Abstract Title of Dissertation: THE RADIO-FREQUENCY SINGLE-ELECTRON TRANSISTOR DISPLACEMENT DETECTOR Matthew D. LaHaye, Doctor of Philosophy, 2005 Dissertation directed by: Adjunct Professor Keith C. Schwab Department of Physics For more than two decades, the standard quantum limit (SQL) has served as a benchmark for researchers involved in ultra-sensitive force and displacement detection. In this thesis, I discuss a novel displacement detection technique which we have implemented that has allowed us to come within a factor of 4.3 from the limit, closer than any previous e?ort. Additionally, I show that we were able to use this nearly quantum-limited scheme to observe the thermal motion of a 19.7 MHz in-plane mode of a nanomechanical resonator down to a temperature of 56 mK. At this temperature, the corresponding thermal occupation number of the mode was ?n th ??60. This is the lowest thermal occupation number ever demonstrated for a nanomechanical (or larger) device. We believe that the combination of these two results has important and promising implications for the future study of nanoelec- tromechanical systems (NEMS) at the quantum limit. The detection scheme that we used was based upon the single-electron tran- sistor (SET). The SET has been demonstrated to be the world?s most sensitive electrometer and is considered to be a near-ideal linear amplifier. We used stan- dard lithographic techniques for the on-chip integration of the SET with both a microwave-matching network and nanomechanical resonator. The SET served as a transducer of the resonator?s motion: fluctuations in the resonator?s position mod- ulated the SET impedance. The microwave-matching circuit allowed us to read-out the modulation of the SET?s impedance with ? 75 MHz bandwidth. The com- bination of microwave-matching circuit and SET is known as the radio-frequency single-electron transistor (RFSET). Including the nanomechanical resonator, the configuration is called the radio-frequency single-electron transistor displacement detector. In this thesis, I discuss the basics of quantum-limited measurement and some of the subtleties of observing mechanical quantum phenomena. I then discuss the basics of the RFSET displacement detector, its ultimate limits, its engineering and operation, the first generation results, and finally what improvements could be made to future generation devices. THE RADIO-FREQUENCY SINGLE-ELECTRON TRANSISTOR DISPLACEMENT DETECTOR by Matthew D. LaHaye Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements of the degree of Doctor of Philosophy 2005 Advisory Committee: Professor Fredrick C. Wellstood, Chair Professor Keith C. Schwab, Co-Chair Professor Christopher J. Lobb Professor James R. Anderson Professor Donald L. DeVoe c? Copyright by Matthew D. LaHaye 2005 Dedication ...tomylovelywife,Juhee,and herheart ofgold. ii Acknowledgements This thesis is dedicated to my wife, Juhee Kothari LaHaye, for her enduring support and patience throughout these six years. She bore the brunt of the long hours, lost weekends, and endless uncertainty, yet still kept a smile on her face and on mine. Juhee is the binding that holds this book together. I am forever grateful to my family, Mom, Dad, Mark, and Chris. They gave me the world, taught me the essentials, and then let me go on my way. With each success and blunder, they have stood behind me the same. I have had a great life, and they made that possible. I am truly fortunate to have such wonderful in-laws. Mom, Dad, and Pooja have o?ered to me nothing but love and enlightenment since the day we met. I am especially indebted to my advisor Professor Keith Schwab. Keith has been an inspiration, a mentor, and a friend. Along with the opportunity to travel and meet other leading researchers in the field, he provided a lab full of equipment and five years of ideas. From scratch, he built the LPS Nanomechanics group into a first-class research team. I am fortunate and proud to have been a part of this experience. Also, thanks to Drs. Bruce Kane, Keith Miller, and Marc Manheimer for their support over the years and the hard work they put into the quantum computing group. I owe many thanks to my labmates, fellow students, and sta? in both the iii physics department at the University of Maryland and at the Laboratory for Physical Sciences. Thanks to the the nanomechanics group, a very unique collection of talented and friendly individuals: Emrah Altunkaya, Harish Bhaskaran, Dr. Olivier Buu, Benedetta Camarota, Jared Hertzberg, Akshay Naik, and Patrick Truitt. Particularly, I would like to thank Dr. Olivier Buu. Olivier provided me with advice on everything from Matlab to cryogenics to life in general. He is a selfless and tireless individual who contributed greatly to my education as an experimental scientist. He played an integral role in these experiments, assembling the dilution refrigerator and SET-Gain feedback circuit, and sharing the measurement responsi- bilities with me. Also, I would like to thank Benedetta Camarota for many, many insightful conversations, much sound advice, fabrication support, and half-a-million laughs. Benedetta also played an essential role in developing the device fabrication recipe. Finally, thanks to Jared Hertzberg for reading drafts of this thesis and always o?ering his opinion or a helping hand. I must also take the opportunity to thank Dan Sullivan, Carlos Sanchez, and Kenton Brown for their support, integrity, and friendship. Essentially, we all started the program together and have been through many years of ups and downs. Dan?s humor, wit and insight into everything from physics to politics to sports will not be forgotten. Carlos was always willing to share his extensive knowledge, whether it was music, politics, electronics or fabrication. I must particularly recognize the advice he provided for SET fabrication. Kenton?s intelligence and generosity will iv also be greatly missed. I would like to thank Josh Higgins, Ben Palmer, Jonghee Lee, and Nathan Siwak for both friendly and technical conversations over the years. I am grateful for the technical support and machine work that Les Lorenz, J.B Dottellis, and Russell Frizzell provided. As well, I must thank them for many pleasant and informative conversations. Thanks to Toby Olver for keeping the cleanroom running and being so patient with my requests. Thanks to Butch Bilger for keeping the Helium flowing. I would also like to thank Jane Hessing for her work over the years keeping me up-to-date on paperwork, informed of deadlines, and answering my questions. As well I must thank Margaret Lukomska for her work expediting purchase orders, travel requests and various other questions. I would like to thank Profs. Miles Blencowe and Aashish Clerk for their help clarifying the subtleties of SET back action and the quantum limit. Finally, I would like to sincerely thank the members of my thesis committee, Profs. Keith Schwab, Fred Wellstood, Bob Anderson, Chris Lobb, and Don DeVoe for taking the time to read through 200 pages of technical information and run-on sentences, as well as to participate in my defense and provide very helpful feed- back. I am especially grateful to Prof. Wellstood for accepting the responsibility of ?Chairman? and meeting with me on multiple occasions to discuss corrections. v TABLE OF CONTENTS List of Figures ix List of Tables xiv List of Symbols xv 1Overview 1 1.1 ContextandMotivation ......................... 1 1.2 StructureoftheThesis.......................... 8 2 Introduction 11 2.1 TheQuantumLimitI:ThermalNoise.................. 12 2.2 TheQuantumLimitII:IdealDetection................. 20 2.3 Nanomechanical RFSET Displacement Detection . . . . . . . . . . . 38 3 Design and Fabrication 56 3.1 TheWafers ................................ 56 3.2 Silicon Nitride Membrane Fabrication . . . . . . . . . . . . . . . . . . 56 3.3 Fabrication of the Bond Pads and Tank Circuit . . . . . . . . . . . . 59 3.4 Fabrication of the Nanomechanical Resonator and SET . . . . . . . . 70 4 Apparatus 82 4.1 TheSamplePackage ........................... 82 4.2 TheDilutionRefrigeratorandShieldedRoom ............. 84 4.3 RefrigeratorWiring............................ 90 vi 5 The RFSET 102 5.1 RFSETReflectometry ..........................102 5.2 Gain Stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 6 RFSET Displacement Detection 124 6.1 Methodology ...............................126 6.2 MechanicalNoiseThermometry.....................140 6.3 DiscussionofSETBackAction .....................157 7 Conclusions 169 7.1 ShotNoiseLimitedDetection ......................169 7.2 SampleThermalization..........................174 7.3 PartingMotivation ............................187 A Useful Mechanics Information 188 A.1 Euler-Bernoulli Theory and The Simple Harmonic Oscillator . . . . . 188 A.2 SpringConstantsforSETDetection...................193 A.3 CorrectionstoFrequencyDuetoTension................197 A.4 The Driven-Damped Harmonic Oscillator . . . . . . . . . . . . . . . . 201 A.5 TheMagnetomotiveTechnique .....................203 B Useful SET and RFSET Information 208 B.1 SETParameters..............................208 B.2 MeasurementCircuitParameters ....................227 B.3 CalibrationofChargeSensitivity ....................232 vii Bibliography 238 viii LIST OF FIGURES 1.1 Generic circuit schematic and SEM image of an RFSET displacement detector................................... 6 2.1 Thermal occupation factor for typical nanoresonators . . . . . . . . . 14 2.2 Estimated decay time for a nanoresonator in a superposition state . . 16 2.3 Generic circuit schematic and SEM image of an RFSET displacement detector................................... 40 2.4 Ultimate detection limit of a normal-state SET displacement detector 51 2.5 Ultimate detection limit of a normal-state SET displacement detector 53 3.1 Silicon nitride membrane fabrication . . . . . . . . . . . . . . . . . . 57 3.2 SEM image of RFSET displacement detector: tank circuit and bond pads .................................... 67 3.3 SEM image of RFSET tank circuit inductor . . . . . . . . . . . . . . 68 3.4 Double-angleevaporation......................... 76 3.5 Fabricationofthenanoresonator..................... 79 3.6 SEM image of two RFSET displacement detectors . . . . . . . . . . . 80 3.7 SEM image of an RFSET displacement detector: Device 2 . . . . . . 81 4.1 Imageofasamplepackage......................... 83 4.2 Image of the dilution refrigerator . . . . . . . . . . . . . . . . . . . . 85 4.3 Images of the dilution refrigerator mixing chamber and sample stage . 86 ix 4.4 Picture of shielded room and optical table. The 4 He dewar is visible behind the front pillar. A winch is used to raise the dewar over the refrigerator. ................................ 87 4.5 Picture of the top of the optical table. DC connections at the top of the fridge are made through ?-powder filters. Battery-powered pre- amps and voltage sources as well as break-out boxes are shelved on theelectronicsrack............................. 88 4.6 Image of ?-powderfilterandtransfercharacteristics.......... 89 4.7 Circuit schematic of wiring in the shielded room and refrigerator . . . 91 4.8 Transfercharacteristicsofpowderfilters ................ 92 4.9 Transfercharacteristicofgateline.................... 93 4.10 Image and transfer characteristics of microstrip heat sink . . . . . . . 96 4.11 Transfer characteristics of the microwave circuit: input. . . . . . . . . 98 4.12 Transfer characteristics of microwave circuit: output. . . . . . . . . . 99 5.1 CircuitschematicforRFSETreflectometry...............103 5.2 Plot of reflection coe?cient and I SD versus Q g .............106 5.3 RFSET reflectometry: plot of 1 MHz sideband modulation . . . . . . 110 5.4 I SD V SD V g map in the superconducting state. . . . . . . . . . . . . . . 113 5.5 ThederivativemapandRFSETgain..................114 5.6 Plot of sideband dependence on V g ...................117 5.7 Modulation of sideband amplitude for RFSET gain stabilization . . . 118 5.8 Circuit schematic for RFSET gain stabilization . . . . . . . . . . . . 120 x 5.9 Circuitschematicforaudiofeedbackcircuit ..............121 5.10 Plot demonstrating RFSET gain stabilization . . . . . . . . . . . . . 123 6.1 Circuit schematic of RFSET displacement detector . . . . . . . . . . 125 6.2 Circuit schematic for RFSET detection of capacitivelyexcited nanores- onators...................................130 6.3 Response of capacitively excited nanoresonator measured with RF- SET,Device2...............................133 6.4 I SD V SD V NR mapofDevice2 ......................134 6.5 Measurement of time-domain response of a nanoresonator, Device 2 . 137 6.6 FFToftime-domainresponse,Device2.................138 6.7 Power spectrum measurements of a nanoresonators thermal motion . 145 6.8 Plot of a nanoresonator?s integrated thermal response versus refrig- eratortemperature ............................147 6.9 Plot of a nanoresonator?s integrated thermal response versus refrig- erator temperature and coupling voltage: Device 2 . . . . . . . . . . . 149 6.10 Plot of a nanoresonator?s integrated thermal response versus refrig- erator temperature and coupling voltage: Device 1 . . . . . . . . . . . 150 6.11 E?ective resonator temperature, T e =56mK:Device2........151 6.12 Plot of the integrated resonator response scaled with coupling and measurementcircuitgain:Devices1and2 ...............153 6.13 Plot of the lowest noise temperature achieved for RFSET displace- mentdetection:Device2.........................155 xi 6.14 Plot of noise temperature versus coupling voltage: Device 1 . . . . . . 158 6.15 Plot of noise temperature versus coupling voltage: Device 2 . . . . . . 159 6.16 Plot of nanoresonator frequency shift versus coupling voltage . . . . . 161 6.17 Plots of nanoresonator e?ective damping versus temperature . . . . . 163 6.18 Plot of damping versus coupling: Devices 1 and 2 . . . . . . . . . . . 165 7.1 Noise temperature versus coupling: the shot noise limit. . . . . . . . . 170 7.2 Illustration of Nb-Nb semi-rigid coax for heat-flow calculation. . . . . 176 7.3 Thermal circuit for the SET and resonator on the SiN membrane . . 181 7.4 Numerical calculation of the temperature of phonons near the nanores- onator as a function of bath temperature and power dissipated in the SET ....................................183 A.1 Schematic of a prismatic, doubly-clamped nanoresonator . . . . . . . 189 A.2 Plot of the first four transverse modes of a doubly-clamped resonator 191 A.3 Plot of the e?ect of tension on the frequency of the first six modes of adoubly-clampednanoresonator.....................199 A.4 Schematic of magnetomotive detection . . . . . . . . . . . . . . . . . 204 A.5 Plot of the response of a nanoresonator measured using magnetomo- tivedetection................................206 B.1 Circuit diagram for I SD V SD V g measurement..............209 B.2 Normal-state I SD V SD V g map ......................210 B.3 Extraction of SET capacitance parameters: normal-state . . . . . . . 215 xii B.4 ExtractionofSETresistanceparameters ................217 B.5 Failure of the normal-state SET parameter extraction . . . . . . . . . 220 B.6 Superconducting I SD V SD V g map ....................222 B.7 Extraction of SET junction capacitance: superconducting state . . . . 224 B.8 Circuit schematic for the measurement of the shot noise ring-up of theRFtankcircuit............................228 B.9 Tankcircuitresponse...........................230 B.10Experimentaldeterminationofchargesensitivity............233 B.11 Charge sensitivity calibration: the Bessel function fit . . . . . . . . . 236 xiii LIST OF TABLES 3.1 Tableoftankcircuitparameters:Devices1-4............. 69 3.2 Table of geometric parameters: Devices 1 - 4 . . . . . . . . . . . . . . 73 3.3 Table of e?ective spring constants: Devices 1 - 4 . . . . . . . . . . . . 73 3.4 TableofSETjunctionparameters:Devices1-4............. 77 3.5 Table of SET charging energy and gap energy: Devices 1-4 . . . . . . 77 4.1 Table of measurement circuit parameters: Devices 1 - 4 . . . . . . . . 100 A.1 Table of geometric constants for first six modes of a doubly-clamped nanoresonator...............................192 A.2 Table of geometric parameters: Devices 1 - 4 . . . . . . . . . . . . . . 196 A.3 Table of e?ective spring constants: Devices 1 - 4 . . . . . . . . . . . . 196 B.1 TableofSETjunctionparameters:Devices1-4.............226 B.2 Table of SET charging energy and gap energy: Devices 1-4 . . . . . . 226 B.3 Table of measurement circuit characteristics: Devices 1 - 4 . . . . . . 231 B.4 Tableofchargemodulation:Devices1-4 ...............237 B.5 Tableofchargesensitivities:Devices1-4 ................237 xiv LIST OF SYMBOLS Symbol Definition a 1 Relationship between resonator?s mid-point motion and the average motion over the length of the SET island ?a ?,? ,?a ? ?,? Lowering and raising operators of the input signal into ideal linear amplifier A Interaction strength between resonator and detector; also charge-to-voltage ratio for SET gate in electrons; total linear amplifier noise contribution in units of quanta; am- plitude of resonator?s thermal response; cross-sectional area of resonator A i Linear amplifier noise contribution referred to input for each quadrature in units of quanta; also cross-sectional area of SET junctions A Q Charge gain of RFSET electrometer based upon the 1 MHz calibration signal A r Amplitude of reflected microwave signal b Geometric constant for the first derivative of C NR with respect to y avg b Geometric constant for the first derivative of C NR with respect to y avg B Bandwidth; magnetic field magnitude xv Symbol Definition ? b ?,? , ? b ? ?,? Lowering and raising operators of the output signal from ideal linear amplifier ?b Total rms noise in the amplitude at the output of ideal linear amplifier ?b i Total rms noise for each quadrature at the output of ideal linear amplifier C g SET gate capacitance C gnr Capacitance between SET gate and resonator C i Individual SET junction capacitances C n E?ective electromechanical capacitance for magnetomo- tive detection, n th mode C NR Capacitance of resonator to SET island C T Tank circuit capacitance C ? Total SET island capacitance ? Quasiparticle gap energy of SET leads and islands (as- sumed to be equal) d gnr Spatial separation between SET gate and resonator d NR Spatial separation between SET island and resonator DJQP Double Josephson Quasiparticle Resonance ?epsilon1 Energy sensitivity of RFSET plus measurement circuit ?epsilon1 SET Energy sensitivity of SET xvi Symbol Definition e magnitude of electron charge ? 1 Projection of force on resonator?s fundamental mode E Young?s Modulus of resonator E C Charging energy of SET E J Josephson energy of SET junctions ?f Half-width of tank circuit resonance, bandwidth of RF- SET detection scheme; also used as noise equivalent bandwidth of the resonator F i Noise contribution to each quadrature of input signal from ideal linear amplifier ?F i RMS noise contribution to each quadrature of input sig- nal from ideal linear amplifier ?F ? i Change in free-energy of SET circuit when an electron tunnels on or o? of the SET island ? b Damping of the resonator?s fundamental mode due to dissipation in the thermal bath T bath ? d Damping of the resonator?s fundamental mode due to dissipation in the e?etive thermal bath T det ? e Total e?ective damping including both damping from en- vironment and SET xvii Symbol Definition ? Reflection coe?cient; also decoherence time of superpo- sition of coherent states ? max Reflection coe?cient at maximum SET conductance ? ? i Tunneling rates onto and o? of SET island through SET junctions G Gain of measurement circuit including pre-amplifiers and loss ?h Planck?s constant divided by 2? I Moment of inertia for transverse, in-plane vibration of rectangular bar IR Integrated resonator response I SD SET source-drain current JQP Josephson Quasiparticle Resonance K m Spring constant of resonator defined with respect to av- erage motion over the length of the SET island for the fundamental mode k B Boltzman?s constant k n Roots of the characteristic equation for the transverse vibration of a doubly-clamped rectangular bar K 1 Spring constant of resonator defined with respect to mid- point motion for the fundamental mode xviii Symbol Definition ? Admittance or response of detector current to resonator displacement; also wavelength of microwave applied to RFSET; also roots for characteristic equation for the transverse vibration of a doubly-clamped rectangular bar under tension L Length of resonator; also, gain of linear amplifier in num- ber quanta L n E?ective electromechanical inductance for magnetomo- tive detection, n th mode L T Tank circuit inductance m Geometric mass of the resonator M Depth of modulation of reflection coe?cient M m E?ective mass of resonator with respect to average dis- placement over the length of the SET island for the fun- damental mode m eff E?ective mass of resonator with respect to mid-point mo- tion for the fundamental mode m i Slopes of JQP resonance; and slope of the onset of tun- neling in the normal state n Number of electrons on SET island; also mode number of transverse, in-plane resonator motion xix Symbol Definition N a Number of quanta in the input signal to ideal linear am- plifier N b Number of quanta in the output signal from ideal linear amplifier ?n th ? Thermal occupation number of resonator?s fundamental mode ? Volume of SET island ? 1 Resonator?s fundamental mode frequency multiplied by 2? ? T Resonant frequency of tank circuit multiplied by 2? P Resonator?s thermal response P o Background or noise-floor of detection scheme P 1MHz Power in 1 MHz sideband gain calibration ? Potential of SET island Q Total e?ective quality factor of tank circuit ? Q Heat transfer down Nb microwave coax and dissipation in SET ? Q 1 Power conducted to bath through SiN membrane ? Q 2 Power conducted through SiN membrane from SET to region near nanoresonator xx Symbol Definition ? Q 3 Power conducted through SiN membrane from region near nanoresonator to bath ? Q 4 Power conducted through electron-phonon coupling to electrons in Au film on top of the nanoresonator Q b Quality factor of the resonator?s fundamental mode, ac- counting only for ? b Q d Quality factor of the resonator?s fundamental mode, ac- counting only for ? d Q e Total e?ective resonator quality factor Q g Polarization charge on SET gate electrode ?Q g Charge modulation on SET gate Q NR Polarization charge on resonator electrode Q S Quality factor of tank circuit due to loading from SET Q T Quality factor of tank circuit due to loading from trans- mission line RFSET Radio Frequency Single-Electron Transistor ? Mass density of resonator R i Individual SET junction resistances R n Real-component of electromechanical impedance for magnetomotive detection, n th mode R Q Resistance quantum xxi Symbol Definition R ? Total resistance of SET junctions at large V SD R S Total resistance of SET junctions SET Single-Electron Transistor ? S F Symmetrized back action noise spectral density of linear detector ? S F b , S f Symmetrized thermal noise spectral density S y F Back action contribution to the spectral displacement sensitivity ? S I Symmetrized forward coupling noise spectral density of linear detector ? Electron-phonon coupling ? int Intrinsic axial stress in SiN film S II SET shot noise spectral density S I? Correlations between SET shot noise and island-potential fluctuations S y I Shot noise contribution to the spectral displacement sen- sitivity SNR Signal-to-noise ratio of 1 MHz sideband gain calibration S ?? SET island-potential fluctuations spectral density S Q , S Q NR Spectral density of charge sensitivity in units of e 2 /Hz S y Total spectral displacement sensitivity xxii Symbol Definition ? E?ective resonator damping time; also used as measure- ment time T Intrinsic tension on SiN resonator t Au Thickness of resonator?s Au layer T b Temperature of the thermal bath, the?environment? or ?external bath? T d E?ective temperature of the SET noise environment T e Total e?ective temperature of the resonator due to con- tact to T bath and T det T island SET island electron temperature T N Noise temperature of displacement detection scheme T det n Noise temperature contribution of measurement circuit including pre-amplifier and circuit loss T o Noise temperature of measurement circuit plus the SET shot noise contribution T QL Quantum limit of noise temperature for continuous linear detection T S Temperature of the dilution refrigerator?s sample stage T SET Noise temperature contribution of the SET shot noise t SiN Thickness of resonator?s SiN layer v c Incident microwave signal, or carrier xxiii Symbol Definition V D Capacitive excitation voltage V g SET gate bias ?V g Voltage modulation on SET gate v n Average transverse, in-plane velocity of doubly-clamped rectangular bar over the bar?s length, n th mode V NR Resonator bias, also referred to as coupling voltage v r Reflected microwave signal V SD SET source-drain bias V ?,i t Threshold voltage for tunneling of electrons onto (o? of) the SET island through junctions i, normal state SET w Width of resonator ? frequency in rad/s ? 