 50

gate voltage settings. As in Fig. 3.4(a), a clear asymmetry in the noise factor is ob-

served, associated with enhanced noise reduction near 0.7 × 2e2/h. For this device, N

remains finite on conductance plateaus, showing super-linear dependence on plateau index.

This is consistent with bias-dependent thermal noise resulting from electron heating. Fol-

lowing Ref. [71], we incorporate into our model the bias-dependent electron temperature

T ∗e (Vsd) =√T 2e + (24/π2)(g/gm)(1 + 2g/gm)(eVsd/2kB)2, where gm is the parallel conduc-

tance of the reservoirs connecting to the QPC. This expression [71] models diffusion by

Wiedemann-Franz thermal conduction of the heat flux gV 2sd/2 on each side of the QPC and

of Joule heating in the reservoirs, assuming ohmic contacts thermalized to the lattice at Te.

51

Page 66: Waves, Particles, and Interactions in Reduced Dimensions

In the absence of independent measurements of reservoir and ohmic contact resistances, we

treat 1/gm as a single free parameter.

Theoretical N curves including effects of bias-dependent heating are obtained from fits

to Eq. (3.1) of calculated SI(Vsd, T∗e (Vsd))−4kBTeg(Vsd). Parameters ωx,n = 1.35, 1.13, 0.86 meV

and γn = 0.019, 0.008, 0 e2/4πε0 for the first three modes (in increasing order) are extracted

from conductance data. To avoid complications arising from a zero-bias anomaly [101]

present in this device, γ0 is extracted from the splitting of the first sub-band edge in the

transconductance image [109], rather than from linear conductance. Other parameters are

extracted in the same way as for QPC 1. As shown in Fig. 3.6, quantitative agreement with

the N data is obtained over the three conductance steps with 1/gm = 75 Ω.

3.4 Conclusion and acknowledgements

We have presented measurements of current noise in quantum point contacts as a function of

source-drain bias, gate voltage, and in-plane magnetic field. We have observed a shot-noise

signature of the 0.7 structure at zero field, and investigated its evolution with increasing field

into the signature of spin-resolved transmission. Comparison to a phenomenological model

with density-dependent level splitting yielded quantitative agreement, and a device-specific

contribution to bias-dependent noise was shown to be consistent with electron heating.

We thank H.-A. Engel, M. Heiblum, L. Levitov, and A. Yacoby for valuable discussions,

and S. K. Slater, E. Onitskansky, N. J. Craig, and J. B. Miller for device fabrication. We

acknowledge support from NSF-NSEC, ARO/ARDA/DTO, and Harvard University.

52

Page 67: Waves, Particles, and Interactions in Reduced Dimensions

Chapter 4

Tunable noise cross-correlations ina double quantum dot

D. T. McClure, L. DiCarlo, Y. Zhang, H.-A. Engel, C. M. MarcusDepartment of Physics, Harvard University, Cambridge, Massachusetts 02138

M. P. Hanson, A. C. GossardDepartment of Materials, University of California, Santa Barbara, California 93106

We report measurements of the cross-correlation between temporal current fluctuations

in two capacitively coupled quantum dots in the Coulomb blockade regime. The sign of

the cross-spectral density is found to be tunable by gate voltage and source-drain bias.

We find good agreement with the data by including inter-dot Coulomb interaction in a

sequential-tunneling model.1

1This chapter is adapted with permission from Phys. Rev. Lett. 98, 056801 (2007). c©(2007) by the American Physical Society.

53

Page 68: Waves, Particles, and Interactions in Reduced Dimensions

4.1 Introduction

Current noise cross-correlation in mesoscopic electronics, the fermionic counterpart of intensity-

intensity correlation in quantum optics, is sensitive to quantum indistinguishability as well

as many-body interactions [19, 29, 78, 79, 80, 81]. A distinctive feature of fermionic sys-

tems is that in the absence of interactions, noise cross-correlation is expected to always

be negative [54, 55]. Experimentally, negative correlations have been observed in several

solid-state Hanbury-Brown and Twiss-type noise measurements [85, 88, 86]. Since no sign

constraint exists for interacting systems, a positive noise cross-correlation in a fermi system

is a characteristic signature of interactions.

Sign reversal of noise cross-correlation has been the focus of recent theory and ex-

periment [118, 119, 120, 121, 122, 123, 124, 125, 126, 127, 128, 129, 87, 130]. Theory

indicates that positive cross-correlations can arise in the presence of BCS-like interaction

[118, 119, 120], dynamical screening [121, 124], dynamical channel blockade [125, 126], and

strong inelastic scattering [127, 124, 128, 129]. Experimentally, sign reversal of noise cross-

correlation has been realized using a voltage probe to induce inelastic scattering [87], and

in a beam-splitter geometry, where the sign reversal was linked to a crossover from sub-

to super-Poissonian noise in a tunnel-barrier source [130]. This crossover was attributed

to Coulomb interaction between naturally-occurring localized states in the tunnel barrier

[113], as has been done in experiments on GaAs MESFETs [43] and stacked, self-assembled

quantum dots [44].

In this chapter, we investigate gate-controlled sign reversal of noise cross-correlation

in a simple four-terminal device. The structure consists of a parallel, capacitively coupled

double quantum dot operated in the Coulomb blockade regime. In this configuration, the

double dot acts as a pair of tunable interacting localized states, enabling a systematic study

54

Page 69: Waves, Particles, and Interactions in Reduced Dimensions

280 mK 4.2 K

Digitize &Analyze

R L C

60 dB

R L60 dB

Vtl Vtc Vtr

Vbl

Vr

Vt

Vl

(a)

It

IbVb

500 nm

gb

Vbc Vbr

Stb

St

Sb

50Ω

50Ω

50Ω

50Ω

gt

(M,

N)(M,

N+1)

(M+1,

N+1)

(M+1,

N)

Figure 4.1: (a) Scanning electron micrograph of the double-dot device, and equivalentcircuit at 2 MHz of the noise detection system measuring the power spectral densities andcross spectral density of fluctuations in currents It and Ib. (b) Differential conductances gt

(yellow) and gb (magenta) as a function of Vtc and Vbc over a few Coulomb blockade peaksin each dot, at Vt = Vb = 0. Black regions correspond to well-defined charge states in thedouble-dot system. Superimposed white lines indicate the honeycomb structure resultingfrom the finite inter-dot capacitive coupling. (c) Zero-bias (thermal) noise Sb (black dots,right axis), conductance gb (magenta curve, left axis), and calculated 4kBTegb (magentacurve, right axis) as a function of gate voltage Vbc, with Vtc = −852.2 mV.

of Coulomb-induced correlation. Turning off inter-dot tunneling by electrically depleting

the connection between dots ensures that indistinguishability (i.e., fermi statistics) alone

cannot induce any cross-correlation; any cross-correlation, positive or negative, requires

inter-dot Coulomb interaction. We find good agreement between the experimental results

and a sequential-tunneling model of capacitively coupled single-level dots.

55

Page 70: Waves, Particles, and Interactions in Reduced Dimensions

4.2 Device

The four-terminal double-dot device [see Fig. 4.1(a)] is defined by top gates on a GaAs/AlGaAs

heterostructure grown by molecular beam epitaxy. The two-dimensional electron gas 100 nm

below the surface has density 2 × 1011 cm−2 and mobility 2 × 105 cm2/Vs. Gate voltages

Vl = Vr = −1420 mV fully deplete the central point contact, preventing inter-dot tunneling.

Gate voltages Vtl (Vbl) and Vtr (Vbr) control the tunnel barrier between the top (bottom)

dot and its left and right leads. Plunger gate voltage Vtc (Vbc) controls the electron number

M (N) in the top (bottom) dot; for this experiment M ∼ N ∼ 100. The lithographic area

of each dot is 0.15 µm2. We estimate level spacing ∆t(b) ≈ 70 µeV in each dot, for ∼ 100 nm

depletion around the gates.

4.3 Methods

Measurements are performed in a 3He cryostat using a two-channel noise measurement

system [Fig. 4.1(a)] [131]. A voltage bias Vt (Vb) is applied to the left lead of the top

(bottom) dot, with right leads grounded. Separate resistor-inductor-capacitor resonators

(R = 5 kΩ, L = 66 µH, C = 96 pF) convert fluctuations in currents It and Ib through

the top and bottom dots around 2 MHz into voltage fluctuations on gates of high electron

mobility transistors (HEMTs) at 4.2 K, which in turn produce current fluctuations in two

50 Ω coaxial lines extending to room temperature, where further amplification is performed.

These signals are then simultaneously digitized at 10 MHz, their fast Fourier transforms

calculated, and the current noise power spectral densities St, Sb and cross-spectral density

Stb extracted following 15 s of integration, except for the data in Fig. 4.1(c), which was

averaged for 50 s per point. The total gain of each amplification line and the base electron

56

Page 71: Waves, Particles, and Interactions in Reduced Dimensions

temperature Te = 280 mK are calibrated in situ using Johnson-noise thermometry at base

temperature and 1.6 K with the device configured as two point contacts [131]. Differential

conductance gt (gb) through the top (bottom) dot is measured using standard lock-in tech-

niques with an excitation of 25 (30) µVrms at 677 (1000) Hz. Ohmic contact resistances of

roughly a few kΩ, much less than the dot resistances, are not subtracted.

4.4 Double-dot characterization

Superposed top- and bottom-dot conductances gt and gb as a function of plunger voltages

Vtc and Vbc form the characteristic double-dot honeycomb pattern [132, 133], with dark

regions corresponding to well-defined electron number in each dot, denoted (M,N) (first

index for top dot), as shown in Fig. 4.1(b). Horizontal (vertical) features in gt (gb) are

Coulomb blockade (CB) conductance peaks [134], across which M (N) increases by one

as Vtc (Vbc) is raised. The distance between triple points, i.e., the length of the short

edge of the hexagon, provides a measure of the mutual charging energy U due to inter-dot

capacitive coupling. By comparing this distance to the CB peak spacing, and using the

single-dot charging energy EC = 600 µeV extracted from finite bias CB diamonds (not

shown), we estimate U ≈ 60 µeV [133]. We refer to the midpoint of the short edge of a

hexagon, midway between triple points, as a “honeycomb vertex.” Current noise Sb and

conductance gb, measured simultaneously at zero dc bias, over a CB peak in the bottom

dot (with the top dot in a CB valley) are shown in Fig. 4.1(c). Agreement between the

measured Sb and the Johnson-Nyquist thermal noise value 4kBTegb is observed.

57

Page 72: Waves, Particles, and Interactions in Reduced Dimensions

Figure 4.2: Measured (a) and simulated (b) cross-spectral density Stb near a honeycombvertex, with applied bias Vt = Vb = −100 µV (e|Vt(b)| ≈ 4kBTe ≈ EC/6). Blue regions(lower-left and upper-right) indicate negative Stb, while red regions indicate positive Stb.

4.5 Sign-reversal of noise cross correlation

Turning now to finite-bias noise measurements, Fig. 4.2(a) shows the measured cross-

correlation Stb as a function of plunger gate voltages Vtc and Vbc, in the vicinity of a

honeycomb vertex, with voltage bias of −100 µV applied to both dots. The plot reveals

a characteristic quadrupole pattern of cross-correlation centered on the honeycomb vertex,

comprising regions of both negative and positive cross-correlation. Similar patterns are

observed at all other honeycomb vertices. The precise symmetry of the pattern is found to

depend rather sensitively on the relative transparency of each dot’s left and right tunnel

barriers. Away from the vertices, noise cross-correlation vanishes.

58

Page 73: Waves, Particles, and Interactions in Reduced Dimensions

4.6 Master equation simulation

To better understand this experimental result, we model the system as single-level dots

capacitively coupled by a mutual charging energy U , each with weak tunneling to the

leads. The energy needed to add electron M + 1 to the top dot depends on the two

plunger gate voltages as well as the electron number n ∈ N,N + 1 on the bottom dot:

Et = αtVtc + βtVbc + U · n + const., where lever arms αt and βt are obtained from the

honeycomb plot in Fig. 4.1(b) [132] and the measured EC . The energy Eb to add electron

N + 1 to the bottom dot is given by an analogous formula. Occupation probabilities for

charge states (M,N), (M +1, N), (M,N +1), and (M +1, N +1) are given by the diagonal

elements of the density matrix, ρ = (ρ00, ρ10, ρ01, ρ11)T . The time evolution of ρ is given by

a master equation dρ/dt =Mρ, where

M =

−W out00 W00←10 W00←01 0

W10←00 −W out10 0 W01←11

W01←00 0 −W out01 W10←11

0 W11←10 W11←01 −W out11

. (4.1)

Each diagonal term of M gives the total loss rate for the corresponding state: W outα =∑

βWβ←α. Off-diagonal terms give total rates for transitions between two states. For

example, W10←00 = W l10←00 + W r

10←00 is the total tunneling rate into (M + 1, N) from

(M,N), combining contributions from the top-left and top-right leads.

Rates for tunneling between a dot and either of its leads i ∈ tl, tr, bl, br depend on

both the transparency Γi of the tunnel barrier to lead i and the Fermi function fi(ε) =

1 + exp [(ε− µi)/kBTe]−1 evaluated at ε = Et(b), where µi is the chemical potential in

lead i. For example, the rates for tunneling into and out of the top dot from/to the left

lead are given by W l10←00 = Γltflt(Et) and W l

00←10 = Γlt [1− flt(Et)], respectively. As Et is

59

Page 74: Waves, Particles, and Interactions in Reduced Dimensions

lowered across µlt, W l10←00 increases from 0 to Γlt over a range of a few kBTe, while W l

00←10

does the opposite.

We obtain the steady-state value of ρ, denoted ρ, by solving Mρ = 0. Following

Refs. [135, 136, 137], we define current matrices J tr and Jbr for the top- and bottom-right

leads, with elements J trmn,m′n′ = |e|δnn′(m − m′)W rmn←m′n′ and Jbrmn,m′n′ = |e|δmm′(n −

n′)W rmn←m′n′ . We next obtain the average currents 〈It(b)〉 =

∑i

[J t(b)rρ

]i

and the cor-

relator 〈It(τ)Ib(0)〉 =∑

i

[θ(τ)J treMτJbrρ+ θ(−τ)JbreMτJ trρ

]i

(θ is the Heaviside step

function). The cross-spectral density in the low-frequency limit is then given by Stb =

2∫∞−∞ [〈It(τ)Ib(0)〉 − 〈It〉〈Ib〉] dτ.

Simulation results for cross-correlation Stb as a function of plunger gate voltages are

shown in Fig. 4.2(b), with all parameters of the model extracted from experiment: U =

60 µeV, Te = 280 mK, Γtl = Γtr = 1.5×1010 s−1, and Γbl = Γbr = 7.2×109 s−1. The Γi were

estimated from the zero-bias conductance peak height using Eq. (6.3) of Ref. [138], taking

left and right barriers equal. The simulation shows the characteristic quadrupole pattern of

positive and negative cross-correlation, as observed experimentally. We note that the model

underestimates Stb by roughly a factor of two. This may be due to transport processes not

accounted for in the model. For instance, elastic cotunneling should be present since the

Γi are comparable to kBTe/~. Also, since the voltage-bias energy |eVt(b)| is greater than

the level spacing ∆t(b), transport may occur via multiple levels [139, 125, 126, 46, 33] and

inelastic cotunneling [140, 141, 45].

4.7 Intuitive explanation

Intuition for how Coulomb interaction in the form of capacitive inter-dot coupling can lead

to the observed noise cross-correlation pattern can be gained by examining energy levels in

60

Page 75: Waves, Particles, and Interactions in Reduced Dimensions

U

(b)

(c) (d)

(a)

Vbc

Vtc

eVt

eVb

kT

kT

µtl

µtr

µbl

µbr

U

Figure 4.3: Energy level diagrams in the vicinity of a honeycomb vertex, with biases Vt(b) =−100 µV. (The various energies are shown roughly to scale.) The solid horizontal line inthe top (bottom) dot represents the energy Et(b) required to add electron M + 1 (N + 1)when the bottom (top) dot has N (M) electrons. The dashed horizontal line, higher thanthe solid line by U , represents Et(b) when the bottom (top) dot has N+1 (M+1) electrons.In each dot, the rate of either tunneling-in from the left or tunneling-out to the right issignificantly affected by this difference in the energy level, taking on either a slow value(red arrow) or a fast value (green arrow) depending on the electron number in the otherdot. In (a) and (d), where the occurrence of each U -sensitive process enhances the rate ofthe other, we find positive cross-correlation. In (b) and (c), where the occurrence of eachU -sensitive process suppresses the rate of the other, we find negative cross-correlation.

both dots in the space of plunger gate voltages, as shown in Fig. 4.3. With both dots tuned

near Coulomb blockade peaks, the fluctuations by one in the electron number of each dot,

caused by the sequential tunneling of electrons through that dot, cause the energy level of

the other dot to fluctuate between two values separated by U . These fluctuations can raise

and lower the level across the chemical potential in one of the leads of the dot, strongly

affecting either the tunnel-in rate (from the left, for the case illustrated in Fig. 4.3) or the

tunnel-out rate (to the right) of that dot. Specifically, the rate of the “U -sensitive” process

61

Page 76: Waves, Particles, and Interactions in Reduced Dimensions

in each dot fluctuates between a slow rate (red arrow), suppressed well below Γi, and a

fast rate (green arrow), comparable to Γi. For balanced right and left Γi in each dot, the

U -sensitive process becomes the transport bottleneck when its rate is suppressed.

These U -sensitive processes correlate transport through the dots. In region (b) of

Fig. 4.3, for instance, where Stb is negative, the U -sensitive process in each dot is tunneling-

out. Here and in (c), where the U -sensitive process in each dot is tunneling-in, the U -

sensitive processes compete: occurrence of one suppresses the other, leading to negative

Stb. Conversely, in region (a) [(d)], where Stb is positive, the top [bottom] dot’s U -sensitive

process is tunneling-out, but the bottom [top] dot’s is tunneling-in. Here, the U -sensitive

processes cooperate: occurrence of one lifts the suppression of the other, leading to positive

Stb.

4.8 Some additional checks

The arguments above also apply when one or both biases are reversed. When both are

reversed, we find both experimentally and in the model that the same cross-correlation

pattern as in Fig. 4.2 appears (not shown). When only one of the biases is reversed, we find

both experimentally [as shown in Fig. 4.4(a)] and in the model that the pattern reverses

sign. In the absence of any bias, cross-correlation vanishes both experimentally [as shown

in Fig. 4.4(b)] and in the model, despite the fact that noise in the individual dots remains

finite [as seen in Fig. 4.1(c)].

62

Page 77: Waves, Particles, and Interactions in Reduced Dimensions

Figure 4.4: (a) Measured Stb near a honeycomb vertex, with opposite biases Vt = −Vb =−100 µV. Note that the pattern is reversed from Fig. 4.2(a): negative cross-correlation(blue) is now found in the upper-left and lower-right regions, while positive cross-correlation(red) is now found in the lower-left and upper-right. (b) Measured Stb near a honeycombvertex, with Vt = Vb = 0. Cross-correlation vanishes at zero bias, though the noise in eachdot is finite.

4.9 Conclusion and acknowledgements

We have observed gate-controlled sign reversal of noise cross-correlation in a double quan-

tum dot in the Coulomb blockade regime with purely capacitive inter-dot coupling. Ex-

perimental observations are in good agreement with a sequential-tunneling model, and can

be understood from an intuitive picture of mutual charge-state-dependent tunneling. This

study, notable for the simplicity and controllability of the device, may be particularly useful

for understanding current noise in systems where interacting localized states occur naturally

and uncontrollably.

We thank N. J. Craig for device fabrication and M. Eto, W. Belzig, C. Bruder, E. Sukho-

rukov, and L. Levitov for valuable discussions. We acknowledge support from the NSF

through the Harvard NSEC, PHYS 01-17795, DMR-05-41988, DMR-0501796, as well as

support from NSA/DTO and Harvard University.

63

Page 78: Waves, Particles, and Interactions in Reduced Dimensions

Chapter 5

Noise correlations in a Coulombblockaded quantum dot

Yiming Zhang, L. DiCarlo, D. T. McClure, C. M. MarcusDepartment of Physics, Harvard University, Cambridge, Massachusetts 02138, USA

M. Yamamoto, S. TaruchaDepartment of Applied Physics, University of Tokyo, Bunkyoku, Tokyo 113-8656, Japan

ICORP-JST, Atsugi-shi, Kanagawa 243-0198, JapanM. P. Hanson, A. C. Gossard

Department of Materials, University of California, Santa Barbara, California 93106, USA

We report measurements of current noise auto- and cross correlation in a tunable quan-

tum dot with two or three leads. As the Coulomb blockade is lifted at finite source-drain

bias, the auto-correlation evolves from super-Poissonian to sub-Poissonian in the two-lead

case, and the cross correlation evolves from positive to negative in the three-lead case, con-

sistent with transport through multiple levels. Cross correlations in the three-lead dot are

found to be proportional to the noise in excess of the Poissonian value in the limit of weak

output tunneling.1

1This chapter is adapted with permission from Phys. Rev. Lett. 99, 036603 (2007). c©(2007) by the American Physical Society.

64

Page 79: Waves, Particles, and Interactions in Reduced Dimensions

5.1 Introduction

Considered individually, Coulomb repulsion and Fermi statistics both tend to smooth elec-

tron flow, thereby reducing shot noise below the uncorrelated Poissonian limit [19, 29, 54,

55]. For similar reasons, Fermi statistics without interactions also induces a negative noise

cross correlation in multiterminal devices [19, 54, 55, 85, 88]. It is therefore surprising that

under certain conditions, the interplay between Fermi statistics and Coulomb interaction

can lead to electron bunching, i.e., super-Poissonian auto-correlation and positive cross

correlation of electronic noise.

The specific conditions under which such positive noise correlations can arise has been

the subject of numerous theoretical [142, 127, 124, 128, 129, 140, 137, 139, 125, 126, 141]

and experimental [142, 43, 44, 113, 130, 87, 45, 46, 143, 34] studies in the past few years.

Super-Poissonian noise observed in metal-semiconductor field effect transistors [43], tunnel

barriers [113] and self-assembled stacked quantum dots [44] has been attributed to interact-

ing localized states [136, 137, 43] occurring naturally in these devices. In more controlled

geometries, super-Poissonian noise has been associated with inelastic cotunneling [140] in

a nanotube quantum dot [45], and with dynamical channel blockade [139, 125, 126] in

GaAs/AlGaAs quantum dots in the weak-tunneling [46] and quantum Hall regimes [143].

Positive noise cross correlation has been observed in a capacitively coupled double dot [34]

as well as in electronic beam splitters following either an inelastic voltage probe [87, 127,

124, 128, 129] or a super-Poissonian noise source [130]. The predicted positive noise cross

correlation in a three-lead quantum dot [125, 126] has not been reported experimentally to

our knowledge.

This chapter describes measurement of current noise auto- and cross correlation in a

Coulomb-blockaded quantum dot configured to have either two or three leads. As a function

65

Page 80: Waves, Particles, and Interactions in Reduced Dimensions

of gate voltage and bias, regions of super- and sub-Poissonian noise, as well as positive and

negative noise cross correlation, are identified. Results are in good agreement with a multi-

level sequential-tunneling model in which electron bunching arises from dynamical channel

blockade [139, 125, 126]. For weak-tunneling output leads, noise cross correlation in the

three-lead configuration is found to be proportional to the deviation of the auto-correlation

from the Poissonian value (either positive or negative) similar to the relation found in

electronic Hanbury Brown–Twiss (HBT)–type experiments [85, 88, 130].

5.2 Device

The quantum dot is defined by gates on the surface of a GaAs/Al0.3Ga0.7As heterostructure

[Fig. 5.1(a)]. The two-dimensional electron gas 100 nm below the surface has density 2 ×

1011 cm−2 and mobility 2 × 105 cm2/Vs. Leads formed by gate pairs Vl-Vbl, Vr-Vbr, and

Vl-Vr connect the dot to three reservoirs labeled 0, 1, and 2, respectively. Plunger gate

voltage Vbc controls the electron number in the dot, which we estimate to be ∼ 100. The

constriction formed by Vtl-Vl is closed.

5.3 Methods

A 3He cryostat is configured to allow simultaneous conductance measurement near dc and

noise measurement near 2 MHz [131]. For dc measurements, the three reservoirs are each

connected to a voltage amplifier, a current source, and a resistor to ground (r = 5 kΩ).

The resistor r converts the current Iα out of reservoir α to a voltage signal measured by

the voltage amplifier; it also converts the current from the current source to a voltage

excitation Vα applied at reservoir α. The nine raw differential conductance matrix elements

66

Page 81: Waves, Particles, and Interactions in Reduced Dimensions

(a)

Digitize&

AnalyzeS12

S2

S1

0.3 K 4.2 K

60 dB

60 dB

Vtl

Vbl

VrVl

I2

I1

Vbc Vbr

V0I0

500 nm

R

R

Figure 5.1: (a) Micrograph of the device and equivalent circuit near 2 MHz of the noisedetection system (see text for equivalent circuit near dc). For the data in Figs. 5.1 and 5.2,the Vl-Vr constriction is closed and the dot is connected only to reservoirs 0 and 1. (b, c)Differential conductance g01 and current noise spectral density S1, respectively, as a functionof V0 and Vbc. (d) S1 versus |I1| data (circles) and multi-level simulation (solid curves) alongthe four cuts indicated in (b) and (c) with corresponding colors. Black solid (dashed) lineindicates S1 = 2e|I1| (S1 = 1e|I1|). (e) Data (diamonds) and multi-level simulation (solidcurves) of the modified Fano factor F along the same cuts as taken in (d). Inset: detail ofF at high |V0|.

gαβ = dIβ/dVα are measured simultaneously with lock-in excitations of 20 µVrms at 44, 20

and 36 Hz on reservoirs 0, 1 and 2, respectively. Subtracting r from the matrix g yields the

67

Page 82: Waves, Particles, and Interactions in Reduced Dimensions

intrinsic conductance matrix2 g = [E + rg]−1 · g, where E is the identity matrix. Ohmic

contact resistances (∼ 103 Ω) are small compared to dot resistances (& 105 Ω), and are

neglected in the analysis. Values for the currents Iα with bias V0 applied to reservoir 0 are

obtained by numerically integrating g0α.

Fluctuations in currents I1 and I2 are extracted from voltage fluctuations around 2 MHz

across separate resistor-inductor-capacitor (RLC) resonators [Fig. 5.1(a)]. Power spectral

densities SV1,2 and cross-spectral density SV12 of these voltage fluctuations [131] are av-

eraged over 20 s, except where noted. Following the calibration of amplifier gains and

electron temperature Te using noise thermometry [131], the dot’s intrinsic current noise

power spectral densities S1,2 and cross-spectral density S12 are extracted by solving the

Langevin [19] equations that take into account the feedback [128] and thermal noise from

the finite-impedance external circuit3:

S1 = a211SV 1 + a2

21SV 2 + 2a11a21SV 12 − 4kBTe/R

S2 = a212SV 1 + a2

22SV 2 + 2a12a22SV 12 − 4kBTe/R

S12 = a11a12SV 1 + a21a22SV 2 + (a11a22 + a12a21)SV 12,

where a11(22) = 1/R − g11(22), a12(21) = −g12(21) and R is the RLC resonator parallel

resistance.

5.4 Noise in the two-lead configuration

Figure 5.1(b) shows conductance g01 as a function of Vbc and V0 in a two-lead configuration,

i.e., with the Vl-Vr constriction closed. The characteristic Coulomb blockade (CB) diamond

2See derivation in Sec. C.1 of Ap. C.

3See derivation in Sec. C.2 of Ap. C.

68

Page 83: Waves, Particles, and Interactions in Reduced Dimensions

structure yields a charging energy EC = 0.8 meV and lever arm for the plunger gate

ηbc = ∆εd/(e∆Vbc) = 0.069, where εd is the dot energy. The diamond tilt ηbc/(1/2 − η0)

gives the lever arm for reservoir 0: η0 = ∆εd/(e∆V0) = 0.3. As shown in Fig. 5.1(d), current

noise S1 along selected cuts close to the zero-bias CB peak (red, orange cuts) is below the

Poissonian value 2e|I1| at all biases |I1|, while cuts that pass inside the CB diamond (green,

blue cuts) exceed 2e|I1| at low currents, then drop below 2e|I1| at high currents. At finite

Te, the current noise SP1 = 2eI1 coth(eV0/2kBTe) of an ideal Poissonian noise source at bias

V0 may exceed 2e|I1| due to the thermal (Johnson) noise contribution [140]. Accordingly, we

define a modified Fano factor F ≡ S1/SP1 . Figure 5.1(e) shows regions of super-Poissonian

noise (F > 1) when the green and blue cuts are within the CB diamond. For all cuts, F

approaches 1/2 at large bias.

Current noise can also be identified as sub- or super-Poissonian from the excess Pois-

sonian noise SEP1 ≡ S1 − SP

1 being negative or positive, respectively. Unlike F , SEP1 does

not have divergent error bars inside the CB diamond, where currents vanish. As shown in

Fig. 5.2(a), in regions where both I1 and S1 vanish, SEP1 also vanishes. Far outside the CB

diamonds, SEP1 is negative, indicating sub-Poissonian noise. However, SEP

1 becomes positive

along the diamond edges, indicating super-Poissonian noise in these regions.

We next compare our experimental results to single-level and multi-level sequential-

tunneling models of CB transport. The single-level model yields exact expressions for

average current and noise [19, 29, 141, 144]: I1 = (e/h)∫dεγ0γ1(f1 − f0)/[(γ1 + γ0)2/4 +

(ε − εd)2], S1 = (2e2/h)∫dεγ2

0γ21 · [f0(1 − f0) + f1(1 − f1)] + γ0γ1[(γ1 − γ0)2/4 + (ε −

εd)2] · [f0(1 − f1) + f1(1 − f0)]/[(γ1 + γ0)2/4 + (ε − εd)2]2, where γ0(1) is the tunneling

rate to reservoir 0(1) and f0(1) is the Fermi function in reservoir 0(1). The dot energy εd

is controlled by gate and bias voltages: εd = −eVbcηbc − eV0η0 − eV1η1 + const. For the

69

Page 84: Waves, Particles, and Interactions in Reduced Dimensions

multi-level sequential-tunneling model, a master equation is used to calculate current and

noise4, following Refs. [139, 125, 126, 135]. To model transport, we assume simple filling

of orbital levels and consider transitions to and from N -electron states that differ in the

occupation of at most n levels above (indexed 1 through n) and m levels below (indexed

−1 through −m) the highest occupied level in the (N + 1)-electron ground state (level 0).

For computational reasons, we limit the calculation to n = m = 3. For simplicity, we

assume equal level spacings, symmetric tunnel barriers, and an exponential dependence of

the tunneling rates on level energy: ∆εl ≡ εld−ε0d = l×δ and γl0 = γl1 = Γ exp(κ∆εl), where

l = −3, ..., 0, ..., 3 is the level index, εld is the energy of level l, and γl0(1) is the tunneling rate

from level l to reservoir 0(1). We choose δ = 150 µeV, Γ = 15 GHz and κ = 0.001 (µeV)−1

to fit the data in Figs. 5.1(d) and 5.1(e).

Super-Poissonian noise in the multi-level model arises from dynamical channel block-

ade [139, 125, 126], illustrated in the diagrams in Fig. 5.2. Consider, for example, the energy

levels and transport processes shown in the green-framed diagram, which corresponds to the

location of the green dot on the lower-right edge in Fig. 5.2(c). Along that edge, the trans-

port involves transitions between the N -electron ground state and (N + 1)-electron ground

or excited states. When an electron occupies level 0, it will have a relatively long lifetime,

as tunneling out is suppressed by the finite electron occupation in reservoir 1 at that en-

ergy. During this time, transport is blocked since the large charging energy prevents more

than one non-negative-indexed level from being occupied at a time. This blockade happens

dynamically during transport, leading to electron bunching and thus to super-Poissonian

noise. At the location of the pink dot on the lower-left edge in Fig. 5.2(c), the transport

involves transitions between the (N + 1)-electron ground state and N -electron ground or

4The simulation source routines are provided in Ap. E.

70

Page 85: Waves, Particles, and Interactions in Reduced Dimensions

-1.0

-0.5

0.0

0.5

1.0

V0

[mV

]

-1.74-1.77 Vbc [V]

-8 -4 0 4 8SEP [10-28A2/ Hz]

SUB-POISSONIAN | SUPER-POISSONIAN

EXP. (a)

1

-1.0

-0.5

0.0

0.5

1.0

V0

[mV

]

-1.745-1.76 Vbc [V]

SIM. (S.L.) (b)

-1.745-1.76 Vbc [V]

SIM. (M.L.) (c)

+3+2+10-1-2-3

Figure 5.2: (a) Excess Poissonian noise SEP1 as a function of V0 and Vbc. Red (blue)

regions indicate super(sub)-Poissonian noise. (b, c) Single-level (S.L.) and multi-level (M.L.)simulation of SEP

1 , respectively, corresponding to the data region enclosed by the whitedashed parallelogram in (a). At the four colored dots superimposed on (c), where SEP

1

is most positive, energy diagrams are illustrated in the correspondingly colored frames atthe bottom. In these diagrams, black (white) arrows indicate electron (hole) transport;the greyscale color in the reservoirs and inside the circles on each level indicates electronpopulation, the darker the higher.

excited states; a similar dynamical blockade occurs in a complementary hole transport pic-

ture. The hole transport through level 0 is slowed down by the finite hole occupation in

reservoir 0, modulating the hole transport through negative-indexed levels, thus leading to

hole bunching and super-Poissonian noise. Transport at the blue (orange) dot is similar

71

Page 86: Waves, Particles, and Interactions in Reduced Dimensions

to transport at the green (pink) dot, but with the chemical potentials in reservoirs 0 and

1 swapped. Both experimentally and in the multi-level simulation, SEP1 is stronger along

electron edges than along hole edges. This is due to the energy dependence of the tunnel-

ing rates: since the positive-indexed electron levels have higher tunneling rates than the

negative-indexed hole levels, the dynamical modulation is stronger for electron transport

than for hole transport.

5.5 Noise in the three-lead configuration

We next investigate the three-lead configuration, obtained by opening lead 2 [Fig. 5.3(a)].

At zero bias, thermal noise cross correlation is found to be in good agreement with the

theoretical value5, S12 = −4kBTeg12, as seen in Fig. 5.3(b).

To minimize this thermal contribution to S12, output leads are subsequently tuned to

weaker tunneling than the input lead (g01 ∼ g02 ∼ 4g12), for reasons discussed below. Note

that as a function of Vbc and V0, S12 [Fig. 5.3(c)] looks similar to SEP1 [Fig. 5.2(a)] in the

two-lead configuration. The slightly positive S12 (∼ 0.2×10−28A2/Hz) inside the rightmost

diamond is due to a small drift in the residual background of SV 12 over the 13 h of data

acquisition for Fig. 3(c). Without drift, as in the shorter measurement of Fig. 5.3(b), S12

approaches 0 at zero bias as g12 vanishes.

Both the single-level and multi-level models can be extended to include the third

lead [144, 125, 126]. Figures 5.3(d) and 5.3(e) show the single-level and multi-level simula-

tions of S12, respectively. Similar to the two-lead case, only the multi-level model reproduces

the positive cross correlation along the diamond edges.

5At zero bias, the fluctuation-dissipation theorem requires S12 = −2kBTe(g12 + g21), butg12 = g21 at zero bias and zero magnetic field.

72

Page 87: Waves, Particles, and Interactions in Reduced Dimensions

R

R

Vtl

Vbl

VrVl

I2

I1

Vbc Vbr

V0I0

500 nm

(a)

-0.3

-0.2

-0.1

0.0

[10-2

8 A2 / H

z]

-1.512 -1.509Vbc [V]

-0.04

-0.02

0.00

[e2/ h x 4k

B Te ]

S12 -4kBTe g12

(b)

-1.49-1.50 Vbc [V]

SIM. (M.L.) (e)

-1.0

-0.5

0.0

0.5

1.0

V0

[mV

]

-1.49-1.50 Vbc [V]

SIM. (S.L.) (d)

-1.0

-0.5

0.0

0.5

1.0

V0

[mV

]

-1.48-1.49-1.50-1.51 Vbc [V]

-2 -1 0 1 2S12 [10-28A2/ Hz]

EXP. (c)

Figure 5.3: (a) The device in the three-lead configuration, in which the data for this figureand for Fig. 5.4 are taken. (b) S12, integrated for 200 s, and −4kBTeg12 over a CB peak atzero bias. Left and right axes are in different units but both apply to the data. (c) S12 as afunction of V0 and Vbc. Red (blue) regions indicate positive (negative) cross correlation. (d,e) Single-level (S.L.) and multi-level (M.L.) simulation of S12, respectively, correspondingto the data region enclosed by the white dashed parallelogram in (c).

To further investigate the relationship between noise auto- and cross correlation, we com-

pare S12 to the total excess Poissonian noise, SEP ≡ S1+S2+2S12−2e(I1+I2) coth(eV0/2kBTe),

measured in the same three-lead configuration. Figure 5.4 shows SEP and S12, measured

at fixed bias V0 = +0.5 mV. The observed proportionality S12 ∼ SEP/4 is reminiscent

73

Page 88: Waves, Particles, and Interactions in Reduced Dimensions

-1

0

1

[10-2

8 A2 / H

z]

-1.48-1.49-1.50-1.51Vbc [V]

S12 SEP/ 4

(a)

-4 0 4SEP [10-28A2/ Hz]

S12 vs. SEP

Slope = 1/4

(b)

Figure 5.4: (a) S12 (green) and SEP/4 (blue) as a function of Vbc at V0 = +0.5 mV [greenhorizontal line in Fig. 5.3(c)]. (b) Parametric plot of S12 (green circles) versus SEP for thesame data as in (a). The solid black line has a slope of 1/4, the value expected for a 50/50beam splitter.

of electronic HBT-type experiments [85, 88, 130], where noise cross correlation following a

beam splitter was found to be proportional to the total output current noise in excess of the

Poissonian value, with a ratio of 1/4 for a 50/50 beam splitter. In simulation, we find that

this HBT-like relationship holds in the limit g01 ∼ g02 g12 (recall that g01 ∼ g02 ∼ 4g12

in the experiment); on the other hand, when g01 ∼ g02 ∼ g12, thermal noise gives a negative

contribution that lowers S12 below SEP/4, as we have also observed experimentally (not

shown). The implications are that first, with weak-tunneling output leads, the three-lead

dot behaves as a two-lead dot followed by an ideal beam splitter, and second, the dynamical

channel blockade that leads to super-Poissonian noise in the two-lead dot also gives rise to

positive cross correlation in the three-lead dot.

5.6 Acknowledgements

We thank N. J. Craig for device fabrication and H.-A. Engel for valuable discussions. We

acknowledge support from the NSF through the Harvard NSEC, PHYS 01-17795, DMR-

05-41988, DMR-0501796. M. Yamamoto and S. Tarucha acknowledge support from the

74

Page 89: Waves, Particles, and Interactions in Reduced Dimensions

DARPA QuIST program, the Grant-in-Aid for Scientific Research A (No. 40302799), the

MEXT IT Program and the Murata Science Foundation.

75

Page 90: Waves, Particles, and Interactions in Reduced Dimensions

Chapter 6

Shot noise in graphene

L. DiCarlo†, J. R. Williams‡, Yiming Zhang†, D. T. McClure†, C. M. Marcus††Department of Physics, Harvard University, Cambridge, Massachusetts 02138, USA‡School of Engineering and Applied Sciences, Harvard University, Cambridge,

Massachusetts 02138, USA

We report measurements of current noise in single- and multi-layer graphene devices.

In four single-layer devices, including a p-n junction, the Fano factor remains constant to

within ±10% upon varying carrier type and density, and averages between 0.35 and 0.38.

The Fano factor in a multi-layer device is found to decrease from a maximal value of 0.33

at the charge-neutrality point to 0.25 at high carrier density. These results are compared

to theories for shot noise in ballistic and disordered graphene.1

1This chapter is adapted with permission from Phys. Rev. Lett. 100, 156801 (2008). c©(2008) by the American Physical Society.

76

Page 91: Waves, Particles, and Interactions in Reduced Dimensions

6.1 Introduction

Shot noise, the temporal fluctuation of electric current out of equilibrium, originates from

the partial transmission of quantized charge [19]. Mechanisms that can lead to shot noise

in mesoscopic conductors include tunneling, quantum interference, and scattering from im-

purities and lattice defects. Shot noise yields information about transmission that is not

available from the dc current alone.

In graphene [145, 146], a zero-gap two-dimensional semi-metal in which carrier type and

density can be controlled by gate voltages [147], density-dependent shot-noise signatures

under various conditions have been investigated theoretically [148, 149]. For wide samples

of ballistic graphene (width-to-length ratio W/L & 4) the Fano factor, F , i. e., the current

noise normalized to the noise of Poissonian transmission statistics, is predicted to be 1/3 at

the charge-neutrality point and ∼ 0.12 in both electron (n) and hole (p) regimes [148]. The

value F = 1−1/√

2 ≈ 0.29 is predicted for shot noise across a ballistic p-n junction [149]. For

strong, smooth “charge-puddle” disorder, theory predicts F ≈ 0.30 both at and away from

the charge-neutrality point, for all W/L & 1 [150]. Disorder may thus have a similar effect on

noise in graphene as in diffusive metals, where F is universally 1/3 [151, 152, 153, 73, 74, 82]

regardless of shape and carrier density. Recent theory investigates numerically the evolution

from a density-dependent to a density-independent F with increasing disorder [154]. To

our knowledge, experimental data for shot noise in graphene has not yet been reported.

This chapter presents an experimental study of shot noise in graphene at low tempera-

tures and zero magnetic field. Data for five devices, including a locally gated p-n junction,

are presented. For three globally-gated, single-layer samples, we find F ∼ 0.35 − 0.37 in

both electron and hole doping regions, with essentially no dependence on electronic sheet

density, ns, in the range |ns| . 1012 cm−2. Similar values are obtained for a locally-gated

77

Page 92: Waves, Particles, and Interactions in Reduced Dimensions

Digitize& FFT

60 dB

60 dB CrossSpectrum

0.3 K

2 µmI0

4.2 K (c)

Vbg

50Ω

50Ω 50Ω

50Ω

r

Figure 6.1: (a) Differential resistance R of sample A1 as a function of back-gate voltage Vbg

at electron temperature Te = 0.3 K, perpendicular field B⊥ = 0, and source-drain voltageVsd = 0. (b) Differential two-terminal conductance g(Vsd = 0) as a function of B⊥ and Vbg

in the quantum Hall regime, after subtracting a quadratic fit at each B⊥. Lines of constantfilling factors 6, 10, 14, and 18 (dashed lines) indicate a single-layer sample. (c) Equivalentcircuit near 1.5 MHz of the system measuring current noise using cross correlation of twochannels [131]. Current bias Io contains a 7.5 nArms, 20 Hz part for lock-in measurementsand a controllable dc part generating the dc component of Vsd via the shunt resistancer = 5 kΩ. False-color scanning electron micrograph of a three-lead pattern defining twodevices similar to A1 and A2. Purple indicates single-layer graphene and gold indicatesmetallic contacts.

single-layer p-n junction in both unipolar (n-n or p-p) and bipolar (p-n or n-p) regimes.

In a multi-layer sample, the observed F evolves from 0.33 at the charge-neutrality point to

0.25 at ns ∼ 6× 1012 cm−2.

6.2 Methods

Devices were fabricated by mechanical exfoliation of highly-oriented pyrolytic graphite [147].

Exfoliated sheets were deposited on a degenerately-doped Si substrate capped with 300 nm

of thermally grown SiO2. Regions identified by optical microscopy as potential single-layer

78

Page 93: Waves, Particles, and Interactions in Reduced Dimensions

graphene were contacted with thermally evaporated Ti/Au leads (5/40 nm) patterned by

electron-beam lithography. Additional steps in the fabrication of the p-n junction device are

detailed in Ref. [155]. Devices were measured in two 3He cryostats, one allowing dc (lock-

in) transport measurements in fields |B⊥| ≤ 8 T perpendicular to the graphene plane, and

another allowing simultaneous measurements of dc transport and noise [131] near 1.5 MHz,

but limited to B⊥ ∼ 0.

6.3 Shot noise in single-layer devices

Differential resistance R = dVsd/dI (I is the current, and Vsd is the source-drain voltage) of a

wide, short sample [A1, (W,L) = (2.0, 0.35) µm] is shown as a function of back-gate voltage

Vbg at Vsd = 0 and B⊥ = 0 in Fig. 6.1(a). While the width of the peak is consistent with

A1 being single-layer graphene [156, 157], more direct evidence is obtained from the QH

signature shown in Fig. 6.1(b). The grayscale image shows differential conductance g = 1/R

as a function of Vbg and B⊥, following subtraction of the best-fit quadratic polynomial

to g(Vbg) at each B⊥ setting to maximize contrast. Dashed lines correspond to filling

factors nsh/eB⊥ = 6, 10, 14, and 18, with ns = α(Vbg + 1.1 V) and lever arm α =

6.7×1010 cm−2/V. Their alignment with local minima in δg(Vbg) identifies A1 as single-layer

graphene [158, 159]. The Drude mean free path ` = h/2e2 ·σ/kF [160], where kF =√π|ns|,

is found to be ∼ 40 nm away from the charge-neutrality point using the B⊥ = 0 conductivity

σ = (RW/L)−1 [Fig. 6.2(a) inset].

Current noise spectral density SI is measured using a cross-correlation technique de-

scribed in Ref. [131] [see Fig. 6.1(c)]. Following calibration of amplifier gains and electron

temperature Te using Johnson noise thermometry (JNT) for each cooldown, the excess

noise SeI ≡ SI − 4kBTeg(Vsd) is extracted. Se

I(Vsd) for sample A1 is shown in Fig. 6.2(a).

79

Page 94: Waves, Particles, and Interactions in Reduced Dimensions

Figure 6.2: (a) Inset: Conductivity σ = (RW/L)−1 calculated using R(Vbg) data inFig. 6.1(a) and W/L = 5.7. Solid black circles correspond to σ(Vsd = 0) at the Vbg settingsof noise measurements shown in (b). Main: Excess noise Se

I as function of Vsd near thecharge-neutrality point, Vbg = −0.75 V. The solid red curve is the single-parameter bestfit to Eq. (6.1), giving Fano factor F = 0.349 (using Te = 303 mK as calibrated by JNT).(b) Best-fit F at 25 Vbg settings across the charge-neutrality point for electron and holedensities reaching |ns| ∼ 1.4× 1012 cm−2. (c) R (left axis) and σ (right axis) of sample A2as a function of Vbg (W/L = 1.4), with Vsd = 0, at 0.3 K (solid markers) and at 1.1 K (openmarkers). (d), (e) Crossover width Tw (normalized to JNT-calibrated Te) and F , obtainedfrom best-fits using Eq. (6.1) to Se

I(Vsd) data over |Vsd| ≤ 350(650) µV for Te = 0.3(1.1) K.

80

Page 95: Waves, Particles, and Interactions in Reduced Dimensions

Linearity of SeI at high bias indicates negligible extrinsic (1/f or telegraph) resistance fluc-

tuations within the measurement bandwidth. For these data, a single-parameter fit to the

scattering-theory form (for energy-independent transmission) [97, 55],

SeI = 2eIF

[coth

(eVsd

2kBTe

)− 2kBTe

eVsd

], (6.1)

gives a best-fit Fano factor F = 0.349. Simultaneously measured conductance g ≈ 22.2 e2/h

was independent of bias within ±0.5% (not shown) in the |Vsd| ≤ 350 µV range used for the

fit. Note that the observed quadratic-to-linear crossover agrees well with that in the curve

fit, indicating weak inelastic scattering in A1 [73, 74], and negligible series resistance (e. g.,

from contacts), which would broaden the crossover by reducing the effective Vsd across the

sample.

Figure 6.2(b) shows similarly measured values for F as a function of Vbg. F is observed

to remain nearly constant for |ns| . 1012 cm−2. Over this density range, the average F is

0.35 with standard deviation 0.01. The estimated error in the best-fit F at each Vbg setting

is ±0.002, comparable to the marker size and smaller than the variation in F near Vbg = 0,

which we believe results from mesoscopic fluctuations of F . Nearly identical noise results

(not shown) were found for a similar sample (B), with dimensions (2.0, 0.3) µm and a QH

signature consistent with a single layer.

Transport and noise data for a more square single-layer sample [A2, patterned on the

same graphene sheet as A1, with dimensions (1.8, 1.3) µm] at Te = 0.3 K (solid circles) and

Te = 1.1 K (open circles) are shown in Figs. 6.2(c-e). At both temperatures, the conductivity

shows σmin ≈ 1.5 e2/h and gives ` ∼ 25 nm away from the charge-neutrality point. That

these two values differ from those in sample A1 is particularly notable as samples A1 and

A2 were patterned on the same piece of graphene. Results of fitting Eq. (6.1) to SeI(Vsd)

for sample A2 are shown in Figs. 6.2(d) and 6.2(e). To allow for possible broadening of

81

Page 96: Waves, Particles, and Interactions in Reduced Dimensions

the quadratic-to-linear crossover by series resistance and/or inelastic scattering, we treat

electron temperature as a second fit parameter (along with F) and compare the best-fit

value, Tw, with the Te obtained from Johnson noise. Figure 6.2(d) shows Tw tracking the

calibrated Te at both temperatures. Small deviations of Tw/Te from unity near the charge-

neutrality point at Te = 0.3 K can be attributed to conductance variations up to ±20%

in the fit range |Vsd| ≤ 350 µV at these values of Vbg. As in sample A1, F is found to be

independent of carrier type and density over |ns| . 1012 cm−2, averaging 0.37(0.36) with

standard deviation 0.02(0.02) at Te = 0.3(1.1) K. Evidently, despite its different aspect

ratio, A2 exhibits a noise signature similar to that of A1.

6.4 Shot noise in a p-n junction

Transport and noise measurements for a single-layer graphene p-n junction [155], sample

C, are shown in Fig. 6.3. The color image in Fig. 6.3(a) shows differential resistance R as

a function of Vbg and local top-gate voltage Vtg. The two gates allow independent control

of charge densities in adjacent regions of the device [see Fig. 6.3(c) inset]. In the bipolar

regime, the best-fit F shows little density dependence and averages 0.38, equal to the average

value deep in the unipolar regime, and similar to results for the back-gate-only single-layer

samples (A1, A2 and B). Close to charge neutrality in either region (though particularly

in the region under the top gate), SeI(Vsd) deviates from the form of Eq. (6.1) (data not

shown). This is presumably due to resistance fluctuation near charge neutrality, probably

due mostly to mobile traps in the Al2O3 insulator beneath the top gate.

82

Page 97: Waves, Particles, and Interactions in Reduced Dimensions

12

Figure 6.3: (a) Differential resistance R of sample C, a single-layer p-n junction, as afunction of back-gate voltage Vbg and top-gate voltage Vtg. The skewed-cross pattern definesquadrants of n and p carriers in regions 1 and 2. Red lines indicate charge-neutrality linesin region 1 (dotted) and region 2 (dashed). (b) Se

I(Vsd) measured in n-p regime with(Vbg, Vtg) = (5,−4) V (solid dots) and best fit to Eq. (6.1) (red curve), with F = 0.36. (c)Main: Best-fit F along the cuts shown in (a), at which ns1 ∼ ns2 (purple) and ns1 ∼ −4 ns2(black). Inset: Schematic of the device. The top gate covers region 2 and one of thecontacts.

6.5 Shot noise in a multi-layer device

Measurements at 0.3 K and at 1.1 K for sample D, of dimensions (1.8, 1.0) µm, are shown

in Fig. 6.4. A ∼ 3 nm step height between SiO2 and carbon surfaces measured by atomic

force microscopy prior to electron-beam lithography [161] suggests this device is likely multi-

layer. Further indications include the broad R(Vbg) peak [162] and the large minimum

conductivity, σmin ∼ 8 e2/h at B⊥ = 0 [Fig. 6.4(a)], as well as the absence of QH signature

for |B⊥| ≤ 8 T at 250 mK (not shown). Two-parameter fits of SeI(Vsd) data to Eq. (6.1)

show three notable differences from results in the single-layer samples [Figs. 6.4(b) and

83

Page 98: Waves, Particles, and Interactions in Reduced Dimensions

Figure 6.4: (a) Differential resistance R (left axis) and conductivity σ (right axis) of sampleD as a function of Vbg, with Vsd = 0, at 0.3 K (solid markers) and at 1.1 K (open markers).(b),(c) Best-fit Tw (normalized to JNT-calibrated Te) and F to Se

I(Vsd) data over |Vsd| ≤0.5(1) mV for Te = 0.3(1.1) K. Inset: Sublinear dependence of Se

I on Vsd is evident in datataken over a larger bias range. Solid red curve is the two-parameter best fit of Eq. (6.1)over |Vsd| ≤ 0.5 mV.

6.4(c)]: First, F shows a measurable dependence on back-gate voltage, decreasing from

0.33 at the charge-neutrality point to 0.25 at ns ∼ 6 × 1012 cm−2 for Te = 0.3 K; Second,

F decreases with increasing temperature; Finally, Tw/Te is 1.3-1.6 instead of very close to

1. We interpret the last two differences, as well as the sublinear dependence of SeI on Vsd

(see Fig. 6.4 inset) as indicating sizable inelastic scattering [151, 152] in sample D. (An

alternative explanation in terms of series resistance would require it to be density, bias, and

84

Page 99: Waves, Particles, and Interactions in Reduced Dimensions

temperature dependent, which is inconsistent with the independence of g on Vsd and Te).

6.6 Summary and acknowledgements

Summarizing the experimental results, we find that in four single-layer samples, F is in-

sensitive to carrier type and density, temperature, aspect ratio, and the presence of a p-n

junction. In one multi-layer sample, F does depend on density and temperature, and SeI(Vsd)

shows a broadened quadratic-to-linear crossover and is sublinear in Vsd at high bias. We

may now compare these results to expectations based on theoretical and numerical results

for ballistic and disordered graphene.

Theory for ballistic single-layer graphene with W/L & 4 gives a universal F = 1/3 at

the charge-neutrality point, where transmission is evanescent, and F ∼ 0.12 for |ns| & π/L2,

where propagating modes dominate transmission [148]. While the measured F at the charge-

neutrality point in samples A1 and B (W/L = 5.7 and 6.7, respectively) is consistent with

this prediction, the absence of density dependence is not: π/L2 ∼ 3×109 cm−2 is well within

the range of carrier densities covered in the measurements. Theory for ballistic graphene

p-n junctions [149] predicts F ≈ 0.29, lower than the value ∼ 0.38 observed in sample C

in both p-n and n-p regimes. We speculate that these discrepancies likely arise from the

presence of disorder. Numerical results for strong, smooth disorder [150] predict a constant

F at and away from the charge-neutrality point for W/L & 1, consistent with experiment.

However, the predicted value F ≈ 0.30 is ∼ 20% lower than observed in all single-layer

devices. Recent numerical simulations [154] of small samples (L = W ∼ 10 nm) investigate

the vanishing of carrier dependence in F with increasing disorder strength. In the regime

where disorder makes F density-independent, the value F ∼ 0.35− 0.40 is found to depend

weakly on disorder strength and sample size.

85

Page 100: Waves, Particles, and Interactions in Reduced Dimensions

Since theory for an arbitrary number of layers is not available for comparison to noise

results in the multi-layer sample D, we compare only to existing theory for ballistic bi-layer

graphene [163]. It predicts F = 1/3 over a much narrower density range than for the single

layer, and abrupt features in F at finite density due to transmission resonances. A noise

theory beyond the bi-layer ballistic regime may thus be necessary to explain the observed

smooth decrease of F with increasing density in sample D.

We thank C. H. Lewenkopf, L. S. Levitov, and D. A. Abanin for useful discussions.

Research supported in part by the IBM Ph.D. Fellowship program (L.D.C.), INDEX, an

NRI Center, and Harvard NSEC.

86

Page 101: Waves, Particles, and Interactions in Reduced Dimensions

Chapter 7

Distinct Signatures For CoulombBlockade and Aharonov-BohmInterference in ElectronicFabry-Perot Interferometers

Yiming Zhang, D. T. McClure, E. M. Levenson-Falk, C. M. MarcusDepartment of Physics, Harvard University, Cambridge, Massachusetts 02138, USA

L. N. Pfeiffer, K. W. WestBell Laboratories, Alcatel-Lucent Technologies, Murray Hill, New Jersey 07974, USA

Two distinct types of magnetoresistance oscillations are observed in two electronic

Fabry-Perot interferometers of different sizes in the integer quantum Hall regime. Mea-

suring these oscillations as a function of magnetic field and gate voltages, we describe three

signatures that distinguish the two types. The oscillations observed in a 2.0 µm2 device

are understood to arise from a Coulomb blockade mechanism, and those observed in an

18 µm2 device from an Aharonov-Bohm mechanism. This work clarifies, provides ways to

distinguish, and demonstrates control over these distinct mechanisms of oscillations seen in

electronic Fabry-Perot interferometers. 1

1This chapter is adapted from Ref. [22], which has been accepted for publication inPhys. Rev. B as a Rapid Communication.

87

Page 102: Waves, Particles, and Interactions in Reduced Dimensions

7.1 Introduction

Mesoscopic electronics can exhibit wave-like interference effects [4, 10, 11, 164], particle-like

charging effects [3], or a complex mix of both [20]. Experiments over the past two decades

have investigated the competition between wave and particle properties [21], as well as

regimes where they coexist [14, 15, 165, 20]. The electronic Fabry-Perot interferometer

(FPI)— a planar two-contact quantum dot operating in the quantum Hall regime—is a

system where both interference and Coulomb interactions can play important roles. This

device has attracted particular interest recently due to predicted signatures of fractional [62]

and non-Abelian [64, 63, 166] statistics. The interpretation of experiments, however, is

subtle, and must account for the interplay of charging and interference effects in these

coherent confined structures.

The pioneering experimental investigation of resistance oscillations in an electronic FPI

[167] interpreted the oscillations in terms of an Aharonov-Bohm (AB) interference of edge

states, attributing the magnetic field dependence of the field-oscillation period to a chang-

ing effective dot area. More recent experiments [168, 169, 170, 171, 69] have observed

frequencies of integer multiples of the fundamental AB frequency; in particular, a pro-

portionality of field frequency to the number of fully-occupied Landau levels (LL’s) has

been well established [172, 170, 171, 69] in devices up to a few µm2 in size. Both experi-

mental [173, 168, 170, 171, 69] and theoretical [172, 174, 175] investigations indicate that

Coulomb interaction plays a critical role in these previously observed oscillations—as a

function of both magnetic field and electrostatic gate voltage—suggesting an interpretation

in terms of field- or gate-controlled Coulomb blockade (CB). The questions of whether it

is even possible to observe resistance oscillations that arise from pure AB interference in

FPI’s, and if so, in what regime, and how to distinguish the two mechanisms, have yet to

88

Page 103: Waves, Particles, and Interactions in Reduced Dimensions

be answered to our knowledge.