1 fundamental mode frequency in rad/s X i Signal quadratures at the input of ideal linear amplifier Y n Eigenfunctions of the wave-equation for in-plane dis- placement of the resonator ?y QL Quantum limit for continuous linear position detection ?y ZP Zero-point fluctuations of mechanical resonator for in- plane, fundamental mode y 1 In-plane displacement of the resonator?s mid-point for the fundamental mode xxiv Symbol Definition y 1,m , y m Average, in-plane displacement of resonator over the length of the SET island for the fundamental mode ?y, ?y m Total displacement sensitivity in resonator?s noise- equivalent bandwidth Z o 50 ? Transmission line impedance Z LC Characteristic impedance of tank circuit Z LCR Impedance of tank circuit and SET at the resonant fre- quency of the tank circuit Z m Electromechanical impedance for magnetomotive detec- tion xxv Chapter 1 Overview 1.1 Context and Motivation The Heisenberg uncertainty principle [1] places a limit on the precision with which one can measure an object?s position [2]. For the case of two successive measurements of a mass M undergoing simple harmonic motion, this limit, known as the ?standard quantum limit?, is neatly expressed as [2] ?y SQL = radicalBigg ?h 2M? , (1.1) where ?/2? is the frequency with which the mass oscillates, and ?h is Planck?s con- stant. Since the 1970?s, researchers have been engaged in both theoretical and experi- mental e?orts to understand and implement mechanical detectors and measurement strategies for displacement detection at (or even below) the standard quantum limit [2-17]. Inititally, the impetus for quantum-limited displacement detection arose out of the hunt for gravitational waves [2]. Through nearly three decades of e?ort, the gravitational-wave community has moved quantum-limited detectors from mere thought-experiments to nearly practicable measurement devices. For example, the 4 km L1 interferometer of the LIGO I project has demonstrated a sensitivity, at 200 Hz, of ?y ? 150?y SQL [11] for the displacement detection of its 10 kg test- 1 masses. Researchers at Laboratoire Kastler Brossel, employing a tabletop Fabry- Perot interferometer, recently achieved a displacement sensitivity of ?y ? 25 ?y SQL for the read-out of the 2 MHz surface modes of a silica mirror [7]. The SQUID-based amplifiers developed for the Auriga project have demonstrated noise temperatures of ? 10?s ?K, corresponding to sensitivities of ?y ? 100 - 200 ?y SQL for the read- out of the vibrational modes of ? 2000 kg acoustic bar resonators [9]. As well, researchers in the Supeconductivity Center at the University of Maryland used a scheme based on a Paik-style transducer [19] to achieve a noise temperature of ? 1 mK at 900 Hz, yielding ?y ? 200 ?y SQL [8]. In the last decade, the development of nanoelectromechanical systems (NEMS) has generated a second wave of interest in the standard quantum limit. Driven by potential applications to ultra-sensitive imaging [20] [21], mass detection [22], and quantum computing [23] [24], as well as, ultimately, the possibility to study mechanical quantum systems in the macroscopic limit [23-33], the NEMS community has rapidly pushed mechanical transduction to the quantum frontier. In the last year alone, several important results have been generated. For example, researchers at IBM used magnetic resonance force microscopy (MRFM) [20] to detect the spin of a single electron [21]. Using a magnetomotive technique [22], researchers at the California Institute of Technology demonstrated mass sensivity on the order of zeptograms, su?cient for the detection of a single molecule [34]. Finally, our group in the Laboratory for Physical Sciences at the University of Maryland used the radio- frequency single-electron transistor displacement detector [12] [13] to demonstrate both displacement sensitivity approaching closer to the standard quantum limit 2 than any previous measurement scheme (?y ? 5.8 ?y SQL 1 ) [15] and an approach to low thermal occupation numbers (?n th ??60) for a 20 MHz nanomechanical resonator[15]. It is important to note that the achievement of low thermal occupation num- bers is a general and significant point of distinction between NEMS devices and the resonators used in gravitational-wave detection. We can see why this is by first looking at the definition of the thermal occupation number. For a resonant mode with frequency ? in thermal equilibrium with a bath of temperature T, the mode?s thermal occupation number is given by [35] ?n th ? = 1 2 +(e ?h?/k B T ?1) ?1 , (1.2) where k B is Boltzman?s constant and ? 1 2 ? accounts for the mode?s zero-point fluc- tuations. This quantity provides a simple ?rule-of-thumb? for gauging whether one should be able to observe a mode?s quantum properties: k B T ?h? ? 1, (1.3) If Eq. 1.3 is satisfied, the mode is said to be ?frozen out?. That is, the mode is in it?s ground state and the contribution of the thermal energy to the mode?s total energy is comparable to or less than the zero-point contribution. As k B T/?h? grows, so too does the contribution from thermal fluctuations, making it more di?cult to observe the mode?s quantum attributes. There is no general prescription for how small the ratio k B T/?h? must be before quantum behavior becomes observable (see Chapter 1 In Chapter 2, I make the distinction between ?y SQL and ?y QL . In terms of ?y QL ,amore appropriate gauge for continuous position detection, we achieved ?y ? 4.3?y QL 3 2). For the purpose of observing the resonator in a pure quantum state such as a Fock state or superposition state, or for detection of the resonator?s zero-point fluctuations, the smaller the ratio the better (see Chapter 2). For the above-mentioned gravitational-wave detectors, the operating temper- atures were su?ciently high (> 1 K) and resonators? frequencies were su?ciently low (< 5MHz), that, at a minimum, ?n th ??3x10 6 (the Auriga project at 1.5 K and the Fabry-Perot scheme at 300 K). In contrast, because NEMS devices have demonstrated resonant frequencies as high as ? 1 GHz [36] [23] and are routinely installed on cryogenic probes for measurement at mK temperatures, it should be possible for researchers to observe ?n th ??1. The demonstration of nearly quantum-limited position detection and low ther- mal occupation numbers promises NEMS researchers the opportunity to push the study of quantum mechanics to a significantly larger realm. For example, one recent proposal to prepare and measure a nanomechanical resonator mode in a superposi- tion of position states could be implemented if one could cool the mode to ?n th ??50 [25] (please see Refs. [23-33] for other recent proposals). This is signficant because, while NEMS devices are, by definition, nanoscopic, they are typically composed of ? 10 10 atoms. With a few exceptions, such as the measurement of the quantum of thermal conductance [37], previous demonstrations of mechanical quantum phe- nomena have been limited to the scale of molecules and atoms (for exmaple Refs. [38] [39] [40]). In this thesis, I discuss the details of the first generation of radio-frequency single-electron transistor (RFSET) displacement detectors. The technique was first 4 proposed by Miles Blencowe and Martin Wybourne [12] and utilizes the RFSET?s large bandwidth (demonstrated to be > 100 MHz) [41] and near-ideal noise char- acteristics [42] to perform displacement detection near the quantum limit. Figure 2.3(a) shows an SEM image of an RFSET displacement detector, and Fig. 2.3(b) shows a generic circuit schematic for the transduction process. Here, a metallized SiN nanoresonator is positioned within 1 ?m of an SET island, resulting in a cou- pling capacitance C NR on the order of 10?s aF. Displacement of the nanoresonator from its equilibrium position linearly modulates the coupling capacitance through ?C NR ? C NR d NR ?y, (1.4) where d NR is the separation between the nanoresonator and the SET island, and ?y lessmuch d NR is the displacement of the resonator from equilibrium. Establishing a voltage V NR between the resonator and the SET converts the capacitance fluctuations into charge fluctuations: ?Q NR ? C NR V NR d NR ?y. (1.5) The charge fluctuations modulate the SET impedance which is then monitored by performing microwave reflectometry [41]. The use of an on-chip tank circuit (L T and C T in Fig. 2.3(b)) allows for matching between the large SET impedance (typically 10?s k?) and 50 ? transmission line. Ultimately, the sensitivity of the RFSET displacement detector is limited by the intrinsic noise of the SET [12] [13]. This is composed of two sources [43] (1) the SET shot noise and (2) the potential fluctuations of the SET island. The SET shot noise is forward coupling. That is, it simply adds to the signal, resulting in 5 (a) (b) Figure 1.1: (a) SEM image of the RFSET displacement detector and (b) Circuit schematic 6 a contribution to the total displacement noise that is inversely proportional to the coupling C NR V NR /d NR . The island-potential noise is back-acting. That is, the island-potential fluctuations couple to the resonator through C NR and drive it, re- sulting in a contribution to the total displacement noise that is linearly proportional to the coupling. A minimum in the total displacement noise is found at a coupling strength where the two sources contribute equally. For such optimal coupling, and typcial device parameters (see Chapter 2), the total displacement noise has been predicted to be ?y ? 2?y QL [13]. In the measurement of the first generation of RFSET displacement detectors (LPS), we were not limited by the intrinsic noise of the SET. Instead, we were lim- ited by the 80 pV/ ? Hz noise (referred to the input) of our cryogenic pre-amplifier (see Chapter 4 and Chapter 7), which set our charge sensitivity at approximately a factor of 4 - 6 from the SET?s intrinsic shot noise limit. Consequently, the lowest dis- placement sensitivity which we observed was on the on the order of a factor of 4 from the quantum limit [15]. Nevertheless, this is the closest approach to the quantum limit of displacement detection that anyone before or since has demonstrated, and marks a factor of 30 improvement over the SET-mixer technique previously demon- strated by Robert Knobel and Andrew Cleland at the University of California, Santa Barbara [14]. An additional improvement of the LPS detectors over the Santa Barbara SET- mixer technique was the ? 75 MHz bandwidth provided by the rf-matching network. In contrast, at best, the maximum bandwidth of the Santa Barbara technique would have been on the order of kHz, either limited by the DCSET or the dc electronics 7 at room temperature. Either way, as the quality factor and resonant frequency of the resonator were ? 1.5 x 10 3 and 116 MHz respectively [14], the half-width of the resonator?s spectral response was ? 10 5 and thus much larger than the detection bandwidth. The large bandwidth of the RFSET technique allowed us to observe the res- onator?s full spectral response, facilitating the detection of the resonator?s thermal motion. In the end, we were able observe the thermal motion of the nanoresonator down to a temperature as low as ? 56 mK, corresponding to a thermal occupation number of ?n th ??60, and demonstrating, that, indeed, NEMS is on the verge of the quantum regime. 1.2 Structure of the Thesis The structure of this thesis is as follows. Chapter 2 provides the basic definitions and theoretical concepts upon which the rest of the text is based. First, the quantum limits of a mechanical resonator are defined, and the criteria for reaching these limits are presented. This is followed by the introduction of the RFSET displacement detector and a discussion of its basic operating principles. In the final section, the intrinsic noise properties of the SET are reviewed and used to demonstrate that, in principle, the RFSET is capable of performing as a nearly quantum-limited displacement detector. Chapter 3 presents a detailed account of the fabrication steps we developed and followed to produce our first generation of RFSET displacement detectors. 8 Chapter 4 discusses the details of the apparatus which we constructed for the measurement of our samples. Chapter 5 presents and explains the RFSET reflectometry technique, the back- bone of the detection scheme. Chapter 6 describes the implementation of the RFSET displacement detection technique and presents our main research results. Relying heavily on the results of Chapter 5, it begins with a treatment of the basic methodology. This is followed by a discussion of the RFSET detection of capacitively driven nanoresonators. Next, the topic of nanomechanical noise thermometry is introduced. It is in this section that the central results of the thesis are put forth. Finally, the chapter finishes by addressing the issue of SET back action. Chapter 7 concludes the main body of the thesis with a discussion of the technical improvements and future prospects. The remaining chapters I label as Appendix A and Appendix B. They contain information that I think is either essential for understanding the basic concepts and limitations of the RFSET displacement detector or is useful for the actual implementation. Included in these chapters are tables of the various parameters for the devices around which this thesis is built. Two of the devices, Device 3 and Device 4, are included even though they are not discussed in the main body of the dissertation. Initially, my intent was to pro- duce a work that fully addresses the noise characteristics of the RFSET displacement detector, including the SET back action. Devices 1 and 2 were to be used for treat- ing the forward-coupling limit. Devices 3 and 4 were to be used for discussing the 9 back action limit. However, the physics involved with SET back action, particularly the superconducting SET, are more complicated and interesting than I originally imagined, and their investigation would constitute an entire thesis. Furthermore, we do not understand all the observations that we have made of Devices 3 and 4. I have left Devices 3 and 4 in the thesis mainly for illustrative purposes and for technical explanations of useful information (ie. RFSET gain calibration and RF tank-circuit characterization). Additionally, I would like to have the parameters and characteristics of all four devices and accompanying measurement circuits cataloged in one place. Finally, Device X and Device Y, devices which are not in any of tables, I have also used for illustrative purposes in Chapter 5. The nanoresonator in Device X met an early demise, however, the data taken for the gain-feedback circuit and sideband amplitude versus V g is the best data I have to illustrate these techniques. Device Y is actually from the latest generation of devices (courtesy of Akshay Naik). I used this data to illustrate the equivalence of the reflection map and the numerical derivative of the IV map. In the earlier devices, either this data is incomplete (for Devices 1 and 2 I have no simultaneous measurements of reflection map and numerical derivative) or the IV maps were less ?photogenic? (for Devices 3 and 4 the DJQP and JQP resonances are either smeared or faint). I also used data from Device Y to help illustrate the principle of amplitude modulation. 10 Chapter 2 Introduction At first glance, it might not be obvious why one can treat a nanomechanical device, such as a doubly-clamped resonator, as a simple harmonic oscillator. After all, a typical structure might have dimensions ranging from nanometers to microns, and be comprised of tens of billions of atoms and three times as many normal vibrational modes. The situation is simplified, though, if one is only interested in the lowest- frequency transverse modes. In this case, the ratio of the wavelength-to-lattice spacing is su?ciently large, ? 10 4 , that deformation of the lattice occurs slowly over the length of the device, allowing for the use of continuum elasticity theory to model the mode?s behavior [44] [45]. For deformations smaller than a critical amplitude [46], non-linear e?ects are negligible. Below the critical amplitude, the system can be reduced to a simple harmonic oscillator with an e?ective mass and spring constant determined by the mode shape and the portion of the oscillating structure that one considers (see Appendix A). The critical amplitude for the resonators measured in this research can be calculated to be ? nm?s [46]. The typical displacements we measure are ? pm?s. Peering at such a structure, for example, through an optical microscope, if our eyes and brains had the temporal resolution, we would expect to see it jumping 11 about, its motion driven by thermal fluctuations and other classical interactions. We might not expect to observe any deviations from the classical behavior we are so familiar with from our daily experiences. However, the question arises: what would it take to observe one of these structures exhibiting quantum behavior? In this chapter, I present some basic criteria which, when met, could allow for the observation of quantum phenomena in macroscopic mechanical resonators [4]. The first criteria, which I call Quantum Limit I, establishes an approximate level to which classical interactions must be reduced in order to observe the resonator?s quantum dynamics. It is implicit in my discussion that thermal fluctuations are the biggest problem and that all other classical forces are negligible. The second criteria, which I call Quantum Limit II, establishes the characteristics that a linear amplifier must possess in order that it minimally disturb the resonator during the process of measurement. It is seen that quantum mechanics requires such an amplifier to add a minimum of one-half quanta of noise power in the bandwidth of the signal. In the final section, I present and discuss the basics of the radio-frequency single- electron transistor (RFSET) displacement detector, a detection scheme which we have implemented and which has allowed us to come closer than any previous scheme to satisfying both criteria. 2.1 The Quantum Limit I: Thermal Noise The question of how cold a mechanical mode must be before thermal fluctua- tions are reduced to a level that does not obscure the mode?s quantum dynamics is 12 rather subtle. To thoroughly treat the topic is beyond the scope of this section and thesis. However, here, I present some basic, case-specific constraints on temperature with which I can later guage our experimental results (Chapter 5 and Chapter 6). I first discuss the freeze-out of a mechanical mode to its ground state. This is the simplest case to treat and provides a back-of-the-envelope estimate of how ?quan- tum? a particular mode at a given temperature is (ie. whether or not the mode?s dynamics can be described by classical equations of motion). Second, I briefly ex- amine the issue of decoherence. In particular, I discuss the decoherence of a pure harmonic oscillator state due to linear coupling to a thermal bath, and present an expression for the decoherence rate of a superposition of position states in terms of the mode?s temperature. Finally, I consider the detection of a mechanical mode?s zero-point motion in the presence of thermal noise. I show that, even if k B T b greatermuch ?h? 1 , depending on the duration of the measurement and the coupling of the resonator to the thermal bath T b , it is possible to reduce the change in amplitude due to thermal fluctuations below that due to zero-point fluctuations. Freeze-Out The simplest constraint to consider is a resonator?s ?freeze-out? to the ground state. This is equivalent to determining the temperature at which a mode?s thermal occupation number is reduced signifcantly below 1. The average thermal occupation of an oscillator mode with frequency ? 1 /2? is given by [35] ?n th ? =(e ?h? 1 /k B T b ?1) ?1 , (2.1) 13 Figure 2.1: Thermal occupation number ?n th ? plotted as a function of temperature for a range of nanomechanical resonant frequencies. The dashed lines represent the large-T b limit given of Eq. 2.1. Note that the zero-point contribution of 1 2 has not been included. 14 where ?h is Planck?s constant, k B is the Boltzman constant, T b is the temperature of the mode, and the criteria for freeze-out is just k B T b ?h? 1 ? 1. (2.2) Note that I have neglected the zero-point contribution of 1 2 . Figure 2.1 shows the thermal occupation number for mode frequencies ranging from 100 KHz to 1 GHz. This range roughly represents the realm of demonstrated doubly-clamped, nanomechanical mode frequencies. Examination of the plot re- veals that achieving freeze-out with passive refrigeration techniques (eg. dilution refrigeration) requires working with resonant frequencies in excess of 100 MHz. Of course with adiabatic demagnetization, the limit could be pulled down toward 10 MHz. Note, though, that I have not taken into account the issue of thermalization of the mechanical mode of interest. Whether a mechanical mode at 100 MHz can be tightly coupled to, say, the mixing chamber of a dilution refrigerator is a compli- cated problem that depends on both the experimental apparatus (eg. connections, ?heat leaks?, etc.) as well as the resonator?s geometry and material (essentially the parameters that determine the resonator?s quality factor), and one that I address in Section 7.2. Decoherence of a Mechanical Superposition The temperature constraint for the observation of a mechanical superposition state depends on the quality factor of the resonator under measurement and the desired duration of the superposition. A theoretical treatment of the harmonic 15 Figure 2.2: Decay time of a superposition of coherent states versus temperature for a range of nanomechanical resonant frequencies and quality factors. It is assumed that Gaussian peaks of the coherent states are separated by 2?y zp . 16 oscillator states suggests that a mechanical superposition of two coherent states with spatial separation ?y between the Gaussian peaks will preferentially decay to a single coherent state at a rate given by [47] ?= 4k B T b ?hQ b parenleftBigg ?y ?y zp parenrightBigg 2 , (2.3) where Q b = ? 1 ? is the resonator?s quality factor, assumed to be determined strictly from coupling to the thermal bath, ?y zp = radicalBig ?h/2M m ? 1 is the resonator?s zero-point deviation, and M m is the resonator?s e?ective mass. Figure 2.2 displays a plot of the inverse of the decay rate versus temperature for resonators with quality factors in the range of 10 2 to 10 5 . Here I assume that the superposition has been prepared so that ?y =2?y zp for each case. Thus the resonant frequency does not factor into Eq. 2.3. However, the quality factor for each of the resonators has been chosen to roughly reflect what has been demonstrated experimentally with real nanomechanical resonators. Ideally, one would want to engineer a nanoresonator with both large Q b andhighfrequency,sayQ b ? 10 5 and f 1 ? 1 GHz, so that the decay would occur over many cycles at 50 - 100 mK. From Fig. 2.2, for such a device at 50 mK, the decay time would be on the order of 10 3 cycles. In practice, achieving such a large quality factor and high resonant frequency might prove di?cult. To date, the only published, doubly-clamped 1 GHz resonator demonstrated a quality factor of approximately 10 2 [36], which would yield one cycle over the decay time at 50 mK. On the other hand, a 10 MHz resonator with quality factor in excess of 10 5 has recently been demonstrated [48], which would yield ? 10 coherent cycles at 50 mK. 17 Thermal Amplitude Flutuations In this subsection I estimate the temperature below which thermal fluctuations in a resonator?s amplitude become negligible with respect to the resonator?s zero- point motion. The amplitude of a resonator in contact with a thermal bath T b is seen to undergo a ?random-walk? with a variance approximated by [49] ?y 2 m ?? k B T b M m ? 2 1 parenleftBig 1?e ?t/? parenrightBig , (2.4) where ?=Q b /? 1 is the resonator?s thermal-relaxation time and I have assumed that at time t = 0 that the amplitude is known precisely, ie.that?y 2 m ? = 0. I note that the subscript ?m? is used for consistency with later portions of the thesis. It denotes the mean displacement of the neutral surface over the segment of the nanoresonator that couples to the SET detector, essentially the length of the SET island. For times t greatermuch ?, Eq. 2.4 reduces to the standard equipartition relationship. In this case, I expect thermal fluctuations of the amplitude to become small with respect to the resonator?s zero-point fluctuations when [4] k B T b M m ? 1 2 ? (?y zp ) 2 (2.5) or k B T b ? ?h? 1 2 . (2.6) This is a rather strict condition, and nearly identical to the criteria for freeze-out. On the other hand, for t lessmuch ?, the fluctuations in the resonator?s amplitude 18 are seen to increase linearly with t: ?y 2 m ?? k B T b M m ? 