In this chapter, we report two different types of resistance oscillations as a function of

perpendicular magnetic field, B, and gate voltage in FPI’s of two different sizes. The type

observed in the smaller (2.0 µm2) device, similar to previous results [167, 173, 168, 169,

170, 171, 69], is consistent with the interacting CB interpretation, while that observed in

the larger (18 µm2) device is consistent with noninteracting AB interference. Specifically,

three signatures that distinguish the two types of oscillations are presented: The magnetic

field period is inversely proportional to the number of fully occupied LL’s for CB, but field-

independent for AB; The gate-voltage period is field-independent for CB, but inversely

proportional to B for AB; Resistance stripes in the two-dimensional plane of B and gate

voltage have a positive (negative) slope in the CB (AB) regime.

7.2 Device and measurement

The devices were fabricated on a high-mobility two-dimensional electron gas (2DEG) resid-

ing in a 30 nm wide GaAs/AlGaAs quantum well 200 nm below the chip surface, with Si

δ-doping layers 100 nm below and above the quantum well. The mobility is ∼ 2, 000 m2/Vs

measured in the dark, and the density is 2.6×1015 m−2. Surface gates that define the FPI’s

are patterned using electron-beam lithography on wet-etched Hall bars [see Fig. 7.1(a)].

These gates come in from top left and bottom right, converging near the middle of the Hall

bar. Figures 7.1(b) and (c) show gate layouts for the 2.0 µm2 and 18 µm2 interferome-

ters. All gate voltages except VC are set around ∼ −3 V (depletion occurs at ∼ −1.6 V).

Voltages, VC, on the center gates are set near 0 V to allow fine tuning of density and area.

Measurements are made using a current bias I = 400 pA, with B oriented into the

2DEG plane as shown in Fig. 7.1(a). The diagonal resistance, RD ≡ dVD/dI is related to

89

Page 104: Waves, Particles, and Interactions in Reduced Dimensions

(a) B

VTVTL VTR VR

VBVL VC1 µm

2.0 µm2 device(b)

1 µm

VTVTL VTRVL

VBRVB VRVBLVC

18 µm2 device(c)

RD

I

Figure 7.1: Measurement setup and devices. (a) Diagram of the wet-etched Hall bar, surfacegates, and measurement configuration. Diagonal resistance, RD, is measured directly acrossthe Hall bar, with current bias, I. Subsequent zoom-ins of the surface gates are also shown;the red box encloses the detailed gate layouts for the device shown in (c). (b,c) Gate layoutsfor the 2.0 µm2 and 18 µm2 devices, respectively. The areas quoted refer to those underVC.

the dimensionless conductance of the device g = (h/e2)/RD [176]. Here, VD is the voltage

difference between edge states entering from the top right and bottom left of the device.

7.3 Resistance oscillations in the 2.0 µm2 device

Figure 7.2(a) shows RD as a function of B measured in the 2.0 µm2 device, displaying

several quantized integer plateaus. Figures 7.2(b) and (c) show the zoom-ins below the

g = 1 and 2 plateaus, respectively, displaying oscillations in RD as a function of B, with

periods ∆B = 2.1 mT and 1.1 mT. This ∆B of 2.1 mT corresponds to one flux quantum,

90

Page 105: Waves, Particles, and Interactions in Reduced Dimensions

1

1/2

1/3

1/41/51/61/8

RD [

h / e

2 ]

87654321B [ T ]

f 0 = 4 3 2 1

(a)

0.50

0.47

RD [ h / e

2 ]

3.703.66B [ T ]

ΔB = 1.1 mT

(c)

0.98

0.94R

D [ h / e2 ]

7.307.26 B [ T ]

ΔB = 2.1 mT

(b)

Figure 7.2: Oscillations in RD as a function of magnetic field, B, for the 2.0 µm2 device.(a) RD as a function of B, showing well-quantized integer plateaus. Different colored back-grounds indicate different numbers of fully-occupied LL’s, f0, through the device. (b, c)Zoom-ins of the data in (a), at f0 = 1 and 2, respectively, showing oscillations in RD, andtheir B periods, ∆B.

φ0 ≡ h/e, through an area A = 2.0 µm2, which matches the device design; hence 1.1 mT

corresponds to φ0/2 through about the same area. This is indeed the field-period scaling

observed previously [167, 170, 171, 69], where for f0 number of fully occupied LL’s in the

constrictions, ∆B is expected to be given by (φ0/A)/f0. Thus, in Fig. 7.3(a) we show ∆B

at each 1/f0, and a linear fit constrained through the origin, demonstrating the expected

relationship.

We emphasize that this field-period scaling is inconsistent with simple AB oscillations,

which would give a constant ∆B corresponding to one flux quantum through the area of the

device. This can, however, be understood within an intuitive picture presented in a recent

theoretical analysis [174] that considers a dominant Coulomb interaction within the device.

In this picture, on the riser of RD where f0 < g < f0 + 1, the (f0 + 1)th and higher LL’s

91

Page 106: Waves, Particles, and Interactions in Reduced Dimensions

2.0 μm2 device

2.0

1.5

1.0

0.5

0.0

Δ B [

mT

]

ΔB = 2.1 mT • [ ]

(a)

3

2

1

0

ΔVT, Δ

VC [

mV

]

ΔVT = 3.1 mV ΔVC = 0.18 mV

(e) 0.98

0.94

RD [ h / e

2 ]

210ΔVC = 0.17 mV

(f) B = 7.2 T

0.47

0.45

RD [ h / e

2 ]

210ΔVC = 0.18 mV

(g) B = 3.6 T

0.24

0.22

RD [ h / e

2 ]

210VC [ mV ]

ΔVC = 0.18 mV

(h) B = 1.8 T

0.96

0.92R

D [ h / e2 ]

7.137.12ΔB = 2.1 mT

(b)

0.42

0.40

RD [ h / e

2 ]

3.443.43ΔB = 1.1 mT

(c)

0.24

0.23

RD [ h / e

2 ]

1.821.81B [ T ]

ΔB = 0.54 mT

(d)

112

14

16

180

f 0

1

f 0

1

112

14

16

180

f 0

1

Figure 7.3: Magnetic field and gate voltage periods at various f0, for the 2.0 µm2 device.(a) ∆B as a function of 1/f0, and a best-fit line constrained through the origin. (b-d) RD

oscillations as a function of B, at f0 = 1, 2, and 4, respectively. (e) ∆VT (diamonds) and∆VC (circles) as a function of 1/f0, and their averages indicated by horizontal lines. (f-h)RD oscillations as a function of VC, at f0 = 1, 2, and 4, respectively.

will form a quasi-isolated island inside the device that will give rise to Coulomb blockade

92

Page 107: Waves, Particles, and Interactions in Reduced Dimensions

effects for sufficiently large charging energy,

EC =1

2C(ef0 ·BA/φ0 + eN − CgVgate)2, (7.1)

where N is the number of electrons on the island, C is the total capacitance, and Cg is the

capacitance between the gate and the dot. The magnetic field couples electrostatically to

the island through the underlying LL’s: when B increases by φ0/A, the number of electrons

in each of the f0 underlying LL’s will increase by one. These LL’s will act as gates to

the isolated island: Coulomb repulsion favors a constant total electron number inside the

device, so N will decrease by f0 for every φ0/A change in B, giving rise to f0 resistance

oscillations.

Further evidence for the CB mechanism in the 2.0 µm2 device is found in the resistance

oscillations as a function of gate voltages. Figures 7.3(f-h) show RD as a function of center

gate voltage VC, for f0 = 1, 2 and 4, respectively. Figure 3(e) summarizes gate voltage

periods ∆VT and ∆VC at various f0, and shows they are independent of f0. This behavior

is consistent with the CB mechanism, because, as can be inferred from Eq. (7.1), gate-

voltage periods are determined by the capacitance Cg, which should be independent of

f0.

7.4 Resistance oscillations in the 18 µm2 device

Having identified CB as the dominant mechanism2 for resistance oscillations in the 2.0 µm2

device, we fabricated and measured an 18 µm2 device, an order of magnitude larger in size,

2Although the existence of interference in small devices cannot be ruled out, we emphasizethat Coulomb charging alone is sufficient to explain all data observed in small devices.A recent preprint (Ref. [177]) interprets magneto-oscillations in a small dot in terms ofan interfering AB path, quantized to enclose an integer N . However, as will be seen inFig. 7.4(e), gate voltage periods can change continuously by an order of magnitude in theAB regime, suggesting that N is not quantized in the AB regime.

93

Page 108: Waves, Particles, and Interactions in Reduced Dimensions

18 μm2 device0.3

0.2

0.1

0.0

ΔB [

mT

]

2.01.51.00.50.01 / B [ T-1 ]

ΔB = 0.244 mT

(a)

14

12

10

8

6

4

2

0

ΔVT, Δ

VC [

mV

]

2.01.51.00.50.01 / B [ T-1 ]

2.34

2.32

RD [ h / e

2 ]

-3.004 -3.000ΔVT = 0.74 mV

(f) B = 6.2 T

0.562

0.559

RD [ h / e

2 ]

-3.00 -2.98ΔVT = 3.0 mV

(g) B = 2.5 T

0.184

0.180

RD [ h / e

2 ]

-3.03 -3.00 -2.97VT [ V ]

ΔVT = 9.4 mV

(h) B = 0.72 T

2.27

2.24R

D [ h / e2 ]

6.1986.194ΔB = 0.24 mT

(b)

0.555

0.552

RD [ h / e

2 ]

2.5142.510

ΔB = 0.24 mT

(c)

0.185

0.181

RD [ h / e

2 ]

0.7280.724B [ T ]

ΔB = 0.25 mT

(d)

ΔVT = 6.7 mV • [ ]

ΔVC = 1.3 mV • [ ]

(e) 1TB1TB

Figure 7.4: Magnetic field and gate voltage periods at various B, for the 18 µm2 device.(a) ∆B as a function of 1/B, and their average indicated by a horizontal line. (b-d) RD

oscillations as a function of B, over three magnetic field ranges. (e) ∆VT (diamonds) and∆VC (circles) as a function of 1/B, and best-fit lines constrained through the origin. (f-h)RD oscillations as a function of VT, at B = 6.2 T, 2.5 T, and 0.72 T, respectively.

hence an order of magnitude smaller in charging energy. The center gate covering the whole

device, not present in previous experiments [167, 173, 168, 169, 170, 171, 69], also serves

94

Page 109: Waves, Particles, and Interactions in Reduced Dimensions

to reduce the charging energy. In this device, RD as a function of B at three different

fields is plotted in Figs. 7.4(b-d), showing nearly constant ∆B. The summary of data in

Fig. 7.4(a) shows that ∆B, measured at 10 different fields ranging from 0.5 to 6.2 T, is

indeed independent of B; its average value of 0.244 mT corresponds to one φ0 through an

area of 17 µm2, close to the designed area. This is in contrast to the behavior observed in

the 2.0 µm2 device, and is consistent with simple AB interference. Gate voltage periods are

also studied, as has been done in the 2.0 µm2 device. Figures 7.4(f-h) show RD as a function

of VT at three different fields, and Fig. 7.4(e) shows both ∆VT and ∆VC as a function of

1/B. In contrast to the behavior observed in the 2.0 µm2 device, ∆VT and ∆VC are no

longer independent of B, but proportional to 1/B. This behavior is consistent with AB

interference, because the total flux is given by φ = B · A and the flux period is always φ0;

assuming that the area changes linearly with gate voltage, gate-voltage periods would scale

as 1/B for AB.

7.5 One more signature

As shown above, the magnetic field and gate voltage periods have qualitatively different B

dependence in the 2.0 µm2 and 18 µm2 devices, the former consistent with CB, and the

latter consistent with AB interference. Based on these physical pictures, one can make

another prediction in which these two mechanisms will lead to opposite behaviors. In the

CB case, increasing B increases the electron number in the underlying LL’s, thus reducing

the electron number in the isolated island via Coulomb repulsion. This is equivalent to

applying more negative gate voltage to the device. On the other hand, for the AB case,

increasing B increases the total flux through the interferometer, and applying more positive

gate voltage increases the area, thus also the total flux; therefore, higher B is equivalent to

95

Page 110: Waves, Particles, and Interactions in Reduced Dimensions

1.0

0.8

0.6

0.4

0.2

0.0

VC [

mV

]

1.7661.7651.7641.763B [ T ]

-5

0

5

δRD [ 10

-3 h / e2 ]

2.0 μm2 device

-8

-6

-4

-2

0

2

4

VC [

mV

]

0.5460.5450.544B [ T ]

2

0

-2

δRD [ 10

-3 h / e2 ]

18 μm2 device

(a)

(b)

Figure 7.5: (a) δRD, i.e. RD with a smooth background subtracted, as a function of Band VC, for the 2.0 µm2 device. (b) Same as in (a), but for the 18 µm2 device.

more positive gate voltage. As a result, if RD is plotted in a plane of gate voltage and B,

we expect stripes with a positive slope in the CB case and a negative slope in the AB case.

Figures 7.5(a,b) show RD as a function of VC and B for the 2.0 µm2 and 18 µm2

devices, respectively. As anticipated, the stripes from the 2.0 µm2 device have a positive

96

Page 111: Waves, Particles, and Interactions in Reduced Dimensions

slope, consistent with the CB mechanism, while stripes from the 18 µm2 device have a

negative slope, consistent with AB interference. This difference can serve to determine the

origin of resistance oscillations without the need to change magnetic field significantly.

7.6 Discussion

The three distinct signatures that we observe between CB and AB interference in this work

can also shed light on some of the previous experiments and their interpretations. A few

recent experiments studying fractional charge and statistics in FPI’s [65, 66, 67, 68] interpret

resistance oscillations as arising from AB interference while taking each gate-voltage period

as indicating a change of a quantized charge. However, as shown in Fig. 7.4(e), the gate

voltage periods observed in the big device change by more than an order of magnitude over

the field range that we study, and are inversely proportional to 1/B, suggesting that charge

is not quantized in the AB regime. Also in Ref. [67], the authors have observed that the

magnetic field period stay constant between filling factor 1 and 1/3, but the gate voltage

period at filling factor 1/3 is only 1/3 the size at filling factor 1. Although these observations

can be interpreted as a result of fractional statistics, as the authors have done, there are at

least two other possible interpretations: integer AB interference and CB with a charge of

e/3. We consider clear identification of the mechanisms leading to oscillations—for instance

using the method of Fig. 7.5—to be crucial for interpreting future experiments, particularly

as the quantum states under investigation become more subtle.

97

Page 112: Waves, Particles, and Interactions in Reduced Dimensions

7.7 Acknowledgements

We acknowledge J. B. Miller for device fabrication and discussion, R. Heeres for his work

on the cryostat, and I. P. Radu, M. A. Kastner, B. Rosenow, N. Ofek, I. Neder and B. I.

Halperin for helpful discussions. This research is supported in part by Microsoft Corporation

Project Q, IBM, NSF (DMR-0501796), and the Center for Nanoscale Systems at Harvard

University.

98

Page 113: Waves, Particles, and Interactions in Reduced Dimensions

Chapter 8

Edge-State Velocity and Coherencein a Quantum HallFabry-Perot Interferometer

D. T. McClure†, Yiming Zhang†, B. Rosenow†‡, E. M. Levenson-Falk†, C. M. Marcus††Department of Physics, Harvard University, Cambridge, Massachusetts 02138, USA‡Max-Planck-Institute for Solid State Research, Heisenbergstr. 1, D-70569 Stuttgart,

GermanyL. N. Pfeiffer, K. W. West

Bell Laboratories, Alcatel-Lucent Technologies, Murray Hill, New Jersey 07974, USA

We investigate nonlinear transport in electronic Fabry-Perot interferometers in the in-

teger quantum Hall regime. For interferometers sufficiently large that Coulomb blockade

effects are absent, a checkerboard-like pattern of conductance oscillations as a function of

dc bias and perpendicular magnetic field is observed. Edge-state velocities extracted from

the checkerboard data are compared to model calculations and found to be consistent with

a crossover from skipping orbits at low fields to ~E × ~B drift at high fields. Suppression of

visibility as a function of bias and magnetic field is accounted for by including energy- and

field-dependent dephasing of edge electrons. 1

1This chapter is adapted from Ref. [178].

99

Page 114: Waves, Particles, and Interactions in Reduced Dimensions

8.1 Introduction

The electronic Fabry-Perot interferometer (FPI), implemented as a quantum dot in the

quantum Hall (QH) regime, has attracted theoretical [62, 63, 64, 174, 166] and experimen-

tal [66, 179, 67, 69, 68, 22] interest recently, especially in light of the possibility of observing

fractional [62] or non-Abelian [64, 63, 166, 26] statistics in this geometry. Earlier experi-

ments reveal that Coulomb [167, 173, 180] and Kondo [181, 182] physics can play important

roles, as well. With such a rich spectrum of physics in these devices, a thorough under-

standing of the mechanisms governing transport even in the integer QH regime remains

elusive.

While most work on electronic FPI’s to date has focused on transport at zero dc bias,

finite-bias measurements have proved to be a useful tool in understanding the physical

mechanisms important in other interferometer geometries. In metallic [9] and semiconduct-

ing [132] rings interrupted by tunnel barriers, oscillations in transmission as a function of

magnetic field and dc bias, forming a checkerboard pattern, have been observed. These fea-

tures, attributed to the electrostatic Aharonov-Bohm (AB) effect [183, 184, 185], were used

to measure the time of flight and dephasing in these devices. Similar checkerboard-like lobe

structures have also been observed in Mach-Zehnder interferometers [186, 164, 187]. In that

case, the pattern of oscillations is not readily explained within a single-particle picture and

remains the subject of continued theoretical study [188, 189, 190, 191]. In electronic FPI’s,

conductance oscillations as a function of dc bias have been investigated theoretically [62]

and provide a means of extracting the edge-state velocity from the period in dc bias. Edge-

state velocity measurement without the use of high-bandwidth measurements [192, 193]

will likely be useful in determining appropriate device parameters to probe exotic statistics

beyond the integer regime. This approach was recently used [179] to measure the edge-state

100

Page 115: Waves, Particles, and Interactions in Reduced Dimensions

velocity at ν = 1/3, though in a small (∼ 1 µm2) device where Coulomb interactions, absent

in the theory, may be expected to play a dominant role [69, 22].

In this chapter, we present measurements of finite-bias conductance oscillations in an

18 µm2 electronic FPI whose zero-bias behavior is consistent with AB interference without

significant Coulomb effects [22]. We find a checkerboard-like pattern of conductance oscil-

lations as a function of dc bias and magnetic field, in agreement with the predictions of

Chamon et al. [62]. Measuring the period in dc bias allows the velocity of the tunneling

edge state to be extracted over a range of magnetic fields, yielding a low-field saturation

consistent with a crossover from ~E × ~B drift to skipping orbits. High-bias fading in the

checkerboard pattern is quantitatively consistent with a dephasing rate proportional to en-

ergy and magnetic field. Zero-bias oscillations in a 2 µm2 device of similar design, where

Coulomb effects are significant [22], do not evolve periodically with dc bias; instead, plots of

conductance versus bias and magnetic field reveal diamond-like regions of blockaded trans-

port in the weak-forward-tunneling regime that become more smeared out with stronger

forward tunneling.

8.2 Device and measurement

Devices are fabricated on GaAs/AlGaAs quantum-well structures with a two-dimensional

electron gas (2DEG) of density n = 2.7× 1015 m−2 and mobility µ = 2, 000 m2/Vs located

200 nm below the surface. Hall bars are wet-etched as shown in Fig. 8.1(a), and metal

surface gates are patterned by electron-beam lithography as in Fig. 8.1(b). Interferometers

are defined by negative voltages (∼ −3V) applied to all gates except VC, and samples are

cooled in a dilution refrigerator to ∼ 20 mK. A current bias I, consisting of a dc component

of up to 30 nA and a 135-Hz component of 400 pA, gives rise to the diagonal voltage VD

101

Page 116: Waves, Particles, and Interactions in Reduced Dimensions

(a) B VD

I

(b)

1 µm

VTVTL VTRVL

VBRVB VVBLVC R

(c)

a

Figure 8.1: Measurement setup and the electronic Fabry-Perot device (a) With a currentbias I applied at one end of the Hall bar, voltage VD is measured directly across its width.Surface gates are shown in increasing detail, with a red box indicating the region shownin (b). (b) Gate layout of the 18 µm2 device, which is operated as an interferometer bydepleting all gates except VC. (c) Schematic diagram of possible transmission paths throughthe device in the quantum Hall regime.

across the device, measured directly across the width of the Hall bar [Fig. 8.1(a)]. Lock-

in measurements of diagonal conductance, GD ≡ dI/dVD, are used to study changes in

interferometer transmission as a function of both VD and perpendicular magnetic field B.

As shown in Fig. 8.1(c), the current-carrying chiral edge states can be partially reflected

at each constriction, leading to interference between the different possible trajectories as a

function of the phase accumulated by encircling the interferometer.

102

Page 117: Waves, Particles, and Interactions in Reduced Dimensions

8.3 Checkerboard pattern and interpretation

A typical measurement of GD as a function of B and VD in the 18 µm2 device is shown in

Fig. 8.2(a), where a smooth background has been subtracted. A checkerboard-like pattern

of oscillations periodic in both B and VD is observed, with reduced amplitude at high bias.

Similar patterns are seen at fields B = 0.22−1.26 T; over this range the Landau level index,

N , of the tunneling edge ranges from 4 to 1, but the field period of oscillations is always

∆B ≈ 0.25 mT, independent of both field and bias.

Magnetoconductance oscillations in this device reflect AB interference of partially trans-

mitted edge states [22], with a phase shift ∆ϕ = 2πΦ/Φ0, where Φ = BA is the flux enclosed

(in area A) by the interfering edge, and Φ0 ≡ h/e is the magnetic flux quantum. The ob-

served field period corresponds to A ≈ 17 µm2, consistent with the dot area after subtracting

a depletion length of roughly the 2DEG depth. The sinusoidal lineshape of the oscillations

seen here suggests that coherent transport is dominated by two trajectories that differ in

length by one traversal of the dot perimeter.

When a dc bias is added to VD, an additional phase shift appears between interfering

trajectories, associated with the energy-dependent wave vector of the contributing edge-

state electrons; we will refer to this as the Fabry-Perot phase. The wave vector changes

with energy as δk = δε/~v, where v denotes the edge-state velocity. Following the analysis

of non-interacting electrons in Ref. [62], in which bias is assumed to affect mainly the

chemical potential, we assign an additional relative phase 2aε/~v to an electron traversing

the perimeter at energy ε above the zero bias Fermi level, where a ∼ 2√A = 8.2 µm

denotes the path length between constrictions [Fig. 8.1(c)]. For a symmetrically applied

dc bias (relative to the gate voltages), and neglecting contributions from multiply-reflected

103

Page 118: Waves, Particles, and Interactions in Reduced Dimensions

-60

-40

-20

0

20

40

60

VD [ m

V ]

1.51.00.50.0

d B [ mT ]

Calc

ula

ted d

GD [ a

. u. ]

(b)

-60

-40

-20

0

20

40

60

VD [ m

V ]

1.51.00.50.0

d B [ mT ]

-0.05

0.00

0.05

dG

D [ e

2 / h ]

(a)

Figure 8.2: Nonlinear magnetoconductance in an Fabry-Perot interferometer (a) GD as afunction of B and VD in the 18 µm2 device near B = 0.47 T, with a smooth backgroundsubtracted. (b) δGD calculated from Eq. (8.1), multiplied by the damping factor fromEq. (8.2), with ∆B = 0.25 mT, ∆VD = 56 µV, and α = 0.2.

trajectories, the expected differential conductance has the form

δGD(Φ, VD) = δG0 cos(2πΦ/Φ0) cos(eVDa/v~), (8.1)

where the amplitude δG0 does not depend on field or dc bias. Note that in this model,

the contributions of AB and Fabry-Perot phase separate into a product of two cosines,

yielding a checkerboard pattern, as observed in the experimental data, Fig. 8.2(a). Ref. [62]

104

Page 119: Waves, Particles, and Interactions in Reduced Dimensions

predicts that when the bias is only applied to one contact, with the other contact held at

ground (again, relative to the gates), the two phase contributions from bias and field instead

appear as arguments of a single cosine, yielding a diagonal stripe pattern. Experimentally,

the bias is always applied only at one end of the Hall bar, with the other end grounded;

however, interaction effects within the dot are likely to effectively symmetrize the applied

bias [194]. Alternatively, a model in which the bias mainly affects the electrostatic (rather

than chemical) potential [195] also yields Eq. (8.1) without the need for a symmetric bias. In

either interpretation, the bias period corresponds to the edge velocity via ∆VD = (h/e)(v/a).

We account for the reduced amplitude of oscillations at high bias by multiplying the right

side of Eq. (8.1) by a damping factor, e−2πα |VD|/∆VD , where (2πα)−1 gives the number of

periods over which the amplitude falls to 1/e of its zero-bias value. Lacking theory for edge-

state dephasing in FPI’s, this form is motivated by the observation in related experiments

of a dephasing rate proportional to energy [132, 196]. We thus identify a voltage-dependent

dephasing rate, τ−1ϕ (VD) = α|eVD|/2~, which reduces amplitude by e−2to/τϕ , where 2to =

2a/v is the time of flight around the interferometer. To extract interference and dephasing

parameters, the form

δG(VD) = δG0e−2πα |δx| cos(2π δx), (8.2)

where δx = (VD−Voff)/∆VD and Voff is a bias offset, is fit to cuts of the data in Fig. 8.2(a),

which yields a period ∆VD = 56 µV and dephasing parameter α = 0.2. These values,

along with ∆B = 0.25 mT are then used to produce the plot shown in Fig. 8.2(b). Fig-

ures 8.3(a,b) show vertical cuts from data along with fits of Eq. (8.2) at B = 0.22 T and

1.26 T, respectively, representing a trend toward smaller ∆VD and larger α at higher fields,

the details of which we now study.

105

Page 120: Waves, Particles, and Interactions in Reduced Dimensions

Figure 8.3: Magnetic field dependence of extracted velocity and damping factor (a) GD

as a function of VD (black dots) at a field of B = 0.22 T, with a fit of Eq. (8.2) (redcurve) yielding ∆VD = 76 µV and α = 0.063. (b) Same as (a) but at B = 1.26 T andyielding ∆VD = 47 µV and α = 0.34. (c) Black dots indicate edge velocities (left axis)determined from measured ∆VD (right axis) as a function of 1/B. The red curves indicatetheoretical calculations: at low 1/B, the diagonal dashed line indicates the drift velocitycorresponding to E = 8× 104 V/m; at high 1/B, the top and bottom solid curves indicatethe predicted skipping-orbit velocities corresponding to the lowest and highest constrictiondensities, respectively. (d) Best-fit damping parameter α as a function of B, with a linearfit of slope 0.26 T−1 constrained through the origin. Inset: γ = 2πα/e∆VD as a function ofB, with a linear fit of slope 31 (meV · T)−1 constrained through the origin.

8.4 Edge-state velocity and energy-dependent dephasing

The black circles in Fig. 8.3(c) indicate the best-fit ∆VD (right axis) and corresponding edge

velocity (left axis) as a function of 1/B. The velocities appear roughly proportional to 1/B

106

Page 121: Waves, Particles, and Interactions in Reduced Dimensions

before saturating at v ∼ 1.5 × 105 m/s for 1/B & 2 T−1. Red curves indicate calculations

based on single-particle models of edge velocities in two regimes. In the high-field limit,

where the cyclotron radius is much smaller than the length scale on which the confining

potential changes by the cyclotron gap, ~E × ~B drift gives a velocity vd = E/B, where E

is the local slope of the confining potential. The data in this regime are consistent with

a value E ∼ 8 × 104 V/m, which is reasonable given device parameters. At low fields,

where the cyclotron radius exceeds the length scale set by E, electron velocities can be

estimated from a skipping-orbit model. For hard-wall confinement, the skipping velocity

would be proportional to the cyclotron frequency and radius: vs ∼ ωcrc. Here, we have

performed a detailed semi-classical calculation assuming a more realistic confining potential

that vanishes in the bulk and grows linearly near the edge. In this regime, the predicted

velocity depends on not only B and E but also on the Landau level index, N , resulting in

a discrete jump in velocity for every change in N . Since the density in the constrictions

(which along with B determines N) varies over the course of the experiment, two theoretical

curves are plotted in this regime: the top one corresponds to the lowest observed constriction

density of 2.8× 1014 m−2, and the bottom one corresponds to the highest, 9.5× 1014 m−2,

both estimated from GD and B.

Figure 8.3(d) shows the best-fit damping parameter α as a function of B, revealing

rough proportionality: a straight line constrained to cross the origin describes the data well

with a best-fit slope of 0.26 T−1. In analogy to dephasing in 2D diffusive systems [197],

we suggest that coupling to compressible regions in the bulk may lead to dephasing with

the VD-dependence τ−1ϕ ∝ RVD, where R is the resistance per square in the bulk. Over

the field range of our data, the bulk longitudinal resistivity RXX (not shown) is on average

roughly proportional to B; taking RXX as an estimate of R would then lead to a predicted

107

Page 122: Waves, Particles, and Interactions in Reduced Dimensions

dephasing rate proportional to both energy and magnetic field, consistent with the data.

Despite this agreement, we emphasize that Ref. [197] was not developed for edge states or

FPI’s, and a theory of dephasing in this regime remains lacking.

Alternatively, defining the damping factor as simply e−γ|eVD|, one also finds rough pro-

portionality between γ and B, as shown in the inset of Fig. 8.3(d). Here the best-fit slope

for a straight line constrained through the origin is 31 (meV ·T)−1. The damping parameter

γ is related to α and to the dephasing length, `ϕ = vτϕ, by γ = αto/~ = 2a/|eVD|`ϕ; there-

fore, since to varies with field, at most one of α and γ can be proportional to B. Physically,

the latter case would correspond to `−1ϕ being the quantity that is linear in B instead of

τ−1ϕ . Experimental scatter prevents us from distinguishing these two possibilities.

8.5 Nonlinear magnetoconductance in a 2 µm2 device

Measurements on a 2 µm2 device of similar design, whose zero-bias oscillations have previ-

ously been demonstrated as consistent with Coulomb-dominated behavior [22], do not yield

regular oscillations as a function of bias. Figure 8.4 shows GD as a function of B and VD

in a regime of weak forward-tunneling, where diamond-like features appear. Interpreting

these features as the result of Coulomb blockade yields a charging energy of roughly 25 µeV,

reasonable given the device size, 2DEG depth, and the large capacitance afforded by the top

gate. In regimes of stronger forward tunneling, the diamond edges become more smeared

out, but in contrast to the behavior in the 18 µm2 device, periodic oscillations as a function

of dc bias are not seen.

108

Page 123: Waves, Particles, and Interactions in Reduced Dimensions

-50

0

50

VD [ m

V ]

3.7723.7703.7683.7663.764

B [ T ]

2.15

2.10

2.05

GD [ e

2 / h ]

Figure 8.4: GD as a function of B and VD in the 2 µm2 device.

8.6 Conclusion

In conclusion, quantum Hall FPI’s large enough that Coulomb charging is negligible are

found to display both AB and Fabry-Perot conductance oscillations. The combination of

these two effects yields a checkerboard-like pattern of oscillations from which the edge-

state velocity and dephasing rate can be extracted, and both are found to be consistent

with theoretical calculations. Although this pattern resembles that seen in Mach-Zehnder

interferometers, the dependence of its characteristics on magnetic field is evidently quite

different from what has been observed in those devices [186, 187], providing experimental

evidence that the underlying mechanisms for oscillations with bias in the two types of

devices may be quite different.

109

Page 124: Waves, Particles, and Interactions in Reduced Dimensions

8.7 Acknowledgements

We are grateful to R. Heeres for technical assistance and to B. I. Halperin, M. A. Kastner,

C. de C. Chamon, A. Stern, R. Gerhardts, J. B. Miller and I. P. Radu for enlightening

discussions. This research has been funded in part by Microsoft Corporation Project Q,

IBM, NSF (DMR-0501796), Harvard University, and the Heisenberg program of DFG.

110

Page 125: Waves, Particles, and Interactions in Reduced Dimensions

Chapter 9

Unpublished results

In this last chapter, I will be describing several results that are not published nor submitted

for publication. These results are not fully understood and further experimental studies

are needed to bring them to the quality level considered publishable in the Marcus group.

Although they will generally not appear in a Ph.D. dissertation, I believe they are well

worth the presentation here because, on one hand, it can still be of value to publicize them

in case some one seeing them would be able to help uncover the mystery underlying these

data; on the other hand, I think it can be instructive for new graduate students to see some

unpublished data and to learn the specific reasons why some results remain unpublished.

Since these results are still for one reason or another incomplete, and not submitted

for publication, their presentation is much more casual, and the references to relevant work

are not as rigorously checked as those published results. To some degree, they are more

or less like peeks into an experimentalist’s lab notebook. I would also like to thank my

collaborators, Doug McClure, Leo DiCarlo, Eli Levenson-Falk, as well as my advisor Charlie

Marcus for their work on these projects.

In the following sections, I will be describing four such results: the first section stud-

ies current noise through a quantum dot, modulated by the charge noise of another dot

nearby, showing huge super-Poissonian noise; the second section studies temperature and

111

Page 126: Waves, Particles, and Interactions in Reduced Dimensions

bias dependent transport through a quantum point contact in the quantum Hall regime

between filling factors 2 and 3, observing both quantized fractional plateaus and quasi-

particle tunneling; the third section reports the experimental observation of an unexpected

3/2 quantized conductance plateau in a constriction; and the final section studies high-bias

magnetoresistance in the bulk, showing rich structures that might be related to reentrant

or anisotropic integer quantum Hall effects.

9.1 Current noise modulated by charge noise

In non-interacting electronic systems, current noise is expected to be always below the

Poissonian value due to Fermi statistics [19, 29, 54, 55]. Interactions, however, can raise

the current noise above the Poissonian limit, and the specific conditions under which such

super-Poissonian noise can arise has been the subject of numerous studies recently [43,

113, 44, 136, 137, 140, 45, 139, 125, 126, 46, 143, 33]. Super-Poissonian noise observed

in metal-semiconductor field effect transistors [43], tunnel barriers [113] and self-assembled

stacked quantum dots [44] has been attributed to interacting localized states [136, 137, 43]

occurring naturally in these devices. In more controlled geometries, super-Poissonian noise

has been associated with inelastic cotunneling [140] in a nanotube quantum dot [45], and

with dynamical channel blockade [139, 125, 126] in GaAs/AlGaAs quantum dots [46, 143,

33].

In essence, these types of super-Poissonian noise arise because the current through

the device is switching between two or more levels, leading to random telegraph noise

(RTN). The most natural system to realize such RTN in a controllable fashion is perhaps

a charge sensor measuring charge fluctuations of a nearby quantum dot. Charge sensors

can be realized in the form of quantum dots (also called single electron transistors) [12,

112

Page 127: Waves, Particles, and Interactions in Reduced Dimensions

13, 198], or quantum point contacts [199, 200, 201]. In this section, we have realized such

controlled system using a double quantum dot, with one dot acting as a charge sensor, and

the other dot acting as a controlled charge noise generator. We have observed huge super-

Poissonian current noise through the charge sensing dot, when the other dot is near the

charge degeneracy point, and the current noise depends non-monotonically on the tunneling

rates of the charge noise generator.

The double quantum dot device is the same one as that used in Ch. 4, with the bottom

dot acting as a charge sensor, and the top dot acting as a charge noise generator and always

operating at zero dc bias [see Fig. 9.1(a)]. The central point contact is always depleted

by gate voltages Vl = Vr = −1420 mV, preventing inter-dot tunneling. Gate voltages Vtl

(Vbl) and Vtr (Vbr) control the tunnel barrier between the top (bottom) dot and its left and

right leads. Plunger gate voltage Vtc (Vbc) controls the electron number in the top (bottom)

dot. For convenience, the variable, Vts ≡ [(Vtr + 614 mV) + (Vtl + 577 mV)]/√

2, is used to

control the left and right barriers of the top dot simultaneously. Differential conductance

gt (gb) through the top (bottom) dot is measured using standard lock-in techniques with

an excitation of 25 (30) µVrms at 677 (1000) Hz, and the current noise through the bottom

dot, Sb, is measured near 2 MHz [131]. The dc current Ib through the bottom dot at finite

Vb is obtained by numerically integrating gb.

Shown in the inset of Fig. 9.1(b) are conductances gt and gb, as a function of Vts,

which controls the barrier transparencies of the top dot, as well as its energy. When Vts is

tuned more negative, the top dot tunnel barriers are getting higher, and electrons are being

pushed out one by one, leading to a series of Coulomb blockade (CB) peaks in gt that are

getting smaller in height. In the meantime, as the bottom dot is configured near a charge

degeneracy point, its conductance also evolves through a full CB peak due to the coupling

113

Page 128: Waves, Particles, and Interactions in Reduced Dimensions

[ mV ]

280 mK 4.2 K

Digitize &

Analyze

R R L L C C

60 dB

R R L L 60 dB

Vtl Vtc Vtr

Vbl

Vr

Vt

Vl

(a)

(b)

It

Ib Vb

500 nm

gb

Vbc Vbr

Stb

St

Sb

50Ω

50Ω

50Ω

50Ω

gt

1.5

1.0

0.5

0.0

g [e

2 / h]

-40 -30 -20 -10 0 10Vts [ mV ]

gt gb

-850

-800

-750

Vtc

[ m

V ]

-100 -80 -60 -40 -20 0Vts

0.60.40.20.0gb [e

2 / h]

1.51.00.50.0 gt [e2 / h]

gt

gb

Figure 9.1: (a) Micrograph of the device and equivalent circuit near 2 MHz of the noisedetection system. (b) Conductance through the top and bottom dot, gt and gb, as a functionof Vts and Vtc (see main text for the definition of Vts). Inset: gt and gb as a function of Vts.

between Vts and the bottom dot. In addition, whenever the electron number in the top dot

changes by one, gb exhibits a jump, even when the CB peak in gt is getting unmeasurably

small. This demonstrates the charge sensing of the top dot by the bottom dot.

Figure 9.1(b) shows gt and gb as a function of Vts and Vtc, tracking three consecutive

CB peaks in gt, while the top dot barrier is getting higher with more negative Vts. The

plunger gate Vtc is used to compensate the energy shift caused by changing Vts so that the

dot electron numbers remain unchanged. Although the peaks in gt is vanishing into the

background noise passing Vts ∼ −15 mV, the conductance jumps in gb clearly tracks the

114

Page 129: Waves, Particles, and Interactions in Reduced Dimensions

-848

-847

-846

-845

-844

Vtc

[ m

V ]

0.030.00 gt [ e2 / h ] 0.100.02 gb [ e2 / h ]

-848

-847

-846

-845

-844

Vtc

[ m

V ]

-1257 -1256 -1255 -1254Vbc [ mV ]

1.00.0 Sb [ nA x 2e ]

-1257 -1256 -1255 -1254Vbc [ mV ]

0.5-0.5 SEPb [ nA x 2e ]

(a)

(c)

(b)

(d)

Figure 9.2: Conductance and current noise near a honey-comb vertex: gt (a), gb (b), Sb

(c), and SEPb (d), as a function of Vtc and Vbc.

charge transitions of the top dot. In later studies of the dependence of Sb on Vts, Vtc is

always used to compensate the energy shift by Vts in this way, making sure that the same

CB peak is studied.

Setting the top dot so that its conductance is barely measurable, at Vts = −15 mV, and

applying 100 µV dc bias to the bottom dot and zero bias to the top dot, both conductance

and bottom-dot current noise are measured as a function of Vtc and Vbc, as shown in

Fig. 9.2. We find that Sb is greatly enhanced at two points in the plane of Vtc and Vbc [see

Fig. 9.2(c)], corresponding to the two sides of the CB peak in gb, where gb is most sensitive

to nearby charge fluctuations, intersecting with the CB peak in gt. Plotting in Fig. 9.2(d)

the excess Poissonian noise, SEPb ≡ Sb − 2e|Ib|, through the bottom dot, we find that the

115

Page 130: Waves, Particles, and Interactions in Reduced Dimensions

two Sb maxima indeed exceeds the Poissonian limit1. The right maximum of Sb is as large

as 1.2 nA× 2e, while the dc current at that point is only 0.2 nA, corresponding to a Fano

factor of 6.

The data suggest that the enhanced noise in Sb to be RTN induced by the top-dot charge

noise: when the top dot comes into the charge degeneracy point, its charge fluctuates

between two possible values, leading to the bottom-dot current switching between two

values. The power spectrum for RTN associated with two-level switching is given by the

Lorentzian form [36]: SRTN(f) = (∆I)2 ·4Γ01Γ10/(Γ01 +Γ10)/[(Γ01 +Γ10)2 +(2πf)2], where

Γ01 (Γ10) is the transition rate from state 0 (1) to state 1 (0), and ∆I is the difference in

current between state 0 and state 1. Since the top dot is operated at zero-bias, we can write

Γ01 = ftΓt and Γ10 = (1 − ft)Γt, where ft is the Fermi function of the top-dot reservoirs

evaluated at the dot energy level, and Γt = Γtl + Γtr is the total tunneling rate to both

reservoirs of the top dot. The maximum noise occurs when ft = 1/2, corresponding to a

top-dot CB peak, and the maximally enhanced Sb is simply given by:

Smaxb (f) = (∆Ib)2 · Γt/[Γ2

t + (2πf)2]. (9.1)

Extraction of ∆Ib can be performed by measuring Ib as a function of Vtc, at the Vbc

setting that corresponds to the right maximum in Sb, as shown in the inset of Fig. 9.3.

Without any fitting parameters, the theoretical prediction for Smaxb (f = 2 MHz), using

Eq. (9.1), as a function of Γt is shown in Fig. 9.3(b). The predicted Smaxb depends non-

monotonically on Γt, with a maximum at Γt = 2π · 2 MHz. The maximum Sb as a function

of the top-dot barrier transparencies is studied by changing Vts while compensating the

1Super-Poissonian Sb is also observed away from the top-dot charge degeneracy point.This is the related to dynamical channel blockade from multi-level transport of the bottomdot, as has been studied in Ch. 5. Here, we focus on the super-Poissonian noise induced bythe top-dot charge noise, much larger than the intrinsic noise of the bottom dot.

116

Page 131: Waves, Particles, and Interactions in Reduced Dimensions

8

6

4

2

0S

bmax

[ nA

x 2

e ]

-100 -80 -60 -40 -20 0Vts [ mV ]

(a) Data8

6

4

2

0104 106 108 1010

Γt [ Hz ]

(b) Theory

ΔIb = 232 pA

0.4

0.2

0.0

I b [

nA ]

-848 -844Vtc [ mV ]

Figure 9.3: (a) Measured maximum current noise, Smaxb , as a function of Vts. (b) Theory

for Smaxb as a function of top dot tunneling rate Γt. Inset: Ib as a function of Vtc, and

measurement of ∆Ib.

energy shift with Vtc, as has been done in Fig. 9.1(b). Shown in Fig. 9.3(a) are Smaxb at

13 settings of Vts from 0 to −110 mV. Indeed, Smaxb depends non-monotonically on Vts,

and approaching the single-dot values at high and low limits of barrier transparencies. Yet

unexpectedly, Smaxb exhibits two peaks for intermediate values of Vts, one near −40 mV and

the other near −70 mV. So far, this has remained a mystery to us.

In conclusion, we have observed huge super-Poissonian noise through the bottom quan-

tum dot when charge sensing the top dot. The enhanced noise depends non-monotonically

on the barrier transparencies of the top dot. It is also the first time that we have studied

frequency-dependent current noise, when the time scale of the dynamics is comparable or

slower than our noise measurement frequency at 2 MHz. In contrast to the simple expecta-

tion for RTN, however, Smaxb exhibits two peaks as a function of Vts, which remains to be

understood.

117

Page 132: Waves, Particles, and Interactions in Reduced Dimensions

9.2 Quasi-particle tunneling between filling factor 2 and 3 in

a constriction

This section is mainly a follow-up experiment of the quasi-particle tunneling experiment

by Radu et al. [202]. Radu et al. have found strong nonlinear I − V characteristics in

a quantum point contact (QPC) at filling factor 5/2, associating them with tunneling of

the 5/2 quasi-particles. Both the width and height of the tunneling conductance peaks are

found to scale with temperature, allowing an extraction of quasi-particle charge e∗ and the

interaction parameter g within the weak tunneling framework [203].

This section reports a similar measurement of conductance through a QPC in the quan-

tum Hall regime between filling factors 2 and 3. The new contributions of this section

include the observation of exact quantized plateaus at filling factors of 7/3, 5/2, and 8/3, as

well as two groups of tunneling peaks between these three plateaus. Both groups of peaks

are studied as a function of bias, temperature, and the high-bias background resistance,

and they are found to scale with temperature. Following the methods used in Ref. [202], e∗

and g are extracted as a function of the high-bias background resistance.

The QPC device is fabricated on a high-mobility two-dimensional electron gas (2DEG)

residing in a 30 nm wide GaAs/AlGaAs quantum well 200 nm below the chip surface.

The 2DEG mobility is ∼ 2, 000 m2/Vs measured in the dark, and the density is 2.6 ×

1015 m−2. Surface gates that define the QPC have a separation of 1.2 µm, and are patterned

using electron-beam lithography on wet-etched Hall bars. Bulk Hall resistance RXY, bulk

longitudinal resistance RXX, and diagonal resistance across the device RD are measured

with current bias I applied through the entire Hall bar and perpendicular magnetic field B

oriented into the 2DEG plane [see Fig. 9.4]. In the quantum Hall regime, RXY (= 1/νB·h/e2)

118

Page 133: Waves, Particles, and Interactions in Reduced Dimensions

RDRXY

RXX

VG

VG 150 µm

I

B

Figure 9.4: Optical micrograph of the device and measurement setup. The QPC understudy is highlighted in orange, and has a width of 1.2 µm. Resistances RXY, RXX, andRD are measured across the indicated pairs of ohmic contacts, with current bias I appliedthrough the entire Hall bar. (Image courtesy of Ref. [176]).

and RD (= 1/νD ·h/e2) give independent measurements of the filling factor in the bulk, νB,

and that in the constriction, νD [176].

Shown in Fig. 9.5 are RD as a function of dc I, at B values from 3.0 to 4.4 T and

evenly spaced by 10 mT, for 7 different temperatures from T = 13 mK to 66 mK. This

magnetic field range covers the range in νD from 2 to 3. In this representation, resistance

plateaus as a function of B would appear as accumulation of traces. As can be seen from

Fig. 9.5, in addition to two strong integer plateaus at νD = 2 and 3, fractional plateaus at

7/3 and 5/2 appear at all temperatures, and the plateau at 8/3 is also visible at the lowest

temperatures. Between these three fractional plateaus, two groups of resistance peaks show

up as a function of I, with the peaks getting higher and narrower at lower temperatures.

We denote the group of peaks between 3/7 and 5/2 plateaus “Group 1”, and those between

5/2 and 8/3 “Group 2”.

Following Ref. [202], we interpret these peaks as tunneling between counter-propagating

119

Page 134: Waves, Particles, and Interactions in Reduced Dimensions

R

1/2

1/3

2/5

3/7

3/8

D [

h / e

2 ]

-10 -5 0 5 10I [nA]

13 mK

-10 -5 0 5 10I [nA]

26 mK

-10 -5 0 5 10I [nA]

33 mK

-10 -5 0 5 10I [nA]

40 mK

-10 -5 0 5 10I [nA]

50 mK

-10 -5 0 5 10I [nA]

58 mK

-10 -5 0 5 10I [nA]

0.50

0.48

0.46

0.44

0.42

0.40

0.38

0.36

0.34

66 mK

Gro

up 1

Gro

up 2

Figure 9.5: Diagonal resistance RD as a function of I, at magnetic fields evenly spaced by10 mT, for 7 different temperatures from 13 mK to 66 mK.

fractional edge states, and use the weak tunneling formula [203] to analyze them. Due to

some small but random device drift over the time of these measurements, instead of B, the

high-bias limit RD value, R0D, is chosen as the metric for picking the tunneling peaks of

the same constriction filling for different temperatures. This choice is justified, because not

only it eliminates the small drift, but also R0D is found to be proportional to B in the range

of interest, thus would reflect the true constriction filling without the effect of tunneling

near zero bias.

The red curves shown in Fig. 9.6 are resistance peaks of all temperatures at R0D =

0.414 h/e2, plotted as a function of the diagonal voltage VD, which is extracted by numerical

integration of RD(I). The data set at each R0D are then fit to the weak tunneling formula:

RD = R0D +A · T 2g−2 · F

(e∗VD

kBT, g

), (9.2)

120

Page 135: Waves, Particles, and Interactions in Reduced Dimensions

0.424

0.422

0.420

0.418

0.416

0.414

RD [

h / e

2 ]

-60 0 60VD [ µV ]

13 mK

-60 0 60VD [ µV ]

26 mK

-60 0 60VD [ µV ]

33 mK

-60 0 60VD [ µV ]

40 mK

-60 0 60VD [ µV ]

50 mK

-60 0 60VD [ µV ]

58 mK

-60 0 60VD [ µV ]

66 mK

Figure 9.6: An example of the best fit (black), using the weak tunneling formula [Eqs. (9.2)and (9.3)], to bias- and temperature-dependent RD data (red) at R0

D = 0.414 h/e2, givinge∗ = 0.23 e and g = 0.44.

F(x, g) = B(g + i

x

2π, g − i x

)·π cosh

(x2

)− 2 sinh

(x2

)· Im

[Ψ(g + i

x

)], (9.3)

where A is the tunneling amplitude, e∗ and g are the quasi-particle charge and the interac-

tion parameter, respectively. B(x, y) is the Euler beta function, and Ψ(x) is the digamma

function.

Treating A, e∗ and g as free parameters, and simultaneously fitting to all seven data

curves shown in Fig. 9.6, the weak tunneling formula yields a satisfactory fit, and gives

best-fit e∗ = 0.23 e and g = 0.44. Again following Ref. [202], we can prefix e∗ and g over

a grid of possible values, and only allow A as the free parameter to fit to the data. Shown

in Fig. 9.7 is the mean squared fit error as a function of prefixed e∗ and g, when fitting

to the same data with R0D = 0.414 h/e2. As discussed in Ref. [202], this plot yields more

information on the uncertainty of these measured e∗ and g values, and would facilitate

comparison to various theories.

The constriction filling that corresponds to R0D = 0.414 h/e2 is near midway between

7/3 and 5/2 states, and we would like to analyze the peaks with different constriction fillings

121

Page 136: Waves, Particles, and Interactions in Reduced Dimensions

0.8

0.6

0.4

0.2

g [d

imle

ss]

0.60.50.40.30.20.1e* [e]

321 Fit error [ a. u. ]

Figure 9.7: Mean squared fit error as a function of prefixed e∗ and g, for fitting to the datashown in Fig. 9.6.

0.35

0.30

0.25

0.20

e* [e

]

Group2 Group1

0.60

0.55

0.50

0.45g [d

imle

ss]

0.420.410.400.39RD

0 [ h / e2 ]

Group2 Group1

Figure 9.8: Best-fit e∗ and g as a function of R0D for both groups of tunneling peaks.

within “Group 1” in the same way, and also compare to the peaks within “Group 2”. Shown

in Fig. 9.8 are the best-fit e∗ and g, as a function of R0D, from both groups of peaks.

In a simplistic picture, the peaks within “Group 1” corresponds to either backscattering

of the 5/2 edge across the 5/2 liquid, or forward tunneling of the 5/2 edge over the 7/3

122

Page 137: Waves, Particles, and Interactions in Reduced Dimensions

liquid; the peaks within “Group 2” corresponds to either backscattering of the 8/3 edge

across the 8/3 liquid, or forward tunneling of the 8/3 edge over the 5/2 liquid. Further

assuming that tunneling is enhanced at zero-bias and suppressed at high bias, since peaks

in RD correspond to increased backscattering, we can eliminate the possibility of forward

tunneling, which would lead to dips in RD. Therefore, we expect the peaks in “Group

1” to exhibit the characteristics of the 5/2 state, and those in “Group 2” to exhibit the

characteristics of the 8/3 state. As shown in Fig. 9.8, in “Group 1”, the extracted e∗ is

between 0.2 and 0.25 e, and g ranges from 0.44 to 0.54, in reasonable agreement with the

predictions e∗ = 0.25 e and g = 0.5 of the anti-Pfaffian state [204, 205, 202]. In “Group 2”,

however, the extracted e∗ and g show a large variation at different R0D, making it hard to

draw any conclusions.

In conclusion, we have observed well quantized fractional plateaus at filling factors of

7/3, 5/2, and 8/3, as well as two groups of tunneling peaks between them. Best-fit e∗

and g as a function of constriction filling are extracted using the weak tunneling formula,

for both groups of peaks. The extracted values from peaks in “Group 1”, assumed to be

related to tunneling of the 5/2 edge, is consistent with predictions of the anti-Pfaffian state;

however, the extracted values from peaks in “Group 2”, presumably related to tunneling of

the 8/3 edge, show a large variation. Another drawback of this analysis is that the peaks

can overshoot the value of the next plateau at the lowest temperatures [see the 13 mK data

in Fig. 9.5], putting the theoretical assumption of weak tunneling into question.

123

Page 138: Waves, Particles, and Interactions in Reduced Dimensions

9.3 The 3/2 quantized plateau in quantum point contacts

Since the original discovery of the 1/3 fractional quantum Hall effect over two decades

ago [27], a wealth of fractional quantum Hall states have been discovered, and vast majority

of them are odd-denominator fractions [206, 24]. The odd denominator is a result of Fermi

statistics, which requires the wave function to be antisymmetric under particle exchange;

consequently, any fraction with an even denominator would require an explanation quite

different from those with odd denominators. The 5/2 [207] and 7/2 states in the first excited

Landau level (LL) are the only known even-denominator states in a single-layer system; yet

their origin, possibly being a paired state [25], is still under investigation. In bilayer systems,

1/2 [208, 209] and 3/2 [210] states in the lowest LL have been observed and understood to

arise from the so-called (3, 3, 1) state [211, 212].

In this section, we report observation of the 3/2 plateau, precisely quantized to within

0.02 %, through quantum point contacts (QPCs). Evidence show that they possess a

different origin from that of the single-layer 5/2 and 7/2 states, or of the bilayer 1/2 and

3/2 states: they are never seen in the bulk and only through the QPCs; they appear when

the constriction filling is at 5/3, suggesting their relationship with the 5/3 state. Studying

their temperature and bias dependence, we found that they survive up to 80 mK or over

8 nA.

The QPC devices are formed with top gates in two devices, labeled as device 1 and

device 2, shown in the lower right insets of Figs. 9.9(a) and (b), respectively. These two

devices are fabricated on different two-dimensional electron gas (2DEG) wafers of the same

design2, with a mobility of ∼ 2, 000 m2/Vs and a density of 2.6 × 1015 m−2. As explained

2Device 1 is the 2.0 µm2 dot also used in Chs. 7 and 8. Device 2 is fabricated on thesame wafer as the 18 µm2 dot studied in Chs. 7 and 8.

124

Page 139: Waves, Particles, and Interactions in Reduced Dimensions

1

1/2

1/3

1/41/51/61/8

3/4

3/52/3

R [

h / e

2 ]

108642 B [ T ]

(a) Device 1: RXY RD

0.70

0.68

0.66

0.64

RD [

h / e

2 ]6.66.4

B [ T ]

2/3

1

108642 B [ T ]

(b) Device 2: RXY RD

0.68

0.66

0.64

0.62

RD [

h / e

2 ]

7.06.86.6B [ T ]

2/3x 1.26

VTVTL VTR VR

VBVL VC1 µm 2 µm

VC VRVBRVBLVB

VL VT VTRVTL

Figure 9.9: (a) Bulk Hall resistance RXY (red) and diagonal resistance RD (black) as afunction of magnetic field B measured from device 1, whose gate layout is shown in thelower right inset. The QPC is formed with VT = VL = −1.06 V. The grey curve is the RD

data scaled horizontally by a factor of 1.26, to match the filling factor in the bulk. Theupper left inset shows that RD exhibits a 3/2 quantized plateau. (b) Similar data takenfrom device 2. The QPC is formed with VT = VB = −2.15 V and VC = +0.6 V, where theconstriction density matches that of the bulk.

in the previous section and in Chs. 7 and 8, bulk Hall resistance RXY (= 1/νB · h/e2) and

diagonal resistance RD (= 1/νD · h/e2) give independent measurements of the filling factor

in the bulk, νB, and that in the constriction, νD, with the perpendicular magnetic field B

applied into the 2DEG plane.

Figure 9.9(a) shows RXY and RD as a function of B measured from device 1. In

addition to well quantized integer plateaus, RXY shows fractional plateaus at 5/3, 4/3, etc.;

yet, RD shows a plateau that is quantized at 2/3 h/e2, corresponding to a conductance

of 3/2 channels through the QPC. This 3/2 plateau is precisely quantized to within the

measurement noise of 0.02 %, as shown in the upper left inset of Fig. 9.9(a). The plateaus

in RXY and in RD are not aligned in B due to different densities in the bulk and in the

QPC; therefore, we have scaled the RD data horizontally by a factor of 1.26, to align it with

the RXY trace. Surprisingly, the 3/2 plateau in RD appears at the same filling factor as the

125

Page 140: Waves, Particles, and Interactions in Reduced Dimensions

1/2

3/4

3/5

2/3

RD [

h / e

2 ]

-8 -4 0 4 8I [ nA ]

(b)

1/2

3/5

2/3

RD [

h / e

2 ]

5.04.94.84.74.64.5B [ T ]

13 mK 20 mK 40 mK 60 mK 80 mK

(a)

Figure 9.10: (a) Temperature dependence of the 3/2 plateau in a QPC formed with VB =VTL = −1.6 V in device 1. (b) RD as a function of I, at magnetic fields evenly spaced by20 mT from 5.0 to 6.6 T, measured in a QPC formed with VB = VTL = −2.7 V in device 1.

5/3 plateau in RXY, suggesting that they are somehow related.

Figure 9.9(b) shows RXY and RD as a function of B measured from device 2. This

QPC is formed with negative voltages on two facing gates VT = VB = −2.15 V, and positive

voltage on a center gate VC = +0.6 V, matching the density in the QPC to the bulk. Again,

RD as a function B shows a 3/2 plateau, also precisely quantized to within the measurement

noise. Similar to the behavior of device 1, the position of the 3/2 plateau in RD is the same

as the 5/3 plateau in RXY.