1 2 t ? . (2.7) The condition for thermal fluctuations to be small with respect to ground-state uncertainty in position is then [4] k B T b M m ? 2 1 t ? ? ?h 2M m ? 1 (2.8) or T b ? ?hQ b 2k B 1 t . (2.9) Clearly this is a less stringent requirement; and it implies that, if one could prepare an high-Q resonator in a well known position at time t =0andthenmakea measurement in a time t lessmuch 1/?, the exchange of energy between the resonator and the thermal bath would be a fraction t/? smaller than k B T b . Strictly speaking, then, the temperature to which one would have to cool a particular resonator for thermal fluctuations to become negligible would be inversely proportional to how quickly one could make a measurement of the resonator?s position and linearly proportional to the quality factor. This is just an order-of-magnitude analysis, and it begs a couple of questions: can one specify the position of the resonator with ?y m 2 ? = 0? And, what is the e?ect of the detector on the resonator during the measurement process? These are questions that I address in the following section. 19 2.2 The Quantum Limit II: Ideal Detection In this section I consider a second aspect of the quantum limit dealing with optimizing measurement precision. Ultimately, quantum mechanics places a limit on the precision with which certain information (ie. conjugate coordinates) can be extracted from the measurement of an object 1 ignore here, and throughout the thesis. This fundamental measurement limit is a direct result of the Heisenberg uncertainty relations for both the measured object?s coordinates of interest (eg.?y and ?p y of an oscillator) and the measurement device?s detection coordinates (eg. ? I and ? V of transistor). The purpose of this section, then, is to develop an understanding of such constraints in the context of the measurement of the displacement of a mechanical mode, and determine the conditions necessary to perform detection at this fundamental limit. Initially, I consider the simple case of ?quick?, repeated measurements of an harmonic oscillator?s position coordinate ?y, and derive the so-called Standard Quan- tum Limit for position detection. I then discuss the case of continuous linear de- tection of a generic narrow band signal, and derive the corresponding quantum constraints on amplifier noise temperature. Finally, I use linear response theory to phrase the quantum constraints on position detection in terms of an amplifier?s intrinsic noise characteristics. The result is thus a prescription which an amplifier must fulfill in order to operate as a quantum-limited position detector; and fur- 1 Techniques (eg. squeezed states, QND measurement, and contractive states) have been pro- posed for beating the quantum limit (for example, see Refs. [2] [4] [5] [6] [17]). However, these advanced measurement strategies are beyond the scope of the research presented here. 20 thermore, a guage by which I can assess our experimental results (Chapter 6 and Chapter 7). The Standard Quantum Limit Following reference [2], I start with a crude derivation of the Standard Quan- tum Limit. Consider a mechanical mode undergoing simple harmonic motion with frequency ? 1 . The hamiltonian for such a system is given by H = p 2 y 2M m + 1 2 M m ? 2 1 y 2 m , (2.10) where p y and y m are the conjugate momentum and displacement of the resonator and M m is the e?ective mass for the motion of interest (See Appendix A). I note that, as in the previous section, y m is used for consistency with later sections in which it denotes the average displacement of the neutral surface over the length of the SET detector. The task at hand is to determine how precisely one can measure y m by mak- ing two measurements such that the measurement time ? lessmuch 1/? 1 . For a classical resonator, in principle, there is no limit on how precisely one can measure y m or p y . However, for a quantum resonator, the resonator?s position and momentum are de- scribed by the operators ?y m and ?p y , which are constrained through the commutation relation [50] [?y m , ?p y ]=i?h 21 to obey the Heisenberg uncertainty principle ?y m ?p y ? ?h 2 . (2.11) Simply put, the more precisely one specifies ??y m ?, the less precisely one can know ??p y ?. This is not really a concern for one quick measurement of ??y m ?;in principle, it can be done with arbitrary precision. However, if one intends to make two or more measurements of ??y m ? with the highest precision possible, the e?ect of the measurement on ?p y , or the quantum mechanical back action, must be taken into account. From Eq. 2.10, in the Heisenberg representation, the equations of motion for ?y m and ?p y are given by [51] ?y m (t)=?y m (0)cos ? 1 t + ?p y (0) M m ? 1 sin? 1 t (2.12) and ?p y (t)=?M m ? 1 ?y m (0)sin ? 1 t +?p y (0)cos ? 1 t. (2.13) If the resonator is not in an energy eigenstate, then the expectation values ??y m (t)? and ??p y (t)? will be oscillatory functions of time with the respective variances given by [2] (?y m (t)) 2 =(?y m (0)) 2 cos 2 ? 1 t + parenleftBigg ?p y (0) M m ? 1 parenrightBigg 2 sin 2 ? 1 t (2.14) and (?p y (t)) 2 =(?y m (0)M m ? 1 ) 2 sin 2 ? 1 t +?p y (0) 2 cos 2 ? 1 t. (2.15) I see that if, at time t = 0, I make an initial measurement??y m (0)? with precision ?y m (0), the uncertainty in the resonator?s position due to the initial measurement 22 at a time t later is (?y m (t)) 2 ? (?y m (0)) 2 cos 2 ? 1 t + parenleftBigg ?h 2M m ? 1 parenrightBigg 2 sin 2 ? 1 t, (2.16) where I have assumed that there is no correlation between the uncertainties in ??y m ? and ??p y ?, only that the rms amplitudes are related through the uncertainty principle, Equation 2.11. To minimize the uncertainty in position due to the initial measurement, it is clear that I must have ?y m (0) = ?h ?y m (0)2M m ? 1 (2.17) or ?y m (0) = radicalBigg ?h 2M m ? 1 . (2.18) This is known as the Standard Quantum Limit (SQL) for position detection [4] [2]. From Equation 2.16, for such a measurement, ?y m (t) is constant in time, implying a resonator state with phase-insensitive noise. One set of phase-insensitive states, with the additional stipulation that the equality in Eq. 2.16 be satisfied, is the set of coherent states [52]. Thus I can conclude this section by stating that, to minimize the error in each of two consectutive quick measurements of ??y m (t)?, it is necessary that the first measurement projects the resonator into a minimum uncertainty state. The Ideal Linear Amplifier While the analysis of the preceding section provides us with an idea of the role of the Heisenberg uncertainty principle in measurement, it is unsatisfactory 23 for at least two reasons. First, one is not always interested in simply making two consecutive quick measurements of a system. For example, the results presented in this thesis were obtained in the continuous measurement limit (ie. the limit in which the time interval between measurements becomes small with respect to the time scale of the dynamics of the measured system). Second, the analysis makes no reference to a measuring device, relying only upon the uncertainty relation for the measured oscillator, or, essentially, its wave nature. Which begs the question: what istheroleofthedetector? In this section, I paraphrase a work of Carleton Caves [3] and derive the quan- tum measurement limit for the case of a quantum signal continuously measured by a linear quantum amplifier. It is seen that such a detection scheme necessarily adds a minimum of one-half of a quanta of noise to the measured signal. As this minimum is imposed only by the assumptions of linearity and the appropriate commutation relations invoked by unitarity, the limit is known as the ideal linear amplifier limit, and such an amplifier is referred to as an ideal linear amplifier. In Cave?s model [3], the input signal and the amplifier are represented by Bosonic modes with noise power per unit bandwidth per mode given in terms of the number quanta ? N a =?a ? ?a ? ? and N b = ? b ? ? b ? ? respectively. Here ?a ? ,?a ? ? and ? b ? , ? b ? ? are annihilation and creation operators for the respective modes of the oscillator and detector, and obey the commuation relations bracketleftBig ?a ? , ?a ? ? bracketrightBig = ? ? ? , bracketleftBig ? b ? , ? b ? ? bracketrightBig = ? ? ? (2.19) 24 and [?a ? , ?a ? ]=0, bracketleftBig ? b ? , ? b ? bracketrightBig =0. (2.20) The analysis proceeds in the Heisenberg representation where it is assumed that the evolution of the output (detector) operators can be expressed as a linear superposition of the input (oscillator) modes [3]: ? b ? = summationdisplay ? parenleftBig M ?? ?a ? + L ?? ?a ? ? parenrightBig + ? F ? (2.21) and ? b ? ? = summationdisplay ? parenleftBig ?a ? ? M ? ?? +?a ? L ? ?? parenrightBig + ? F ? ? , (2.22) where M ?? and L ?? are matrices related to the amplifier?s gain and ? F ? is an operator representing the amplifier?s noise contribution, which is assumed to be random in time with a Gaussian distribution. It is further assumed that ? F depends only on the internal modes, or the internal state, of the amplifier and thus commutes with the input mode operators. It turns out that this assumption has rather important consequences, which I will discuss in the end. For a less ideal amplifier M ?? and L ?? would be replaced by operators to account for any time dependence in the gain (ie. gain fluctuations). Next several assumptions are made. First, the analysis is restricted to the case of single mode detection 2 so that Eqs. 2.21 and 2.22 simplify to ? b = parenleftBig M?a + L?a ? parenrightBig + ? F (2.23) 2 I note that Caves also treats the more general multi-mode case. The purpose of this section, however, is to give a brief demonstration of how the quantum limit arises in the context of continous measurement. For this purpose, presentation of the single-mode analysis is su?cient. 25 and ? b ? = parenleftBig ?a ? M ? +?aL ? parenrightBig + ? F ? . (2.24) Second, it is assumed that the amplifier is phase-conjugating, ie. that a phase- shift in the input signal generates the opposite sign phase-shift in the output signal. That is, if ?a prime =?ae ?i? , then ? b prime ? ? F = ? be i? ? ? F, This requires that M =0.ThusIamleftwith ? b =?a ? L + ? F (2.25) and ? b ? = L ? ?a + ? F ? . (2.26) This assumption is made arbitrarily. I could have just as easily assumed phase- preseving. In the end, Caves demonstrates that for large gain amplifiers, the ultimate limit is the same. Finally, it is assumed that the amplifier noise is phase-sensitive. That is, the amplifier?s rms noise contribution is split unequally between the input signal?s quadratures. Thus one must break up the input and output signals into their re- spective quadratures: ?a = ? X 1 + i ? X 2 (2.27) 26 and ? b = ? b 1 + i ? b 2 , (2.28) where ? b 1 = L ? X 1 + ? F 1 ? b 2 = L ? X 2 + ? F 2 , (2.29) where ? F= ? F 1 +i ? F 2 . To recover the phase-insensitive amplifier, simply set F 1 = F 2 . With these assumptions, one can now express the total output noise for each quadrature as [3] (?b 1 ) 2 = |L| 2 (?X 1 ) 2 +(?F 1 ) 2 , (2.30) (?b 2 ) 2 = |L| 2 (?X 2 ) 2 +(?F 2 ) 2 , (2.31) where ?X 1 (?X 2 )and?F 1 (?F 2 ) are the signal and detector rms noise contribu- tions to the quadratures respectively. The amplifier contribution referred to the input for each quadrature is thus [3] A 1 = (?F 1 ) 2 |L| 2 , (2.32) A 2 = (?F 2 ) 2 |L| 2 , (2.33) where |L| 2 plays the role of the amplifier?s power gain in number of quanta. 27 Using the Schwartz inequality [3], ?F 1 ?F 2 ? 1 2 |? bracketleftBig ? F 1 , ? F 2 bracketrightBig ?|, (2.34) and the relation bracketleftBig ? F, ? F ? bracketrightBig = ?2i bracketleftBig ? F 1 , ? F 2 bracketrightBig , (2.35) one arrives at the uncertainty relation for a phase-sensitive linear amplifier [3] A 1 A 2 ? 1 16 |?[ ? F, ? F ? ]?| 2 , (2.36) and the total amplifier noise contribution A = A 1 + A 2 ? 1 2 |?[ ? F, ? F ? ]?|. (2.37) As they stand, Eqs. 2.36 and 2.37 are not very illuminating. However, from the commutation relation for ? b, Eq. 2.20, one finds [3] [ ? F, ? F ? ]=1+|L| 2 . (2.38) Thus radicalBig A 1 A 2 ? 1 4 (1 +|L| ?2 ), (2.39) and A ? 1 2 (1 +|L| ?2 ). (2.40) For large gain, |L| 2 greatermuch 1, Eqs. 2.39 and 2.40 tell us two things: noise in one quadrature can only be reduced at the expense of signal-to-noise degradation in the other quadrature [3]; and the absolute minimum total noise power per unit bandwidth that an amplifier can add to a narrow band signal is one-half quanta [3]. 28 Using Eqs. 2.30, 2.31, 2.32, 2.33, and 2.37, one can re-express the total output noise as ?b 2 = |L| 2 parenleftBig ?X 2 + A parenrightBig . (2.41) If the signal contributes one-half quanta of noise, ie.?X 2 = 1 2 ,then ?b 2 = |L| 2 parenleftbigg 1 2 + A parenrightbigg . (2.42) From Eq. 2.40, this then yields |?b| 2 ? 1 2 |L| 2 + 1 2 (1 +|L| 2 )=|L| 2 + 1 2 , (2.43) which simply states that, for large gain, the minimum total noise at the output of an ideal amplifier is composed of two parts: one-half quanta contributed by the internal amplifier modes, and one-half quanta contributed by the input mode; both of which are amplified by |L| 2 [3]. The fact that the two noise sources add in quadrature is a consequence of the assumption that the internal states of the amplifier and the initial input signal state are independent. As a result, their fluctuations are uncorrelated. Finally, Caves defines the noise temperature, T QL , of the ideal linear amplifier by assuming that the total input noise is given by the Planck distribution (plus the zero-point energy), |?a| 2 = 1 2 coth parenleftBigg ?h? 1 2k B T b parenrightBigg , (2.44) and asking: by how much would one have to increase T b to observe |?b| 2 at the output of the amplifier? In the limit of large gain, after working through the algebra, 29 Caves finds: T QL = ?h? 1 k B ln(3) , (2.45) if T b =0; and T QL = ?h? 1 2k B , (2.46) if k B T b greatermuch ?h? 1 . This minimum is imposed only by the assumptions of linearity and the appro- priate commutation relations invoked by unitarity, and is known as the ideal linear amplifier limit. An amplifier that meets this condition is referred to as an ideal linear amplifier. In the low-T b limit, then, the resulting minimum position sensitivity is ?y QL = radicalBigg T QL k B K m = radicalBigg ?h ln(3)M m ? 1 , (2.47) which is greater than the standard quantum limit, Equation 2.18: ?y QL = radicalBigg 2 ln(3) ?y SQL . (2.48) Quantum-Limited Position Detection In the previous sections it was demonstrated that quantum mechanics places a limit on the minimum rms uncertainty in the knowledge of a resonator?s position; it was also shown that quantum mechanics requires that there be an additional minimum noise contribution from the amplifier itself. 3 However, the discussion up until this point has been rather abstract; it is not obvious how to extend the formalism or the results to a solid-state position amplifier such as the SET. 3 I implicitly mean an amplifier with linear, time-indepedent coupling to the resonator?s position. 30 In this section, I present and discuss the results of Aashish Clerk?s linear response approach to quantum-limited position detection [16]. This approach ar- rives at the same conclusion as Cave?s Bosonic-mode model, with the advantage of phrasing the quantum constraints on continuous linear amplification in terms of an amplifier?s intrinsic noise properties. Clerk considers a resonator with conjugate momentum, ?p y , and displacement, ?y m , and an amplifier with input and output characterized by the hermitian operators ? F and ? I respectively. He further assumes that the resonator is coupled to both an equilibrium bath with temperature T b and to the amplifier via the interaction H int = ?A ? F ? ?y m , (2.49) where A sets the strength of the interaction and ? F can be thought of as the inter- action force or equivalently the back action of the amplifier on the resonator. The analysis is restricted to the case of weak coupling (H int ? 0) the relevant parameter regime for our experiments (see Section 2.3). There are two consequences of this. First, the response of the output of the amplifier ? ? I? to a small change ??y m ? can be determined using linear response theory [53]. From Liouville?s theorem, to first-order in H int , Clerk finds that the amplifier?s output response is given by ?? ? I(t)? = Tr( ? I??(t)) = A integraldisplay ? ?? dt prime ?(t?t prime )??y m (t)?, (2.50) where ??(t) is the first-order density matrix term in the iterative solution of the Liouville equation. 31 The admittance or amplifier gain, ?, is given by [16] ?(t?t prime )= 1 i?h ?(t?t prime )? bracketleftBig ? I(t), ? F(t prime ) bracketrightBig ?; (2.51) and the expansion is done about the amplifier?s zero-coupling configuration. The second consequence of the weak-coupling assumption is that the equation of motion for ??y m (t)? reduces to a ?Langevin-like? expression. Specifically, Clerk finds that M m ? 2 ??y m (t)? ?t 2 = ?M m ? 1 2 ??y m (t)??? b ???y m (t)? ?t ? (2.52) ?A 2 integraldisplay dt prime ?(t?t prime ) ???y m (t prime )? ?t prime + F b (t)+A ?F(t), (2.53) where ? b and F b describe the damping and fluctuating forces provided by the bath respectively and are related through the fluctuation-dissipation theorem: ? S F b = ? bath ?h? coth( ?h? k b T b ). (2.54) The detector?s influence is manifest in the damping term ?(t ? t prime )andthe back action force F(t). In the limit where the resonator?s frequency is small with respect to the intrinsic time-scale of the amplifier, Clerk demonstrates that the detector-induced damping and back action force are related in a manner simliar to the fluctuation-dissipation theorem: 2k B T d = ? S F ? . (2.55) 32 Here, ? S F = lim ??0 S F (?)+S F (??) 2 (2.56) and ? = lim ??0 S F (?)?S F (??) 2?h? , (2.57) where, S F (??)= integraldisplay ? ?? dt? ? F(t) ? F(0)?e ?i?t (2.58) are the positive and negative frequency components of the amplifier?s back action noise spectral density, with (+) referring to energy transfer from the resonator to the amplifier, and (-) referring to energy transfer from the amplifier to the resonator. See reference [54] for a nice explanation of positive and negative frequencies in quantum noise. Classically, ?F(t)F(0)? =?F(0)F(t)?,sothatS F (?)= S F (??). However, this is not generally true for a quantum mechanical system, ie. bracketleftBig ? F(t), ? F(0) bracketrightBig negationslash= 0 [54] . It is convenient then to use the relation ? F(t) ? F(0) = 1 2 parenleftBigbraceleftBig ? F(t), ? F(0) bracerightBig + bracketleftBig ? F(t), ? F(0) bracketrightBigparenrightBig to break-up Eq. 2.58 into two components: a real component representing the total force spectral density experienced by the resonator due to the amplifier?s back action noise (Eq. 2.56); and an imaginary component representing the energy-loss rate of the resonator due to the interaction with the amplifier, or detector-induced damping (Eq. 2.57). With these definitions, the e?ective amplifier temperature T d is thus interpreted as gauging the asymmetry between the amplifier?s positive and negative frequency back action noise. 33 For the case of normal-state SET?s and tunnel-junctions, it has been found that T d is positive and propotional to the average energy lost by an electron as it traverses the device?s junctions [55] [56] [57]. On the other hand, due to the myriad tunnelling processes, the case of the superconducting SET (SSET) is much more complicated [58] [59] [60]. For example, both positive and negative e?ective temperature and dissipation are possible when the SSET is biased near the single and double Cooper-pair resonances (see Appendix B for brief discussion of the SSET); the ?direction? of the exchange of energy depending on whether energy needs to be removed or added for the resonant tunneling of Cooper-pairs to occur . Regardless of whether T d is positive or negative, the total e?ective tempera- ture, T e , of the resonator is given by the sum of T d and T b , weighted by the respective coupling to each reservoir [16] [55]: T e = 1 ? e (? b T b + ? d T d ), (2.59) where ? e = ? b + ? d = M m ? 1 Q e , (2.60) ? d = A 2 ?, (2.61) and Q e is the e?ective quality factor of the resonator due to damping induced from both the detector and the environment. From the above definitions, and Eqs. 2.54 and 2.55, Clerk expresses the spec- tral density of the resonator?s motion as 34 ? S y (?)= ? S F b (?)+A 2 ? S F (?) |M m (? 2 ?? 1 2 + i?? 1 /Q e )| 2 = = |g(?)| 2 parenleftBig ? S F b (?)+A 2 ? S F (?) parenrightBig . (2.62) Using Eqs. 2.51 and 2.62, the total noise-power density at the output of the detector is thus ? S I,tot (?)= ? S I (?)+A 2 |?(?)| 2 ? S y (?)?2A 2 Re bracketleftBig ?(?) ? g ? (?) ? S IF bracketrightBig , (2.63) where ? S I (?) is the symmetrized spectral density of the amplifier?s forward coupling noise, ? S IF (?) is the symmetrized spectral density of the cross-correlations between forward and back-acting noise sources. Finally, Clerk shows that Eq. 2.63 can be converted into an equivalent dis- placement noise density, referred to the input of the amplifier: S y,tot (?)= ? S I (?) |?(?)| 2 A 2 + A 2 |g(?)| 2 ? S F (?)? 2Re bracketleftBig ? ? (?)g ? (?) ? S IF (?) bracketrightBig |?(?)| 2 + +|g(?)| 2 ? S F,b (?). (2.64) The first three terms represent the amplifier?s contribution to the total dis- placement noise; whereas the last term represents resonator fluctuations due strictly to the equilibrium bath. Minimization of Eq. 2.64 is a rather involved process, requiring the optimiza- tion of the noise sources ? S I (?), ? S F (?), and ? S IF (?) and the coupling A. However, 35 Clerk imposes two important constraints that allow for the determination of a min- inum on resonance, ?=? 1 . First, ? S I (?)and ? S F (?) are constrained by the amplifier uncertainty principle: ? S I (?) ? S F (?) ? ?h 2 4 (Re [?(?)]) 2 + parenleftBig Re bracketleftBig ? S IF (?) bracketrightBigparenrightBig 2 (2.65) with the equality fulfilled for the case of a quantum-limited amplifier. Strictly speaking, Eq. 2.65 states that in the presence of gain, even an ideal or quantum-limited amplifier must add a minimal amount of back-acting and forward- coupling gain. An alternative interpretation of the equality in Equation 2.65 is that no signal information is lost in the process of measurement [61]. One can see this by recognizing that the rate at which information is attained from the output of the detector, ? meas , is inversely proportional to ? S I (?)(ie. the smaller ? S I (?), the better the signal-to-noise, and the less time for which one needs to integrate); whereas, the rate at which information ?enters? the detector, ? ? , is proportional to the interaction S F . At the quantum-limit, Clerk et al. demonstrated that ? meas =? ? , implying a tight coupling between the amplifier input and output degrees of freedom. In a sub-ideal amplifier, ? ? > ? meas , implying that some information about the input signal is lost to internal degrees of freedom which do not influence the amplifier?s output. Second, Clerk demands that the total power available at the output of the amplifier be much greater than the total power delivered from the amplifier to the resonator. He shows that, for a quantum-limited amplifier, this is equivalent to 36 requiring that k B T d ?h? 1 greatermuch 1. (2.66) With these two constraints, the requisite conditions for minimization can be stated. First, it is necessary that the amplifier noise terms satisfy the equality in Equation 2.65. That is, the amplifier must be quantum-limited. Second, the symmetrized cross-correlation term ? S IF (?) must vanish, minimiz- ing the product ? S I (?) ? S F (?). Third, the back action and forward-coupling must contribute equally to the total displacement noise. This requirement falls out of the optimization of the coupling A, A opt = radicaltp radicalvertex radicalvertex radicalbt ? S I (?) |?(?)g(?)| 2 ? S F (?) , (2.67) and can be thought of as being analagous to noise impedance matching for opti- mization of signal-to-noise. A consequence of this third condition is that the detector-induced damping, ? d , must be small with respect to ? bath to ensure that the resonator is more tightly cou- pled to T b than T d (a consequence of the second constraint and Eq. 2.59). Explicitly, Clerk shows that the third condition implies A opt 2 ? ? b + A opt 2 ? = ?h? 1 4k B T d , (2.68) which is necessarily much less than one due to the assumptions I have made. If all three conditions are satisfied, Clerk shows, then, that an amplifier must contribute at least the equivalent of the resonator?s zero-point contribution to the measured signal. From Eqs. 2.68 and 2.59, it is evident that half of the contribution 37 is in the form of back action or heating of the resonator: T e = A opt 2 ?T d + ? b T b A opt 2 ? + ? b = ?h? 1 4k B + T b , (2.69) as one would expect, having optimized with respect to the coupling A. The other half of the amplifier contribution is necessarily forward-coupling noise. Thus, the noise temperature of a such an optimized detector is T QL = ?h? 1 2k B . (2.70) Essentially, this is the same result derived in the previous section: an amplifier must add at least one-half quanta of noise power per unit bandwidth to the mea- sured signal. However, the advantage of the present approach is that it provides a prescription (ie. the three conditions listed above) which an amplifier must fulfill in order for quantum-limited displacement detection to be possible. Additionally, the e?ect of the amplifier?s back action on the resonator is explicitly manifest as heating of the resonator by one-half its zero-point energy. 2.3 Nanomechanical RFSET Displacement Detection With the development of the RFSET [41] [62] [63], came the suggestion that, it could be implemented as a nearly quantum-limited nanomechanical displacement transducer [12] [13]. The realization was spurred by a combination of factors. For one, theoretical treatments of the SET suggested that the electrometer could achieve near-ideal noise characteristics required for quantum measurement schemes [43] [42] [64]. Secondly, the RFSET had been demonstrated to be capable of operating with 38 over 100 MHz bandwidth [41], a pre-requisite for the read-out of the high frequency nanoresonators thought to be necessary to demonstrate freeze-out. And third, the similarities in size-scale and fabrication between SETs and nanoresonators suggested that sub-micron positioning of the devices and, hence, tight coupling should be possible. In this section I first review the basic idea behind nanomechanical RFSET displacement detection. I then review the theoretical work on the noise characteris- tics of the SET and apply the results of the previous section to discuss the ultimate limits of the detection scheme. The RFSET Displacement Detector In essence, the RFSET displacement detector is a capacitive microphone: me- chanical fluctuations are converted into an electrical signal via the capacitive mod- ulation of an SET?s di?erential resistance; the di?erential resistance of the SET is then read-out using microwave reflectometry. A generic circuit schematic for the transduction process and an SEM micrograph of an RFSET displacement detector are displayed in Fig. 2.3. By application of a large DC bias, V NR , between the nanoresonator and the SET island, mechanical displacement of the resonator, y m , results in modulation of the polarization charge on the SET island through the relation ?Q NR = ?C NR ?y m V NR y m ? b C NR V NR d NR y m , (2.71) where C NR and d NR are the capacitive-coupling and spatial separation between the 39 (a) (b) Figure 2.3: (a) Circuit schematic and (b) SEM image of the RFSET displacement detector 40 resonator and the SET island respectively. In principle, C NR can be found to high precision from measurements of the SET conductance versus V NR (see Appendix B). However, the derivative of C NR with respect to y m must be calculated numerically. Thus, in the last step in Eq. 2.71, I have used a capacitance extraction program [65] and found, for typical device parameters, that ?C NR /?y m ? bC NR /d NR where b is of order unity. In Fig. 2.3(a), y is the in-plane displacement of the mid-point of the neutral surface (solid line) from the equilibrium position (dashed line). The quantity y m is defined as the average displacement of the neutral surface over the region a to b, the length of the SET island (the relationship between y m and y is calculated in Appendix A). It is straight-forward to show that the resonator?s fundamental mode couples most strongly to the SET island (ie.bothy n,m and ?C NR /?y n,m decrease with increasing number, n, of resonator nodes). For the remainder of the thesis, I will assume that y m represents the average displacement of resonator?s fundamental in-plane mode over the length of the SET island. It should also be noted that, in Fig. 2.3(b), the resonator?s displacement, the lengths of the SET island and the resonator, and the separation between the resonator and SET are not drawn to scale. For our samples the resonator?s displace- ment is about 10 ?6 ? d NR , the length of the SET island is about 0.33 - 0.5 times the length of the resonator, and d NR is typically about 0.02 - 0.05 times the length of the resonator. The modulation of the SET-island charge by ?Q NR results in the modula- tion of the SET?s di?erential resistance, R S . For a normal-state SET, biased at 41 the edge of the Coulomb blockade, the relationship between Q NR and R S can be approximated by (see Chapter 5 and Appendix B) 4 R S ? V SD I SD ? 2R ? sin(?Q NR /e)+1 , (2.72) where V SD and I SD are the source-drain voltage bias and current respectively. For small displacement and, hence, small charge modulation, ?Q NR lessmuch e,the maximum modulation of R S is given by ?R S ??2?R ? ?Q NR e ??b 2?R ? C NR V NR ed NR y m , (2.73) where R ? is the SET?s di?erential resistance at large source-drain bias V SD and e is the magnitude of the electron charge. The modulated di?erential resistance ?R S is measured by applying a mi- crowave signal v c (t) to the SET drain and measuring the modulation in the reflected- signal (sideband microwave reflectometry is discussed in Chapter 5). Because R S greatermuch Z o ,whereZ o is the characteristic transmission line impedance of 50 ?, an LC circuit is inserted in series with the SET for impedance matching. Ideally, the values of L T and C T are chosen so that, at the carrier frequency ? T =1/ ? L T C T , the impedance of the L T C T R S circuit is Z LCR = L T R S C T = Z o . (2.74) 4 As is discussed in Chapter 5 for both normal-state and superconducting-state SET?s no analytic expression for the di?erential resistance for an arbitray bias point is known. One must either solve the SET master equation numerically or use a measured IV curve and take the numerical derivative to find the relationship between ?Q NR and ?R S . 42 Equivalently, this can be seen as transforming the transmission line impedance so that Q 2 T Z o = R S where Q T = ? T L T /Z o is the external quality factor. The resulting modulation of the reflected-signal ?v r (t) is well approximated by linearizing it with respect to ?R S (t) [66], so that, on resonance ? T , ?v r (t) ? v c (t)??(t), (2.75) where ??(t) ??b Q 2 T Z o C NR V NR R ? ? ? 2ed NR y m (t) (2.76) at a bias-point of maximum Q NR -response (see Chapter 5 for details). As stated, in Eq. 2.76, I have assumed that the reflected-signal frequency is at the tank-circuit resonance, ? T , so that the Z LRC = L T /R S C T . Typically, the carrier frequency is tuned to ? T . However, y m (t) might be modulated at, say, 5 MHz. The reflected signal will then have sidebands at ? T ?(2??5MHz). The magnitude of the sidebands will depend on the half-width of the L T C T resonance. This is determined by loading from both Z o and R S : ?f = ? T 4?Q (2.77) where 1 Q = 1 Q T + 1 Q S = Z o ? T L T + ? T L T R S . (2.78) ?f is essentially the bandwidth of the RFSET. For optimal matching, it reduces to ?f = ? T 2?Q T . (2.79) In practice, the desired bandwidth, along with the matching-condition, sets the choice of the tank-circuit inductor and capacitor values. 43 The role of the tank circuit is clear from Eq. 2.76: near resonance, it e?ectively serves to amplify the modulation of the reflected-signal by Q 2 T . Without the tank- circuit ?? ? Z o /R S ? 10 ?3 , and thus ?v r ??10 ?3 b? ? 2 C NR V NR e y m d NR v c . (2.80) On the otherhand, with the tank circuit, and for optimal matching, Q 2 T Z o = R S ,I have ?v r ?? b? ? 2 C NR V NR e y m d NR v c . (2.81) Finally, I close this subsection with some remarks about optimizing the reflected- signal modulation response. First, it is obvious that maximizing the reflected-signal response requires optimizing the impedance matching. However, it also requires op- timizing v c and the coupling C NR V NR /d NR . The optimal carrier amplitude depends on the bias-point, tank-circuit quality-factor Q [66] [67], as well as whether the SET is superconducting or normal. In practice, it is simplest to determine the optimal value by tuning the amplitude manually and looking for the maximum response. The optimization of the coupling is more subtle. This is because I am not simply interested in maximizing the SET reponse to fluctuations in position of a nearby resonator. I am primarily interested in optimizing the SET displacement sensitivity, which, because of the SET back action, is a separate issue. In the following sec- tion, then, I consider the intrinsic SET noise, address the issue of optimal coupling C NR V NR /d NR , and present the predictions for the ultimate limit to RFSET position sensitivity. 44 The Ultimate Limit For simplicity, I develop the ultimate limit of RFSET displacement detection by considering the intrinsic noise of a normal-state DCSET. At the end of the section, I briefly address the much more complex intrinsic noise limits of the superconducting SET (SSET). Also, I assume that the di?erence in position sensitivity between the dc and rf modes of operation can be accounted for by applying the predicted reduction factor of 1.4 - 1.9 for the RFSET?s optimal intrinsic sensitivity [67]. The authors in Reference [67] state that the degradation of ultimate sensitivity in the rf mode compared to the dc mode is simply a result of the increased bandwidth of the rf mode. I assume that the orthodox theory (see Reference [68] and Appendix B) is applicable, and thus neglect the e?ects of co-tunneling [69]. Additionally, I assume that k B T b lessmuch E c , and neglect any thermal contributions to SET tunneling. Finally, I assume that the frequency range of interest is above the 1/f noise tail, ? 10 kHz, and below the intrinsic SET tunneling rate, (R ? C ? ) ?1 ? 1 - 100 GHz. In the relevant limits, the intrinsic noise of the SET is due to two white sources [43] [42]: the shot noise in the source-drain current I SD ; and fluctuations in the SET- island potential, ?. The origins of both sources arise from the stochastic nature of electron tunneling events, of which I SD is composed (see Appendix B). First, consider the shot noise. From the Orthodox Model, if one could in- sert an ammeter at each of the SET junctions, the tunneling-events would appear as delta-function peaks, separated in time according to a correlated-Poisson distri- 45 bution. The correlations arise from the fact that the probabilities for successive tunneling events are not independent but related through the accompanying change in the SET?s free-energy. To calculate the spectral density of the current fluctua- tions, one must solve the SET master equation and calculate the auto-correlation function, taking into account all the relevant tunneling processes. However, near the Coulomb-blockade threshold (the sequential tunneling regime), an approximate analytic expression for the shot noise spectral density exists 5 , and is given by [43] S II (?)=? I 2eI SD , (2.82) where ? I = ? 2 1 +? 2 2 (? 1 +? 2 ) 2 . (2.83) Here ? I accounts for the correlations between tunneling events, and ? 1 and ? 2 are the tunneling rates through junctions 1 and 2 respectively. It is S II (?)thatultimately sets the limit of the SET charge sensitivity, S Q NR (?)=S II (?) parenleftBigg ?I SD ?Q NR parenrightBigg ?2 . (2.84) In terms of the SET parameters, for symmetric junctions, this is expressed as [42] S Q NR (?) similarequal (1?? 2 )(1 + ? 2 ) 8? 2 eV SD R ? C 2 ? , (2.85) where ? = (2C NR V NR ?e) C ? V SD . (2.86) 5 From the initial assumptions, this is a classical analysis of the amplifier?s noise sources, and, thus, it follows that there is no need to symmetrize the sources as in Section 2.2. 46 Here, the parameter ? specifies I SD and the set of (V NR , V SD )biaspointswhich yield that particular value of current. The expresion above is valid provided that one restrict the bias-points to non-degeneracy points, 0 100 mK, we found IR exhibited a linear dependence on T S .The y-intercept was within measurement error of the origin (Fig. 6.8). Furthermore, the data scaled with V 2 NR , Figs. 6.9 and 6.10. That is, when divided by the square of the coupling voltage, for a given T e , IR exhibited no dependence on V 2 NR .These observations were su?cient evidence to conclude that IR was an accurate measure- ment of the temperature of the fundamental mode of the resonator, and that the mode was in thermal equilibrium with the sample holder and RuO 2 thermometer (ie. T e = T S ). Accordingly, in this temperature regime, the slope of IR versus T S could be used as a calibration for performing noise thermometry. It is evident in both Fig. 6.9 and Fig. 6.10 that there was scatter of 10 - 20% in some of the data 146 Figure 6.8: Plot demonstrating the integrated resonator response, IR, versus T S at a coupling voltage of V NR = 4 V. Data is for device 2. 147 points. I address this issue at the end of the section when I discuss evidence for back action. From Fig. 6.9 and Fig. 6.10, it is clear that, for T S < 100 mk, the data did not exhibit a linear dependence on sample-stage temperature. From Eq. 6.12, several microvolts at the SET gate could have driven the resonators to an rms amplitude of ? 200 fm - approximately the thermal amplitude of device 2 at 50 mK. However, this can be ruled out based on several facts. First, the data for 100 mK and above fit to a straight line through the origin. Second, the integrated power data, both above and below 100 mK, exhibited no obvious dependence on V 4 NR , as one would expect if the resonator was driven by a capacitively coupled signal. And third, we knew from transmission measurements that the attenuation down the gate lead was around - 20 dB at 20 MHz, and, thus, the noise at the input to the fridge would have to have been ? 10?s ?V rms , which was much greater than the expected Johnson noise from the resistors in the voltage dividers (10?s nV/ ? Hz at most) or the output noise of the optical isolators (also 10?s nV/ ? Hz at 20 MHz). It was more likely the result of either power from the RFSET line or dissipation in the SET heating the resonator - in Chapter 7, I address these possible reasons and solutions for the hang-up. Regardless of the source of the heating, we could use the calibration of IR at and above 100 mK to determine the e?ective temperature T e below 100 mK. For example, T e , at a sample-stage temperature of T S = 35 mK was determined by dividing the integrated response at 35 mK by the integrated response at 100 mK, Fig. 6.11. For the left peak, IR 35mK = 423 ? 43 ?e 2 /V 2 , and for the right peak, 148 Figure 6.9: A log-log plot demonstrating the integrated resonator response, IR, versus T S temperature scaled by V 2 NR for Device 2. Using the data from 100 mK and above as a calibration, the minimum temperature of the resonator?s fundamental mode is found to be 56 ? 7mK. 149 Figure 6.10: A log-log plot demonstrating the integrated resonator response, IR, versus T S scaled by V 2 NR for Device 1. Using the data from 100 mk and above as a calibration, the minimum temperature of the resonator?s fundamental mode is found to be 99 ? 4mK. 150 Figure 6.11: Using the data from the 100 mK peak as a calibration, the integrated response at 35 mK is found to correspond to T e = 56 mK. The data is for Device 2. Please note that the 100 mK peak has been shifted by 1.0 kHz for clarity. 151 IR 100mK = 747 ? 47 ?e 2 /V 2 . Thus we obtained T e = IR 35mK IR 100mK =56?7mK., (6.32) Incidentally, T e = 56 mK was the lowest mode temperature that we measured. Using the the Planck distribution function, ?n th ? =(e ?h? 1 /k B T eff ?1) ?1 (6.33) we calculate that this corresponds to a thermal occupation number of 59 ? 7. This is the lowest thermal occupation number ever measured for a collective mechanical mode [15]. Finally, I note that the slope of the IR/V 2 NR versus T S for Devices 1 and 2 di?er by approximately a factor of 7 (see Figs. 6.9 and 6.10). This is a result of several factors: (1) the increased coupling capacitance of Device 1 compared to Device 2 (61 aF compared to 27 aF); (2) the increased bandwidth of the spectrum analyzer for the measurement of Device 1 (a factor of 1.7); and (3) di?erent spring constants for the two devices (19 N/m for Device 1 compared with 15 N/m for Device 2). Figure 6.12 shows a plot of the integrated resonator response versus sample- stage temperature for both Devices 1 and 2. The data has been scaled with respect to the derivative of the capacitive coupling, ?C NR /?y m and the e?ective spring con- stant K m for each device. That is, I have plotted IR/(K m ?C NR V 2 NR /?y m ). Note that the data for the two devices fall on the same line, confirming that we understand the basic principles of the detection scheme, and have taken into account the dominant parameters - eg. C NR , K m . However, the fact that the slope of the scaled response deviates from k B by a factor of ? 3 tells us that there is systematic uncertainty in 152 Figure 6.12: A plot demonstrating that the integrated resonator response, IR, for Devices 1 and 2 collapse onto the same line when the data for each device is scaled by the corresponding A Q , V NR , C NR ,andK m . 153 our estimate of the measurement circuit gain. This discrepancy is consistent with the deviation in the calculation of the capacitively driven resonator response from the measured response, Section 6.1, and the deviation between the measured and calculated charge sensitivity. Also note that the errors bars on the data points in Fig. 6.12 are considerably larger then the error bars in the previous integrated power plots. The main source of the error is from uncertainty in the spring constant, ? 50%. I have not included uncertainty in the gain of the measurement circuit. Noise Temperature The noise performance of the displacement detection scheme can be evaluated by defining the noise temperature T N . This quantity correponds to the e?ective resonator mode temperature, T e , at which the resulting thermal displacement can be transduced and detected with a signal-to-noise ratio of 1. In other words, it is the temperature at which the rms amplitude of the resonator response A is equal to the rms background level P o . Figure 6.13 displays a plot of the resonator response (Device 2) for sample- stage temperature T S = 35 mK and coupling voltage V NR = 15 V. The data was fit to an harmonic oscillator response, and the frequency, quality factor, amplitude, background, and integrated response (IR), were extracted. Using the integrated resonator response versus T S (7.3 ? .1 ?e 2 /V 2 ) as the calibration (Fig. 6.9), the integrated response of the peak (535 ? 24 ?e 2 /V 2 ) was found to correspond to a resonator mode temperature of T e =73? 2 mK. From the ratio of the amplitude 154 Figure 6.13: A plot demonstrating the lowest noise temperature, T N achieved by RFSET displacement detection. From the slope of the integrated response versus sample-stage temperature, T e = 73 mK . The ratio of the amplitude to the back- ground yields a noise temperature of 15.5 ? .4 mK. Data is taken at V NR =15V, and is for device 2. 155 to the background (4.71 ? .01) the noise temperature for the measurement of the peak was found to be T N = T e 4.71 =15.5 ?.4mK. (6.34) Similiarly, we found, for Device 1, that the minimum noise temperature achieved was 43 ? 2mK. From T N and the equipartition relation, a rough estimate of the corresponding displacement noise spectral density within the resonator?s noise equivalent band- width can be made: S y = k B T N K m 4Q e ? 1 . (6.35) Thus, a noise temperature of T N =15.5?.4 mK corresponded to a displacement sensitivity of 3.8 ? .9 fm/ ? Hz,forQ e ? 3.5 x 10 4 . For Device 1, the minimum T N corresponded to a displacment sensitivity of 7.5 ? 2fm/ ? Hz. Notice, though, that because the calculation of the displacement sensitivity requires knowledge of the spring constant, K m , the error in the estimate was roughly 25 % for both devices. To find out how close our detection scheme was to the ideal, we expressed the displacement sensitivity in terms of the quantum limit for each device, Chapter 2: parenleftBigg ?y m ?y QL parenrightBigg 2 = T N T QL = ln3k B ?h? 1 T N . (6.36) Notice that the dependence on K m has dropped out. For Device 1 then, ?y m ?y QL =7.4?.2, (6.37) 156 For Device 2, ?y m ?y QL =4.3?.3. (6.38) These numbers represented the closest approach to the quantum limit, to date, achieved in the read-out of the displacement of a mechanical system [15]. Finally, Figs. 6.14 and 6.15 show the noise temperature, T N (left axis) and mean-square displacement noise (right axis) of Devices 1 and 2 as a function of V 2 NR . The displacement noise was normalized with respect to the quantum limit for each device, ?y m /?y QL . Also plotted in the figure are lines (dashed) representing the expected displacment sensitivity for a measurement circuit charge sensitivity of 10 and 20 ?e rms / ? Hz. Thus as we increased the coupling voltage, the noise temperature improved linearly with V 2 NR , as expected from Eq. 6.7. 6.3 Discussion of SET Back Action From the discussion in Chapter 2, SET back action produces three e?ects in the measurement of a nanomechanical resonator?s displacement: a frequency shift, damping, and displacement fluctuations. In this section, I argue that there is no clear evidence of any of these e?ects in the measurement of Device 1 or Device 2. First, I note that the SET-induced frequency shift and damping arise as a result of the dependence of the SET-island potential ? on resonator position. A change in the resonator?s position alters the island potential, which changes the electrostatic force between the SET island and the resonator. The in-phase component of the response shifts the resonator?s frequency ac- 157 Figure 6.14: A plot demonstrating the noise temperature, T N , of the RFSET dis- placement detection scheme as a function of V 2 NR for device 1. The right axis is the corresponding square of the position sensitivity normalized with respect to the quantum limit. A minimum noise temperature of 43 ? 2 mK was achieved. This corresponds to a displacement sensivity of a factor of 7.4 ? .2 from the quantum limit, or 7.5 ? 2fm/ ? Hz. 158 Figure 6.15: A plot demonstrating the noise temperature of the RFSET displace- ment detection scheme as a function of V 2 NR for device 2. The right axis is the corre- sponding square of the position sensitivity normalized with respect to the quantum limit. A minimum noise temperature of 15.5 ? .4 mK was achieved. This corre- sponds to a displacement sensivity a factor of 4.3 ? .3 from the quantum limit, or 3.8 ? .9 fm/ ? Hz. 159 cording to ?? 1 ? 1 ?? C NR V 2 NR K m d 2 NR C NR 2C ? . (6.39) Comparison with Eq. 6.16 shows that this shift in the resonator?s frequency should be ? C NR C ? smaller than the frequency shift due strictly to the electrostatic softening from V NR . For Devices 1 and 2, the ratio is ? 0.1 - 0.2, so the e?ect should provide a small correction. Figure 6.16 shows the frequency shift of the nanoresonators from Device 1 and Device 2 versus V 2 NR . 1 From a linear fit to the data, a slope of 72 and 124 Hz/V 2 were obtained respectively. These values are to be compared with the estimates of 140 and 100 Hz/V 2 provided by Eq. 6.16. From Eq. 6.39, I expect that the e?ect of the back action should have been about 10 - 20% of these values or ? 10 Hz/V 2 , which was approximately the magni- tude of the scatter in the data points. Little more can be said as we lacked the data to make a more precise determination of the slope. Furthermore, for this small of an e?ect, I would need to develop a more detailed model of both the frequency shift due strictly to V 2 NR (ie. calculate numerically ? 2 C NR /?y 2 m ) and the frequency shift due to the back action (ie. calculate correlations between tunneling and position fluctuations). Thus, the frequency shift of the resonator cannot be used as a gauge of the level of SET back action in the measurement. Second, I note that the out-of-phase component of the SET response produces 1 The data for Device 1 excludes the 6 V coupling data as it was taken on a separate cool-down, and exhibited a shift of 15 kHz. 160 (a) (b) Figure 6.16: Plot of the frequency shift versus V 2 NR for (a) Device 1 at 35 mK and (b) Device 2 at 100 mK. The shift is measured with respect to the lowest voltage data point. 161 damping: ? det ? ? 1 Q d = parenleftBigg bC NR V NR C ? V SD parenrightBigg 2 e 2 R ? 2M m d 2 NR = A NR V 2 NR , (6.40) where A NR /2? ? 0.02 and 0.003 Hz/V 2 for Devices 1 and 2 respectively. The total e?ective resonator damping is thus ? eff = ? 1 Q eff = ? bath + AV 2 NR . (6.41) It is assumed that A NR is independent of the bath temperature, T b . I expect, then, that the temperature dependence of ? e should follow the temperature depen- dence of ? b . While the sources responsible for ? b in nanoresonators are not well understood, ? b has generally been observed to obey a power-law dependence of T a b in several di?erent materials, with a ? 