Temperature and bias dependence of the 3/2 plateau are also studied in a QPC formed

with gates VB and VTL in device 1. Figure 9.10(a) shows RD as a function of B at five

different temperatures up to 80 mK, showing that the 3/2 plateau can survive up to 80 mK.

Figure 9.10(b) shows RD as a function of dc I, at B settings from 5.0 to 6.6 T evenly spaced

by 20 mT. The 3/2 plateau, showing up as accumulation of traces at RD = 2/3 h/e2, persists

to at least 8 nA, or equivalently 140 µV.

The origin of the 3/2 plateau in RD, and indeed how it is related to the 5/3 plateau in

126

Page 141: Waves, Particles, and Interactions in Reduced Dimensions

RXY are quite puzzling, especially because the 3/2 plateau has an even denominator. Some

possibilities, however, can be ruled out. Edge reconstruction of the possibly complicated

5/3 edge does not seem to be able to explain the even denominator of the 3/2 plateau. The

precision of quantization also eliminates the possibility of it being some other nearby state

with an odd denominator. Because the 3/2 plateau is absent in the bulk, and appears at

the same filling factor as the 5/3 state in the bulk, it is unlikely to share the same origin

as the 5/2 and 7/2 states, nor as the 1/2 and 3/2 states in bilayer systems. Since the 3/2

plateau has been reproduced in three different QPCs on two wafers, we do not believe it

to be some form of experimental artifacts, and further experimental and theoretical studies

would be needed to reveal the physics underlying this intriguing finding.

127

Page 142: Waves, Particles, and Interactions in Reduced Dimensions

9.4 Non-linear transport in N ≥ 2 Landau levels

It has been known for a over decade that the N ≥ 2 Landau levels (LL’s) exhibit large

anisotropy and reentrant integer states [213, 214]. Their origin, although believed to be

charge density waves with stripe or bubble phases [215, 216, 217], remains an open question.

This section provides nonlinear transport data in this regime, showing rich structures that

are present only at high bias. These data provide new information on these anisotropic and

reentrant states, and should help resolve their precise nature.

The two Hall bars studied in this section are fabricated on the GaAs/AlGaAs two-

dimensional electron gas (2DEG) wafers also used in the previous two sections and in Chs. 7

and 8. These two Hall bars are oriented perpendicular to each other with respect to the

GaAs/AlGaAs crystal direction. The bulk Hall resistance, RXY, and the bulk longitudinal

resistance, RXX, are the quantities of interest here.

We first reproduce the results of Refs. [213, 214] in Fig. 9.11. The bulk transport, RXX

and RXY as a function of B, covering filling factors νB = 2 to 8, are studied as a function

of crystal direction and temperature. The data in Figs. 9.11(a) and (b) [(c) and (d)] are

measured from a Hall bar oriented with crystal direction 1 (2), which are perpendicular

to each other. The blue traces in Fig. 9.11 are obtained at the fridge base temperature of

13 mK, while the red traces are obtained when the fridge were just cooled down and the

electrons are not yet thermalized to the fridge. Comparing the RXX data near 5/2 to those

in Ref. [176], the electron temperature for the red traces should be around 40 − 50 mK.

Similar to previous results [213, 214], reentrant features in RXY are clearly visible for N ≥ 2

(νB ≥ 4)when the electrons are warm, but they merge with the main plateaus and RXY show

only simple steps when the electrons are cold. In the meantime, RXX show large anisotropic

behaviors near 9/2, 11/2, etc.: when the electrons are cold, RXX almost vanishes for crystal

128

Page 143: Waves, Particles, and Interactions in Reduced Dimensions

0.04

0.03

0.02

0.01

RX

X [

h / e

2 ]

Warm Cold

(a)

0.001/2

1/3

1/4

1/51/61/8

RX

Y [ h

/ e2 ]

5432 B [ T ]

Warm Cold

(b)

0.08

0.06

0.04

0.02

0.00

Warm Cold

(c)

1/2

1/3

1/4

1/51/61/8

5432 B [ T ]

Warm Cold

(d)

Crystal Direction 1 Crystal Direction 2

Figure 9.11: RXX (a,c) and RXY (b,d) as a function of B, at the base temperature (blue)and an elevated temperature (red). The data in (a,b) are measured from a Hall bar that isoriented perpendicular to the Hall bar used for obtaining the data in (c,d).

direction 1, but shows sharp peaks for crystal direction 2; when the electrons are warm, the

RXX data exhibit side peaks around their minima near half fillings for crystal direction 1,

but they show peaks at half fillings for crystal direction 2.

We move on now to nonlinear transport measurements. Figure 9.12 show RXX and RXY

as a function of dc I and B, at filling factors νB = 4 to 9, for crystal direction 1 and cold

electrons. Although at zero bias, RXX almost vanishes, and RXY only show simple steps,

several intriguing features appear in both RXX and RXY at high bias. Around 11/2, for

example, four leaf-shaped regions of non-zero RXX appear in each bias direction. A similar

behavior is seen near 9/23, but there are only two such regions, which are getting closer

at lower B, around 13/2, 15/2, and 17/2. The boundaries of these non-zero RXX regions

correspond to sharp transitions in RXY. Furthermore, there are small and quite regularly

3There is possibly another leaf (in each bias direction) that the measurement did notcover at higher B.

129

Page 144: Waves, Particles, and Interactions in Reduced Dimensions

1/4

1/5

1/6

1/7

1/8

RX

Y [ h / e2 ]

2.42.22.01.81.61.4 B [ T ]-100

-50

0

50

100

I [ n

A ]

0.26

0.24

0.22

0.20

0.18

0.16

0.14

0.12

RX

Y [ h / e2 ]

-100

-50

0

50

100

I [ n

A ]

0.025

0.020

0.015

0.010

0.005

0.000

RX

X [ h / e2 ]

8 7 6 5 4

(a)

(b)

Figure 9.12: RXX (a) and RXY (b) as a function of I and B at the base temperature, withcrystal direction 1.

spaced bright spots in RXX along these boundaries, and corresponding features are also

present in RXY.

So far, the nonlinear transport data are obtained only for crystal direction 1 and cold

electrons, and it would be very interesting to study the temperature- and crystal-direction-

dependence of such data. It would be equally interesting to extend the bias range since the

observed features seem to extend far beyond the bias range of study. At this point, I am

not aware of any interpretations for these features, but my guess is that they are somehow

related to the anisotropic and reentrant states for N ≥ 2 LL’s. Considering the amount of

information nonlinear transport gives at these high LL’s, similar measurements for N = 0

or N = 1 LL’s, possibly with fractional states, may give surprising results as well.

130

Page 145: Waves, Particles, and Interactions in Reduced Dimensions

Appendix A

Fridge Wiring: Thermal Anchoringand Filtering

Well designed and implemented electrical wiring is an essential part of low-temperature

transport experiments, yet wiring is often quite tricky as well, because:

• when transmitting electrical signals to and from the sample, the wires can also bring

down significant amount of heat to the lowest temperature stage of the fridge and the

sample, raising base fridge and electron temperatures;

• long and exposed dc wires are like antennas, picking up radiation and unwanted noise,

which would degrade the signal, and also heat up the electrons.

The usual approach [218] for dc wiring used in the Marcus lab has been to use long re-

sistive wires, eg. twisted-pair constantan (55% Cu, 45% Ni) wires from Oxford Instruments,

thermally anchored at each stage of the fridge with copper posts and GE varnish. At the

lowest-temperature stage, 3He pot or mixing chamber, a bank of resistors are inserted to

the lines. The resistors and stray capacitance from the wires to ground form simple RC

filters and effectively filter out most of the radiation from room temperature.

This simple and robust approach works quite well for bare dc lines down to 50 mK

or so, yet for the noise measurement circuit in the 3He fridge, and for the dc lines in the

131

Page 146: Waves, Particles, and Interactions in Reduced Dimensions

Figure A.1: Photograph of a bank of simple RC filters

Microsoft dilution fridge that we aim at an electron temperature of 20 mK, additional

thermal anchoring and filtering are needed. Specifically, for the noise measurement circuit,

each gate line needs to be heavily filtered to prevent the thermal noise from dc wires to

couple to the 2DEG at 2 MHz, for which we used simple RC filters. We have also used

sapphire heat sinks and circuits boards for both the noise measurement circuits and the dc

lines in the Microsoft dilution refrigerator. In addition, we have used Mini-Cuicuits VLFX

filters in the Microsoft fridge with great success. Recently, we have replaced bare dc wires

with thermocoax lines in the Microsoft fridge. I will describe each of these techniques in

detail below.

A.1 Simple RC filters

For the noise measurement operating near 2 MHz, one source of extraneous noise is from

the thermal noise of gate lines, coupled capacitively to the 2DEG. Here, we simply used RC

filters, with resistance R = 5 kΩ, and capacitance C = 22 nF, giving a 3 dB cut-off frequency

of 1.4 kHz. Due to the large number of lines that need to be filtered, we have designed a

132

Page 147: Waves, Particles, and Interactions in Reduced Dimensions

compact bank of RC filters, with four layers of 0805-sized surface mount components on

three boards stacked together, and soldered to two 37-pin D-sub connectors, as shown in

Fig. A.1.

A.2 Sapphire heat sinks and circuit boards

The inner conductor of the coaxial lines used for the noise circuit is insulated from the outer

conductor in teflon, and has to be thermalized before it reaches the sample. As described

in Ch. 2, we insert thin meandering lines evaporated on sapphire boards to heat sink them

without adding much resistance to the lines. In addition, all circuit boards inside the 3He

fridge are made on sapphire boards, for maximal thermalization of coaxial lines at each

stage. Given the success with sapphire heat sinks used for the noise circuit, we also make

use of it for heat sinking all dc lines in the Microsoft fridge.

Sapphire has been chosen because it has among the best thermal conductivity for insu-

lators at low temperatures [92], second only to quartz below 10 K, yet it is a lot less brittle

than quartz, making it much easier to work with than quartz. Unlike conventional circuit

boards, often made commercially, we fabricate sapphire circuit boards in house starting

with blank sapphire pieces. The latest recipe of fabrication is detailed below:

1. Blank sapphire pieces are first cut into the desired dimensions with Automatic Dicing

Saw (CNS facility at Harvard) at the lowest cutting speed of 0.1 mm/s;

2. Then 30/300 nm of Cr/Au or Ti/Au are evaporated on the top surface, and annealed

in the Rapid Thermal Annealer for better adhesion and releasing any strain in the

film. We have simply used the same annealing recipe as that for making ohmics. The

ideal temperature and annealing time have not been investigated, and is probably not

133

Page 148: Waves, Particles, and Interactions in Reduced Dimensions

so important anyway;

3. There are several ways to pattern the traces. The simplest one, which works well for

large features, is to just draw traces with a Sharpie. For better control and smaller

features, one can print the designed pattern on a special toner transfer paper (Pulsar

toner transfer mask), and then iron the pattern onto the surface. For very fine features

and best results, one can use photo lithography to define the patterns;

4. After the traces are covered with masks (from Sharpie ink to photo resist), the rest

of the metal films are etched in Au and Cr (or Ti) etchants. After etching, the masks

can be easily washed off in acetone;

5. Cr/Au or Ti/Au (30/300 nm) are then evaporated on the back surface;

6. Surface-mount components are soldered to the circuit traces using In solder at the

lowest soldering temperature (350 F). Soldering these tiny components with In solder

on gold traces are tricky, because: In can easily creep over and create shorts between

traces; In does not stick to gold surfaces very well, and may develop breaks over

time; higher temperatures can readily peel off the gold traces on sapphire. Therefore,

extreme care must be taken and a lot of practice are needed. Sometimes when a break

does develop, which occurs quite often, it can be fixed with a tiny touch of silver paint

or Transene silver bond. An easier and more robust recipe is surely desired, but has

not been developed at this point;

7. Finally, the back surface are soldered to copper braids, or silver epoxied to the desired

surface, for best thermalization of the sapphire boards.

Some examples of sapphire circuit boards and heat sinks are shown in Fig. A.2.

134

Page 149: Waves, Particles, and Interactions in Reduced Dimensions

(a) (d)

(b)

(c)

Figure A.2: Photographs of sapphire circuit boards and heat sinks: (a) the HEMT circuitboard, (b) the SINK board, (c) the RES circuit board, and (d) the sapphire heat sink fordc lines used in the Microsoft fridge, with the pattern defined by photo lithography.

A.3 Mini-circuit VLFX filters

When pushing to the lowest electron temperature of a dilution refrigerator, say below 30 mK,

the main source of heating is high-frequency (GHz or higher) radiation carried by dc lines

from higher-temperature stages. Simple RLC filters usually stop working at these fre-

quencies due to their parasitic capacitance and inductance. Metal powder filters are often

used [219, 220] for this purpose, but they are usually home-made and sometimes tend to

break over time after a few thermal cycles. Another type of filters, called Frossati filters, are

effective, robust, and commercially available from Leiden Cryogenics, yet they cost a few

hundred dollar per filter, making the cost prohibitively high. Another approach is to use a

special type of coax from Thermocoax Co., which will be described in the next section.

Here, we have used a type of low-cost ($40 a piece) commercially available LC filter

that works very well for blocking radiation: Mini-Circuits VLFX filters. These VLFX filters

feature 21 sections of filtering, guaranteed 40 dB of isolation up to 20 GHz, and temperature

stable structure. Shown in Fig. A.3 are 18 Mini-circuits VLFX-80 filters installed with the

cold finger. These filters, together with the sapphire heat sinks described in the previous

section, have allowed the electrons to be cooled to below 20 mK, as inferred from the bulk

transport data shown in Fig. A.4, in the Microsoft fridge at a fridge temperature of 13 mK.

135

Page 150: Waves, Particles, and Interactions in Reduced Dimensions

(a) (b)

Figure A.3: 18 Mini-circuits VLFX-80 filters assembled with the cold finger: (a) side view,(b) top view.

1/2

1/3

2/5

3/7

3/8

RX

Y [

h / e

2 ]

RXY

0.10

0.08

0.06

0.04

0.02

0.00

RX

X [

h / e

2 ]

5.04.84.64.44.24.03.8Bperp [ T ]

RXX

Figure A.4: Quantum Hall bulk transport, showing well developed 7/3, 5/2, 8/3 states,and reentrant integer features, consistent with an electron temperature below 20 mK whencompared to the data in Ref. [176].

A.4 Thermocoax cables

Recently, we have decided to try a special type of coaxial cable [221] from Thermocoax

Co., to replace the dc lines in the Microsoft fridge. The cable used has an outer diameter

of 0.5 mm, NiCr (80/20) alloy inner conductor, stainless steel outer conductor, and highly

compacted magnesium oxide powder as the dielectric material. Its mechanism for blocking

high-frequency radiation is the same as that for metal powder filters: skin-effect damping in

136

Page 151: Waves, Particles, and Interactions in Reduced Dimensions

the very large surface area of the powder. The additional benefits include that the filtering

is built into the cable itself, thus no additional filters are needed, and the coaxial geometry

prevents pickups that bare dc lines are prone to.

Soldering connectors to thermocoax cables need special attention because the stainless

steel outer conductor is hard to solder to without stainless steel flux, and the powder

dielectric tend to absorb moisture and flux, creating dc voltage or MΩ leak between the

inner and outer conductors. After doing a lot of practice and learning from errors, Doug

and I have developed a specific set of steps for soldering connectors to thermocoax cables.

The following steps are for soldering SMP connectors, but they are similar for other types

of connectors as well:

1. Use a 30-gauge wire-stripper to cut the outer connector roughly 5 mm from the end;

rotate the cutter in a circle to be sure the outer is completely scored. Use a pair of

pliers to grab the section of cable between the cut and the end. Holding the cable

with two fingers just on the other side of the cut, use the pliers to bend the end back

and forth until you feel the outer completely breaks;

2. Squeeze the end section with the pliers to flatten it; repeat at a 90-degree angle to

flatten it the other way. Doing this, you should notice some white powder coming

out. Squeeze the end just enough to make it roughly round again, and try to pull it

straight off with the pliers. If this fails, repeat the previous step and then try again;

3. Heat the exposed end for a minute or so with a heat gun above 100 C to evaporate

any water that has been absorbed. Immediately seal the exposed dielectric using some

epoxy, eg. the Huber-Suhner blue epoxy. Test for shorts using a DMM, and allow the

epoxy to completely set before proceeding;

137

Page 152: Waves, Particles, and Interactions in Reduced Dimensions

4. Use a sharp blade to gently clean the inner conductor. Hold the center pin of the

SMP connector with a little clamp or vice, heat and let a small amount of solder to

flow into the small hole at the back end of the center pin. Make sure the outer surface

of the center pin is clean. With the soldering iron still heating the center pin, push

the exposed inner conductor into the small hole; take off the iron, and let the solder

to cool and get hardened.

5. Hold the cable vertically with the center pin pointing upwards. Use a Q-tip to apply

a small amount of stainless steel flux to the outer connector ∼ 2 mm below the cut.

Make sure that the flux is not reaching the exposed powder. Apply some flux to the

tip of the solder wire as well, then tin the outer conductor. Clean the outer with

Q-tips and IPA.

6. Push the cable and center pin into the main part of the SMP connector, and solder

the outer conductor to the SMP connector outer. This should complete the process,

and check again with a DMM to see if there is any voltage or MΩ leak between the

inner and outer.

Shown in Fig. A.5 are 28 Thermocoax cables, terminated with SMP connectors and

assembled with Doug’s new silver colder finger on the Microsoft fridge. At the time of

writing, initial tests suggest that the electrons are getting close to the fridge temperature,

yet the fridge base temperature is higher than before, at ∼ 25 mK. The specific reasons for

the higher base temperature is still under investigation.

138

Page 153: Waves, Particles, and Interactions in Reduced Dimensions

Figure A.5: Thermalcoax cables terminated with SMP connectors and assembled on theMicrosoft fridge

139

Page 154: Waves, Particles, and Interactions in Reduced Dimensions

Appendix B

Igor implementation of virtualDACs

As Charlie once boasted in one group meeting, the use of “Igor Pro” has been a secret

weapon of the Marcus lab. Indeed, this little known yet powerful program has been the

primary platform for almost all computer related tasks for experiments, from controlling

instruments, data acquisition and processing, to making illustrations for final publication.

Even some theoretical simulations done in this thesis have made use of Igor Pro.

While most people in the group are using the ”Alex Igor Suite”, developed by Alex

Johnson [218], for controlling experiments and acquiring data, the ”Noise Team”, and later

the “5/2 Team” (excluding Jeff) have been using our own set of Igor routines. Many features

are shared between Alex’s procedures and ours, including universal interfaces of acquiring

various types of data and controlling independent parameters, universal 1d and 2d sweeps,

and automatically saving data waves and related parameters, etc. One feature that has

been greatly enhanced in our procedures is what we call “virtual DACs”, which I would like

to share here in my thesis. It is a framework that allows easy definition and simultaneous

control of a linear combination of multiple independent parameters, and is designed to be

140

Page 155: Waves, Particles, and Interactions in Reduced Dimensions

AO6 DacUnitsPoint DacNames Defined-inRow

Used-inRow

0 Vbias 0.000 mV 0 -1

1 Vc -0.000 mV -1 -1

2 Vcfine 0.000 mV -1 -1

3 Vrt 0.000 mV 3 -1

4 Vrb -0.000 mV 3 -1

5 Vlt 0.000 mV 1 -1

6 Vmc 0.000 mV -1 -1

7 Vstill 0.000 mV -1 -1

8 Vlb -0.000 mV 1 -1

9 Vt 0.000 mV 2 -1

10 Vb -0.000 mV 2 -1

11 Vl1 0.000 mV 5 -1

12 Vl2 0.000 mV 6 -1

13 BVout 0.000 mV 0 -1

14 Vr1 -0.000 mV 5 -1

15 Vr2 -0.000 mV 6 -1

16 Bperp 0.000 mT -1 -3

17 Vls -0.000 mV 4 1

18 Vla 0.000 mV -1 1

19 Vs 0.000 mV -1 2

20 Va 0.000 mV -1 2

21 Vrs 0.000 mV 4 3

22 Vra 0.000 mV -1 3

23 Vss 0.000 mV -1 4

24 Vsa -0.000 mV -1 4

25 V1s 0.000 mV -1 5

26 V1a 0.000 mV -1 5

27 V2s -0.000 mV -1 6

28 V2a 0.000 mV -1 6

29 DeltaB 0.000 mT -1 0

30 Idc 0.000 nA -1 0

Figure B.1: Channel definition table

fully compatible with any previously written codes.1

B.1 Igor implementation of virtual DACs

For historical reasons, the digital-to-analog converter (DAC) outputs are stored in a wave

named “AO6” [see Fig. B.1], and are controlled by the function setdac(chan,value). The

idea is to overwrite this function, and define virtual DAC channels to allow controlling a

combination of real or virtual channels. Another type, called special channel, can also be

defined by providing the names of “set” and “get” functions in the text waves DACSetFunc-

1One only needs to change setdac() to sd() and rampdac() to rd() in any old procedures.

141

Page 156: Waves, Particles, and Interactions in Reduced Dimensions

tion and DACGetFunction, for the ultimate flexibility. Channel 16, Bperp, which controls

the perpendicular magnetic field, is one example of such special channel.

Each pair of new virtual DAC channels, VChan1 and VChan2, is defined by a linear

combination of two other channels, RChan1 and RChan2, as follows:

V Chan1 = (RChan1−ROfs1)× V 1R1 + (RChan2−ROfs2)× V 1R2

V Chan2 = (RChan1−ROfs1)× V 2R1 + (RChan2−ROfs2)× V 2R2

RChan1 = V Chan1×R1V 1 + V Chan2×R1V 2 +ROfs1

RChan2 = V Chan1×R2V 1 + V Chan2×R2V 2 +ROfs2

where, V1R1, V1R2, V2R1, V2R2, ROfs1, and ROfs2 are user-given coefficients, and

are used to calculate values of R1V1, R1V2, R2V1, and R2V2. All coefficients for defining

one pair of virtual channels are stored in one row of the 2d wave, VDACparams, as shown

in Fig. B.2.

After initializing the procedures by running SetupVDACs(), and setting up default pa-

rameters by SetDefaultParams(), a pair of new virtual DACs can be defined by the function

DefineNewVDAC(VChan1,VChan2,RChan1,RChan2,VChan1Name,VChan2Name). By de-

fault, VChan1 is defined as the average of RChan1 and RChan2, and VChan2 is the differ-

ence between RChan1 and RChan2, but one can easily assign any conversion coefficients by

using SetNewCoef(row,V1R1,V1R2,V2R1,V2R2,ROfs1,ROfs2) for the pair of virtual DACs

defined in the row “row” of the wave VDACparams. To remove the definition of a pair,

simply use RemoveRow(row).

These definitions can be cascaded so that simultaneous control of many channels can be

142

Page 157: Waves, Particles, and Interactions in Reduced Dimensions

Row x y VirtualCh1 VirtualCh2 RealCh1 RealCh2 V1R1 V1R2 V2R1 V2R2 Real ofs1 Real ofs2 R1V1 R1V2 R2V1 R2V20 30 29 0 13 0.01 0 0 -0.002316 0 0 100 -0 -0 -431.648

1 17 18 5 8 0.5 0.5 1 -1 0 0 1 0.5 1 -0.5

2 19 20 9 10 0.5 0.5 1 -1 0 0 1 0.5 1 -0.5

3 21 22 3 4 0.5 0.5 1 -1 0 0 1 0.5 1 -0.5

4 23 24 17 21 0.5 0.5 1 -1 0 0 1 0.5 1 -0.5

5 25 26 11 14 0.5 0.5 1 -1 0 0 1 0.5 1 -0.5

6 27 28 12 15 0.5 0.5 1 -1 0 0 1 0.5 1 -0.5

7 -1 -1 -1 -1 0.5 0.5 1 -1 0 0 1 0.5 1 -0.5

Figure B.2: Virtual DACs parameters table

easily achieved. For example, as shown in Figs. B.1 and B.2, the virtual channels Vls/Vla

(Vrs/Vra) are defined from Vlt/Vlb (Vrt/Vrb), and then Vls/Vrs are used to define Vss/Vsa.

All interdependent channels are updated automatically when the value of any channel is

changed. For example, when Vls is changed, the channels Vlt/Vlb as well as Vss/Vsa are

updated.

Here, I provide the source codes for these procedures:

/////////////////////////////////////////////////////////////////////////////////////

//Functions and operations for working with user defined virtual DAC channels

/////////////////////////////////////////////////////////////////////////////////////

constant VDAC_VCHAN1 = 0 //Virtual DAC channel 1

constant VDAC_VCHAN2 = 1 //Virtual DAC channel 2

constant VDAC_RCHAN1 = 2 //Real DAC channel 1

constant VDAC_RCHAN2 = 3 //Real DAC channel 2

constant VDAC_V1R1 = 4 //Real 1 to virtual 1 conversion coefficient

constant VDAC_V1R2 = 5 //Real 2 to virtual 1 conversion coefficient

constant VDAC_V2R1 = 6 //Real 1 to virtual 2 conversion coefficient

constant VDAC_V2R2 = 7 //Real 2 to virtual 2 conversion coefficient

constant VDAC_ROFS1 = 8 //Real DAC channel 1 offset

constant VDAC_ROFS2 = 9 //Real DAC channel 2 offset

constant VDAC_R1V1 = 10 //Virtual 1 to real 1 conversion coefficient

constant VDAC_R1V2 = 11 //Virtual 2 to real 1 conversion coefficient

constant VDAC_R2V1 = 12 //Virtual 1 to real 2 conversion coefficient

constant VDAC_R2V2 = 13 //Virtual 2 to real 2 conversion coefficient

//Conversion rules:

//VChan1 = (RChan1 - ROfs1) * V1R1 + (RChan2 - ROfs2) * V1R2

//VChan2 = (RChan1 - ROfs1) * V2R1 + (RChan2 - ROfs2) * V2R2

//RChan1 = VChan1 * R1V1 + VChan2 * R1V2 + ROFS1

//RChan2 = VChan1 * R2V1 + VChan2 * R2V2 + ROFS2

//R1V1, R1V2, R2V1 and R2V2 are calculated from V1R1, V1R2, V2R1 and V2R2

//Setup waves for defining and using virtual DACs

//Each row of VDACparams defines a pair of virtual DACs from a pair of real or virtual DACs

//UsedinRow specifies which row the channel is used to define virtual DACs; if the channel

// is not used for defining virtual channels, UsedinRow = -1.