0.2 [121]. Assuming that the power-law holds down to mK temperatures, if ? d is comparable to or greater than ? b , then its e?ect should be evident in the deviation of ? e from the T 1/5 b dependence; the deviation becoming more pronounced at lower T b as ? b decreases and ? d remains constant. Figure 6.17 shows plots of ? e /2? versus T S for V NR = 6 V (Device 1) and 10 V (Device 2). The data sets each represent the largest bias voltage for each device for which complete data sets (35 mK to 500 mK) were taken. If back action is a factor, it should be most pronounced here. The solid line in each plot denotes T 1/5 S dependence, and was generated by forcing a fit of the data to ? eff = C + DT .2 S , (6.42) where A and B are o?-set and slope parameters. It is clear from the plots that the e?ective damping does not saturate as T S decreases to 35 mK. 2 2 The other complete data sets for both Device 1 and Device 2 exhibit similiar behavior. 162 (a) (b) Figure 6.17: Plots of ? e /2? Vs. T S for (a) Device 1 at V NR =6Vand(b)Device 2atV NR =10V. 163 Furthermore, at the lowest T S , it is seen that there was no obvious dependence of damping on V NR (Fig. 6.18). For both devices, ? e /2? was scattered about 500 Hz, ranging from 400 to 900 Hz. Using the estimate of ? d from Eq. 6.40 and the parameters for Devices 1 and 2, I find that ? d /2? should have become comparable to this range of values when V NR > 100 V, which is well above the parameter range explored. Finally I turn to the third back action e?ect: position-fluctuations. In Chapter 2, I showed that, in the absence of coupling to any other environment, the SET will drive the measured resonator, resulting in fluctuations in the resonator?s position with a variance given by ?y 2 m ? = k B T d K m , (6.43) where T d is considered to be a measure of the asymmetry in the SET?s quantum noise. If, in addition, the resonator is coupled to a thermal bath, then the resonator?s variance will be given by ?y 2 m ? = k B K m T e , (6.44) where T e = (? b T b + ? d T d ) ? e . (6.45) For small V NR , one expects T e = T b . However, as one increases V NR , it is expected that the dependence of T e on T b will become weaker until ? d T d greatermuch ? b T b ,atwhich point the resonator will hang-up at T e = T d . Turning back to Figs. 6.10 and 6.9, there is no discernible evidence for this 164 (a) (b) Figure 6.18: A plot of damping versus coupling for (a) Device 1 and (b) Device 2. 165 e?ect in either Device 1 or Device 2. As V NR was increased, the dependence of T e on T S (or T b above 100 mK) remained constant. From Eq. 2.96 and the parameters for Device 2, it is stright-forward to estimate the expected contribution of the SET back action to T e at 15 V coupling. An order of magnitude estimate of the SET potential fluctuations yields S 1/2 ?? (?) ? 1nV/ ? Hz. The SET-island potential fluctuations should have then driven the resonator to an rms amplitude of 30 fm rms . This would have corresponded to an e?ective heating of ? 100 ?K, and, at T b = T S = 100 mK, would have been a 0.1 % e?ect. I close this chapter with a few comments. First, it is clear that, for each of the three quantities (?f 1 , ? e ,andT e ), there is scatter beyond the statistical uncertainty determined from the least-squares fit of the power spectrum data and error propagation. For example, in Fig. 6.9, at 175 mK, the three data points for 10 V coupling each have error bars representing ? 2% relative uncertainty. However, the scatter about the mean of the three points is ? 10 - 15 %. While it is not shown, there may have been a correlation between the scatter in T e and the scatter in ? e ;thatis,forasetofdatapointsataparticularcoupling voltage and temperature, the data points which exhibited larger T e -ascompared to the other data points in the set - also exhibited larger ? e ,andviceversa. Additionally, for Device 2, the scatter in the data at 10 V and 15 V coupling was accompanied by fluctuations in the RFSET gain. These fluctuations were sub- stantial, and resulted in the RFSET gain-feedback unlocking. The frequency and magnitude of the fluctuations appeared to increase with V NR . Two-level charge fluctuators and back action are both possible explanations for 166 the observed scatter in the ? f 1 , ? e ,andT e . The possibility of charge fluctuations is bolstered by the observation that the magnitude of scatter appeared to increase with increasing coupling, and that it was accompanied by RFSET gain fluctuations. The possibility of back action being the culprit is weakened by the estimates above which demonstrate that all three back action manifestations should be small with respect to thermal noise and other factors. Of course, until the back action is measured, limited confidence can be placed in these estimates. Finally, in all of the back action estimates, we used normal-state SET ap- proximations. Recent theoretical [59] [60] and experimental [73] investigations of the SSET back action near the JQP resonances demonstrate significantly di?erent behavior than what is predicted for a simple normal-state SET. For instance, both the magnitude and the sign of T d and ? d are very sensitive functions of the SET?s detuning from the JQP (DJQP) resonance ridges. Furthermore, it has been found that the SET must be biased o? of the center of the JQP ridge and the RFSET carrier amplitude (v c ) must be reduced to a fraction of the JQP resonance-width in order to avoid sampling both the stable (negative ? d , negative detuning) and unstable (positive ? d , positive detuning) regimes. At the time the measurements of Device 1 and Device 2 were made, we were not aware of these details. We were also not particularly careful with maintaining a consistent SET bias point. Typically, the intent was to choose whichever bias point maximized the RFSET gain. The records we have for the bias points (Fig. 6.3 for example) demonstrate that at least some of the time we were biased near the top of the JQP ridge. As well, the peak-peak amplitude of v c at the SET corresponded to 167 ? 2Q T 8 ?V ? 140 ?V, which is on the order of the half-width of the JQP resonance. There are significant di?erences though between the present work (Devices 3,4,Y) and the work I report in this Chapter (Devices 1 and 2). The level of cou- pling is much greater. For the present generations devices, d NR , the resonator/SET spacing, has been decreased to ? 100 nm. This is to be compared with the 600 nm spacing of Devices 1 and Devices 2. Additionally, the spring constants of the present generations devices have been reduced by as much as a factor of 2 - 3, making the resonator more ?susceptible? to the SET?s back action forces. Finally, the evidence for back action in the more recent samples, while not yet fully understood, is much greater. For example, in Device 3, the e?ective quality factor has been observed to decrease with V 2 NR dependence from above 1 x 10 5 at 1Vcouplingto2x10 4 at 12 V coupling. For the same span in coupling voltage, the relationship between T e and T S is shown to go from directly proportional at 1 V to independent at 12 V. Finally, in an even more recent sample (Device Y), both positive and negative damping have been observed in the vicinity of the both the JQP and DJQP resonances. In the light of these facts, while we cannot completely rule out the influence of SET back action in Devices 1 and 2, it is clearly not a significant e?ect. 168 Chapter 7 Conclusions So far, I have demonstrated that we are capable of approaching the quantum limit on two fronts: we have implementeda near-ideal displacement detection scheme with sensitivity ?y m =4.3?y QL ; and we have cooled a mode of a nanomechanical oscillator to ?n th ??60. These observations are the closest approach to the quantum limit for a nanomechanical or macroscopic object to date. In this final chapter, I discuss possible reasons why our observations were limited to these values and suggest several technical improvements to push even closer to the quantum limit in future work. 7.1 Shot Noise Limited Detection Figure 7.1 demonstrates the deviations from ideality of the displacement de- tection scheme for the measurement of Device 2. We found that as we increased V 2 NR , the noise temperature T N decreased linearly with a slope (dashed line) deter- mined by S Q ? 10?e ? Hz. This is at odds with the charge sensitivity of 30?e/ ? Hz we measured from the 1 MHz gain calibration (Chapter 5 and Appendix II). The discrepancy could be a result of the improper calibration of charge sensitivity during the measurement of Device 2. At that time we were not aware of the Bessel func- tion calibration method. It could also be the case that the charge sensitivity was 169 Figure 7.1: A plot of T N Vs. V 2 NR for Device 2. The solid line is the total noise temperature including both SET shot noise and back action. 170 better at 20 MHz than at 1 MHz. Nevertheless, the solid lines denote the estimated SET shot noise and back action contribution to the noise temperature, assuming a normal-state SET. It is clear that there is room for a factor of 5 reduction in the forward-coupling measurement circuit contribution before the detection scheme becomes limited by the SET shot noise. To achieve this, there are at least three improvements that we could make. First, we could account for and reduce the 4 - 5 dB attenuation in the portion of the microwave circuit between the sample and the HEMT pre-amplifier. 1 From a simple consideration of the measurement circuit noise performance, one can see that a loss of 4 - 5 dB between the RFSET and the HEMT pre-amplifier is a significant contribution to the mesasurement circuit noise temperature T det n ? T L + T HEMT 10 ?L/10 ? 11 ?15 K, (7.1) where T L =4(10 L/10 ?1) is the equivalent noise temperature for the section of the circuit where the 4 - 5 dB is lost, L = 4 - 5 dB is the loss, and T HEMT ? 2Kisthe equivalent noise temperature of the pre-amplifier. Thus, while it is still a significant factor, the ultra-low noise HEMT accounts for less than 20% of the measurement circuit noise temperature. Second, we could replace the HEMT with a better cryogenic amplifier and use the HEMT as a follower. One possibile replacement would be the nearly-quantum limited microstrip SQUID amplifier [119], which has been demonstrated with an 1 For Devices 1 and 2, we do not have a record of the loss in this portion of the circuit. Consid- ering that T det n ? 20 K for Device 2, the attenuation was probably closer to 6 dB. 171 equivalent noise temperature a factor of 2 from the quantum limit at 500 MHz [119]. It has been used routinely with a noise temperature of 100 mK and gain of 20 dB up to 500 MHz. These amplifiers can operate below 100 mK, which would allow for placement very close to the sample, and thus reduce the possibility of signal-loss in the coupled-portion of the microwave circuit (See Chapter 5 and Appendix II). I can calculate the overall improvement in the noise temperature of the mea- surment circuit for the case that the RFSET is read-out with a microstrip SQUID. Assuming a gain of 20 dB and noise temperature of 100 mK for the SQUID, and using the 2K HEMT 2 as a follower, I calculate T det n ? 120 mK, (7.2) without loss in the circuit. With 5 dB of loss following the SQUID, I calculate T det n ? 250 mK. (7.3) This is a factor of at least 40 improvement in the noise temperature of the measurement circuit. Assuming, an SET with the identical parameters as Device 1 or Device 2, this should result in the reduction of the charge sensitivity to the SET shot noise limit. There are several concerns with using the microstrip SQUID amplifier. One is that it might require the use of a lower carrier frequency for the RFSET, and hence lower bandwidth. This problem could probably be circumvented by detuning 2 There is a HEMT amplifier available from Reference[107] with a quoted noise temperature of .9 K at 650 MHz which could be used as a follower for the SQUID. 172 the carrier frequency from the tank circuit resonance by the expected mechanical resonator frequency. The second concern deals with dynamic range. It isn?t clear whether the reflected signal from the RFSET would swamp the SQUID amplifier. For example, in Reference[120], the author calculates that the maximum output power a typical SQUID in open-loop configuration could supply to a 50 ?load is approximately 3 nW. Based on this calculation, if the SQUID amplifier has a gain of 20 dB, then the maximum input power is limited to approximately 30 pW. For RFSET operation, depending on the bias point, the reflected power could be as large as several 100 pW. This requires further invsetigation. Finally, the third improvement that we could make would be to the matching characteristics of the RFSET LC circuit. For Devices 1 - 4, the transformed SET impedance on resonance, Z LRC , was approximately 0.04 - 0.12 Z o .Asaresult,the reflection coe?cients at maximum conductance, ? max , ranged from ? 0.8 - 0.92, and yielded maximum depths of modulation M = 20log(? max )of? 0.7 - 2.0 dB. The obvious solution to this is to begin using larger inductance coils. Keeping C T fixed at ? 250 fF would require increasing the inductance up to 100?s nH for optimal matching. This would also have the desired a?ect of reducing the tank- circuit frequency down to an acceptable range for the microstrip SQUID amplifier. Again, though, we would pay the price in bandwidth. For C T ? 250 fF, optimal matching at 400 MHz, would require L T ? 600 nH. This fixes the tank-circuit quality factor and bandwidth to be ? 15 and 27 MHz respectively. Obviously, the optimization is tricky. One might not need to implement perfect matching. For example, increasing L T so that Z LRC ? 0.5 Z o would increase M by 6 - 8 dB, and 173 hence improve the charge sensitivity by approximately a factor of 2. 7.2 Sample Thermalization Ultimately, the base temperature of the dilution refrigerator is ? 9mK.This has been confirmed using nuclear orientation thermometry. In the previous chapter, I showed that the minimum resonator mode temperature that we measured was around 60 mK. Thus, it is clear that considerable improvement could be made to bring the mixing chamber and the nanoresonator?s fundamental mode into thermal equilibrium. There are at least three components to this problem: (1) minimizing the thermal impedance between the mixing chamber and the sample stage; (2) min- imizing the thermal impedance between the nanoresonator?s mode and the sample stage; and (3) reducing heat to the device. I assume that the first component, while not trivial, can be made negligible using the proper materials and connections. The second component is also not trivial and should depend on both the material out of which the samples are made and the geometry and mode of the nanoresonator. While, for a given nanoresonator mode, the thermal coupling between the mode and the substrate ?bath? can be inferred from measurements of the resonator?s quality factory, the nature of the coupling is not well understood [121] and is deserving of an entire thesis. Thus, in the section, I focus on the third component. In particular, I discuss two possible sources of heating: the Nb-Nb microwave coax and the SET. 174 Heat Transfer Through Nb-Nb Coax It is possible that heat was conducted via the microwave coaxial cable to the sample stage. From 1 K down to the mixing-chamber, we used Nb-Nb UT-85 coax (see Chapter 4) [106]. Gold-plated copper clamps were used to thermalize the outer-shield of the coax at the still-stage and cold-plate. At the mixing-chamber, the coax connected to a bias tee. The connection from the bias tee to the sample holder was made via a Cu UT-85 semi-rigid coax. From the clamps and the low thermal conductivity of the superconducting Nb [111], it seems unlikely, then, that heat flow through the outer-shield of the coax was responsible for heating of the sample. However, because of the Teflon insulation between the outer shield and inner conductor, it is possible that heat transfer through the coax?s center conductor could have resulted in the center conductor being out of equilibrium with the shield and the mixing chamber. To estimate the temperature di?erence between the inner and outer conductor of the coax at the mixing chamber, I consider the coax to be a cylinder of length L and composed of three concentric regions (Fig. 7.2): (1) an inner Nb conductor with thermal conductivity ? 1 and radius r 1 ; (2) a Teflon insulator with thermal conductivity ? o ; and (3) an outer Nb shield with thermal conductivity ? 2 = ? 1 , inner diameter 2?r 2 ,andthicknesst. I assume that the inner and outer conductors are in thermal equilibrium at the 1 K pot (z = 0). On our dilution fridge, a typical 1 K pot temperature was 1.7 K. Thus T 1 (0) = T 2 (0) = 1.7 K. I further assume that, at the mixing chamber (z = L), the shield and the mixing chamber are in thermal 175 Figure 7.2: An illustration of the Nb-Nb coax semi-rigid coax for heat-flow calcula- tion. It is assumed that the center conductor and shield are in thermal equilibrium at 1 K (T 1K ) and that the mixing-chamber end of the shield is in thermal equilbrium with the mixing chamber. 176 equilbrium. The problem is to find the temperature T 1 (L) of the inner conductor at the mixing chamber. As a first approximation, I assume that the heat transfer in the Teflon is purely radial (ie. the heat transfer along the length of the coax between 1.7 K and the mixing chamber is dominated by the Nb conductors, which have a much larger thermal conductance due to ? 1,2 /? o greatermuch 1 below 1 K [111]). Thus, I write the heat transfer per unit length between the inner conductor and the shield at a position z along the length of the coax as [122] ? Q = 2?? o ln(r 2 /r 1 ) (T 1 (z)?T 2 (z)). (7.4) I also assume that the heat transfer is purely axial (along z) within the in- ner conductor and within the shield. Fourier?s law [122] for a segment dz of each conductor thus yields ? 1 ?r 2 1 ?T 1 ?z vextendsingle vextendsingle vextendsingle vextendsingle z ? ? Qdz = ? 1 ?r 2 1 ?T 1 ?z vextendsingle vextendsingle vextendsingle vextendsingle z+dz (7.5) and 2?r 2 t? 2 ?T 2 ?z vextendsingle vextendsingle vextendsingle vextendsingle z + ? Qdz =2?r 2 t? 2 ?T 2 ?z vextendsingle vextendsingle vextendsingle vextendsingle z+dz , (7.6) where I have assumed that r 2 greatermuch t. Expanding about z, the steady-state temperature profile of the inner conductor is found from ?r 2 1 ? 1 ? 2 T 1 (z) ?z 2 + 2?? o ln(r 2 /r 1 ) (T 2 (z)?T 1 (z)) = 0. (7.7) 177 As well, the steady-state temperature profile of the shield is found from 2?r 2 t? 2 ? 2 T 2 (z) ?z 2 + 2?? o ln(r 2 /r 1 ) (T 1 (z)?T 2 (z)) = 0. (7.8) If I assume that the thermal conductivities ? 1 , ? 2 ,and? o are temperature independent and that ? 1 = ? 2 = ?, then the temperature di?erence between the inner conductor and shield is given by T 1 (z)?T 2 (z)=(T 1K ?T m/c ) r 2 1 +2r 2 t r 2 1 sinh(L/?)+(2r 2 tL/?)cosh(L/?) sinh(z/?), (7.9) where ? = radicaltp radicalvertex radicalvertex radicalbt parenleftBigg ?ln(r 2 /r 1 ) ? o parenrightBigg r 2 1 r 2 t r 2 1 +2r 2 t . (7.10) Using the ratio ?/? o ? 1x10 3 at 100 mK [111], and measuring r 1 ? 0.25 mm, r 2 ? 3.5r 1 and t ? 0.75r 1 , I calculate that, for T m/c =.01KandT 1K =1.7K, T 1 (L) ? 30 mK, (7.11) where I have assumed that L = 0.5 m. This analysis suggests that the inner-conductor of the microwave coax is heated by approximately 20 mK above the mixing chamber. However, the assumption that the thermal conductivities of Teflon and Nb are independent of temperature between 1 K and 10 mK (and that their ratio is given by their values at 100 mK) is not accurate. In fact, both materials exhibit a strong temperature dependence below 1 K. Specifically, ? ? T 3 and ? o ? T 2 [111]. Thus, the ratio ?/? o is a function of position along the length of the cable, decreasing from ? 10 4 at 1.7 K to ? 10 2 at the mixing-chamber. It is likely, then, that T 1 (L) could be heated less than 178 indicated by the above considerations. To estimate a lower bound, I assume that the ratio of the thermal conductivities is given by their values at 10 mK (? 100). I calculate, then, that T 1 (L) ? 17 mK. To conclude, I note that, at a minimum, the center conductor should have been heated by approximately 10 mK above the mixing chamber. However, a more detailed analysis taking into account the temperature dependence of the thermal conductivity of each component of the coax must be done. Dissipation in the SET It is also possible that the heating of the resonator mode was due to the dissipation of power in the SET. I can estimate the phonon temperature in the vicinity of the resonator using a steady-state thermal circuit model (Fig. 7.3). I make several assumptions. First, I assume that the power dissipated in the SET was determined by the dc current I SD and the SET resistance R ? : ? Q ? I 2 SD R ? = 400 fW (7.12) for I SD =2nAandR ? = 100 k?. Second, I assume that the power dissipated in the SET must have been con- ducted from the electrons in the SET island to the phonons in the SiN membrane beneath the SET via the electron-phonon coupling for normal metals [123] [124] ? Q =? 1 V 1 parenleftBig T 5 1 ?T 5 2 parenrightBig , (7.13) where ? 1 similarequal 2x10 9 nW/m 3 K, V 1 similarequal 5x10 ?21 m 3 is the volume of the SET island, T 1 is the temperature of the electrons in the SET island, and T 2 is the temperature 179 of the phonons in the membrane in the vicinity of the SET. It is not known how accurate this assumption is for superconducting metals. Next, I assume that the phonon transport in the SiN membrane was di?usive [125] and that the power delivered through the SiN membrane (beneath the SET) to the Si substrate can be written as [125] ? Q 1 = .0145 A 3L parenleftBig T 3 2 ?T 3 0 parenrightBig , (7.14) where A similarequal 5x10 ?12 m 2 is the cross-sectional area of the SiN membrane between the SET and the edge of the mebrane, L =25?m is the distance between the SET and the edge of the membrane, and T 0 = 30 mK is the bath temperature, assumed to be the temperature of the Si substrate and sample stage. Similarly, I write the power delivered through the SiN membrane, from the SET island to the phonons in the vicinity of the resonator, as [125] ? Q 2 = .0145 A prime 3L prime parenleftBig T 3 2 ?T 3 3 parenrightBig , (7.15) where L prime = 600 nm is the distance between the SET island and the resonator, A prime similarequal 1x10 ?12 m 2 is the cross-sectional area of the SiN membrane between the SET and the resonator, and T 3 is the phonon temperature near the resonator. Finally, I assume that there were two ?paths? between the region around the resonator and the sample-stage bath: (1) di?usive transport through the SiN membrane [125] ? Q 3 = .0145 A 3L parenleftBig T 3 3 ?T 3 0 parenrightBig ; (7.16) and (2) electron-phonon coupling [123] [124] between the phonons in the SiN mem- brane near the resonator and the electrons in the Au film, which are thermally 180 Figure 7.3: Thermal circuit for the SET and resonator on the SiN membrane. Dissi- pation in the SET ( ? Q) results in the heating of the phonon temperature around the resonator to T 3 . The thermal resistances in the circuit are: (R EP ) electron-phonon resistance for the Al SET island; (R 2D ) 2D thermal resistance of the SiN membrane from the SET to the bath; (R? 2D ) 2D thermal resistance of SiN membrane between the SET and resonator; (R? EP ) electron-phonon coupling for the Au layer of the resonator; and (R WF ) Weidemann-Franz resistance of the Au layer of the resonator to the bath. 181 connected to the bath through electron scattering (Weidemann-Franz) [111], ? Q 4 =? 2 V 2 parenleftBig T 5 3 ?T 5 4 parenrightBig = A primeprime L primeprime ?L o parenleftBig T 2 4 ?T 2 0 parenrightBig , (7.17) where, for simplicity, I set ? 2 =? 1 Here, V 2 similarequal 1x10 ?18 m 3 is the volume of the Au on the membrane (this does not include the Au film on top of the resonator, just the Au film leading to the resonator), A primeprime =2x10 ?14 m 2 is the area of the interface between the Au layer and the SiN membrane, L primeprime =2?25 ?m is the length of the Au film (essentially from the ends of the resonator to the edges of the membrane), T 4 is the temperature of the electrons in the Au film layer, ? similarequal 1x10 8 1/?m is the conductivity of Au, and L o =2.4x10 ?8 W?/K 2 is the Lorenz number [111]. With the additional assumption that Kircho??s law applies (ie. ? Q = ? Q 1 + ? Q 2 and ? Q 2 = ? Q 3 + ? Q 4 ), I use Matlab to solve Eqs. 7.13, 7.14, 7.15, 7.16, and 7.17 for T 3 . I find that, for T 0 =30mKand ? Q = 400 fW , approximately 200 fW is delivered to the SiN membrane near the resonator. This results in the heating of the resonator region to T 3 ? 