//DefinedinRow specifies which row the virtual channel is defined; if the channel is a real

// DAC, DefinedinRow = -1; if the channel is a real special channel, DefinedinRow = -3; if

// the channel is not defined, DefinedinRow = -2

function SetupVDACs ( )

variable/G NumRowVDAC = 32

make/d/o/n=(NumRowVDAC,14) VDACparams //Each row defines a conversion pair

SetDimLabel 1, VDAC_VCHAN1, $"VirtualCh1", VDACparams

143

Page 158: Waves, Particles, and Interactions in Reduced Dimensions

SetDimLabel 1, VDAC_VCHAN2, $"VirtualCh2", VDACparams

SetDimLabel 1, VDAC_RCHAN1, $"RealCh1", VDACparams

SetDimLabel 1, VDAC_RCHAN2, $"RealCh2", VDACparams

SetDimLabel 1, VDAC_V1R1, $"V1R1", VDACparams

SetDimLabel 1, VDAC_V1R2, $"V1R2", VDACparams

SetDimLabel 1, VDAC_V2R1, $"V2R1", VDACparams

SetDimLabel 1, VDAC_V2R2, $"V2R2", VDACparams

SetDimLabel 1, VDAC_ROFS1, $"Real ofs1", VDACparams

SetDimLabel 1, VDAC_ROFS2, $"Real ofs2", VDACparams

SetDimLabel 1, VDAC_R1V1, $"R1V1", VDACparams

SetDimLabel 1, VDAC_R1V2, $"R1V2", VDACparams

SetDimLabel 1, VDAC_R2V1, $"R2V1", VDACparams

SetDimLabel 1, VDAC_R2V2, $"R2V2", VDACparams

NVAR NumChan //Number of DAC Channels, defined in ingotDAC.ipf

make/o/n=(NumChan) UsedinRow, DefinedinRow

make/o/n=(NumChan)/t DACSetFunction,DACGetFunction //Only used for real special channels

end

//Set Default Parameters

function SetDefaultParams()

wave VDACparams, UsedinRow, DefinedinRow

VDACparams[][VDAC_VCHAN1] = -1 //Undefined if Chan < 0

VDACparams[][VDAC_VCHAN2] = -1

VDACparams[][VDAC_RCHAN1] = -1

VDACparams[][VDAC_RCHAN2] = -1

VDACparams[][VDAC_V1R1] = 1/2

VDACparams[][VDAC_V1R2] = 1/2

VDACparams[][VDAC_V2R1] = 1

VDACparams[][VDAC_V2R2] = -1

VDACparams[][VDAC_ROFS1] = 0

VDACparams[][VDAC_ROFS2] = 0

CalcBackConversionCoef ( )

UsedinRow = -1 //Not used

DefinedinRow = -2 //Not defined

NVAR NumRealChan

DefinedinRow[0,NumRealChan-1] = -1 //Real channels

end

//Return DAC number for a given DAC name

function Name2Num(channame)

string channame

wave/T DacNames

variable i=0

do

if(stringmatch(DacNames[i],channame))

return i

endif

i+=1

while (i < numpnts(DacNames))

print "ERROR: Unknown channel",channame

Abort

end

//Return 1 if row is used; otherwise, return 0

function RowUsed( row )

variable row

NVAR NumRowVDAC

if( row<0 || row > NumRowVDAC-1 )

printf "Warning! Row %g out of range.\r", row

return 0

endif

144

Page 159: Waves, Particles, and Interactions in Reduced Dimensions

wave VDACparams

return ( VDACparams[row][VDAC_VCHAN1] > 0 && VDACparams[row][VDAC_VCHAN2] > 0 )

end

//Return 1 if chan is defined; otherwise, return 0

function ChannelDefined ( chan )

variable chan

wave DefinedinRow

NVAR NumChan

if( chan<0 || chan > NumChan-1 )

printf "Warning! Channel %g out of range.\r", chan

return 0

endif

return ( DefinedinRow[chan] != -2 )

end

//Return 1 if chan is used to define virtual DACs; otherwise, return 0

function ChannelUsed ( chan )

variable chan

wave UsedinRow

NVAR NumChan

if( chan<0 || chan > NumChan-1 )

printf "Warning! Channel %g out of range.\r", chan

return 0

endif

return ( UsedinRow[chan] > -0.5 )

end

//Check if chan is a real or special channel

function ChannelReal ( chan )

variable chan

return ( ChannelRealDAC ( chan ) || ChannelRealSpecial ( chan ))

end

//Check if chan is a real DAC channel

function ChannelRealDAC ( chan )

variable chan

wave DefinedinRow

NVAR NumChan

if( chan<0 || chan > NumChan-1 )

printf "Warning! Channel %g out of range.\r", chan

return 0

endif

return ( DefinedinRow[chan] == -1)

end

//Check if chan is a special channel

function ChannelRealSpecial ( chan )

variable chan

wave DefinedinRow

NVAR NumChan

if( chan<0 || chan > NumChan-1 )

printf "Warning! Channel %g out of range.\r", chan

return 0

endif

return ( DefinedinRow[chan] == -3)

end

//Define a new pair of virtual DACs

//If successful, return the row number that defines them

//If unsuccessful, a negative error code is returned

function DefineNewVDAC(Vchan1,Vchan2,Rchan1,Rchan2,Vchan1Name,Vchan2Name)

variable Vchan1,Vchan2,Rchan1,Rchan2

string Vchan1Name,Vchan2Name

wave VDACparams

145

Page 160: Waves, Particles, and Interactions in Reduced Dimensions

variable numrows = dimsize(VDACparams,0)

//Make sure Rchan1, Rchan2 are already defined

if( ! ChannelDefined( Rchan1 ) || ! ChannelDefined( Rchan2 ) )

printf "Error! Rchan1 %g or Rchan2 %g are not defined.\r", Rchan1, Rchan2

return -1

endif

//Make sure Rchan1, Rchan2 are not used to define other virtual DACs

if( ChannelUsed( Rchan1 ) || ChannelUsed( Rchan2 ) )

printf "Error! Rchan1 %g or Rchan2 %g are used to define other virtual DACs.\r", Rchan1, Rchan2

return -1

endif

//Make sure Vchan1, Vchan2 are not defined

if( ChannelDefined( Vchan1 ) || ChannelDefined( Vchan2 ) )

printf "Error! Vchan1 %g or Vchan2 %g are already defined.\r", Vchan1, Vchan2

return -2

endif

wave/T DacNames

wave UsedinRow, DefinedinRow

variable row

for(row = 0; row < numrows; row += 1)

if( ! RowUsed( row ) )

VDACparams[row][VDAC_VCHAN1] = Vchan1

VDACparams[row][VDAC_VCHAN2] = Vchan2

VDACparams[row][VDAC_RCHAN1] = Rchan1

VDACparams[row][VDAC_RCHAN2] = Rchan2

UpdateVirtualDACs ( row )

UsedinRow[Rchan1] = row

UsedinRow[Rchan2] = row

DefinedinRow[Vchan1] = row

DefinedinRow[Vchan2] = row

DacNames[Vchan1] = Vchan1Name

DacNames[Vchan2] = Vchan2Name

printf "Success! Vchan1 %g and Vchan2 %g have been defined from Rchan1 %g and

Rchan2 %g in row %g.\r", Vchan1, Vchan2, Rchan1, Rchan2, row

return row

endif

endfor

printf "Error! Cannot find an unused row. Increase the number of rows in SetupVDACs( ).\r"

return -3

end

//Remove a row that defines a pair of Virtual DACs

function RemoveRow( row )

variable row

wave/T DacNames

wave VDACparams, UsedinRow, DefinedinRow, AO6

if( ! RowUsed( row ) ) //Check if the row is used

printf "Warning! Row %g is not used, no need to remove.\r", row

return -1

endif

//Check if the virtual channels defined in this row are used elsewhere

if( ChannelUsed( VDACparams[row][VDAC_VCHAN1] ) || ChannelUsed( VDACparams[row][VDAC_VCHAN1] ) )

printf "Failure! The virtual channels defined in this row %g are used to define other

virtual channels.\r", row

return -2

endif

UsedinRow[ VDACparams[row][VDAC_RCHAN1] ] = -1 //Rchan1 is no longer used

UsedinRow[ VDACparams[row][VDAC_RCHAN2] ] = -1 //Rchan2 is no longer used

146

Page 161: Waves, Particles, and Interactions in Reduced Dimensions

DefinedinRow[ VDACparams[row][VDAC_VCHAN1] ] = -2 //Vchan1 is no longer defined

DefinedinRow[ VDACparams[row][VDAC_VCHAN2] ] = -2 //Vchan2 is no longer defined

DacNames[ VDACparams[row][VDAC_VCHAN1] ] = "" //Remove Vchan1 label

DacNames[ VDACparams[row][VDAC_VCHAN2] ] = "" //Remove Vchan2 label

AO6[ VDACparams[row][VDAC_VCHAN1] ] = nan

AO6[ VDACparams[row][VDAC_VCHAN2] ] = nan

VDACparams[row][VDAC_VCHAN1] = -1 //row is no longer used

VDACparams[row][VDAC_VCHAN2] = -1

VDACparams[row][VDAC_RCHAN1] = -1

VDACparams[row][VDAC_RCHAN2] = -1

end

//Set new coefficients to the pair of virtual DACs defined in row

Function SetNewCoef( row, v1r1, v1r2, v2r1, v2r2, ROfs1, ROfs2 )

variable row, v1r1, v1r2, v2r1, v2r2, ROfs1, ROfs2

if( ! RowUsed( row ) )

printf "Warning! The row %g does not define any virtual DACs. Nothing is updated.\r", row

return -1;

endif

wave VDACparams

VDACparams[row][VDAC_V1R1] = v1r1

VDACparams[row][VDAC_V1R2] = v1r2

VDACparams[row][VDAC_V2R1] = v2r1

VDACparams[row][VDAC_V2R2] = v2r2

VDACparams[row][VDAC_ROFS1] = ROfs1

VDACparams[row][VDAC_ROFS2] = ROfs2

CalcBackConversionCoef ( )

UpdateAllVirtualDACs ( )

end

//Calculate R1V1, R1V2, R2V1 and R2V2 from V1R1, V1R2, V2R1 and V2R2

function CalcBackConversionCoef ( )

wave VDACparams

variable numrows = dimsize(VDACparams,0)

variable row, v1r1, v1r2, v2r1, v2r2, det

for(row=0; row<numrows; row+=1)

v1r1 = VDACparams[row][VDAC_V1R1]

v1r2 = VDACparams[row][VDAC_V1R2]

v2r1 = VDACparams[row][VDAC_V2R1]

v2r2 = VDACparams[row][VDAC_V2R2]

det = v1r1 * v2r2 - v1r2 * v2r1

VDACparams[row][VDAC_R1V1] = v2r2 / det

VDACparams[row][VDAC_R1V2] = - v1r2 / det

VDACparams[row][VDAC_R2V1] = - v2r1 / det

VDACparams[row][VDAC_R2V2] = v1r1 / det

endfor

end

//Update Virtual DAC values

function UpdateVirtualDACs ( row )

variable row

wave VDACparams

wave AO6, UsedinRow

if( ! RowUsed( row ) )

printf "Warning! The row %g does not define any virtual DACs. Nothing is updated.\r", row

return -1;

endif

variable vchan1,vchan2,rchan1,rchan2

vchan1 = VDACparams[row][VDAC_VCHAN1]

147

Page 162: Waves, Particles, and Interactions in Reduced Dimensions

vchan2 = VDACparams[row][VDAC_VCHAN2]

rchan1 = VDACparams[row][VDAC_RCHAN1]

rchan2 = VDACparams[row][VDAC_RCHAN2]

AO6[vchan1] = ( AO6[rchan1] - VDACparams[row][VDAC_ROFS1] ) * VDACparams[row][VDAC_V1R1]

+ ( AO6[rchan2] - VDACparams[row][VDAC_ROFS2] ) * VDACparams[row][VDAC_V1R2]

AO6[vchan2] = ( AO6[rchan1] - VDACparams[row][VDAC_ROFS1] ) * VDACparams[row][VDAC_V2R1]

+ ( AO6[rchan2] - VDACparams[row][VDAC_ROFS2] ) * VDACparams[row][VDAC_V2R2]

//If VChan1 is used to define other virtual channels, need to update them as well

if( ChannelUsed ( vchan1 ) )

UpdateVirtualDACs ( UsedinRow[vchan1] )

endif

//If VChan2 is used to define other virtual channels, need to update them as well

if( ChannelUsed ( vchan2 ) )

UpdateVirtualDACs ( UsedinRow[vchan2] )

endif

end

function UpdateAllVirtualDACs ( )

wave VDACparams

variable numrows = dimsize(VDACparams,0)

variable row

for(row=0; row<numrows; row+=1)

if( RowUsed( row ) )

UpdateVirtualDACs ( row )

endif

endfor

end

function UpdateRealSpecialChan( chan )

variable chan

wave AO6

wave/T DACGetFunction

string cmdstr

variable/g tempvar

if ( ChannelRealSpecial ( chan ) ) //If the channel is a real special channel

cmdstr = "tempvar = " + DACGetFunction[chan] + "()"

Execute cmdstr

AO6[chan] = tempvar

return 1

else

return 0

endif

end

//V1 is the real part, V2 is the imaginary part

function/C RealtoVirtual ( row, r1, r2 )

Variable row, r1, r2

wave VDACparams, AO6

if( ! RowUsed ( row ) ) //If row is not used

printf "Error! Row %g is not used.\r", row

return nan

endif

variable v1, v2

v1 = ( r1 - VDACparams[row][VDAC_ROFS1] ) * VDACparams[row][VDAC_V1R1]

+ ( r2 - VDACparams[row][VDAC_ROFS2] ) * VDACparams[row][VDAC_V1R2]

v2 = ( r1 - VDACparams[row][VDAC_ROFS1] ) * VDACparams[row][VDAC_V2R1]

+ ( r2 - VDACparams[row][VDAC_ROFS2] ) * VDACparams[row][VDAC_V2R2]

return cmplx(v1,v2)

end

//R1 is the real part, R2 is the imaginary part

function/C VirtualToReal ( row, v1, v2 )

148

Page 163: Waves, Particles, and Interactions in Reduced Dimensions

Variable row, v1, v2

wave VDACparams, AO6

if( ! RowUsed ( row ) ) //If row is not used

printf "Error! Row %g is not used.\r", row

return nan

endif

variable r1, r2

r1 = v1 * VDACparams[row][VDAC_R1V1] + v2 * VDACparams[row][VDAC_R1V2] + VDACparams[row][VDAC_ROFS1]

r2 = v1 * VDACparams[row][VDAC_R2V1] + v2 * VDACparams[row][VDAC_R2V2] + VDACparams[row][VDAC_ROFS2]

return cmplx(r1,r2)

end

//Read virtual DAC values, using GetDAC

function ReadVDac(chan)

variable chan

wave DefinedinRow, VDACparams, AO6

variable row = DefinedinRow[chan]

if ( ChannelRealSpecial ( chan ) )

return AO6[chan]

elseif ( ChannelRealDAC ( chan ) )

return GetDAC(chan)

else

variable rchan1 = VDACparams[row][VDAC_RCHAN1]

variable rchan2 = VDACparams[row][VDAC_RCHAN2]

variable vchan1 = VDACparams[row][VDAC_VCHAN1]

variable vchan2 = VDACparams[row][VDAC_VCHAN2]

if( chan == vchan1 ) //VChan1 is to be ramped

return real( RealtoVirtual ( row, ReadVDac(rchan1), ReadVDac(rchan2) ) )

else //Vchan2 is to be ramped

return imag( RealtoVirtual ( row, ReadVDac(rchan1), ReadVDac(rchan2) ) )

endif

endif

end

//Extended version of rampdac(), for ramping virtual DACs as well

function rdEX(chan, destV[, stepsize, tau])

variable chan, destV

variable stepsize, tau

if ( ChannelRealSpecial ( chan ) ) //If it’s a real special channel, call sdEX() directly

sdEX( chan, destV )

return 1

endif

if ( ParamIsDefault ( stepsize ) )

stepsize = 5 //in DacUnits

endif

if ( ParamIsDefault ( tau ) )

tau = 0.1 //in secs

endif

wave AO6

//Change default stepsize and tau for arbitrary channel here

if ( chan == 30 ) //For Idc, go much slower!

stepsize = 40

tau = 0.05

endif

variable stepdelta

if ( destV < AO6[chan] )

stepdelta = - stepsize

149

Page 164: Waves, Particles, and Interactions in Reduced Dimensions

else

stepdelta = stepsize

endif

variable numsteps = floor( abs( AO6[chan] - destV ) / stepsize )

variable v = AO6[chan] + stepdelta

variable i

for ( i = 0; i < numsteps; i += 1 )

sdEx ( chan, v )

wait( tau )

v += stepdelta

endfor

sdEx ( chan, destV )

end

//Extended version of setdac(), for ramping virtual DACs as well

function sdEX(chan, destV)

variable chan, destV

if( ! ChannelDefined ( chan ) ) //Make sure the channel is defined

printf "Channel %g is not defined.\r", chan

return -1

endif

wave AO6, VDACparams, DefinedinRow

variable row = DefinedinRow [chan]

if( ChannelReal ( chan ) ) //If the channel is real

setchan(chan, destV)

UpdateAllVirtualDACs()

else //Otherwise, it is a virtual DAC, defined in DRow

UpdateAllVirtualDACs() //Update first, in case user aborted a ramp

variable rchan1 = VDACparams[row][VDAC_RCHAN1]

variable rchan2 = VDACparams[row][VDAC_RCHAN2]

variable vchan1 = VDACparams[row][VDAC_VCHAN1]

variable vchan2 = VDACparams[row][VDAC_VCHAN2]

variable/c Rdest

if( chan == vchan1 ) //VChan1 is to be ramped

Rdest = VirtualToReal ( row, destV, AO6[vchan2] )

else //Vchan2 is to be ramped

Rdest = VirtualToReal ( row, AO6[vchan1], destV )

endif

sdEX( rchan1, real(Rdest) )

sdEX( rchan2, imag(Rdest) )

UpdateVirtualDACs( row )

endif

end

//Setting real DAC or special channels

function setchan(chan, dest)

variable chan, dest

NVAR NumRealChan

wave AO6

wave/T DACSetFunction

string cmdstr

if ( ChannelRealDAC ( chan ) )

setdac(chan, dest)

elseif ( ChannelRealSpecial ( chan ) )

cmdstr = DACSetFunction[chan] + "(" + num2str(dest) + ")"

Execute cmdstr

AO6[chan] = dest

else //If not real channel, return 0

return 0

endif

150

Page 165: Waves, Particles, and Interactions in Reduced Dimensions

end

//Shortcuts

function sd(chan,value)

variable chan,value

sdEX(chan,value)

end

function rd(chan,value[,stepsize,tau])

variable chan,value

variable stepsize,tau

if ( ! ParamIsDefault ( stepsize ) )

if ( ! ParamIsDefault ( tau ) )

rdEX(chan,value,stepsize=stepsize,tau=tau)

else

rdEX(chan,value,stepsize=stepsize)

endif

else

if ( ! ParamIsDefault ( tau ) )

rdEX(chan,value,tau=tau)

else

rdEX(chan,value)

endif

endif

end

151

Page 166: Waves, Particles, and Interactions in Reduced Dimensions

Appendix C

Effects of external impedance onconductance and noise

C.1 Effects of external impedance on conductance

The effects of finite-impedance external circuits need to be subtracted to obtain the intrinsic

properties of the device. The simplest case is a two-terminal devices in series with an

external resistor, r [see Fig. C.1(a)], and we would like to measure the intrinsic device

resistance, R. As the voltage applied, V is different from the real voltage V dropped across

the device, we will need to subtract the series resistance r from the measured resistance

r +R to obtain the intrinsic device resistance.

A more formal approach is needed to subtract series resistance connected to a multi-lead

device, as shown in Fig. C.1(b). We consider that at lead i, the series resistance is ri, the

out-flowing current is Ii, the real voltage at the lead is Vi, and the applied voltage is Vi. As

in Ch. 5, we denote the raw conductance matrix as g, and the device intrinsic conductance

matrix as g, such that:I1

I2

I3

I4

= g ·

V1

V2

V3

V4

and

I1

I2

I3

I4

= g ·

V1

V2

V3

V4

.

152

Page 167: Waves, Particles, and Interactions in Reduced Dimensions

2 1

43Device

r

R

I

V(b)(a)

V~

r1 I1

V1

V1~

r2I2

V2

V2~

r3I3

V3

V3~

r4 I4

V4

V4~

Figure C.1: Circuit schematics for calculating intrinsic conductance in (a) a two-lead deviceand (b) a multi-lead device.

Since Vi − Vi = riIi, we arrive at:I1

I2

I3

I4

= g·

V1 − r1I1

V2 − r2I2

V3 − r3I3

V4 − r4I4

=⇒

I1

I2

I3

I4

=

E + g ·

r1

r2

r3

r4

−1

·g·

V1

V2

V3

V4

,

where E is an identity matrix. Comparing to the expression for g, we get:

g =

E + g ·

r1

r2

r3

r4

−1

· g. (C.1)

For all ri = r, we simply have g = [E + rg]−1 · g, which is the expression used in Ch. 5.

C.2 Effects of external impedance on current noise

External impedance on current noise has two effects [19, 128]: one is the feedback effect

similar to the conductance case, and the other is the added thermal noise. Here, we again

consider a two-lead device first, as shown in Fig. C.2(a), with both leads connected to

external impedance. Following the approach used in Ref. [19], we model current noise, δIi,

with a current source injected at lead i. They induce voltage fluctuation ∆Vi at lead i, and

current fluctuation ∆Ii through the external resistance ri. What we measure is the voltage

noise SV i ∝ 〈∆V 2i 〉t, and the current noise that we wish to extract is SIi+4kBTe/ri ∝ 〈δI2

i 〉t

and SI12 ∝ 〈δI1δI2〉t, where 〈..〉t denotes time average.

153

Page 168: Waves, Particles, and Interactions in Reduced Dimensions

r1r2ΔI2

δI2

ΔI1

ΔV1ΔV2

δI1

12

43Device

r1r2

R

ΔI2

δI2

ΔI1

ΔV1ΔV2 δI1

(a) (b)

Figure C.2: Circuit schematics for calculating intrinsic current noise in (a) a two-lead deviceand (b) a multi-lead device.

Since ∆Vi = ri∆Ii, we have:(∆I1

∆I2

)=(δI1

δI2

)+(g11 g21

g12 g22

)·(

∆V1

∆V2

)=(

∆V1/r1

∆V2/r2

)

=⇒(δI1

δI2

)=(

1/r1 − g11 −g21

−g12 1/r2 − g22

)·(

∆V1

∆V2

). (C.2)

As in Ch. 5, defining a11(22) = 1/r1(2) − g11(22), a12(21) = −g12(21), we then get

SI1 = a211SV 1 + a2

21SV 2 + 2a11a21SV 12 − 4kBTe/r1 (C.3a)

SI2 = a212SV 1 + a2

22SV 2 + 2a12a22SV 12 − 4kBTe/r2 (C.3b)

SI12 = a11a12SV 1 + a21a22SV 2 + (a11a22 + a12a21)SV 12. (C.3c)

To confirm the correctness of Eq. (C.3), consider the device being a simple resistor,

R, in which case we expect SI1,2 = 4kBTe/R and SI12 = −4kBTe/R. Solve for SV 1 in

terms of SI1,2 and SI12 by inverting Eq. (C.3). After working through some algebra, we get

SV 1 = 4kBTer1(R+ r2)/(r1 + r2 +R)—the expected thermal noise measured at lead 1.

Understanding how to extract intrinsic current noise in a two-lead device, we now move

on the the multi-lead device. Here, however, we only consider the case where two of the leads

are attached to external impedance, and the others are grounded, as shown in Fig. C.2(b).

This is the case we are mostly interested in, since we only have a two-channel noise measure-

ment system so far, but the more general case can also be worked out in a similar fashion.

154

Page 169: Waves, Particles, and Interactions in Reduced Dimensions

Similar to the two-lead case, we can write:∆I1

∆I2

∆I3

∆I4

=

δI1

δI2

δI3

δI4

+

g11 g21 g31 g41

g12 g22 g32 g42

g13 g23 g33 g43

g14 g24 g34 g44

·

∆V1

∆V2

∆V3

∆V4

.

Also, ∆V1,2 = r1,2∆I1,2 and ∆V3,4 = 0. Note here, however, that we will arrive at the

same expression as in Eq. (C.2). Consequently, we will obtain the same expression as in

Eq. (C.3) for extracting the intrinsic current noise of the device, as has been used in Ch. 5.

Since Eq. (C.3) makes use of some elements in the intrinsic conductance matrix that are

not commonly measured, it also motivates the measurement of the full conductance matrix,

for which a multi-channel digital lock-in has been developed and used. I will describe the

details of operations of the multi-channel digital lock-in in the next Appendix.

155

Page 170: Waves, Particles, and Interactions in Reduced Dimensions

Appendix D

Conductance matrix measurementand multi-channel digital lock-in

D.1 Simultaneous conductance matrix and current noise mea-

surement

As we have discussed in the previous Appendix, in order to extract the intrinsic current noise

auto- and cross correlations of a multi-lead device, one needs to know certain elements of

the conductance matrix, eg. g12, which are not measured with the setup described in Ch. 2.

For this purpose, we have developed the circuit that allows simultaneous measurement

of the two-channel current noise, and the full conductance matrix of a multi-lead device.

The circuit, shown in Fig. D.1, combines the effective circuit shown in Fig. C.1(b) for

conductance matrix measurement near dc, and the effective circuit shown in Fig. C.2(b) for

current noise measurement at low MHz.

At low MHz, the effective circuit is similar to what has been described in Ch. 2, except

that the parallel resistance of the resonant circuit is now 5 kΩ in parallel with two other

50 kΩ resistors, making it ∼ 4.2 kΩ. Near dc, in addition to a resistor r = 5 kΩ to ground,

there are two tapped and low pass filtered lines at each lead, one connected to a current

source, and one connected to a single-ended voltage preamplifier. The resistor r converts

156

Page 171: Waves, Particles, and Interactions in Reduced Dimensions

2 1

43Device

50kΩ

50kΩ 5kΩ

5kΩ

5nF

5nF

50kΩ

50kΩ 5kΩ

5kΩ

5nF

5nF

10nF

5kΩ

50kΩ

50kΩ5kΩ

5kΩ

5nF

5nF

10nF

5kΩ

96pF5kΩ10nF

66μH

50kΩ

50kΩ5kΩ

5kΩ

5nF

5nF

96pF5kΩ 10nF

66μH

to HEMTto HEMT

Ibias1

Vprobe1

Vprobe4

Vprobe2

Vprobe3

Ibias4

Ibias2

Ibias3

Figure D.1: Circuit schematics for simultaneously measuring conductance matrix and two-channel current noise.

the current Ii out of lead i to a voltage signal measured by the voltage amplifier; it also

converts the current from the current source to a voltage excitation Vi applied at lead i.

The excitations applied at different leads are at different frequencies, and the each current

measured should contain signals at all these frequencies, which can then be used to measure

the full conductance matrix.

D.2 Multi-channel digital lock-in

Although the conductance matrix can be measured with conventional stand-alone lock-ins,

one would need up to 16 of these single-channel lock-ins to measure the full 4×4 conductance

matrix, making it practically impossible. Therefore, we1 have decided to develop in-house

a multi-channel digital lock-in using a National Instruments AD/DA card.

The latest implementation employs a National Instruments PXI-6259 card, which pro-

vides 32 16-bit analog inputs at 1 MS/s, and four 16-bit analog outputs at 2.8 MS/s. One

1This is a collective work by Doug, Reinier and myself.

157

Page 172: Waves, Particles, and Interactions in Reduced Dimensions

Figure D.2: Multi-channel digital lock-in control panel

such card is sufficient to implement a 4× 42 multi-channel digital lock-in. The four analog

outputs are used to generate four sinusoidal excitations at different frequencies, and they

are also fed into channels 1 through 4 of the analog inputs as reference signals. The response

signals are measured by four voltage preamplifiers, and are then fed into channel 5 through

8 of the analog inputs.

Making use of the NI-DAQmx driver, routines of Igor Pro can directly control and

acquire data from the PXI-6259 card, with most settings controllable from the panel shown

in Fig. D.2. The panel consists of three configuration tabs: the “Lock-in” tab, the “Input

Setup” tab, and the “Output Setup” tab. In the “Lock-in” tab, one can select which

conductance matrix elements are to be measured, the lock-in time constant (meas. tau),

controls to start/stop the lock-in, and perform automatic phase adjustment, etc. In the

“Input Setup” tab, we can set the sampling frequency and the input sensitivity for each

analog inputs. In the “Output Setup” tab, we can set the frequency and amplitude of each

analog outputs, and also controls to start/stop these reference signals.

2Actually, a conductance matrix of size up to 4× 28 can be measured with this card.

158

Page 173: Waves, Particles, and Interactions in Reduced Dimensions

The digital processing implemented here makes use of fast Fourier transform (FFT), in

a similar way as that for calculating current noise auto- and cross correlations. We first

acquire all reference and response signals for a length of time given by the chosen lock-in time

constant. Then, the FFTs of each acquired waves are calculate. The auto-correlation of each

reference signal, as well as pairwise cross correlation between each reference and response

waves are calculated. As shown in the code given below, the amplitude of the reference

wave are calculated by integrating its power spectrum within a narrow bandwidth around

the given reference frequency. The real (imaginary) component of the cross correlation

integrated within the same band and then divided by the amplitude of the reference wave

gives the in-phase (out-of-phase) component of the response wave. When the parameter

pshift is given when the routine is called, it will return the amplitude of the response wave

projected along the angel set by pshift; otherwise, it will return the polar angel set by the

in-phase (real) and out-of-phase (imaginary) component of the response wave relative to

the reference. Both the amplitude of the reference waves, as well as the in-phase component

of the response waves relative to each references are stored in a 2d wave named “currvals”.

//Calculate the amplitude of the spectral component in sigwave at the frequency ’freq’,

// given a phase offset of ’pshift’ between the sigwave and refwave

//If pshift is not given, figure out the phase shift automatically

function DLockinFFT(refwave,sigwave,freq[,pshift])

wave refwave,sigwave

variable freq

variable pshift

variable getphase = 0

if(ParamIsDefault(pshift)) //If pshift is not given, figure out the phase shift

getphase = 1

endif

NVAR BW = root:lockin:measbw //average over the range freq +/- BW, where BW is in points

variable df = deltax(refwave)

variable xv,yv

variable numpts = 2*BW + 1

make/n=(numpts)/o PR,Xr,Xi

Setscale/P x,0,df,PR //Power of the refwave

Setscale/P x,0,df,Xr

Setscale/P x,0,df,Xi

variable lockinpt = x2pnt(refwave,freq) //The center point

159

Page 174: Waves, Particles, and Interactions in Reduced Dimensions

PR = magsqr(refwave[lockinpt+p-BW])

Xr = real(refwave[lockinpt+p-BW])*real(sigwave[lockinpt+p-BW])

+ imag(refwave[lockinpt+p-BW])*imag(sigwave[lockinpt+p-BW])

Xi = real(refwave[lockinpt+p-BW])*imag(sigwave[lockinpt+p-BW])

- imag(refwave[lockinpt+p-BW])*real(sigwave[lockinpt+p-BW])

variable/g refV = sqrt(area(PR)) //reference voltage

variable rv = area(Xr)/refV //in-phase component as the reference wave

variable iv = area(Xi)/refV //out-of-phase componentas the reference wave

if (getphase == 0) //do phase shift

xv = rv*cos(pshift) + iv*sin(pshift)

yv = - rv*sin(pshift) + iv*cos(pshift)

return xv //return the x component

else

return imag(r2polar(cmplx(rv,iv))) //return the phase shift

endif

end

The digital lock-in can run in two different modes, local mode or remote (listen) mode.

In the local mode, each time the function DoLockin() is called, it acquires and processes the

data, and updates the wave “currvals”. In the remote mode, it establishes a serial connection

with another computer, and listens to any commands sent over the serial connection. The

commands can either set parameters of the lock-in, or ask for the most recent “currvals”

wave. In this way, the processing computer, together with the NI PXI-6259 card, serves as

a standalone multi-channel digital lock-in.

160

Page 175: Waves, Particles, and Interactions in Reduced Dimensions

Appendix E

The master equation calculation ofcurrent and noise in a multi-lead,multi-level quantum dot

In this Appendix, I provide the Igor routines for the master equation calculation of cur-

rent and noise in a multi-lead, multi-level quantum dot, mainly following Ref. [135]. These

routines allow setting of arbitrary number of leads and arbitrary number of levels within

the dot, and are used for the calculations in Ch. 5. Also calculated are frequency-dependent

charge and current noise. Since these routines are quite general within the sequential tun-

neling limit, I hope they can also be useful for anyone wishing to learn transport, both

conductance and noise, through a Coulomb blockaded quantum dot.

//////////////////////////////////////////////////////////////////////////////////////////////////

// The Master equation calculation of current and noise in a multi-lead, multi-level quantum dot

// by Yiming Zhang and Leo DiCarlo, Noise Team, Marcus Lab, Harvard University

//////////////////////////////////////////////////////////////////////////////////////////////////

function initmodel()

//Parameters describing system complexity

NVAR nleads,ne,nh,nlevels

nleads=3; //Number of leads

ne = 3; //Number of electron excited levels

nh = 3; //Number of hole excited levels

nlevels = ne + nh + 1

//Constants

variable/g ec = 1.6e-19; variable/g kb = 1.38e-23

variable/g lt = 0; variable/g lb = 1; variable/g rb = 2; variable/g rt = 3

make/d/o/t leadlabels = "lt","lb","rb","rt"

//Physical parameters:

variable/g Te //in K, electron temperature

161

Page 176: Waves, Particles, and Interactions in Reduced Dimensions

variable/g Vg //in V, gate voltage

variable/g Vg0 //in V, gate voltage reference

variable/g alphaVg //lever arm for gate voltage. alphaVg = Cgate / Ctot

variable/g f0 //in Hz, measurement frequency

make/d/o/n=(nlevels,nleads) GammaMat //in Hz, the bare tunneling rates from each level to each lead

make/d/o/n=(nlevels) Elevels //in eV, energy of single particle levels

Elevels = 150e-6 * (ne - p) //The ne^th level is the ground level

make/d/o/n=(nlevels) Edotlevels //in Joule, energy of single particle levels, affected by Vg

make/d/o/n=(nleads) TeL //in K, electron temperature in each lead

make/d/o/n=(nleads) Vbias //in uV, Bias voltages on the leads

make/d/o/n=(nleads) MuLead //in Joule, Chemical potentials in the leads

variable/g Beta //Gating ratio of the bias voltages

make/d/o/n=(nleads) BetaL //Gating ratio of the bias voltages for different leads

variable/g NumStates0,NumStates1,NumStates

NumStates0 = nComb(nh,nlevels) //Number of states with nh electrons

NumStates1 = nComb(nh+1,nlevels) //Number of states with nh+1 electrons

NumStates = NumStates0 + NumStates1

make/d/o/n=(NumStates0,nlevels) StateConfig0 = nan //nh electron state configurations

make/d/o/n=(NumStates1,nlevels) StateConfig1 = nan //nh+1 electron state configurations

genConfig0(0,0,nh,nlevels)

genConfig1(0,0,nh+1,nlevels)

//Auxiliary waves, need to be generated once, and used in some functions

make/d/o/n=(nlevels) diffvec

make/d/o/n=(nleads) MvsLead

make/d/o/n=(1,NumStates) eT = 1

make/d/o/n=(NumStates1,NumStates0) StateDiffmat

StateDiffmat = StateDiff(p,q)

//M-matrix

make/d/o/n=(NumStates,NumStates) Mmat //Note: Mmat[i][j] is the rate j->i

make/d/o/n=(NumStates1,NumStates0,nleads) Mmat0t1 //Mmat subblock, rates from eh to eh+1 electrons

make/d/o/n=(NumStates0,NumStates1,nleads) Mmat1t0 //Mmat subblock, rates from eh+1 to eh electrons

//Current matrices, U, D matrices, and steady state vectors

make/d/o/n=(NumStates,NumStates,nleads) Jcube

make/d/o/n=(NumStates,NumStates) Jmat,Jmat1,Jmat2,U,D

make/d/c/o/n=(NumStates,NumStates) Dp,Dm

make/d/o/n=(NumStates,1) rho_o,eigenM

//Number matrix

make/d/o/n=(NumStates,NumStates) Nd

Nd = p==q ? (p>=NumStates0 ? nh+1 : nh) : 0

SetParams()

end

// Set parameters here

function SetParams()

NVAR Te, Vg, Vg0, alphaVg, f0, Beta, ne

wave GammaMat,Elevels,Vbias,BetaL,TeL

NVAR lt,lb,rb,rt

Te = 0.34 //in K, electron temperature

TeL = Te //in K, electron temperature in each lead

Vg0 = -1.7525 //in V, gate voltage reference

Vg = Vg0 //in V, gate voltage

alphaVg = 0.0692308

f0 = 0 //in Hz, measurement frequency, not used now

GammaMat[][lt] = 10e9 //in Hz, the bare tunneling rates from each level to each lead

GammaMat[][lb] = 10e9

// GammaMat[][rb] = 2.5e9

Elevels = 150e-6 * (ne - p) //The ne^th level is the ground level

Vbias = 0 //in uV, Bias voltages on the leads

162

Page 177: Waves, Particles, and Interactions in Reduced Dimensions

Beta = 0.3 //Gating ratio of the bias voltages

BetaL = Beta //Gating ratio of the bias voltages for different leads

end

//Calculate the M-matrix and steady state vector

function PrepMandRho()

NVAR Te,Vg,Vg0,alphaVg,Beta,f0,nleads,ne,nh,nlevels,NumStates0,NumStates1,NumStates

wave GammaMat,Elevels,Edotlevels,Vbias,MuLead,BetaL,TeL

wave Mmat,Mmat0t1,Mmat1t0,StateDiffmat,eT

wave Jcube,U,D,rho_o,eigenM

wave/c Dp,Dm

wave M_R_eigenVectors,W_eigenValues

variable loc_rho_o, tempvar

variable ec = 1.6e-19

//Measurement frequency

variable Omega

Omega = 2 * pi * f0 //f0 is the main parameter, Omega will be updated according to f0

//Chemical potentials in the leads

MuLead = - 1e-6 * ec * Vbias //in Joule, Chemical potentials in the leads

//Energy levels in the dot

variable Edot0 = - 1e-6 * MatrixDot(Vbias,BetaL) - alphaVg * (Vg - Vg0)

Edotlevels = ec * (Edot0 + Elevels) //in eV, dot levels

TeL = Te

// //Implement bias dependent electron heating here

// TeL = sqrt(Te^2 + 1.7e-8*Vbias[p]^2)

//Prepare Rate matrices

// Is it forbidden ?Yes: [ relavant level ][lead]

Mmat1t0 = StateDiffmat[q][p]==-1 ? 0 :

GammaMat[StateDiffmat[q][p]][r] * (1 - FermiF(Edotlevels[StateDiffmat[q][p]],MuLead[r],TeL[r]))

Mmat0t1 = StateDiffmat[p][q]==-1 ? 0 :

GammaMat[StateDiffmat[p][q]][r] * FermiF(Edotlevels[StateDiffmat[p][q]],MuLead[r],TeL[r])

//Prepare M-matrix, first get off-diagonal elements.

// (( top left block ) || ( bottom right block ) ? 0 :

// ( top right block ? eh+1->eh transition :

// eh->eh+1 transition )

Mmat = ((p<NumStates0 && q<NumStates0) || (p>=NumStates0 && q>=NumStates0)) ? 0 :

(p<NumStates0 && q>=NumStates0 ? sumoverlayer(Mmat1t0,p,q-NumStates0) :

sumoverlayer(Mmat0t1,p-NumStates0,q))

MatrixOp/O sumMvec = eT x Mmat //For a given colomn, sum all the rows of the Mmat

Mmat = p==q ? - (sumMvec[0][q] - Mmat[q][q]) : Mmat[p][q] //Get diagonal elements

// Steady state probability vector

loc_rho_o = null(Mmat) //null also calculates the eigen-values and eigen-vectors

U = M_R_eigenVectors //Matrix U is the modal matrix - its columns are the eigenvectors of Mmat

eigenM = real(W_eigenValues[p]) //eigenM contain the eigen values

rho_o = U[p][loc_rho_o]

tempvar = sum(rho_o)

rho_o /= tempvar

//Prepare matrix D

MatrixOP/O D = Diagonal(eigenM)

variable m=0

do

if(m==loc_rho_o)

D[m][m] = 0

Dp[m][m] = 0

Dm[m][m] = 0

else

Dp[m][m] = 1/cmplx(eigenM[m],Omega)

Dm[m][m] = 1/cmplx(eigenM[m],-Omega)

163

Page 178: Waves, Particles, and Interactions in Reduced Dimensions

D[m][m] = 1/eigenM[m]

endif

m+=1

while(m<NumStates)

//Current matrix

// (( top left block ) || ( bottom right block )) ? 0 :

// ( top right block ? eh+1->eh transition : eh->eh+1 transition )

Jcube = ((p<NumStates0 && q<NumStates0) || (p>=NumStates0 && q>=NumStates0)) ? 0 :

(p<NumStates0 && q>=NumStates0 ? Mmat1t0[p][q-NumStates0][r] : -Mmat0t1[p-NumStates0][q][r])

FastOp Jcube = (-ec) * Jcube

end

//Get DC current, in A

function getI(LeadIndex)

variable LeadIndex

wave Jcube,Jmat,rho_o

Jmat = Jcube[p][q][LeadIndex]

MatrixOP/O tempwave = Jmat x rho_o;

return sum(tempwave)

end

//Get frequency-dependent charge noise spectral density, in e^2/Hz

function getSN()

wave U,Dp,Dm,rho_o,Nd

MatrixOP/O tempwave = Nd x U x Dp x Inv(U) x Nd x rho_o

+ Nd x U x Dm x Inv(U) x Nd x rho_o

return real(-2*sum(tempwave))

end

//Get frequency-dependent current noise auto- and cross-correlation spectral density, in A*2e

function getSI(lead1,lead2)

variable lead1,lead2

wave Jcube,Jmat1,Jmat2,U,D,Dp,Dm,rho_o,Nd

variable ec = 1.6e-19; // in Coulomb

if(lead1==lead2) //auto-correlation

Jmat1 = Jcube[p][q][lead1]

MatrixOP/O tempwave = Jmat1 x U x Dp x Inv(U) x Jmat1 x rho_o

+ Jmat1 x U x Dm x Inv(U) x Jmat1 x rho_o

- ec * (Nd x Jmat1 - Jmat1 x Nd) x rho_o

return real(-2*sum(tempwave)/2/ec)

else //cross-correlation

Jmat1 = Jcube[p][q][lead1]

Jmat2 = Jcube[p][q][lead2]

MatrixOP/O tempwave = Jmat1 x U x Dp x Inv(U) x Jmat2 x rho_o

+ Jmat2 x U x Dm x Inv(U) x Jmat1 x rho_o

return real(-2*sum(tempwave)/2/ec)

endif

end

164

Page 179: Waves, Particles, and Interactions in Reduced Dimensions

Bibliography

[1] C. W. J. Beenakker and H. van Houten, Quantum transport in semiconductor nanos-

tructures, Solid State Physics 44, 1 (1991).

[2] S. Datta, Electronic Transport in Mesoscopic Systems (Cambridge University Press,

Cambridge, 1995).

[3] L. P. Kouwenhoven, C. M. Marcus, P. L. McEuen, S. Tarucha, R. M. Westervelt, and

N. S. Wingreen, Electron Transport in Quantum Dots, NATO ASI Series E, Vol. 345

(Kluwer, 1997).

[4] B. L. Altshuler, P. A. Lee, and R. A. Webb, Mesoscopic Phenomena in Solids (North-

Holland, Amsterdam, 1991).

[5] K. von Klitzing, G. Dorda, and M. Pepper, New method for high accuracy determi-

nation of the fine-structure constant based on quantized Hall resistance, Phys. Rev.

Lett. 45, 494 (1980).

[6] B. J. van Wees, H. van Houten, C. W. J. Beenakker, J. G. Williamson, L. P. Kouwen-

hoven, D. van der Marel, and C. T. Foxon, Quantized conductance of point contacts

in a two-dimensional electron gas, Phys. Rev. Lett. 60, 848 (1988).

165

Page 180: Waves, Particles, and Interactions in Reduced Dimensions

[7] R. A. Webb, S. Washburn, C. P. Umbach, and R. B. Laibowitz, Observation of h/e

Aharonov-Bohm oscillations in normal-metal rings, Phys. Rev. Lett. 54, 2696 (1985).

[8] A. Yacoby, M. Heiblum, D. Mahalu, and H. Shtrikman, Coherence and phase sensitive

measurements in a quantum dot, Phys. Rev. Lett. 74, 4047 (1995).

[9] A. van Oudenaarden, M. H. Devoret, Y. V. Nazarov, and J. E. Mooij, Magneto-electric

Aharonov-Bohm effect in metal rings, Nature 391, 768 (1998).

[10] 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 (2001).

[11] Y. Ji, Y. Chung, D. Sprinzak, M. Heiblum, D. Mahalu, and H. Shtrikman, An elec-

tronic machczehnder interferometer, Nature 422, 415 (2003).

[12] T. A. Fulton and G. J. Dolan, Observation of single-electron charging effects in small

tunnel junctions, Phys. Rev. Lett. 59, 109 (1987).

[13] U. Meirav, M. A. Kastner, and S. J. Wind, Single-electron charging and periodic

conductance resonances in GaAs nanostructures, Phys. Rev. Lett. 65, 771 (1990).

[14] D. C. Glattli, C. Pasquier, U. Meirav, F. I. B. Williams, Y. Jin, and B. Etienne,

Co-tunneling of the charge through a 2-D electron island, Z. Phys. B 85, 375 (1991).

[15] J. A. Folk, S. R. Patel, S. F. Godijn, A. G. Huibers, S. M. Cronenwett, C. M. Marcus,

K. Campman, and A. C. Gossard, Statistics and parametric correlations of Coulomb

blockade peak fluctuations in quantum dots, Phys. Rev. Lett. 76, 1699 (1996).

[16] S. M. Cronenwett, S. M. Maurer, S. R. Patel, C. M. Marcus, C. I. Duruoz, and J. S.

Harris, Mesoscopic Coulomb blockade in one-channel quantum dots, Phys. Rev. Lett.

81, 5904 (1998).

166

Page 181: Waves, Particles, and Interactions in Reduced Dimensions

[17] D. Goldhaber-Gordon, H. Shtrikman, D. Mahalu, D. Abusch-Magder, U. Meirav, and

M. A. Kastner, Kondo effect in a single-electron transistor, Nature 391, 156 (1998).

[18] S. De Franceschi, S. Sasaki, J. M. Elzerman, W. G. van der Wiel, S. Tarucha, and

L. P. Kouwenhoven, Electron cotunneling in a semiconductor quantum dot, Phys. Rev.

Lett. 86, 878 (2001).

[19] Y. M. Blanter and M. Buttiker, Shot noise in mesoscopic conductors, Phys. Rep. 336,

1 (2000).

[20] I. L. Aleiner, P. W. Brouwer, and L. I. Glazman, Quantum effects in Coulomb blockade,

Physics Reports 358, 309 (2002).

[21] E. Buks, R. Schuster, M. Heiblum, D. Mahalu, and V. Umansky, Dephasing in electron

interference by a ’which-path’ detector, Nature 391, 871 (1998).

[22] Y. Zhang, D. T. McClure, E. M. Levenson-Falk, C. M. Marcus, L. N. Pfeiffer, and

K. W. West, Distinct signatures for Coulomb blockade and Aharonov-Bohm interfer-

ence in electronic Fabry-Perot interferometers ArXiv:0901.0127.

[23] K. J. Thomas, J. T. Nicholls, M. Y. Simmons, M. Pepper, D. R. Mace, and D. A.

Ritchie, Possible spin polarization in a one-dimensional electron gas, Phys. Rev. Lett.

77, 135 (1996).

[24] H. L. Stormer, D. C. Tsui, and A. C. Gossard, The fractional quantum Hall effect,

Rev. Mod. Phys. 71, S298 (1999).

[25] G. Moore and N. Read, Nonabelions in the fractional quantum Hall effect .

[26] C. Nayak, S. H. Simon, A. Stern, M. Freedman, and S. D. Sarma, Non-Abelian anyons

and topological quantum computation, Rev. Mod. Phys. 80, 1083 (2008).

167

Page 182: Waves, Particles, and Interactions in Reduced Dimensions

[27] D. C. Tsui, H. L. Stormer, and A. C. Gossard, Two-dimensional magnetotransport in

the extreme quantum limit, Phys. Rev. Lett. 48, 1559 (1982).

[28] D. A. Wharam, T. J. Thornton, R. Newbury, M. Pepper, H. Ahmed, J. E. F. Frost,

D. G. Hasko, D. C. Peacock, D. A. Ritchie, and G. A. C. Jones, One-dimensional

transport and the quantization of the ballistic resistance, J. Phys. C 21, L209 (1988).

[29] Y. M. Blanter, Recent advances in studies of current noise, arXiv:cond-mat/0511478.

[30] T. Martin, Noise in mesoscopic physics, in Nanophysics: coherence and transport, Les

Houches Session LXXXI, edited by H. Bouchiat, Y. Gefen, S. Gueron, G. Montam-

baux, and J. Dalibard (Elsevier, New York, 2005).

[31] L. DiCarlo, Y. Zhang, D. T. McClure, D. J. Reilly, C. M. Marcus, L. N. Pfeiffer,

and K. W. West, Shot-noise signatures of 0.7 structure and spin in a quantum point

contact, Phys. Rev. Lett. 97, 036810 (2006).

[32] L. DiCarlo, Y. Zhang, D. T. McClure, D. J. Reilly, C. M. Marcus, L. N. Pfeiffer, K. W.

West, M. P. Hanson, and A. C. Gossard, Current noise in quantum point contacts,

arXiv:0704.3892.

[33] Y. Zhang, L. DiCarlo, D. T. McClure, M. Yamamoto, S. Tarucha, C. M. Marcus, M. P.

Hanson, and A. C. Gossard, Noise correlations in a Coulomb-blockaded quantum dot,

Phys. Rev. Lett. 99, 036603 (2007).

[34] D. T. McClure, L. DiCarlo, Y. Zhang, H.-A. Engel, C. M. Marcus, M. P. Hanson, and

A. C. Gossard, Tunable noise cross correlations in a double quantum dot, Phys. Rev.

Lett. 98, 056801 (2007).

168

Page 183: Waves, Particles, and Interactions in Reduced Dimensions

[35] L. DiCarlo, J. R. Williams, Y. Zhang, D. T. McClure, and C. M. Marcus, Shot noise

in graphene, Phys. Rev. Lett. 100, 156801 (2008).

[36] S. Machlup, Noise in semiconductors: Spectrum of a two-parameter random signal,

J. Appl. Phys. 25, 341 (1954).

[37] L. Spietz, K. W. Lehnert, I. Siddiqi, and R. J. Schoelkopf, Primary electronic ther-

mometry using the shot noise of a tunnel junction, Science 300, 1929 (2003).

[38] R. de-Picciotto, M. Reznikov, M. Heiblum, V. umansky, G. Bunin, and D. Mahalu,

Direct observation of a fractional charge, Nature 389, 162 (1997).

[39] L. Saminadayar, D. C. Glattli, Y. Jin, and B. Etienne, Observation of the e/3 frac-

tionally charged laughlin quasiparticle, Phys. Rev. Lett. 79, 2526 (1997).

[40] M. Reznikov, R. de Picciotto, T. G. Griffiths, M. Heiblum, and V. Umansky, Obser-

vation of quasiparticles with one-fifth of an electron’s charge, Nature 399, 238 (1999).

[41] M. Dolev, M. Heiblum, V. Umansky, A. Stern, and D. Mahalu, Observation of a

quarter of an electron charge at the ν = 5/2 quantum Hall state, Nature 452, 829

(2008).

[42] I. Neder, N. Ofek, Y. Chung, M. Heiblum, D. Mahalu, and V. Umansky, Interfer-

ence between two independent electrons: observation of two-particle Aharonov-Bohm

interference, Nature 448, 333 (2007).

[43] S. S. Safonov, A. K. Savchenko, D. A. Bagrets, O. N. Jouravlev, Y. V. Nazarov, E. H.

Linfield, and D. A. Ritchie, Enhanced shot noise in resonant tunneling via interacting

localized states, Phys. Rev. Lett. 91, 136801 (2003).

169

Page 184: Waves, Particles, and Interactions in Reduced Dimensions

[44] P. Barthold, F. Hohls, N. Maire, K. Pierz, and R. J. Haug, Enhanced shot noise

in tunneling through a stack of coupled quantum dots, Phys. Rev. Lett. 96, 246804

(2006).

[45] E. Onac, F. Balestro, B. Trauzettel, C. F. J. Lodewijk, and L. P. Kouwenhoven,

Shot-noise detection in a carbon nanotube quantum dot, Phys. Rev. Lett. 96, 026803

(2006).

[46] S. Gustavsson, R. Leturcq, B. Simovic, R. Schleser, P. Studerus, T. Ihn, K. Ensslin,

D. C. Driscoll, and A. C. Gossard, Counting statistics and super-Poissonian noise in

a quantum dot: Time-resolved measurements of electron transport, Phys. Rev. B 74,

195305 (2006).

[47] J. P. Eisenstein, K. B. Cooper, L. N. Pfeiffer, and K. W. West, Insulating and frac-

tional quantum Hall states in the first excited Landau level, Phys. Rev. Lett. 88,

076801 (2002).

[48] J. R. Arthur, Molecular beam epitaxy, Surface Science 500, 189 (2002).

[49] M. Switkes, Decoherence and adiabatic transport in semiconductor quantum dots,

Ph.D. thesis, Stanford University (1999).

[50] H. A. Fertig and B. I. Halperin, Transmission coefficient of an electron through a

saddle-point potential in a magnetic field, Phys. Rev. B 36, 7969 (1987).

[51] M. Buttiker, Quantized transmission of a saddle-point constriction, Phys. Rev. B 41,

7906 (1990).

[52] R. Landauer, Spatial variation of currents and fields due to localized scatterers in

metalic conduction, IBM J. Res. Dev. 1, 233 (1957).

170

Page 185: Waves, Particles, and Interactions in Reduced Dimensions

[53] M. Buttiker, Absence of backscattering in the quantum Hall effect in multiprobe con-

ductors, Phys. Rev. B 38, 9375 (1988).

[54] M. Buttiker, Scattering theory of thermal and excess noise in open conductors, Phys.

Rev. Lett. 65, 2901 (1990).

[55] M. Buttiker, Scattering theory of current and intensity noise correlations in conduc-

tors and wave guides, Phys. Rev. B 46, 12485 (1992).

[56] P. Roche, J. Segala, D. C. Glattli, J. T. Nicholls, M. Pepper, A. C. Graham, K. J.

Thomas, M. Y. Simmons, and D. A. Ritchie, Fano factor reduction on the 0.7 con-

ductance structure of a ballistic one-dimensional wire, Phys. Rev. Lett. 93, 116602

(2004).

[57] H. van Houten, C. W. J. Beenakker, and A. A. M. Staring, Coulomb-blockade os-

cillations in semiconductor nanostructures, in Single Charge Tunneling, NATO ASI

Series B294, edited by H. Grabert and M. H. Devoret (Plenum, New York, 1992).

[58] W. G. van der Wiel, S. De Franceschi, J. M. Elzerman, T. Fujisawa, S. Tarucha, and

L. P. Kouwenhoven, Electron transport through double quantum dots, Rev. Mod. Phys.

75, 1 (2002).

[59] A. T. Johnson, L. P. Kouwenhoven, W. de Jong, N. C. van der Vaart, C. J. P. M.

Harmans, and C. T. Foxon, Zero-dimensional states and single electron charging in

quantum dots, Phys. Rev. Lett. 69, 1592 (1992).

[60] D. R. Stewart, D. Sprinzak, C. M. Marcus, C. I. Duruoz, and J. Harris, J. S., Corre-

lations between ground and excited state spectra of a quantum dot, Science 278, 1784

(1997).

171

Page 186: Waves, Particles, and Interactions in Reduced Dimensions

[61] L. P. Kouwenhoven, T. H. Oosterkamp, M. W. S. Danoesastro, M. Eto, D. G. Austing,

T. Honda, and S. Tarucha, Excitation spectra of circular, few-electron quantum dots,

Science 278, 1788 (1997).

[62] C. de C. Chamon, D. E. Freed, S. A. Kivelson, S. L. Sondhi, and X. G. Wen, Two

point-contact interferometer for quantum Hall systems, Phys. Rev. B 55, 2331 (1997).

[63] P. Bonderson, A. Kitaev, and K. Shtengel, Detecting non-Abelian statistics in the

ν = 5/2 fractional quantum Hall state, Phys. Rev. Lett. 96, 016803 (2006).

[64] A. Stern and B. I. Halperin, Proposed experiments to probe the ν = 5/2 quantum Hall

state, Phys. Rev. Lett. 96, 016802 (2006).

[65] F. E. Camino, W. Zhou, and V. J. Goldman, Realization of a laughlin quasiparticle

interferometer: Observation of fractional statistics, Phys. Rev. B 72, 075342 (2005).

[66] F. E. Camino, W. Zhou, and V. J. Goldman, Aharonov-Bohm superperiod in a laughlin

quasiparticle interferometer, Phys. Rev. Lett. 95, 246802 (2005).

[67] F. E. Camino, W. Zhou, and V. J. Goldman, e/3 Laughlin quasiparticle primary-filling

ν = 1/3 interferometer, Phys. Rev. Lett. 98, 076805 (2007).

[68] R. L. Willett, M. J. Manfra, L. N. Pfeiffer, and K. W. West, Interferometric measure-

ment of filling factor 5/2 quasiparticle charge, arXiv:0807.0221.

[69] M. D. Godfrey, P. Jiang, W. Kang, S. H. Simon, K. W. Baldwin, L. N. Pfeif-

fer, , and K. W. West, Aharonov-Bohm-like oscillations in quantum Hall corrals

ArXiv:0708.2448.

172

Page 187: Waves, Particles, and Interactions in Reduced Dimensions

[70] M. Reznikov, M. Heiblum, H. Shtrikman, and D. Mahalu, Temporal correlation of

electrons: Suppression of shot noise in a ballistic quantum point contact, Phys. Rev.

Lett. 75, 3340 (1995).

[71] A. Kumar, L. Saminadayar, D. C. Glattli, Y. Jin, and B. Etienne, Experimental test

of the quantum shot noise reduction theory, Phys. Rev. Lett. 76, 2778 (1996).

[72] R. C. Liu, B. Odom, Y. Yamamoto, and S. Tarucha, Quantum interference in electron

collision, Nature 391, 263 (1998).

[73] A. H. Steinbach, J. M. Martinis, and M. H. Devoret, Observation of hot-electron shot

noise in a metallic resistor, Phys. Rev. Lett. 76, 3806 (1996).

[74] M. Henny, S. Oberholzer, C. Strunk, and C. Schonenberger, 1/3-shot-noise suppres-

sion in diffusive nanowires, Phys. Rev. B 59, 2871 (1999).

[75] S. Oberholzer, E. V. Sukhorukov, C. Strunk, C. Schonenberger, T. Heinzel, and

M. Holland, Shot noise by quantum scattering in chaotic cavities, Phys. Rev. Lett.

86, 2114 (2001).

[76] X. Jehl, M. Sanquer, R. Calemczuk, and D. Mailly, Detection of doubled shot noise

in short normal-metal/superconductor junctions, Nature 405, 50 (2000).

[77] S. Gustavsson, R. Leturcq, B. Simovic, R. Schleser, T. Ihn, P. Studerus, K. Ensslin,

D. C. Driscoll, and A. C. Gossard, Counting statistics of single electron transport in

a quantum dot, Phys. Rev. Lett. 96, 076605 (2006).

[78] T. Martin, A. Crepieux, and N. Chtchelkatchev, Noise correlations, entanglement, and

Bell inequalities, in Quantum noise in mesoscopic physics, edited by Y. V. Nazarov,

NATO Science Series II, Vol. 97 (Kluwer, Dordrecht, 2003).

173

Page 188: Waves, Particles, and Interactions in Reduced Dimensions

[79] P. Samuelsson, E. V. Sukhorukov, and M. Buttiker, Two-particle Aharonov-Bohm

effect and entanglement in the electronic Hanbury Brown–Twiss setup, Phys. Rev.

Lett. 92, 026805 (2004).

[80] C. W. J. Beenakker, M. Kindermann, C. M. Marcus, and A. Yacoby, Entanglement

production in a chaotic quantum dot, in Fundamental Problems in Mesoscopic Physics,

edited by I. V. Lerner, B. L. Altshuler, and Y. Gefen, NATO Science Series II, Vol.

154 (Kluwer, Dordrecht, 2004).

[81] A. V. Lebedev, G. B. Lesovik, and G. Blatter, Entanglement in a noninteracting

mesoscopic structure, Phys. Rev. B 71, 045306 (2005).

[82] R. J. Schoelkopf, P. J. Burke, A. A. Kozhevnikov, D. E. Prober, and M. J. Rooks,

Frequency dependence of shot noise in a diffusive mesoscopic conductor, Phys. Rev.

Lett. 78, 3370 (1997).