60 mK. (7.18) Figure 7.4 shows a numerical calculation of the local resonator temperature T 3 as a function of both the bath temperature T 0 and the power dissipated in the SET ? Q. For the calculation of the plot in Fig. 7.4(a), I assumed that the total power dissipated in the SET is ? Q = 400 fW. It is seen that T 3 saturates at approximately 60 mK. Above 100 mK, T 3 is linear with T 0 . It appears that this behavior results from the increase in electron-phonon conductance above 100 mK. For example, at 10 mK, ? Q 4 (the power delivered from the phonons in the vicinity of the resonator through 182 (a) (b) Figure 7.4: Numerical calculation of temperature T 3 of phonons near the nanores- onator as a function of (a) bath temperature T 0 for ? Q = 400 fW and (b) power dissipated in the SET ? Q for T 0 =30mK. 183 the electron-phonon coupling to the electrons in the Au film) is approximately 1.4 fW. On the other, at T 0 = 500 mK, ? Q 4 ? 80 fW. The electron-phonon coupling throughout this temperature range (T 0 = 10 - 500 mK) is still weak enough, though, that the electrons in the Au layer stay thermalized with the bath T 0 . I note that the SET electron temperature T 1 saturates at 380 mK below T 0 = 200 mK, rising to approximately 520 mK at T 0 = 500 mK. For the calculation of the plot in Fig. 7.4(b), I assumed that the bath temper- ature T 0 = 30 mK. It is seen that for ? Q<10 fW, the local phonon temperature T 3 is heated by less than 2 mK. For ? Q =1pW,T 3 ? 80 mK. The actual power dissipated in the SETs during measurement is not known with high precision. Typically, the SET was biased near IV features like the JQP and DJQP peak (see Appendix B). However, for Devices 1 and 2, we did not keep a record of both the IV characteristics and the bias point for each measurement. For most of the measurements, we simply recorded the value of V SD and adjusted V g to maximize the gain. I can estimate the order-of-magnitude of the dissipated rf and dc power from the existing I SD V SD V g maps (see Figs. 5.4 and 5.5). The half-width (in V SD ) and height of the JQP peaks were ? 50 - 100 ?VandI SD ? 1- 2 nA respectively. Thus, the dissipated dc power should have been ? Q dc ? 100s fW. Typically, the incident rf signal was on the order of 10s ?V. Thus, the dissipated rf power should have been comparable to the dissipated dc power. To conclude this section I make several remarks. First, the temperature T 3 calculated in the above analysis is not necessarily the e?ective temperature of the resonator?s fundamental flexural mode T e .TocalculateT e , it would be necessary 184 to determine the thermal conductance between the flexural mode and the phonons in the SiN membrane near the resonator. In general, the thermal conductance of a suspended nano-bar is a complicated problem (see references [37] [126] [127]) and is beyond the scope of this thesis. Future work could involve incorporating the existing nanoresonator thermal models into the above circuit analysis and fitting the data for Devices 1 and 2 to the theroetical predictions. This work could be important for understanding the nature of dissipation in nanoresonators. For instance, from such an analysis, I could extract the thermal conductance between the nanoresonator and the bath and compare this with the measured quality factor Q e of the fundamental flexural mode. If I assume that the heat capacity of the flexural mode is given by [35] C v = ?E ?T 0 = k B , (7.19) and that the thermal time constant for the mode is given by ? = Q e ? 1 = R 1 C v , (7.20) where R 1 is the thermal resistance between the fundamental flexural mode and some dissipative environment, then I expect that the thermal conductance should be related to Q e through the relationship g 1 = 1 R 1 = ? 1 k B Q e . (7.21) Of course, other sources of dissipation, such as the SET detector (see Chapters 2 and 6) or charge noise in the substrate (see Chapter 6), might contribute to Q e . From the considerations of Chapter 6, I expect that the back action was a negligible 185 factor in the resonator?s dynamics. However, a comparison of the agreement between Q e predicted from thermal conductance models and Q e measured over a range of coupling voltages V NR might allow for a more precise determination of how small the e?ect was. Second, I note that it is also possible that the heating of the resonator mode could have been a result of dissipation in the Au film on top of the resonator. However, from an analysis similiar to the analysis for heating due to SET dissipation, I have found that this would require the electron temperature in the Au film to be approximately 200 mK, far out of equilibrium with the sample stage. Based upon the thermalization of the wiring for the resonator connection (see Chapter 4), it seems unlikely that the resonator lead on-chip would have been at such a temperature. Future work is necessary, though, to rule out black body radiation from the copper grains in the powder filters at the mixing chamber. Finally, I note that several experimental implementations could be made to determine the nature of the sample?s heating. First, heat sinks (see Chapter 4) could be added to the microwave circuit below the 1 K pot to see if better thermalization of the Nb coax?s center conductor reduces the sample temperature. Second, the SiN membrane geometry could be eliminated or the SETs could be fabricated o? of the membrane to allow for the dissipation from the SET to radiate ballistically to the bath. Third, insight could be gained by operating an SET far from the back action limit and monitoring the e?ective temperature of the resonator?s mode as a function of V SD ,V g , and carrier amplitude v c , at a given bath temperature T 0 . 186 7.3 Parting Motivation Besides serving as a manual for myself and others, it was my hope that this text would motivate the pursuit of quantum mechanics in nanomechanical systems by demonstrating how close we are technologically to this possibility. First, we have demonstrated that the RFSET displacement detector is a near- ideal detector, with sensitivity a factor of 4.3 from the quantum limit. It is thus a promising candidate to be used for advanced measurement techniques such as the quantum squeezing of a mechanical oscillator [29] [30] Second, we have demonstrated that it is possible to cool and measure me- chanical resonators to low thermal occupation numbers, ?n th ??60. With technical improvements and the implementation of feedback cooling [28], or by moving to higher frequency resonators, this number could be reduced toward unity where we could implement techniques to see evidence for quantized harmonic oscillator en- ergy levels and zero-point fluctuations. Even with slight improvement,?n th ??50, a proposal to use a Cooper-pair box to prepare a nanoresonator in a superposition of coherent position states could be implemented [25]. As well, these results open up the possibility of implementing various other proposals that could extend the study of quantum mechanics to much larger size scales [23-33]. 187 Appendix A Useful Mechanics Information In this appendix, classical elasticity theory is used to model the transverse displacement of a nanomechanical beam. In the first section, I make the connection between the vibration of an elastic body and simple harmonic motion. In the second section, I calculate the spring constants for SET detection. In the third secion, I solve for the eigenfrequenices and eigenmodes of an elastic bar under tension. In section four, I discuss the case of an elastic body undergoing damped-driven motion. Finally, in the fifth section, I briefly discuss the magnetomotive detection technique. A.1 Euler-Bernoulli Theory and The Simple Harmonic Oscillator I start by modeling our nanomechanical resonators as prismatic bars, clamped at both ends, and composed of isotropic, linear elastic materials (Fig. A.1). If I consider small displacements from equilibrium 1 and assume the cross-sectional area of the bars remain deformationless and perpendicular to the neutral surface, the Euler-Bernoulli assumptions, I can express the equation of motion for vibration in 1 I consider displacements su?ciently small so that the radius of curvature is large with respect to the transverse dimensions of the resonator. The resonators measured in our research easily satisfy this criteria as a typical ratio of displacement-to-length is ? 10 ?6 . 188 Figure A.1: A schematic of a prismatic, doubly-clamped nanoresonator. the y-direction as [128] ?A ? 2 y ?t 2 + EI ? 4 y ?x 4 =0. (A.1) Equation A.1 equates the inertial force-per-unit-length on a segment of the bar with the elastic restoring force the segment experiences when deformed. Here E is the Young?s modulus of the material, and I = w 3 t/12 is the moment of inertia. Parameters ? and A are the material density and the rectangular cross-sectional area respectively. For composite resonators, such as the metallized resonators in our research, EI should be replace by E 1 I 1 + E 2 I 2 ,whereE 1 ,I 1 and E 2 , I 2 are the Young?s moduli and moments of inertia for the two layers, respectively. Additionally, for such a resonator, ?A should be replaced by ? 1 A 1 +? 2 A 2 ,where? 1 and ? 2 are the layer densities, and A 1 and A 2 are the respective layer cross-sectional areas [129]. 189 Equation A.1 can be solved using separation of variables, y(x,t)=Y (x)[C 1 cos?t + C 2 sin?t](A.2) where C 1 and C 2 are determined from the resonator?s shape and velocity at some initial time, t =0. Substituting Eq. A.2 into Eq. A.1, I eliminate the time depedence, and solve for the normal modes. Assuming clamped-clamped boundary conditions, Y n (0) = Y n (L)=0and dY n (0) dx = dY n (L) dx = 0 , the normal modes are found to be [128] Y n (x)=C n bracketleftbigg (sink n x?sinh k n x)? ? parenleftBigg sink n L ?sinh k n L cosk n L ?cosh k n L parenrightBigg (cos k n x?cosh k n x) bracketrightbigg , (A.3) where the constants C n , Table A.1, are determined by normalizing the neutral sur- face displacement of the mode of interest to unity at maximum displacement. The choice of normalization is arbitrary. I have chosen this particular normalization con- vention so that, for the fundamental mode, the equations of motion represent the motion of the resonator?s mid-point (mid-point with respect to resonator length). Also note that for this normalization convention the mode functions Y n (x)aredi- mensionless. The first four modes are plotted in Fig. A.2. The normal mode frequencies, ? n , are found from ? n 2 = EIk 4 n ?A , (A.4) with k n determined by the roots of the eigenvalue equation cosk n Lcosh k n L =1. (A.5) 190 Figure A.2: The functions Y n (x) for the first four modes of a doubly-clamped res- onator. The x-axis has been normalized by the resonator length in each plot. 191 Table A.1: Geometric Constants: Modes 1-5. n k n C n ? n ? n 1 4.730 .619 .397 .295 2 7.853 .663 .439 .145 3 10.996 .661 .437 .081 4 14.137 .661 .437 .052 5 17.279 .661 .437 .036 6 20.420 .661 .437 .026 The eigenvalue equation can be solved graphically or numerically. Table A.1 lists k n for the first six modes. In our research, we were typically concerned with resonator motion of purely one mode. Using Eqs. A.2 and A.3, the solution for a given mode, n is simply y n (x,t)=Y n (x)[C n,1 cos ?t + C n,2 sin?t] . (A.6) I now make the connection between the dynamics of a particular mode and the simple harmonic oscillator. To calculate the bending energy E n of the mode, I consider the average work done in deforming the bar into the mode shape Y n (x): E n = 1 2 integraldisplay ?d? n M n ?, (A.7) where, M n = EI? 2 y n (x,t)/?x 2 is the mode?s bending moment, and ? n = ?y n (x,t)/?x is the slope of the bar?s deformation. 192 Plugging the expressions for M n and ? n into Eq. A.7, I find ?E n ?? EI 2 integraldisplay L 0 ?( ? 2 y n (x,t) ?x 2 ) 2 ?dx = ? n 2 A? 2 integraldisplay L 0 ?y 2 n (x,t)?dx E n = ? n 2 ?AL? n 2 ?y 2 n (t)? = 1 2 K n ?y 2 n (t)?. (A.8) It is apparent that this is just the potential energy of an object undergoing simple harmonic motion, with e?ective spring constant, K n = m eff,n ? 2 n , and e?ective mass, m eff,n = ? n ?AL.Here? n are dependent on mode shape, and are listed in Table (A.1) for the first 6 modes. For the fundamental mode, the paramater ?y 2 1 (t)? is the mean square amplitude of the resonator?s mid-point (length-wise), with magnitude determined by initial conditions. Finally, using the definitions of K n and m eff,n , I multiply Eq. A.1 by Y n (x) and integrate over x to recover the simple relation m eff,n ? 2 y n (t) ?t 2 = ?K n y n (t). (A.9) For the fundamental mode, this is simply the expression for the harmonic oscillation of the mid-point of the resonator. A.2 Spring Constants for SET Detection While Eq. A.8 expresses the potential energy of a particular resonator in terms of the motion of the resonator?s midpoint, in practice, the SET detector is sensitive to the average displacement of the resonator over the length of the SET island. It would be nice, then, to recast Eq. A.8 in terms of this quantity so that I am 193 able to easily calculate the potential energy of the mode from the observed motion. For example, for a resonator undergoing brownian motion, knowing the relation between the measured displacement and the potential energy of the mode, and using the equipartition theorem, I could calculate the mechcanical mode temperature. Alternatively, knowing the temperature of the mechanical mode, I could calculate the displacement signal I should expect to measure with the detector. Essentially, then, what I want to know is the spring constant K n,m for the mean motion of the resonator over the length of the SET island. I can calculate K n,m from Eq. A.8. To do this, I need to determine the relationship between the mean displacement over the length of the SET and the displacement of the mid-point of the resonator: y n,m (t)=a n y n (t), (A.10) where a n = 1 L 2 ?L 1 integraldisplay L 2 L 1 dxY n (x), (A.11) and L 1 and L 2 define the section of the resonator which corresponds to the length of the SET island. I assume that the SET island is centered about the mid-point of the resonator. I solve Eq. A.11 numerically for the fundamental mode of Devices 1 - 4 (Table A.3). For each sample, the length of the SET island was approximately 5 ?m. The length of the respective resonators is listed in Table A.2. For the first mode, Eq. A.8 becomes E 1 = ? 1 2a 2 1 ?AL? 1 2 ?y 2 1,m (t)? = 1 2 K 1,m ?y 2 1,m (t)?, (A.12) where K 1,m =M 1,m ? 2 1 ,andM 1,m =? 1 /a 2 1 ?AL are the e?ective spring constant and 194 e?ective mass for the motion of the fundamental mode of the resonator averaged over the 5 ?m length of the SET. As I am only concerned with the fundamental mode frequency, the superscript ?1? is dropped from K 1,m and M 1,m for the remainder of the section and througout the text. To calculate K m for each device, I need to know both M m and ? 1 .Toestimate M m , I use values of A = wt and L obtained from scanning electron micrograph (SEM) images and knowledge of the etch rates in the processing of the resonator; I assume typical densities of Au and SiN to be 19.3 x 10 3 kg/m 3 and 3000 kg/m 3 respectively [130]. The raw mass m r = ?ALt is calculated by including both the Au and SiN layers (Table A.2). To estimate ? 1 (Table A.3), I use either Eqs. A.15 and A.16 or Eq. A.17. The additional parameters required for this calculation are the Young moduli E Au and E SiN , which I assume to be approximately 50 GPa [131] and 250 GPa [75] respectively. I have assumed ? 50 GPa uncertainty in the value of the Young?s modulus for SiN to reflect the spread in values found in the literature (see [76]) and neglected the uncertainty in Young?s modulus for the Au layer (50 - 90 GPa reported in Reference [131] depending on grain size and thickness) as its contribution to the total error should be a factor of 5 - 10 times smaller than the contribution from the uncertainty in E SiN (a consequence of t Au ? 0.2t SiN for our samples after etching). 195 Table A.2: Geometry and raw mass, m r , of Devices 1-4. The error in the length and width of the resonator comes from 10% quoted error in the SEM calibration. The error of 30% in the thickness of the Au layer on the resonator comes from the spread in etch rates over time, and is a very rough estimate. Device w(nm) L(?m) t Au (nm) t SiN (nm) m r (pg) 1 300 ?30 10 ?1 30 ?20 100 ?2 2.6 ?1.2 2 200 ?20 8 ?.8 30 ?20 100 ?2 1.4 ?.7 3 200 ?20 15 ?1.5 30 ?20 100 ?2 2.6 ?1.2 4 225 ?23 18 ?1.8 30 ?20 100 ?2 3.2 ?1.5 Table A.3: E?ective masses, M m , and spring constants of Devices 1-4. ?a? corre- sponds to frequency calculated using either Eqs. A.15 and A.16 or Eq. A.17. ?b? corresponds to the frequency measured using SET detection at a temperature of 100 mK. Device a 1 M m (pg) ? 1 /2?(MHz) a K a m (N/m) ? 1 /2? b K b m 1 .838 1.5 ?.7 17 ?5 17 ?8 17.976648(3) 19 ?9 2 .760 .96 ?.45 18 ?6 12 ?6 19.654505(7) 15 ?7 3 .941 1.2 ?.54 6 ?1.8 1.5 ?.7 9.37163340(2) 4 ?2 4 .947 1.4 ?.66 4 ?1 .9 ?.4 4.8976624(2) 1.4 ?.6 196 A.3 Corrections to Frequency Due to Tension The nanomechanical resonators used in our research are made from amorphous silicon nitride, which has been deposited using low pressure chemical vapor deposi- tion (LPCVD). The resulting intrinsic stress ? int in the silicon nitride films is on the order of MPa?s [132], and is largely tensile. We can model the e?ect of the stress by including an e?ective tension T = ? int A in the equation of motion (Eq. A.1): [133] ?A ? 2 y ?t 2 + EI ? 4 y ?x 4 ?T ? 2 y ?x 2 =0. (A.13) Using dimensional-analysis, I can estimate the order of magnitude of the con- tribution of the tension to the restoring force. From Eq. (A.13), the ratio of tensile- to-bending force is TL 2 /EI.ForTL 2 /EI lessmuch 1, we can expect the bending moment to dominate, and the dynamics to be governed by the results of the previous section. For TL 2 /EI greatermuch 1, the tenisle force will dominate, and the dynamics will be similiar to the case of a tensioned string. For silicon nitride nanoresonators, however, the dimensionless ratio can range from TL 2 /EI lessmuch 1toTL 2 /EI ? 1. For this range, it would be helpful then to calculate the corrections to normal mode shape and frequency . Equation (A.13) can be solved exactly using separation of variables, yielding: Y n (x)=C n bracketleftbigg parenleftBigg sin? n x? ? n ? n sinh? n x parenrightBigg ? ? ? ? sin? n L? ? n ? n sinh? n L cos? n L ?cosh ? n L ? ? (cos? n x?cosh? n x) bracketrightbigg , (A.14) 197 where ? n L = k n L bracketleftBig (a 2 +1) 1 2 ?a bracketrightBig1 2 ,? n L = k n L bracketleftBig (a 2 +1) 1 2 + a bracketrightBig1 2 , and a = T 2EIk 2 , with k 4 n = ?A? 2 n EI . (A.15) The values of k are found numerically from the characteristic equation: cos(? n L)cosh(? n L)? 1 2 ( ? n ? n ? ? n ? n ) sinh(? n L)sin(? n L)=1. (A.16) It is straight-forward to verify that, as T ? 0, these expressions reduce to the corresponding expressions in Section A.1. Figure A.3 demonstrates the e?ect of tension on the frequency of the normal modes of a doubly-clamped silicon-nitride resonator. In the Fig. A.3(a), the ratio of frequency calculated with tension, ? n (T), to frequency calculated without tension, ? n (0), is plotted versus the dimensionless correction factor, TL 2 /12EI, for the first six normal modes of a resonator with a fundamental frequency of 13.037 MHz and length of 10 ?m. A bi-morph resonator with cross-sectional area A = 0.0375 ?m 2 , width w = 250 nm, and a 50 nm thick layer of gold as the conducting layer are assumed. Young?s moduli and densities of 300 GPa and 3000 Kg/m 3 and 50 GPa and 19.3 x 10 3 Kg/m 3 are assumed for the silicon nitride and gold respectively. It is apparent that for TL 2 /12EI < 0.01, tension increases the resonant fre- quency of the first six modes by less than 0.1%. For the fundamental mode, this corresponds to a frequency shift of ? 10 kHz. 198 (a) (b) Figure A.3: (a) The normalized mode frequency is plotted versus TL 2 /12EI,for the first six modes of a doubly-clamped resonator with fundamental frequency of 13.037 MHz and a length of 10 ?m. (b) The normalized fundmental mode frequency is plotted versus resonator length for intrinsic stress values of 1, 10, and 100 MPa?s. 199 As the tension is increased to TL 2 /12EI ? 1, the shift in resonant frequency for the fundamental mode grows to ? 15% . The shift in the higher order modes is smaller as they are e?ectively sti?er than the fundamental mode, and ranges from ? 1% to ? 8%. Figure A.3(b) demonstrates the shift in frequency of the fundamental mode of a doubly-clamped resonator as a function of the resonator?s length for ? int = 1, 10, and 100 MPa. For each curve, the tension is held constant, and the length of the resonator is increased. Here, I used the same values of EI, ?, and cross-section as were used in Fig. A.3(a). It is evident, that for silicon nitride resonators with lengths less than 10 ?m, the shift in frequency due to tension can be expected to be less than 10% for the fundamental mode - while not shown, the shift in the higher-order mode frequencies is even smaller. An alternative to solving Eq. A.13 exactly is to use a pertubative technique in which it is assumed that the e?ect of tension on the mode shape is negligible, the so-called Rayleigh Method [46]. Starting with Eq. A.13, I separate variables and solve for ? n : ? n (T)=? n (0) parenleftBigg 1+? n TL 2 12EI parenrightBigg1 2 , (A.17) with ? n (0) = ? n L 2 parenleftBigg EI ?A parenrightBigg1 2 , and ? n = 12 L 2 integraltext L 0 dx(?Y n (x)/?x) 2 integraltext L 0 dx (? 2 Y n (x)/?x 2 ) 2 , 200 where ? n = L 2 parenleftBigg integraltext L 0 dx (? 2 Y n (x)/?x 2 ) 2 integraltext L 0 dxY 2 n (x) parenrightBigg . I then make the approximation that Y n (x) are unaltered by the tension and given by Equation A.3. For this case, ? n are found to be equal to k n determined from Eq. A.5. The first ten ? n are listed in Table A.1. For TL 2 /12EI ? 1, I have found that the agreement between Eq. A.17 and the exact resonant frequency calculated by solving for the roots of Eq. A.16 is better than .01%. This implies that the approximation that the Y n (x) are left una?ected by tension T is a good one, and, thus, throughout the report, I simply use the mode-shapes given by Eq. A.3. A.4 The Driven-Damped Harmonic Oscillator I can append Eq. A.1 to account for external non-dissipative forces by simply adding in a term F(x,t) (in dimensions of Force/Length). I can account for damping by also inserting a phenomenological term proportional to the resonator?s transverse velocity. Dissipation in nanomechanical resonators is not well understood [121], and several mechanisms including thermoelastic loss [134], attachment loss [135] [136], and loss due to the measurement process itself [55] [137] have been proposed. Some authors account for damping e?ects by defining a complex Young?s modulus where dissipation is incorporated in the imaginary component (see [134] [138]). For the purpose of this appendix, however, it is su?cient to account for the damping by 201 inserting a velocity-dependent term, as it captures the general physics. Thus, ?A ? 2 y(x,t) ?t 2 + EI ? 4 y(x,t) ?x 4 + ? ?y(x,t) ?t = F(x,t), (A.18) Following the approach of Reference [139], I assume that the damping and driving force have a negligible e?ect on the mode shapes Y n (x). I then substitute solutions of the form y n (x,t)=Y n (x)y n (t), (A.19) into Eq. A.18, multiply by Y n (x), and integrate over the length of the resonator, obtaining the equation of motion for a mode, n,: m eff,n ? 2 y n (t) ?t 2 + K n y n (t)+? n ?y n (t) ?t = f n (t), (A.20) where ? n = integraldisplay L 0 dxY 2 n (x)? (A.21) and f n (t)= integraldisplay L 0 dxF(x,t)Y n (x). (A.22) Equation A.20 is the familiar damped-driven harmonic oscillator equation of motion. A simple case to treat, and one which will be important for Section A.5, is when F(x,t) is spatially invariant and has an harmonic time-dependence: F(x,t)= F o L e j?t , (A.23) where F o is the magnitude of the force. 202 In this case, f n (t)=? n F o e j?t ,where? n is the average of Y n (x)overthelength of the resonator, ? n = 1 L integraldisplay L 0 dxY n (x), (A.24) and determines the projection of the force on a given mode (Table A.1). The steady- state solutions, found for t/? greatermuch 1, where ? = m eff,n /? n , are then given by: y n (t)=A n e j?t , (A.25) where A n = ? n F o m eff,n ((? n 2 ?? 2 )+j (? n ?/Q e )) . (A.26) For a general force F(x,t), Eq. A.26 is replaced by A n = integraltext L 0 dxY n (x)F(x) m eff,n ((? n 2 ?? 2 )+j (? n ?/Q e )) . (A.27) The resonant frequencies, ? n , are given by Eq. A.4. The e?ective quality factor, Q e is defined as Q e = ? n m eff,n /? n , and sets the width of the resonator?s frequency response. I assume that it could be a result of dissipation from coupling to both the measurement environment and a thermal reservoir. Finally, I define the phase di?erence, ? n , between drive signal and resonator response ? n = arctan parenleftBigg ? n ?/Q e ? n 2 ?? 