[83] D. C. Glattli, P. Jacques, A. Kumar, P. Pari, and L. Saminadayar, A noise detection

scheme with 10 mk noise temperature resolution for semiconductor single electron

tunneling devices, J. Appl. Phys. 81, 7350 (1997).

[84] M. Sampietro, L. Fasoli, and G. Ferrari, Spectrum analyzer with noise reduction by

cross-correlation technique on two channels, Rev. Sci. Instrum. 70, 2520 (1999).

[85] M. Henny, S. Oberholzer, C. Strunk, T. Heinzel, K. Ensslin, M. Holland, and

C. Schonenberger, The fermionic Hanbury Brown and Twiss experiment, Science 284,

296 (1999).

174

Page 189: Waves, Particles, and Interactions in Reduced Dimensions

[86] S. Oberholzer, M. Henny, C. Strunk, C. Schonenberger, T. Heinzel, K. Ensslin, and

M. Holland, The Hanbury Brown and Twiss experiment with fermions, Physica E 6,

314 (2000).

[87] S. Oberholzer, E. Bieri, C. Schonenberger, M. Giovannini, and J. Faist, Positive cross

correlations in a normal-conducting fermionic beam splitter, Phys. Rev. Lett. 96,

046804 (2006).

[88] W. D. Oliver, J. Kim, R. C. Liu, and Y. Yamamoto, Hanbury Brown and Twiss-type

experiment with electrons, Science 284, 299 (1999).

[89] A. T. Lee, Broadband cryogenic preamplifiers incorporating GaAs MESFETs for use

with low-temperature particle detectors, Rev. Sci. Instrum. 60, 3315 (1989).

[90] A. T.-J. Lee, A low-power-dissipation broadband cryogenic preamplifier utilizing GaAs

MESFETs in parallel, Rev. Sci. Instrum. 64, 2373 (1993).

[91] A. M. Robinson and V. I. Talyanskii, Cryogenic amplifier for ∼ 1 Mhz with a high

input impedance using a commercial pseudomorphic high electron mobility transistor,

Rev. Sci. Instrum. 75, 3169 (2004).

[92] F. Pobell, Matter and methods at low temperatures, 2nd Ed. (Springer-Verlag, Berlin,

1996).

[93] A. V. Oppenheim and R. W. Schafer, Discrete-time signal processing (Prentice-Hall,

Englewood Cliffs, 1989).

[94] J. B. Hagen, Radio-Frequency Electronics (Cambridge University Press, Cambridge,

1996).

175

Page 190: Waves, Particles, and Interactions in Reduced Dimensions

[95] M. Frigo and S. G. Johnson, FFTW: an adaptive software architecture for the FFT,

in Proceedings of the IEEE International Conference on Acoustics, Speech, and Signal

Processing, Vol. 3, 1381 (IEEE, New York, 1998).

[96] R. H. Dicke, The measurement of thermal radiation at microwave frequencies, Rev.

Sci. Instrum. 17, 268 (1946).

[97] G. B. Lesovik, Excess quantum noise in 2d ballisic point contacts, Pis’ma Zh. Eksp.

Teor. Fiz. 49, 513 (1989), [JETP Lett. 49, 592 (1989)].

[98] A. Kristensen, H. Bruus, A. E. Hansen, J. B. Jensen, P. E. Lindelof, C. J. Marck-

mann, J. Nygard, C. B. Sørensen, F. Beuscher, A. Forchel, and M. Michel, Bias and

temperature dependence of the 0.7 conductance anomaly in quantum point contacts,

Phys. Rev. B 62, 10950 (2000).

[99] D. J. Reilly, G. R. Facer, A. S. Dzurak, B. E. Kane, R. G. Clark, P. J. Stiles, R. G.

Clark, A. R. Hamilton, J. L. O’Brien, N. E. Lumpkin, L. N. Pfeiffer, and K. W. West,

Many-body spin-related phenomena in ultra low-disorder quantum wires, Phys. Rev.

B 63, 121311(R) (2001).

[100] D. J. Reilly, T. M. Buehler, J. L. O’Brien, A. R. Hamilton, A. S. Dzurak, R. G. Clark,

B. E. Kane, L. N. Pfeiffer, and K. W. West, Density-dependent spin polarization in

ultra-low-disorder quantum wires, Phys. Rev. Lett. 89, 246801 (2002).

[101] S. M. Cronenwett, H. J. Lynch, D. Goldhaber-Gordon, L. P. Kouwenhoven, C. M.

Marcus, K. Hirose, N. S. Wingreen, and V. Umansky, Low-temperature fate of the 0.7

structure in a point contact: A Kondo-like correlated state in an open system, Phys.

Rev. Lett. 88, 226805 (2002).

176

Page 191: Waves, Particles, and Interactions in Reduced Dimensions

[102] W. D. Oliver, The generation and detection of electron entanglement, Ph.D. thesis,

Stanford University (2002).

[103] L. P. Rokhinson, L. N. Pfeiffer, and K. W. West, Spontaneous spin polarization in

quantum point contacts, Phys. Rev. Lett. 96, 156602 (2006).

[104] C.-K. Wang and K.-F. Berggren, Spin splitting of subbands in quasi-one-dimensional

electron quantum channels, Phys. Rev. B 54, 14257(R) (1996).

[105] H. Bruus, V. V. Cheianov, and K. Flensberg, The anomalous 0.5 and 0.7 conductance

plateaus in quantum point contacts, Physica E 10, 97 (2001).

[106] Y. Meir, K. Hirose, and N. S. Wingreen, Kondo model for the 0.7 anomaly in transport

through a quantum point contact, Phys. Rev. Lett. 89, 196802 (2002).

[107] K. A. Matveev, Conductance of a quantum wire in the wigner-crystal regime, Phys.

Rev. Lett. 92, 106801 (2004).

[108] A. Ramsak and J. H. Jefferson, Shot noise reduction in quantum wires with the 0.7

structure, Phys. Rev. B 71, 161311(R) (2005).

[109] D. J. Reilly, Phenomenological model for the 0.7 conductance feature in quantum wires,

Phys. Rev. B 72, 033309 (2005).

[110] M. Avinun-Kalish, M. Heiblum, A. Silva, D. Mahalu, and V. Umansky, Controlled

dephasing of a quantum dot in the Kondo regime, Phys. Rev. Lett. 92, 156801 (2004).

[111] L. P. Kouwenhoven, B. J. van Wees, C. J. P. M. Harmans, J. G. Williamson, H. van

Houten, C. W. J. Beenakker, C. T. Foxon, and J. J. Harris, Nonlinear conductance

of quantum point contacts, Phys. Rev. B 39, 8040(R) (1989).

177

Page 192: Waves, Particles, and Interactions in Reduced Dimensions

[112] N. K. Patel, J. T. Nicholls, L. Martn-Moreno, M. Pepper, J. E. F. Frost, D. A.

Ritchie, and G. A. C. Jones, Evolution of half plateaus as a function of electric field

in a ballistic quasi-one-dimensional constriction, Phys. Rev. B 44, 13549 (1991).

[113] Y. Chen and R. A. Webb, Full shot noise in mesoscopic tunnel barriers, Phys. Rev.

B 73, 035424 (2006).

[114] A. Lassl, P. Schlagheck, and K. Richter, Effects of short-range interactions on trans-

port through quantum point contacts: A numerical approach, Phys. Rev. B 75, 045346

(2007).

[115] P. Jaksch, I. Yakimenko, and K.-F. Berggren, From quantum point contacts to quan-

tum wires: Density-functional calculations with exchange and correlation effects, Phys.

Rev. B 74, 235320 (2006).

[116] T. Rejec and Y. Meir, Magnetic impurity formation in quantum point contacts, Nature

442, 900 (2006).

[117] A. Golub, T. Aono, and Y. Meir, Suppression of shot noise in quantum point contacts

in the ‘0.7 regime’, Phys. Rev. Lett. 97, 186801 (2006).

[118] M. P. Anantram and S. Datta, Current fluctuations in mesoscopic systems with An-

dreev scattering, Phys. Rev. B 53, 16390 (1996).

[119] T. Martin, Wave packet approach to noise in n-s junctions, Phys. Rev. A 220, 137

(1996).

[120] J. Torres and T. Martin, Positive and negative Hanbury-Brown and Twiss correlations

in normal metal-superconducting devices, Eur. Phys. J. B 12, 319 (1999).

178

Page 193: Waves, Particles, and Interactions in Reduced Dimensions

[121] A. M. Martin and M. Buttiker, Coulomb-induced positive current-current correlations

in normal conductors, Phys. Rev. Lett. 84, 3386 (2000).

[122] I. Safi, P. Devillard, and T. Martin, Partition noise and statistics in the fractional

quantum Hall effect, Phys. Rev. Lett. 86, 4628 (2001).

[123] A. Crepieux, R. Guyon, P. Devillard, and T. Martin, Electron injection in a nanotube:

noise correlations and entanglement, Phys. Rev. B 67, 205408 (2003).

[124] M. Buttiker, Reversing the sign of current-current correlations, in Quantum noise

in mesoscopic physics, edited by Y. V. Nazarov, NATO Science Series II, Vol. 97

(Kluwer, Dordrecht, 2003).

[125] A. Cottet, W. Belzig, and C. Bruder, Positive cross correlations in a three-terminal

quantum dot with ferromagnetic contacts, Phys. Rev. Lett. 92, 206801 (2004).

[126] A. Cottet, W. Belzig, and C. Bruder, Positive cross-correlations due to dynamical

channel blockade in a three-terminal quantum dot, Phys. Rev. B 70, 115315 (2004).

[127] C. Texier and M. Buttiker, Effect of incoherent scattering on shot noise correlations

in the quantum Hall regime, Phys. Rev. B 62, 7454 (2000).

[128] S.-T. Wu and S. Yip, Feedback effects on the current correlations in Y-shaped conduc-

tors, Phys. Rev. B 72, 153101 (2005).

[129] V. Rychkov and M. Buttiker, Mesoscopic versus macroscopic division of current fluc-

tuations, Phys. Rev. Lett. 96, 166806 (2006).

[130] Y. Chen and R. A. Webb, Positive current correlations associated with super-

Poissonian shot noise, Phys. Rev. Lett. 97, 066604 (2006).

179

Page 194: Waves, Particles, and Interactions in Reduced Dimensions

[131] L. DiCarlo, Y. Zhang, D. T. McClure, C. M. Marcus, L. N. Pfeiffer, and K. W. West,

System for measuring auto- and cross correlation of current noise at low temperatures,

Rev. Sci. Instrum. 77, 073906 (2006).

[132] W. G. van der Wiel, Y. V. Nazarov, S. De Franceschi, T. Fujisawa, J. M. Elzer-

man, E. W. G. M. Huizeling, S. Tarucha, and L. P. Kouwenhoven, Electromag-

netic Aharonov-Bohm effect in a two-dimensional electron gas ring, Phys. Rev. B

67, 033307 (2003).

[133] I. H. Chan, R. M. Westervelt, K. D. Maranowski, and A. C. Gossard, Strongly capac-

itively coupled quantum dots, Appl. Phys. Lett. 80, 1818 (2002).

[134] L. P. Kouwenhoven, C. M. Marcus, P. L. McEuen, S. Tarucha, R. M. Westervelt, and

N. S. Wingreen, Electron transport in quantum dots, in Mesoscopic Electron Transport,

edited by L. L. Sohn, L. P. Kouwenhoven, and G. Schon.

[135] S. Hershfield, J. H. Davies, P. Hyldgaard, C. J. Stanton, and J. W. Wilkins, Zero-

frequency current noise for the double-tunnel-junction Coulomb blockade, Phys. Rev.

B 47, 1967 (1993).

[136] M. Eto, Nonequilibrium transport properties through parallel quantum dots, Jpn. J.

Appl. Phys. 36, 4004 (1997).

[137] G. Kießlich, A. Wacker, and E. Scholl, Shot noise of coupled semiconductor quantum

dots, Phys. Rev. B 68, 125320 (2003).

[138] C. W. J. Beenakker, Theory of Coulomb-blockade oscillations in the conductance of a

quantum dot, Phys. Rev. B 44, 1646 (1991).

180

Page 195: Waves, Particles, and Interactions in Reduced Dimensions

[139] W. Belzig, Full counting statistics of super-Poissonian shot noise in multilevel quan-

tum dots, Phys. Rev. B 71, 161301(R) (2005).

[140] E. V. Sukhorukov, G. Burkard, and D. Loss, Noise of a quantum dot system in the

cotunneling regime, Phys. Rev. B 63, 125315 (2001).

[141] A. Thielmann, M. H. Hettler, J. Konig, and G. Schon, Cotunneling current and shot

noise in quantum dots, Phys. Rev. Lett. 95, 146806 (2005).

[142] G. Iannaccone, G. Lombardi, M. Macucci, and B. Pellegrini, Enhanced shot noise in

resonant tunneling: Theory and experiment, Phys. Rev. Lett. 80, 1054 (1998).

[143] O. Zarchin, Y. C. Chung, M. Heiblum, D. Rohrlich, and V. Umansky, Electron bunch-

ing in transport through quantum dots in a high magnetic field, Phys. Rev. Lett. 98,

066801 (2007).

[144] H.-A. Engel, Electron spins in dots and rings: coherence, read out, and transport,

Ph.D. thesis, University of Basel (2003).

[145] A. K. Geim and K. S. Novoselov, The rise of graphene, Nat. Mater. 6, 183 (2007).

[146] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim, The

electronic properties of graphene, arXiv:0709.1163.

[147] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos,

I. V. Grigorieva, and A. A. Firsov, Electric field effect in atomically thin carbon films,

Science 306, 666 (2004).

[148] J. Tworzyd lo, B. Trauzettel, M. Titov, A. Rycerz, and C. W. J. Beenakker, Sub-

Poissonian shot noise in graphene, Phys. Rev. Lett. 96, 246802 (2006).

181

Page 196: Waves, Particles, and Interactions in Reduced Dimensions

[149] V. V. Cheianov and V. I. Fal’ko, Selective transmission of Dirac electrons and ballistic

magnetoresistance of n-p junctions in graphene, Phys. Rev. B 74, 041403(R) (2006).

[150] P. San-Jose, E. Prada, and D. S. Golubev, Universal scaling of current fluctuations

in disordered graphene, arXiv:0706.3832.

[151] C. W. J. Beenakker and M. Buttiker, Suppression of shot noise in metallic diffusive

conductors, Phys. Rev. B 46, 1889(R) (1992).

[152] M. J. M. de Jong and C. W. J. Beenakker, Mesoscopic fluctuations in the shot-noise

power of metals, Phys. Rev. B 46, 13400 (1992).

[153] Y. V. Nazarov, Limits of universality in disordered conductors, Phys. Rev. Lett. 73,

134 (1994).

[154] C. H. Lewenkopf, E. R. Mucciolo, and A. H. Castro Neto, Conductivity and Fano

factor in disordered graphene, arXiv:0711.3202.

[155] J. R. Williams, L. DiCarlo, and C. M. Marcus, Quantum Hall effect in a gate-controlled

p-n junction of graphene, Science 317, 638 (2007).

[156] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, M. I. Katsnelson, I. V.

Grigorieva, S. V. Dubonos, and A. A. Firsov, Two-dimensional gas of massless Dirac

fermions in graphene, Nature 438, 197 (2005).

[157] Y. Zhang, Y.-W. Tan, H. L. Stormer, and P. Kim, Experimental observation of the

quantum Hall effect and Berry’s phase in graphene, Nature 438, 201 (2005).

[158] E. McCann and V. I. Fal’ko, Landau-level degeneracy and quantum Hall effect in a

graphite bilayer, Phys. Rev. Lett. 96, 086805 (2006).

182

Page 197: Waves, Particles, and Interactions in Reduced Dimensions

[159] J. Martin, N. Akerman, G. Ulbricht, T. Lohmann, J. H. Smet, K. von Klitzing, and

A. Yacoby, Observation of electron-hole puddles in graphene using a scanning single

electron transistor, arXiv:0705.2180.

[160] A. Rycerz, J. Tworzyd lo, and C. W. J. Beenakker, Anomalously large conductance

fluctuations in weakly disordered graphene, Europhys. Lett. 79, 57003 (2007).

[161] D. Graf, F. Molitor, T. Ihn, and K. Ensslin, Phase-coherent transport measured in a

side-gated mesoscopic graphite wire, Phys. Rev. B 75, 245429 (2007).

[162] Y. Zhang, J. P. Small, M. E. S. Amori, and P. Kim, Electric field modulation of gal-

vanomagnetic properties of mesoscopic graphite, Phys. Rev. Lett. 94, 176803 (2005).

[163] I. Snyman and C. W. J. Beenakker, Ballistic transmission through a graphene bilayer,

Phys. Rev. B 75, 045322 (2007).

[164] P. Roulleau, F. Portier, D. C. Glattli, P. Roche, A. Cavanna, G. Faini, U. Gennser,

and D. Mailly, Finite bias visibility of the electronic mach-zehnder interferometer,

Phys. Rev. B 76, 161309 (2007).

[165] S. M. Cronenwett, S. R. Patel, C. M. Marcus, K. Campman, and A. C. Gossard,

Mesoscopic fluctuations of elastic cotunneling in Coulomb blockaded quantum dots,

Phys. Rev. Lett. 79, 2312 (1997).

[166] R. Ilan, E. Grosfeld, and A. Stern, Coulomb blockade as a probe for non-Abelian

statistics in Read-Rezayi states, Phys. Rev. Lett. 100, 086803 (2008).

[167] B. J. van Wees, L. P. Kouwenhoven, C. J. P. M. Harmans, J. G. Williamson, C. E.

Timmering, M. E. I. Broekaart, C. T. Foxon, and J. J. Harris, Observation of zero-

183

Page 198: Waves, Particles, and Interactions in Reduced Dimensions

dimensional states in a one-dimensional electron interferometer, Phys. Rev. Lett. 62,

2523 (1989).

[168] R. P. Taylor, A. S. Sachrajda, P. Zawadzki, P. T. Coleridge, and J. A. Adams,

Aharonov-Bohm oscillations in the Coulomb blockade regime, Phys. Rev. Lett. 69,

1989 (1992).

[169] J. P. Bird, K. Ishibashi, Y. Aoyagi, and T. Sugano, Precise period doubling of the

Aharonov-Bohm effect in a quantum dot at high magnetic fields, Phys. Rev. B 53,

3642 (1996).

[170] W. Zhou, F. E. Camino, and V. J. Goldman, Quantum transport in an Aharonov-

Bohm electron interferometer, AIP Conf. Proc. 850, 1351 (2006).

[171] F. E. Camino, W. Zhou, and V. J. Goldman, Quantum transport in electron Fabry-

Perot interferometers, Phys. Rev. B 76, 155305 (2007).

[172] M. W. C. Dharma-wardana, R. P. Taylor, and A. S. Sachrajda, The effect of Coulomb

interactions on the magnetoconductance oscillations of quantum dots, Solid State

Commun. 84, 631 (1992).

[173] B. W. Alphenaar, A. A. M. Staring, H. van Houten, M. A. A. Mabesoone, O. J. A.

Buyk, and C. T. Foxon, Influence of adiabatically transmitted edge channels on single-

electron tunneling through a quantum dot, Phys. Rev. B 46, 7236 (1992).

[174] B. Rosenow and B. I. Halperin, Influence of interactions on flux and back-gate period

of quantum Hall interferometers, Phys. Rev. Lett. 98, 106801 (2007).

[175] S. Ihnatsenka and I. V. Zozoulenko, Interacting electrons in the Aharonov-Bohm in-

terferometer, Phys. Rev. B 77, 235304 (2008).

184

Page 199: Waves, Particles, and Interactions in Reduced Dimensions

[176] J. B. Miller, I. P. Radu, D. M. Zumbuhl, E. M. Levenson-Falk, M. A. Kastner, C. M.

Marcus, L. N. Pfeiffer, and K. W. West, Fractional quantum Hall effect in a quantum

point contact at filling fraction 5/2, Nat. Phys. 3, 561 (2007).

[177] P. V. Lin, F. E. Camino, and V. J. Goldman, Electron interferometry in quantum

Hall regime: Aharonov-Bohm effect of interacting electrons ArXiv:0902.0811.

[178] D. T. McClure, Y. Zhang, B. Rosenow, E. M. Levenson-Falk, C. M. Marcus, L. N.

Pfeiffer, and K. W. West, Edge-state velocity and coherence in a quantum Hall Fabry-

Perot interferometer ArXiv:0903.5097.

[179] F. E. Camino, W. Zhou, and V. J. Goldman, Transport in the laughlin quasiparticle

interferometer: Evidence for topological protection in an anyonic qubit, Phys. Rev. B

74, 115301 (2006).

[180] P. L. McEuen, E. B. Foxman, J. Kinaret, U. Meirav, M. A. Kastner, N. S. Wingreen,

and S. J. Wind, Self-consistent addition spectrum of a Coulomb island in the quantum

Hall regime, Phys. Rev. B 45, 11419 (1992).

[181] M. Keller, U. Wilhelm, J. Schmid, J. Weis, K. v. Klitzing, and K. Eberl, Quantum dot

in high magnetic fields: Correlated tunneling of electrons probes the spin configuration

at the edge of the dot, Phys. Rev. B 64, 033302 (2001).

[182] M. Stopa, W. G. van der Wiel, S. D. Franceschi, S. Tarucha, and L. P. Kouwenhoven,

Magnetically induced chessboard pattern in the conductance of a Kondo quantum dot,

Phys. Rev. Lett. 91, 046601 (2003).

[183] Y. Aharonov and D. Bohm, Significance of electromagnetic potentials in the quantum

theory, Phys. Rev. 115, 485 (1959).

185

Page 200: Waves, Particles, and Interactions in Reduced Dimensions

[184] Y. V. Nazarov, Aharonov-Bohm effect in the system of two tunnel junctions, Phys.

Rev. B 47, 2768 (1993).

[185] Y. V. Nazarov, Quantum interference, tunnel junctions and resonant tunneling inter-

ferometer, Physica b 189, 57 (1993).

[186] I. Neder, M. Heiblum, Y. Levinson, D. Mahalu, and V. Umansky, Unexpected behavior

in a two-path electron interferometer, Phys. Rev. Lett. 96, 016804 (2006).

[187] L. V. Litvin, A. Helzel, H.-P. Tranitz, W. Wegscheider, and C. Strunk, Edge-channel

interference controlled by Landau level filling, Phys. Rev. B 78, 075303 (2008).

[188] E. V. Sukhorukov and V. V. Cheianov, Resonant dephasing in the electronic mach-

zehnder interferometer, Phys. Rev. Lett. 99, 156801 (2007).

[189] J. T. Chalker, Y. Gefen, and M. Y. Veillette, Decoherence and interactions in an

electronic mach-zehnder interferometer, Phys. Rev. B 76, 085320 (2007).

[190] I. P. Levkivskyi and E. V. Sukhorukov, Dephasing in the electronic mach-zehnder

interferometer at filling factor ν = 2, Phys. Rev. B 78, 045322 (2008).

[191] I. Neder and E. Ginossar, Behavior of electronic interferometers in the nonlinear

regime, Phys. Rev. Lett. 100, 196806 (2008).

[192] R. C. Ashoori, H. L. Stormer, L. N. Pfeiffer, K. W. Baldwin, and K. West, Edge

magnetoplasmons in the time domain, Phys. Rev. B 45, 3894 (1992).

[193] G. Ernst, N. B. Zhitenev, R. J. Haug, and K. von Klitzing, Probing the edge of a

2DEG by time-resolved transport measurements, Physica E 1], pages = 95, year =

1998.

186

Page 201: Waves, Particles, and Interactions in Reduced Dimensions

[194] B. I. Halperin Private communication.

[195] B. Rosenow et al., unpublished .

[196] P. Roulleau, F. Portier, P. Roche, A. Cavanna, G. Faini, U. Gennser, and D. Mailly,

Direct measurement of the coherence length of edge states in the integer quantum Hall

regime, Phys. Rev. Lett. 100, 126802 (2008).

[197] B. L. Altshuler and A. G. Aronov, in Electron-Electron Interactions in Disordered

Systems, edited by A. L. Efros and M. Pollak (North-Holland, Amsterdam, 1985).

[198] R. J. Schoelkopf, P. Wahlgren, A. A. Kozhevnikov, P. Delsing, and D. E. Prober,

The radio-frequency single-electron transistor (RF-SET): A fast and ultrasensitive

electrometer, Science 280, 1238 (1998).

[199] M. Field, C. G. Smith, M. Pepper, D. A. Ritchie, J. E. F. Frost, G. A. C. Jones, and

D. G. Hasko, Measurements of Coulomb blockade with a noninvasive voltage probe,

Phys. Rev. Lett. 70, 1311 (1993).

[200] L. DiCarlo, H. J. Lynch, A. C. Johnson, L. I. Childress, K. Crockett, C. M. Marcus,

M. P. Hanson, and A. C. Gossard, Differential charge sensing and charge delocalization

in a tunable double quantum dot, Phys. Rev. Lett. 92, 226801 (2004).

[201] D. J. Reilly, C. M. Marcus, M. P. Hanson, and A. C. Gossard, Fast single-charge

sensing with a rf quantum point contact, Appl. Phys. Lett. 91, 162101 (2007).

[202] I. P. Radu, J. B. Miller, C. M. Marcus, M. A. Kastner, L. N. Pfeiffer, and K. W. West,

Quasi-particle properties from tunneling in the ν = 5/2 fractional quantum Hall state,

Science 320, 899 (2008).

187

Page 202: Waves, Particles, and Interactions in Reduced Dimensions

[203] X.-G. Wen, Edge transport properties of the fractional quantum Hall states and weak-

impurity scattering of a one-dimensional charge-density wave, Phys. Rev. B 44, 5708

(1991).

[204] M. Levin, B. I. Halperin, and B. Rosenow, Particle-hole symmetry and the Pfaffian

state, Phys. Rev. Lett. 99, 236806 (2007).

[205] S.-S. Lee, S. Ryu, C. Nayak, and M. P. A. Fisher, Particle-hole symmetry and the

ν = 5/2 quantum Hall state, Phys. Rev. Lett. 99, 236807 (2007).

[206] S. D. Sarma and A. Pinczuk (editors), Perspectives in Quantum Hall Effects (Johm

Wiley, New York, 1997).

[207] R. Willett, J. P. Eisenstein, H. L. Stormer, D. C. Tsui, A. C. Gossard, and J. H. En-

glish, Observation of an even-denominator quantum number in the fractional quantum

Hall effect, Phys. Rev. Lett. 59, 1776 (1987).

[208] Y. W. Suen, L. W. Engel, M. B. Santos, M. Shayegan, and D. C. Tsui, Observation

of a ν = 1/2 fractional quantum Hall state in a double-layer electron system, Phys.

Rev. Lett. 68, 1379 (1992).

[209] J. P. Eisenstein, G. S. Boebinger, L. N. Pfeiffer, K. W. West, and S. He, New fractional

quantum Hall state in double-layer two-dimensional electron systems, Phys. Rev. Lett.

68, 1383 (1992).

[210] Y. W. Suen, H. C. Manoharan, X. Ying, M. B. Santos, and M. Shayegan, Origin of

the ν = 1/2 fractional quantum Hall state in wide single quantum wells, Phys. Rev.

Lett. 72, 3405 (1994).

188

Page 203: Waves, Particles, and Interactions in Reduced Dimensions

[211] B. I. Halperin, Theory of the quantized Hall conductance, Helv. Phys. Acta 56, 75

(1983).

[212] D. Yoshioka, A. H. MacDonald, and S. M. Girvin, Fractional quantum Hall effect in

two-layered systems, Phys. Rev. B 39, 1932 (1989).

[213] M. P. Lilly, K. B. Cooper, J. P. Eisenstein, L. N. Pfeiffer, and K. W. West, Evidence

for an anisotropic state of two-dimensional electrons in high Landau levels, Phys. Rev.

Lett. 82, 394 (1999).

[214] R. R. Du, D. C. Tsui, H. L. Stormer, L. N. Pfeiffer, K. W. Baldwin, and K. W. West,

Strongly anisotropic transport in higher two-dimensional Landau levels, Solid State

Commun. 109, 389 (1999).

[215] A. A. Koulakov, M. M. Fogler, and B. I. Shklovskii, Charge density wave in two-

dimensional electron liquid in weak magnetic field, Phys. Rev. Lett. 76, 499 (1996).

[216] M. M. Fogler, A. A. Koulakov, and B. I. Shklovskii, Ground state of a two-dimensional

electron liquid in a weak magnetic field, Phys. Rev. B 54, 1853 (1996).

[217] R. Moessner and J. T. Chalker, Exact results for interacting electrons in high Landau

levels, Phys. Rev. B 54, 5006 (1996).

[218] A. C. Johnson, Charge sensing and spin dynamics in GaAs quantum dots, Ph.D.

thesis, Harvard University (2005).

[219] J. M. Martinis, M. H. Devoret, and J. Clarke, Experimental tests for the quantum

behavior of a macroscopic degree of freedom: The phase difference across a Josephson

junction, Phys. Rev. B 35, 4682 (1987).

189

Page 204: Waves, Particles, and Interactions in Reduced Dimensions

[220] F. P. Milliken, J. R. Rozen, G. A. Keefe, and R. H. Koch, 50 Ω characteristic

impedance low-pass metal powder filters, Rev. Sci. Instrum. 78, 024701 (2007).

[221] A. B. Zorin, The Thermocoax cable as the microwave frequency filter for single electron

circuits, Rev. Sci. Instrum. 66, 4296 (1995).

190