2 parenrightBigg . (A.28) A.5 The Magnetomotive Technique For the past decade, researchers have used various forms of magnetomotive detection to study the properties of nanomechanical resonators [140] [141] [142] 203 Figure A.4: a). Schematic of Magnetomotive technique. b). Circuit diagram dis- playing electromechanical impedance, Z m , current drive, I, and voltage amplifier - assumed to have infinite input impedance. 204 [36]. In its simplest realization (Fig. A.4(a)), ac current I is applied, length-wise, through a metallized resonator; in the presence of a transverse magnetic field B the resonator is driven by the Lorentz force F = IBL andanEMFepsilon1 n = BLv n , develops across it?s length L. Here, v n = ?y n (t) ?t 1 L integraldisplay L o dxY n (x)=? n ?y n (t) ?t (A.29) is resonator?s mode-dependent velocity. From Eqs. A.25 and A.26, the electromotive response takes the form: epsilon1 n = j? 2 n B 2 L 2 ? m eff,n (? n 2 ?? 2 + j? n ?/Q eff ) I. (A.30) The response of the resonator is measured by sweeping the frequency of the applied current through the mechanical resonance, and simultaneously measuring the induced EMF. The magnetomotive measurement is thus an impedance measure- ment. In fact, it is apparent that the response function is equivalent to a parallel RLC circuit (Fig. A.4(b)) with an electromechanical impedance defined as [137] 1 Z m = Q e j? n ?R n parenleftBig ? 2 n ?? 2 + j? n ?/Q e parenrightBig , (A.31) where R n = ? 2 n Q e B 2 L 2 m eff,n ? n , (A.32) and ? 2 n =(L n C n ) ?1 ,withL n = ?L 2 B 2 /m eff,n ? 2 n ,andC n = m eff,n /? n L 2 B 2 . Figure A.5 demonstrates the magnetomotive measurement of the fundamental mode resonance of device 1. A 200 ?V rms voltage signal was applied through a 10 k? resistance to provide the current I. The data was taken at a mixing chamber 205 Figure A.5: Plot of the response of the fundamental mode resonance of Device 1, measured using magnetomotive detection. A lock-in was used to measure the quadrature components of the resonator response. The response (solid line) was fit to a driven-damped harmonic oscillator response. The data (circles) were taken at a mixing chamber temperature of 15 mK and magnetic field B = 6 T. A drive current of I ? 10 nA was used. The quality factor, resonant frequency, and amplitude were determined to be 10881(3), 17.9756642(3)MHz, and 601.1(1) nV respectively. 206 temperature of 15 mK and a magnetic field of 6 T. A lock-in was used to measure the quadrature components of the resonator?s response. These were then fit to a driven- damped harmonic oscillator response. In the plot, the circles are the amplitude of the response from the measured quadratures, and the line is the amplitude of the least-squares fits to the individual quadratures. The resonant frequency and quality factor were extracted from the fit and determined to be f 1 = 19.976 MHz and Q e = 10.9 x 10 3 . While the resonant frequency agrees very well with results of the SET displacement detection technique, the quality factor is substantially lower. This is a result of the loading from the capacitance of the co-axial cable and the 50 ? amplifier impedance, which is much greater than the loading from SET detection. For more details on the loading e?ect and the magnetomotive measurement in general, please see reference [137]. 207 Appendix B Useful SET and RFSET Information Knowledge of the SET and measurement circuit parameters is essential for both the determination of the operating points of the RF-SET displacement detec- tion scheme and for the characterization of its performance. In this appendix, using the normal state and superconducting state current-voltage (IV) characteristics, I first demonstrate how to extract the coupling-capacitance C NR , the gate capaci- tance C g , the junction capacitances C j , the SET junction resistances R j ,andthe superconducting gap energy ?. I then discuss a technique to evaluate the frequency response of the measurement circuit. Finally, I summarize a method that allows for the calibration of the charge sensitivity. B.1 SET Parameters IV Map Measurement To determine the SET parameters, four DC measurements of the SET source- drain current, I SD , were made (Fig. B.1): in the normal state, I SD as a function of the source-drain bias V SD and the resonator bias V NR ; in the normal state, I SD as a function of V SD and the gate bias V g ; in the superconducting state (SSET), I SD as a function of V SD and V NR 1 ; in the superconducting state, I SD as a function of V SD 1 Both of the leads and the island are superconducting so the SET is in fact an (SSS) SET. 208 Figure B.1: Circuit diagram for IV map measurement . and V g . The voltages are set and swept by a computer-controlled digital-to-analog card. For each increment of the voltages, the current is sensed by a transimpedance amplifier, and the output is measured by a digital voltmeter and recorded by the computer through GPIB. A 1 Tesla magnetic field is applied to operate the SET in the normal state. The Normal-State Characteristics: The Orthodox Theory and Capac- itance Calculations Figure B.2 demonstrates a typical result of a normal-state IV map measure- ment. Coulomb-blockade suppresses I SD for V SD below a threshold voltage V t ,which 209 Figure B.2: Normal state IV map Device 2. 210 is periodic in V NR (V g ) with period e.AboveV t , I SD asymptotes to a linear depen- dence on V SD . The value and periodicity of V t are sensitive functions of C j and C NR (C g ), and the asymptotic behavior of I SD at large V SD is a function of the series combination of the R j ?s. Both limits, the onset of current and large V SD , are described by the orthodox theory of single-electron tunneling [68], and can, in principle, be used to extract the corresponding parameters [143] [144] [145]. In the orthodox theory, it is assumed that the charge state of the SET is- land evolves stochastically through single-electron tunneling events, yielding, at any instant of time, a value of n electrons with steady-state probability P(n)[68]. The tunneling events occur through either of the SET junctions, i,andin either direction, on (+) or o? of (-) the SET island. They are characterized by the tunneling rates ? ? i [68]. The net charge transfer or current through each junction is determined by performing a weighted sum over all charge states, n, of the di?erence between the (+) and (-) tunneling rates [68]: I 1 = ?e ? summationdisplay n=0 P(n) parenleftBig ? + 1 ?? ? 1 parenrightBig I 2 = ?e ? summationdisplay n=0 P(n) parenleftBig ? ? 2 ?? + 2 parenrightBig . As P(n) is assumed to be stationary in time, charge accumulation on the SET island does not occur, and the current through each junction must be equal, yielding [68] I SD = I 1 = I 2 . (B.1) 211 The assumption that P(n) be stationary, also leads to the condition of detailed balance [68], P(n +1)? ? (n +1)=P(n)? + (n), (B.2) where ? ? (n +1)=? ? 1 (n +1)+? ? 2 (n +1) and ? + (n)=? + 1 (n)+? + 2 (n). The tunneling rates ? ? i are calculated using Fermi?s golden rule [50]: ? ? i (n)= ?F ? i (n) e 2 R i 1 e ??F ? i (n) ?1 , (B.3) where ? =k B T and ?F ? i (n) is the change in system free-energy accompanying a particular tunneling event, and given, for the case of an asymmetrically biased SET, by ?F ? 1 (n)=?E c bracketleftbigg 2 parenleftbigg n? C NR V NR e ? C g V g e parenrightbigg ?1+ + 2(C 2 + C NR + C g )V SD e bracketrightbigg (B.4) and ?F ? 2 (n)=?E c bracketleftbigg 2 parenleftbigg n? C NR V NR e ? C g V g e parenrightbigg ?1 ? 2C 1 V SD e bracketrightbigg . (B.5) where E c = e 2 /2C ? is the charging energy or electrostatic cost for the tunneling of a single electron, and C ? = C 1 + C 2 + C NR + C g . Knowing ? ? i , P(n)andI SD can be calculated using Eqs. B.2 and B.1. 212 I can obtain a quantitative understanding of the Coulomb-blockade regime without explicitly solving for P(n). For simplicity?s sake, I first set n =0,V NR =0 ,andV g = 0 in Eqs. B.4 and B.5). For small |V SD |,?F ? i (n) > 0, and the work done by the bias V SD is not enough to overcome the charging energy E c .Consequently all four transisitions (?, junctions 1 and 2) are exponentially suppressed through Eq. B.3, and no current is observed. As V SD is increased from zero bias, eventually one of the transitions becomes energetically favorable. That is, either ?F + 2 (0)=0or?F ? 1 (0) = 0, depending on which threshold is smaller, V +,2 T (0) = e/2C 1 or V ?,1 T (0) = e/(2(C 2 + C NR + C g )). If C 1 > (C 2 +C NR +C g ), then an electron tunnels onto the island. As a result, n =1, and the corresponding discharging step, ?F ? 1 (1), becomes energetically favorable, and an electron tunnels o? through junction 1. After the discharge, n =0,andthe charging step through junction 2 again becomes favorable, and so on. On the other hand, if C 1 < (C 2 + C NR + C g ), then V ?,1 t (0) 0 for all four transisitions (?, junctions 1 and 2) and all n. From Eqs. B.4 and B.5, the width of the blockade in V SD is seen to be periodic in C NR V NR with a period of e: minimized when C NR V NR = en/2; and maximized when C NR V NR = 0 or a multiple of ne. Knowing the blockade to have period e,I can calculate C NR from the relation C NR ?V NR = e, (B.6) where ?V NR is the corresponding periodicity in mV (Fig. B.3). Similarly, I can calculate C g , from the normal-state IV map with V NR =0(notshown), C g ?V g = e. (B.7) 214 V NR (mV) V SD (mV) ?5 0 5 ?0.8 ?0.6 ?0.4 ?0.2 0 0.2 0.4 0.6 0.8 C NR ? V NR = e V +,2 V ?,1 V ?,2 V +,1 Figure B.3: Extracting capacitances from normal-state IV Device 2. 215 An expression for the edge of the blockade V t can be derived by noting that the onset of tunneling occurs when at least one of ?F ? i satisfies ?F ? i =0: V ?,1 t = C NR V NR + C g V g (C NR + C g + C 2 ) ? e(1?2n) 2(C NR + C g + C 2 ) (B.8) and V ?,2 t = ? C NR V NR + C g V g C 1 + e(2n?1) 2C 1 . (B.9) With the knowledge of C NR and C g , the junction capacitances C 1 and C 2 can be extracted by equating the slopes of the experimental tunneling onset (Fig. B.3) with the pre-factor of V NR in Eqs. B.8 and B.9. Normal-State Characteristics: Junction Resistances I cannot simply extract R ? from electrostatic considerations, and must solve for I SD using the detailed balance condition (Eq. B.2) and the definitions of ? ? i .For V SD greatermuch 2E c /e, it is necessary to compute P(n) for several thousand n.IsolveEq. B.2 numerically, and find that the slope of I SD vs. V SD asymptotes to R ?1 ? .The serial resistance, R ? is thus extracted from the normal-state IV map by a linear fit of I SD at large V SD (Fig. B.4). With the knowledge of the junction capacitances C 1 and C 2 and the serial resistance R ? , a rough estimate of the individual junction resistances R 1 and R 2 can be made [145] (Table B.1). Assuming that the thicknesses of the two junctions are equal (both junctions are grown at the same time under similiar conditions of pressure and temperature, see Chapter 3) and that C i ? A i and R i ? 1/A i ,where 216 0 20 40 60 80 100 0 500 1000 1500 V SD (mV) I SD (nA) y = b*x ? a b=13.6 ? .2 R ? = 74 ? 1 k? Figure B.4: Extraction of junction resistance Device 2 217 A i is the cross-sectional area of junction-i, the individual junction resistances can be expressed as R 1 = C 2 C 1 + C 2 R ? (B.10) R 2 = C 1 C 1 + C 2 R ? . (B.11) Failure of Normal-State Extraction Method Two e?ects combine to make the extraction of the junction capacitances from the normal-state IV map unreliable: self heating of the SET island and quantum charge fluctuations (co-tunneling). Self-heating of the SET island results from the combination of dissipation in the SET island and poor thermal coupling between the electron gas and phonon bath [152]. Using the standard model for electron-phonon coupling [123] [124], and assuming that ? 50% of the total power dissipated in the SET is dissipated in the SET island, with the SET leads thermalized at the temperature of the phonon bath, the SET-island electron-bath temperature T island at the onset of tunneling can be estimated [152]: T island = parenleftbigg P island ?? parenrightbigg 1 5 , (B.12) where ? = 0.2 nW/?m 3 K 5 [152] is the electron-phonon coupling for aluminum and ? = 0.05 ?m 2 is the total volume of the SET island. I have assumed that the phonon-bath temperature, T b lessmuch T island .IfP island ? I SD V SD /2, with V SD = e/C ? and I SD ? e/4R j C ? from Eqs. B.3, B.4, and B.5, then T island ? parenleftBigg e 2 8R j C 2 ? ?? parenrightBigg1 5 ? 350 ?400mK (B.13) 218 for parameters C ? =1fFandR j = 50 k?. This estimate does not take into account cooling of the island due to the tunneling of electrons to and from the lower-temperature SET leads and should thus be considered as an upper-bound [152]. Nonetheless, the tunneling threshold is broadened, leaving the onset of current V t indistinct [151]. This is modeled with the orthodox theory. Figures B.4(a) and B.4(b) shows two simulations of the onset of tunneling of a normal-state SET with charging energy E C /k B =1.5KandC 1 = C 2 greatermuch C g =10aF,andC ? = 590 aF. In Fig. B.4(a), T island = 10 mK, and the onset of tunneling is very clear and readily fit, yielding onset contour slopes of 0.033 and thus E C = 1.5 K through Eqs. B.8 and B.9. In Fig. B.4(b), T island = 300 mK, the onset of tunneling is unclear, and di?erent slopes are obtained depending on the contour chosen. In addition to self-heating, quantum charge fluctuations, or co-tunneling, can round the onset of current [69]. In general, co-tunneling is the process of charge transfer through the SET via an energetically unfavorable intermediate virtual state, and is the dominant charge transfer mechanism within the coulomb-blockade regime [69]. From the energy-time uncertainty relation, for example, an electron may tunnel through one junction to a forbidden island state, and dwell there for time ?t = ?h/?E,where?E ? E C is the energy required to make the transition. During ?t, it is energetically favorable for an electron to tunnel o? the island through the second junction, resulting in a finite probability for the net transfer of one electron across the SET. The transition probability rate for the net process, and thus the contribution 219 (a) 0 2 4 6 8 x 10 ?4 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 1 2 3 4 5 6 x 10 ?9 |slope| ? .033 n g (e) V SD (V) T=10mK I SD (A) E C ? 1.5K (b) 0 2 4 6 8 x 10 ?4 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 1 2 3 4 5 6 x 10 ?9 T=300mK V SD (V) n g (e) |slope| ? .123 E C ? . 94 K I SD (A) Figure B.5: Normal-state IV simulation for (a) T = 10 mK and (b) T = 300 mK. 220 of co-tunneling to I SD , can be estimated by multiplying the transition probability rate for the forbidden transisition, ? 1 ? 1/R ? C Sigma , by the probability for the energetically favorable tunneling event ? 1 ?t: ? co?tunneling ? ? 1 R Q R ? , (B.14) where R Q = h/e 2 is the quantum of resistance. The contribution of co-tunneling is thus found to be a fraction R Q /R ? of the sequential-tunneling current. For our devices, R Q R ? ? .2 ? .8. This is expected to result in the renormalization of the charging energy by the same order of magnitude [150]: E prime C ? E C (1 ?4 R Q ? 2 R ? ). (B.15) By not accounting for the e?ect of co-tunneling in the normal-state IV charac- teristics, I thus expect to err by as much as 30% in the determination of the junction capacitances. While modeling of the SET self-heating is straight-forward, account- ing for the co-tunneling processes would require calculating 2nd-order (and higher, depending on the precision desired) transition matrices for the tunneling rates and could become both tedious and di?cult. Rather than going to all that trouble, a simpler method is to use the features of the superconducting IV curve. Superconducting IV Characteristics Figure B.6 demonstrates a typical result of the superconducting IV map mea- surement I SD V SD Vg. The complexity of the map reflects the large variety of tun- neling processes that can occur in an SSET due to the combination of Coulomb 221 Figure B.6: Supreconducting IV map Device 2. 222 blockade, Josephson tunneling e?ects, and the superconducting gap energy [148] [149]. Most prominent here are three distinct processes: the double Josephson- quasiparticle resonance (DJQP) [146], the Josephson-quasiparticle resonance (JQP) [147], and single quasiparticle tunneling. All three processes have been studied ex- tensively, and consideration of the respective dynamics can be used to infer the junction capacitances and the superconducting gap energy, ?. Figure B.7 is a color contour plot of the superconducting IV map in Fig. (B.6. The threshold for single quasiparticle tunneling defines the width of the IV plateau. This threshold corresponds to the energy required for a quasiparticle to overcome both the superconducting gap energy and the Coulomb charging barrier, and is seen from simple electrostatic considerations to vary between 4? and 4? + e/C ? .The minimum width of the plateau is thus 8?. Knowing ?, I can determine the Josephson coupling E j of each junction - I assume ? is equal for the leads and the island, which should be the case con- sidering that they have similiar cross-sections and composition - using the Ambe- gaokar/Barato? relation [144]: E j,i = ?h?? 4e 2 R i . (B.16) Within the plateau, current ridges are evident. Along these ridges, the bias voltage is su?cient for the resonant tunneling of Cooper-pairs, followed by the tun- neling of single quasiparticles [146] [147]. As the processes are initiated with the tunneling of a Cooper-pair, the thresholds for the processes can be determined us- ing electrostatic arguments similiar to the case of normal-state tunneling processes 223 V g (mV) V SD (mV) ?20 ?10 0 10 20 ?1.5 ?1 ?0.5 0 0.5 1 ?8 ?6 ?4 ?2 0 2 4 6 8 I SD (nA) 4E C /e 8E C /e m 1 m 2 8?/e 2E C /e Figure B.7: Extracting junction capacitances from superconducting IV map Device 2. 224 (previous section). From these thresholds, the slope of ridges can be determined: m 1 = C g C 2 + C NR + C g (B.17) m 2 = ? C g C 1 . (B.18) Having already determined C NR and C g ,Icanusem 1 and m 2 to determine the individual junction capacitances, Table B.1. The intersection of the ridges at V SD = e/C ? and V SD =2e/C ? are known as the DJQP and JQP resonance peaks respectively. In the case of the JQP resonance, it is energetically favorable for a Cooper-pair tunneling event to occur through either of the two junctions, followed by two sequential quasiparticle tunneling events through the opposite junction (for example, the charge state of the island, n,goes from 0 ? 2 ? 1 ? 0 and so on). In the case of the DJQP resonance, the Cooper-pair tunneling event is followed, first, by a single quasiparticle, then the tunneling of a Cooper-pair, and finally the tunneling of a single quasiparticle (0 ? 2 ? 1 ? -1 ? 0 and so on). For all four transitions in the DJQP cycle to be energetically favorable, it is necessary for E c > 2?/3. The locations of the JQP and DJQP intersections in terms of V SD give independent determinations of C ? (Table B.2). For Devices 1, 2, 3, I find that C ? ?s calculated from the slope of the current ridges and the C ? ?s calculated from the JQP and the DJQP process agree to within 15%. Comparison of C ? calculated using the JQP and DJQP process shows agree- ment within 7%. For Device 4, the DJQP peak was absent and the ridges of the JQP were very indistinct. The only available calculation of C ? was from the location of the JQP peaks. From the symmetry in the normal-state IV map with respect to V g 225 Table B.1: Junction capacitances and resistances of Devices 1-4. Device C 1 (aF) R 1 (k?) C 2 (aF) R 2 (k?) C g (aF) C NR (aF) 1 81 61 84 59 14 61 2 250 21 100 53 10 26 3 173 47 341 24 14 64 4 ? 600 ? 15 ? 600 ? 15 19 63 Table B.2: Total capacitance, charging energy, gap energy and Josephson energies Devices 1 - 4. (a) Total capacitance found by summing C i , C g , C NR .(b)Total capacitance found from position of JQP peak. (c) Total capacitance found from position of DJQP peak. Device C a ? (aF) C b ? (aF) C c ? (aF) E c e (?V ) ? e (?V ) E j,i e (?V ) 1 241 - 279 287 220 12, 12 2 386 410 435 184 220 37, 14 3 592 577 540 148 200 15, 21 4 - 1310 - 61 180 ? 40, 40 226 for Device 4, we assume that C 1 ? C 2 and R 1 ? R 2 . For Devices 1, 2, 3, I find that E C /e > 2?/3e, consistent with the observation of the presence of the DJQP resonance in the superconducting IV maps for these devices. For Device 4, E C /e ? ?/3, consistent with the absence of the DJQP resonance in the superconducting IV map. For Devices 1, 2, 3, I see that E C /e > E j /e. This is consistent with the absence of a supercurrent in the superconducting IV map for these devices. For Device 4, E C /e ? E j /e. This is consistent with the presence of a supercurrent modulated with V g in the superconducting IV map of Device 4. The reason for the chronologically increasing junction capacitances (decreasing charging energy) is not known, but is consistent with the observed line-widths of the SET leads and islands becoming progressively larger for each successive device (see Chapter 3). As well, it is consistent with the general trend of decreasing R ? and charge sensitivity. B.2 Measurement Circuit Parameters The measurement circuit parameters, including the RF tank circuit, were de- termined by applying a large dc bias (V SD greatermuch V t ) across the SET source-drain, and recording shot noise ring-up of the tank circuit with a spectrum analyzer [63] ( Fig. B.8). For V SD >V t and time-scales slow with respect to the SET tunneling time, ? R ? C ? ? 0.1 GHz ?1 , the spectral density of the SET shot noise is white and given 227 Figure B.8: Circuit schematic for the measurement of the shot noise ring-up of the tank circuit. 228 by S II = eI SD [43]. As the center frequency of the tank circuit was designed to be ? 1GHz< 1/R ? C ? , the shot noise served as a calibrated white noise source with which we could probe the measurement circuit?s frequency response. The resulting noise power density measured at the input of the spectrum analyzer (Fig. B.9) takes the form: P in similarequal GeI SD Z o f 4 o (f 2 T ?f 2 ) 2 +(ff T /Q) 2 . (B.19) Here, Q = parenleftbigg Z LC R ? + Z o Z LC parenrightbigg ?1 = Z LC parenleftbigg Z LCR + Z o parenrightbigg ?1 (B.20) is the loaded quality factor of the tank resonance, Z o is the 50 ? transmission line impedance, Z LC = ? L T /C T is the tank circuit characteristic impedance, Z LCR =L T /(R ? C T ) is the transformed-SET impedance on resonance, and G is the power gain of the measurement circuit. Because the LC circuit was superconducting at the measurement temperature, T ? 50 mK, and because the length of the circuit was at most 0.1? 1GHz ,Ihave assumed that the tank circuit was a dissipation-less, lumped-element LC component. Additionally, I have neglected the e?ect of loading on the resonant frequency, f o = 1/(2? ? L T C T ), as it was of the order Z o /R ? ? 0.001. Including the noise of the measurement circuit, on resonance, the noise power at the input of the spectrum analyzer thus takes the form: P similarequal GB parenleftBig eI SD Q 2 Z o + k B T det n parenrightBig , (B.21) where T det n , and B are measurement circuit noise temperature, and bandwidth re- spectively. I have assumed that T det n is independent of the SET bias point, at large 229 (a) (b) Figure B.9: Tank circuit response Device 3. (a) A lorentzian fit (red line) of the noise power versus frequency for I SD = 120 nA (black circles). (b) A linear fit (red line) of the integrated noise power versus I SD (black circles). 230 Table B.3: Measurement circuit matching, gain, and noise temperature Devices 1-4. T SET measured at I SD ? 120nA. Device Z LCR (?) ? max M(dB) T det n (K) G(dB) T SET (K) 1 - - - - - - 2 2.2-3.4 .87-.92 .72-1.2 20.2-31.7 62-64 4.3-6.6 3 2.5 .90 .82 12.6 67 5.8 4 5.6 .8 2.0 13.4 66.5 3.8 V ds [116]. The equivalent noise temperature of the detection scheme T o is then defined by dividing Eq. B.21 by ?GBk B ?: T o = eI SD Q 2 Z o k B + T det n . (B.22) Figure B.9 displays a typical result of the shot noise measurement. Figure B.9(a) is a plot of the noise spectrum as was measured using the spectrum analyzer. From a fit to Eq. B.19, I can extract the width ?f and center frequency f T of the resonance. From the width, I estimate the quality factor Q = f T /?f,andthe detection bandwidth ?F = ?f/2. From the definitions of f T and Q,IcalculateL T , C T , Z LC , and the impedance of the LRC on resonance Z LCR (Tables 3.1 and B.3). Figure B.9(b) displays the integrated noise power and total noise temperature of the detection scheme T o as a function of I SD . The integration was done over a 1 MHz band about the center frequency. As expected, the noise power increased linearly with I SD . Fitting Eq. B.21 to the data, I estimate G and T det n from the 231 slope and y-intercept respectively (Table B.3). For Devices 2, 3, and 4, T det n = 20.2 K, 12.6 K, 13.4 K respectively. The decrease of ? 40% between Device 2 and Devices 3 and 4 is believed to have been due mainly to the factor of ? 2 improvement in gain of the measurement scheme for the measurement of Devices 3 and 4. For Devices 3 and 4, the calculated gain, G,was? 2 dB less than the gain measured from room temperature transmission measurements (see Chapter (4)). For Device 2, G is ? 5 dB less than expected. Finally, the contribution of the SET shot noise T SET to the overall detection noise can be estimated by subtracting T det n from T o at a particular V SD .At? 120 nA, the smallest I SD at which we measured the output shot noise power, T SET ranged from ? 3.8 K - 5.8 K, or ? 20% - 50% of the total noise. B.3 Calibration of Charge Sensitivity The charge sensitivity of the detection scheme was measured using amplitude- modulated (AM) reflectometry (Chapter 5). Microwaves resonant with the LC cir- cuit were applied to the SET drain, and the reflected signal was recorded. Simulta- neously, a 1 MHz sine wave bias of charge amplitude ?Q g (in units of electrons) was applied to the gate of the SET, modulating the amplitude of the reflected carrier signal, and producing sidebands at f o ? nf,wheref = 1 MHz . The sidebands were recovered using homodyne detection, mixing the reflected signal with the carrier, and measured with a spectrum analyzer. The charge sensitivity was then calculated from the ratio of the background noise power level P Background to the power in the 232 Figure B.10: Illustration of the experimental determination of the charge sensitivity from amplitude-modulated reflectometry. The plot shows a 1 MHz sideband of the measured reflected signal after recovery with homodyne mixing at the carrier frequency. 233 1MHz sideband P 1MHz (Fig. B.10): ? S Q = ?Q g ? B 10 ?SNR/10 , (B.23) where SNR = P 1MHz (dBM)-P Background (dBm) is the signal-to-noise ratio and B is the resolution bandwidth of the spectrum analyzer. In practice, due to losses in the sample lead and losses in the cabling and filters inside the dilution refrigerator, the amplitude of the charge signal applied to the SET gate at 1MHz was not known. However, from the dependence of the reflected signal?s sideband amplitude on the amplitude of the 1 MHz modulation, I can calculate the losses in the circuit and calibrate the charge sensitivity. For a dc gate bias, V g ? 0, the sideband power response can be approximated by [153] P = P o sin(2??Q g sin 2?ft), (B.24) where f=1MHz. This can be expanded in terms of the Bessel functions J n (2??Q g ) [153]: P =2P o ? summationdisplay n=0 J 2n+1 (2??Q g )sin ((2n +1)?t). (B.25) Using lock-in detection, we measured the amplitude of the fundamental of the response, n = 0, as a function of the amplitude ?V g of the 1 MHz modulation at the output of the waveform generator (Fig. B.11). I assume that the relationship between the voltage modulation at the generator and the charge modulation at the device is given by ?Q g = A?V g /e. Fitting the response to J 1 (A?V g ), and knowing that the first zero of J 1 (x) occurs at x =3.832, I extract A and determine the ratio ?Q g /?V g in electrons/volt. From the value of ?V g ,Icalculate?Q g , and, using Eq. 234 B.23, I calculate the charge sensitivity at the operating points of the the Devices 1 - 4 (Table B.5). For Devices 3 and 4, using the Bessel function fit, I find 5 - 7 dB attenuation in the gate line. This is consistent with room temperature transmission measurements of the same line (see Chapter 4). For Devices 1 and 2, we were not aware of the Bessel function calibration technique. Thus the reported values of ? S Q are estimated using a value of 6 dB attenuation in the gate line. Finally, the uncoupled energy sensitivity of the total detection scheme, SET shot noise plus measurement circuit noise, is defined as [116] ?epsilon1 = S Q 2C ? . (B.26) 235 0 0.5 1 1.5 2 2.5 3 3.5 4 ?4 ?2 0 2 4 6 8 x 10 ?4 ?V g (V rms ) Sideband Amplitude (V rm s ) Figure B.11: Bessel function fit to sideband response Device 3. 236 Table B.4: Charge modulation calibration Devices 1 - 4. Note that the attenuation listed is the attenuation of the gate line. This does not include an additional 60 dB of attenuation due to attenuators put in place at the top of the fridge. Device A( e rms V ) ?V g (V rms ) ?Q g (e rms ) Attenuation (dB) 1 - 0.50 ? .015 ? 6 2 - 0.50 ? .015 ? 6 3 .036 1.0 .036 7 4 .062 1.0 .062 5 Table B.5: Charge sensitivity Devices 1 - 4. (a) Measured charge sensitivity. (b) Calculated charge sensitivity from curvature of I SD V SD V g map. Both at 1 MHz Device SNR (dB) ? S Q a (?e rms / ? Hz) ? S Q b ?epsilon1(J/Hz) ?epsilon1 SET (J/Hz) 1 56 20 30 85?h - 2 58 30 20 50?h 10?h 3 52 90 70 1300?h 600?h 4 50 200 150 3750?h 1100?h 237 BIBLIOGRAPHY [1] W. Heisenberg. ?Ueber den anschaulichen Inhalt der quantentheoretischen Kine- matik and Mechanik,? Zeitschrift f?urPhysik43 (1927), 172-198. [2]C.M.Caves,K.S.Thorne,R.W.P.Drever,V.D.Sandberg,andM.Zimmer- man. ?On the Measurement of a Weak Classical Force Coupled to a Quantum- Mechanical Oscillator. I. Issues of Principle,? Review of Modern Physics 52,no. 2 (1980), 341-392. [3] C. M. Caves.?Quantum Limits on Noise in Linear Amplifiers,? Physical Review D 26, no. 8 (1982), 1817-1839. [4] V. B. Braginsky and F. Y. Khalili. Quantum Measurement. Cambridge: Cam- bridge University Press, 1995. [5] H. P. Yuen. ?Contractive States and the Standard Quantum Limit for Monitoring Free-Mass Positions,? Physical Review Letters 51, no. 9 (1983), 719-722. [6] M. F. Bocko and R. Onofrio. ?On the Measurement of a Weak Classical Force Coupled to a Harmonic Oscillator: Experimental Progress,? Review of Modern Physics 68, no.3 (1996), 755-799. [7] Y. Hadjar, P. F. Cohadon, C. G. Amino?, M. Pinard, A. Heidmann. ?High- Sensitivity Optical Measurement of Mechanical Brownian Motion,? Europhysics Letters 47, no. 5 (1999), 545-551. 238 [8] G.M.Harry,I.Jin,H.J.Paik,T.R.Stevenson,andF.C.Wellstood.?Two-Stage Superconducting-Quantum-Interference-Device Amplifier in a High-Q Gravita- tional Wave Transducer,? Applied Physics Letters76, no. 11 (2000), 1446-1448. [9] P. Falferi et al. ?Noise Sources and Dissipation Mechanisms of a 120 ?h SQUID Amplifier,? Applied Physics Letters 82, no. 6 (2003), 931-933. [10] A. Buonanno and Y. Chen. ?Optical Noise Correlations and Beating the Stan- dard Quantum Limit in Advanced Gravitational-Wave Detectors,? Classical and Quantum Gravity 18, no. 15 (2001), L95-L101. [11] B. Abbott et al. ?Limits on Gravitational-Wave Emission from Selected Pulsars Using LIGO Data,? Physical Review Letters 94 (2005), 181103/1-6. [12] M. P. Blencowe and M. N. Wybourne. ?Sensitivity of a Micromechanical Dis- placement Detector Based on the Radio-Frequency Single-Electron Transistor,? Applied Physics Letters 77, no. 23 (2000), 3845-3847. [13] Y. Zhang and M. P. Blencowe. ?Intrinsic Noise of a Micromechanical Displace- ment Detector Based on the Radio-Frequency Single-Electron Transistor,? Jour- nal of Applied Physics 91, no. 7 (2002), 4249-4255. [14] R. G. Knobel and A. N. Cleland. ?Nanometre-Scale Displacement Sensing Using a Single Electron Transistor,? Nature 424 (2003), 291-293. [15] M. D. LaHaye, O. Buu, B. Camarota, and K. C. Schwab. ?Approaching the Quantum Limit of a Nanomechanical Resonator,? Science 304 (2004), 74-77. 239 [16] A. A. Clerk. ?Quantum-Limited Position Detection and Amplification: A Lin- ear Response Perspective,? Physical Review B 70, no. 24 (2004), 245306/1-9. [17] V. Giovannetti, S. Lloyd, L. Maccone. ?Quantum-Enhanced Measurements: Beating the Standard Quantum Limit,? Science 306 (2004), 1330-1336. [18] K. S. Thorne, ?Gravitational Radiation,? in Three Hundred Years of Gravita- tion, ed. S. W. Hawking and W. Israel (Cambridge: Cambridge University Press, 1987), 330-458. [19] H. J. Paik. ?Superconducting Tunable-Diaphragm Transducer For Sensitive Acceleration Measurements,? Journal of Applied Physics 47, no. 3 (1976), 1168- 78. [20] J. A. Sidles et al. ?Magnetic Resonance Force Microscopy,? Reviews of Modern Physics 67, no. 1 (1995), 249-265. [21] D. Rugar, R. Budaklan, H. J. Mamin, and B. W. Chul. ?Single Spin Detection by Magnetic Resonance Force Microscopy,? Nature 430, no. 6997 (2004), 329- 332. [22] K. L. Ekinci, Y. T. Yang, and M. L. Roukes. ?Ultimate Limits to Inertial Mass Sensing Based Upon Nanoelectromechanical Systems,? Journal of Applied Phyiscs 95, no. 5 (2004), 2682-2689. [23] A. N. Cleland and M. R. Geller. ?Superconducting Qubit Storage and Entan- glement with Nanomechanical Resonators,? Physical Review Letters 93,no.7 (2004), 070501/1-4. 240 [24] M. R. Geller and A. N. Cleland. ?Superconducting Qubits Coupled to Nanoelec- tromechanical Resonator: An Architecture for Solid-State Quantum Information Processing,? Physical Review A 71 (2005), 032311/1-17. [25] A. D. Armour, M.P. Blencowe, and K.C. Schwab. ?Entanglement and Deco- herence of a Micromechanical Resonator Via Coupling to a Cooper-Pair Box,? Physical Review Letters 88, no. 14 (2002), 148301/1-4 [26] E. K. Irish and K. C. Schwab. ?Quantum Measurement of a Coupled Nanome- chanical Resonator-Cooper-Pair Box System,? Physical Review B 68 (2003), 155311/1-7. [27] M. L. Roukes. ?Nanoelectromechanical Systems Face the Future,? Physics World 14 (2001), 25-28. [28] A. Hopkins, K. Jacobs, S. Habib, and K. C. Schwab. ?Feedback cooling of a nanomechanical resonator,? Physical Review B 68, no. 23 (2003), 235328/1-10. [29] R. Ruskov, A. N. Korotkov, and K. C. Schwab. ?Quantum Nondemolition Squeezing of a Nanomechanical Resonator,? IEEE Transactions on Nanotech- nology 4, no. 1 (2005), 132-140. [30] M. P. Blencowe and M. N. Wybourne. ?Quantum Squeezing of Mechanical Motion For Micron-Sized Cantilevers,? Physica B 280 (200), 555-556. [31] M. P. Blencowe. ?Quantum Electromechanical Systems,? Physics Reports 395 (2004), 159-222. 241 [32] D. H. Santamore, A. C. Doherty, and M. C. Cross. ?Quantum Nondemolition Measurement of Fock States of Mesoscopic Mechanical Oscillators,? Physical Review B 70 (2004), 144301/1-16. [33] I. Wilson-Rae, P. Zoller, and A. Imamoglu. ?Laser Cooling of a Nanomechanical Resonator Mode to its Ground State,? Physical Review Letters 92, no. 7 (2004), 075507/1-4. [34] M. L. Roukes et al. ?Zeptogram Mass Detection-Weighing Molecules,? Physics News Updates. The AIP Bulletin of Physics News no. 725 (2005). [35] Kittel, C. and H. Kroemer, Thermal Physics. 2nd ed. New York: Freeman, 1980. [36] X. M. H. Huang, C.A. Zorman, M. Mehregany, and M.L. Roukes. ?Nanodevice Motion at Microwave Frequencies,? Nature 421 (2003), 496. [37] K.Schwab,E.A.Henriksen,J.M.Worlock,andM.L.Roukes.?Measurement of the Quantum of Thermal Conductance,? Nature 404, no. 6781 (2000), 974- 977. [38] D. M. Meekhof, C. Monroe, B. E. King, W. M. Itano, and D. J. Wineland. ?Generation of Nonclassical Motional States of a Trapped Atom,? Physical Re- view Letters 76 (1996), 17961799. [39] H. Park et al. ?Nanomechanical Oscillations in a Single-C 60 Transistor,? Nature 407 (2000), 57-60. 242 [40] K. Hornberger et al. ?Collisional Decoherence Observed in Matter Wave Inter- ferometry,? Physical Review Letters 90, no. 16 (2003), 160401/1-4. [41] R. J. Schoelkopf, P. Wahlgren, A. A. Kozhevnikov, P. Delsing, and D. E. Prober. ?The Radio-Frequency Single-Electron Transistor: A Fast and Ultrasensitive Electrometer,? Science 280 (1998), 1238-1242. [42] M. H. Devoret and R. J. Schoelkopf. ?Amplifying Quantum Signals With the Single-Electron Transistor,? Nature 406 (200), 1039-1046. [43] A. N. Korotkov. ?Intrinsic Noise of the Single-Electron Transistor,? Physical Review B 49, no. 15 (1994), 10381-10392. [44] N. W. Ashcroft and N. D. Mermin. Solid State Physics. Orlando: Saunders College Publishing, 1976. [45] A. N. Cleland Foundations of Nanomechanics. New York: Springer, 2003. [46] H. A. C. Tilmans, M. Elwenspoek, and J. H. J. Fluitman. ?Microresonant Force Gauges,? Sensors and Actuators A 30, no. 1-2 (1992), 35-53. [47] W. H. Zurek, S. Habib, J. P. Paz. ?Coherent States Via Decoherence,? Physical Review Letters 70, no. 9 (1993), 1187-1190. [48] M. D. LaHaye, O. Buu, B. Camarota, K. C. Schwab, Unpublished. [49] R. K. Pathria. Statistical Mechanics. 2nd ed. Oxford: Butterworth-Heinemann, 1996. 243 [50] J. S. Townsend. A Modern Approach to Quantum Mechanics.NewYork: McGraw-Hill, 1992. [51] J. J. Sakurai. Modern Quantum Mechanics. Revised ed. New York: Addison- Wesley, 1994. [52] R. J. Glauber. ?Coherent and Incoherent States of the Radiation Field,? Phys- ical Review 131 (1963), 2766-2788. [53] R. Kubo. ?The Fluctuation-Dissipation Theorem,? Reports on Progress in Physics 29 (1966), 255-284. [54] R. J. Schoelkopf, A. A. Clerk, S. M. Girvin, K. W. Lehnert, and M. H. Devoret. ?Qubits As Spectrometers of Quantum Noise.? Quantum Noise In Mesoscopic Physics. Proceedings of the NATO Advance Research Workshop. Dordrecht, Netherlands: Kluwer Academic Publishers, 2003. p. 175-203. [55] A. D. Armour, M. P. Blencowe, and Y. Zhang. ?Classical Dynamics of a Nanomechanical Resonator Coupled to a Single-Electron Transistor,? Physical Review B 69, no. 12 (2004), 125313/1-15. [56] D. Mozyrsky and I. Martin. ?Quantum-Classical Transition Induced by Elec- trical Measurement,? Physical Review Letters 89, no. 1 (2002), 018301/1-4. [57] A. Y. Smirnov, L. G. Mourokh, and N. J. M. Horing. ?Nonequilibrium Fluc- tuations and Decoherence in Nanomechanical Devices Coupled to the Tunnel Junction,? Physical Review B 67, no. 11 (2003), 115312/1-9. 244 [58] A. A. Clerk, S. M. Girvin, A. K. Nguyen, and A. D. Stone. ?Resonant Cooper- Pair Tunneling: Quantum Noise and Measurement Characteristics,? Physical Review Letters 89, no. 17 (2002), 176804/1-4. [59] A. A. Clerk. Unpublished Notes (2004). [60] M. P. Blencowe. Unpublished Notes (2005). [61] A. A. Clerk, S. M. Girvin, and A. D. Stone. ?Quantum-Limited Measurement and Information in Mesoscopic Detectors,? Physical Review B 67, no. 16 (2003), 165324/1-12. [62] A. Aassime, G. Johansson, G. Wendin, R. J. Schoelkopf, and P. Delsing. ?Radio-Frequency Single-Electron Transistor as Readout Device For Qubits: Charge Sensitivity and Backaction,? Physical Review Letters 86, no. 15 (2001), 3376-3379. [63] A. Aassime, D. Gunnarsson, K. Bladh, P. Delsing, and R. Schoelkopf. ?Radio- Frequency Single-Electron Transistor: Toward the Shot-Noise Limit,? Applied Physics Letters 79, no. 24 (2001), 4031-4033. [64] A. Shnirman and G. Schon. ?Quantum Measurements Performed With a Single- Electron Transistor,? Physical Review B 57, no. 15 (1997), 15400/1-7. [65] FastCap UNIX version by RLE Computational Prototyping Group, Massachusetts Institute of Technology Cambridge, MA 02139, see www.fastfieldsolvers.com/links. 245 [66] V. O. Turin and A. N. Korotkov. ?Numerical Analysis of Radio-Frequency Single-Electron Transistor Operation,? Physical Review B 69, no. 19 (2004), 195310/1-13. [67] A. N. Korotkov and M. A. Paalanen. ?Charge Sensitivity of Radio-Frequency Single-Electron Transistor,? Applied Physics Letters 74, no. 26 (1999), 4052- 4054. [68] D. V. Averin and K.K Likharev. ?Coulomb Blockade of Single-electron Tun- neling, and Coherent Oscillations In Small Tunnel Junctions.? Journal of Low Temperature Physics 62, no. 3-4 (1986): 345-373. [69] D. V. Averin and Y. Nazarov, ?Macroscopic Quantum Tunneling of Charge and Co-Tunneling.? Single Charge Tunneling. Eds M. H. Devoret and H. Grabert. New York: Plenum, 1992. [70] D. Song, A. Amar, C. J. Lobb, and F. C. Wellstood. ?Advantages of Super- conducting Coulomb Blockade Electrometers,? IEEE Transactions on Applied Superconductivity 5, no. 2 (1995), 3085-3089. [71] A. N. Korotkov. ?Charge Sensitivity Of Superconducting Single-Electron Tran- sistor?. Applied Physics Letters 69, no. 17 (1996): 2593-2595. [72] M. S. Choi, F. Plastina, and R. Fazio. ?Shot Noise for Resonant Cooper Pair Tunneling,? Physical Review Letters 87, no. 11 (2001), 116601/1-4. [73]O.Buu,M.D.LaHaye,A.Naik,A.D.Armour,A.A.Clerk,K.C.Schwab, unpublished. 246 [74] MEMS and Nanotechnology Exchange, Reston, VA 20191, Tel. (703) 262-5368. [75] C. M. Hernandez, T. W. Murray, S. Krishnaswamy .?Photoacoustic Character- ization of the Mechanical Properties of Thin Flms,? Applied Physics Letters 80, no. 4 (2002): 691693. [76] D. M. Profunser, J. Vollmann, and J. Dual. ?Determination of the Mate- rial Properties of Microstructures by Laser Based Ultrasound,? Ultrasonics 42 (2004): 641-646. [77] Futurrex Inc., Franklin, NJ 07416, Tel. (973) 209-1563. [78] Alfa Aesar, Ward Hill, MA 01835, Tel. (978) 521-6300. [79] Madou, Marc. Fundamentals of Microfabrication. Boca Raton: CRC Press, 1997. [80] The recipe for the tri-layer was provided to us by Pierre Echternach. [81] JEOL USA, Inc., Peabody, MA 01960, Tel. (978) 535-5900. [82] Upperspace Co., Pryor, OK 74361, Tel. (918) 825-4844. [83] JC Nabity Lithography Systems, Bozeman, MT 59717, Tel. (406) 587-0848. [84] MicroChem Co., Newton MA 02464, Tel. (617) 965-5511. [85] J. G. Lu, J. M. Hergenrother, and M. Tinkham. ?E?ect of Island Length on the Coulomb Modulation in Single-Electron Transistors,? Physical Review B 57, no. 8 (1998): 4591-4598. 247 [86] T. A. Fulton and G. J. Dolan. ?Observation of Single-Electron Charging E?ects in Small Tunnel Junctions,? Physical Review Letters 59, no. 1 (1987), 109-12. [87] Arch Chemicals, Inc., Norwalk, CT 06856, Tel.(203) 229-2900 [88] Atomergic Chemetals Co., Farmingdale, NY 11735, Tel. (631) 694-9000. [89] R.D. Mathis Company, Long Beach, CA 90809, Tel. (562) 426-7049. [90] Sierra Proto Express, Sunnyvale, CA 94086, Tel. (408) 735-7137. [91] Technic, Cranston, RI 02910, Tel. (401) 781-6100. [92] Sigma Aldrich Chemicals, St. Louis, MO 63178, Tel. (800) 588-9160. [93] TVZ 0001, GE varnish by Oxford Instruments, Concord, MA 01742, Tel. (978) 369-9933. [94] ISO 100CP by Burr Brown Co., Tuscon, AZ 85734, Tel. (520) 746-1111. [95] DC block 8039 by Inmet, Ann Arbor, MI 48103, Tel. (888) 244-6638. [96] SLP and VLP series by Mini-Circuits, Brooklyn, NY 11235, Tel., (718) 934- 4500. [97] J. M. Martinis, M.H. Devoret, and J. Clarke. ?Experimental Tests For the Quantum Behavior of a Macroscopic Degree of Freedom: The Phase Di?erence Across a Josephson Junction,? Physical Review B 35 (1987), 4682-4698. [98] Pi filters 1293-001 and 1250-059 by Spectrum Control, Columbia, MD 21045 Tel. (866) 281-0989. 248 [99] 28% part B of A Stycast 1266 by Lake Shore Cryotronics, Inc., Westerville OH 43082 Tel. (614) 891-2244. [100] SHP series by Mini-Circuits, Brooklyn, NY 11235, Tel., (718) 934-4500. [101] BW-S40 series by Mini-Circuits, Brooklyn, NY 11235, Tel., (718) 934-4500. [102] cc-ss-500 braided stainless steel coax by Lake Shore Cryotronics, Inc., West- erville OH 43082 Tel. (614) 891-2244. [103] SC-033/50-CN-CN CuNi-CuNi Coax, 0.33 mm outer diameter by Alahlam Limited,3-25-4-101 Matsushiro Tsukuba-shi Ibaraki, 305-0035 Japan. [104] C3203-20 by MAC Technology, Klamath Falls, OR97603, Tel., (800) 428-0341. [105] ZFBT-4R2GW by Mini-Circuits, Brooklyn, NY 11235, Tel., (718) 934-4500. [106] Nb-Nb 0.085 in. OD Coax, PTFE dielectric by KeyCom U.S.A. Co., Valley Stream, NY 11581 USA, Tel. (516) 825-1835. [107] L1.5-30H by Berkshire Technologies, Oakland, CA 94609, Tel. (510) 655-1986. [108] model number 15542 by Mini-Circuits, Brooklyn, NY 11235, Tel., (718) 934- 4500. [109] RX-102a-br by Lake Shore Cryotronics, Inc., Westerville OH 43082 Tel. (614) 891-2244. [110] AVS-47 resistance bridge by RV Elektronikka Oy Veromiehentie(Picowatt) 14 FIN-01510 VANTAA Finland, Tel. 358-9-822087. 249 [111] Pobell, F. Matter and Methods at Low Temperatures. 2nd ed. Berlin: Springer, 1996. [112] Nuclear Orientation Kit by Oxford Instruments, Concord, MA 01742, Tel. (978) 369-9933. [113] Lusher, C.P., J. Li, V.A. Maidanov, M.E. Digby, H. Dyball, A. Casey, J. Nyeki, V.V. Dmitiev, B.P. Cowan, J. Saunders. ?Current Sensing Noise Ther- mometry Using A Low T C DC SQUID Preamplifier.? Measurement Science and Technology 12, (2001): 1-15. [114] Model 550 SQUID system by Quantum Design San Diego, CA 92121 Tel.(858) 481-4400. [115] ZFM-2000 Frequency Mixer by Mini-Circuits, Brooklyn, NY 11235, Tel., (718) 934-4500. [116] L. Roschier, P. Hakonen, K. Bladh, P. Delsing, K. W. Lehnert, L. Spietz, and R. J. Schoelkopf. ?Noise Performance of the Radio-Frequency Single-Electron Transistor,? Journal of Applied Physics 95, no. 3 (2004), 1274-1286. [117] W. H. Press, S.A. Teukolsky, W. T. Vetterling, B. P. Flannery. Numerical Recipes in C. 2nd ed. New York: Cambridge University Press, 1992. [118] E. Fukushima. Experimental Pulse NMR: a Nuts and Bolts Approach. Reading: Addison-Wesley, 1981. 250 [119] M. M?uck, J. B. Kycia, and J. Clarke. ?Superconducting Quantum Interfer- ence Device as a Near-Quantum Limited Amplifier at .5 GHz,? Applied Physics Letters 78, no. 7 (2001), 967-969. [120] M. M?uck. ?Increasing the Dynamic Range of a SQUID Amplifier by Negative Feedback,? 368 (2002),141-145. [121] A. B. Hutchinson, P. A. Truitt, K. C. Schwab, L. Sekaric, J. M. Parpia, H. G. Craighead, J. E. Butler. ?Dissipation in Nanocrystalline-Diamond Nanome- chanical Resonators,? Applied Physics Letters 84, no. 6 (2004), 972-974. [122] J. P. Holman. Heat Transfer. 6th ed. New York: McGraw-Hill, 1986. [123] M. L. Roukes, M. R. Freeman, R. S. Germain, R. C. Richardson, and M. B. Ketchen. ?Hot Electrons and Energy Transport In Metals at MilliKelvin Tem- peratures,? Physical Review Letters 55, no. 4 (1985), 422-425. [124] F. C. Wellstood, C. Urbina, and J. Clarke. ?Hot-Electron E?ects In Metals,? Physical Review B 49, no. 9 (1994), 5942-5955. [125] M. M. Levio and J. P. Pekola. ?Thermal Characteristics of Silicon Nitride Membranes at Sub-Kelvin Temperatures,? Applied Physics Letters 72 (1998), 1305-1307. [126] B. A. Glavin. ?Low-Temperature Heat Transfer in Nanowires,? Physical Re- view Letters 86, no. 19 (2001), 4318-4321. 251 [127] R. Lifshitz and M. C. Cross. ?Elastic Wave Transmission At an Abrupt Junc- tion In a Thin Plate With Application to Heat Transport and Vibrations in Mesoscopic Systems,? Physical Review B 64, no. 8 (2001), 085324/1-22. [128] L. D. Landau and E. M. Lifshitz. Theory Of Elasticity.3rded.Oxford: Butterworth-Heinemann, 1986. [129] Y. T. Yang, K. L. Ekinci, X. M. H. Huang, L. M. Schiavone, M. L. Roukes, C. A. Zorman and M. Mehregany. ?Monocrystalline silicon carbide nanoelectrome- chanical systems,? Applied Physics Letters 78, no. 2 (2001), 162-164. [130] see http://www.memsnet.org/material/ [131] M. A. Haque and M. T. A. Saif. ?Deformation Mechanisms in Free-Standing Nanoscale Thin Films: A Quantitative in situ Transmission Electron Microscope Study,? Proceedings of the National Academy of Sciences of the United States of America 101, no. 17 (2004), 6335-6340. [132] C. A. Zorman and M. Mehregany, ?Materials for Microelectromechanical Sys- tems,? The MEMS Handbook. Ed. M. Gad-El-Hak. Boca Raton: CRC Press, 2001. [133] S. Timoshenko and D. H. Young. Vibration Problems in Engineering. chapter 5, 3rd Ed. New York: D. Van Nostrand Company, 1955. [134] R. Lifshitz and M. L. Roukes. ?Thermoelastic Damping In Micro- and Nanomechanical Systems,? Physical Review B 61, no. 8 (2000), 5600-5609. 252 [135] D. M. Photiadis and J. A. Judge. ?Attachment Losses of High Q Oscillators,? Applied Physics Letters 85, no. 3 (2004), 482-484. [136] Y. Jimbo and K. Itao. ?Attachment Loss of a Vibrating Cantilever,? Journal of Horological Instruments Japan 47, no. 1 (1968), 1-12. [137] A. N. Cleland and M. L. Roukes. ?External Control of Dissipation In a Nanometer-Scale Radiofrequency Mechanical Resonator,? Sensors and Actua- tors A 72, no. 3 (1999), 256-261. [138] S. Rast, C. Wattinger, U. Gysin, and E. Meyer. ?Dynamics of Damped Can- tilevers,? Review of Scientific Instruments 71, no. 7 (2000), 2772-2775. [139] B. Yurke, D. S. Greywall, A. N. Pargellis, and P. A. Busch. ?Theory of Amplifier-Noise Evasion In an Oscillator Employing a Nonlinear Resonator,? Physical Review A 51, no. 5 (1995), 4211-4229. [140] D. S. Greywall, B. Yurke, P. A. Busch, A. N. Pargellis, R. L. Willett. ?Evading Amplifier Noise in Nonlinear Oscillators,? Physical Review Letters 72, no. 19 (1994), 2992-2995. [141] A. N. Cleland and M. L. Roukes. ?Fabrication of High Frequency Nanometer Scale Mechanical Resonators From Bulk Si Crystals,? Applied Physics Letters 69, no. 18 (1996), 2653-2655. [142] K. L. Ekinci, Y. T. Yang, X. M. H. Huang, and M. L. Roukes. ?Balanced Elec- tronic Detection of Displacement In Nanoelectromechanical Systems,? Applied Physics. Letters 81, no. 12 (2002), 2253-2255. 253 [143] J. G. Liu. ?Even-Odd Electron Number E?ect in Small Superconducting Is- lands.? Ph.D. dissertation, Harvard University 1997. [144] M. Tinkham. Introduction to Superconductivity. 2nd Ed. New York: Dover, 1996. [145] D. Song. ?Properties of Coulomb-Blockade Electrometers in the Superconduct- ing State.? Ph.D. dissertation, University of Maryland at College Park, 1997. [146] A. M. Van den Brink, G. Sch?on, and L. J. Geerligs. ?Combined Single-Electron and Coherent-Cooper-Pair Tunneling In Voltage-Biased Josephson Junctions,? Physical Review Letters 67, no. 21 (1991), 3030-3033. [147] T. A. Fulton, P. L. Gammel, D. J. Bishop, and L. N. Dunkelberger. ?Obser- vation of Combined Josephson and Charging E?ects In Small Tunnel Junction Circuits,? Physical Review Letters 63, no. 12 (1989), 1307-1310. [148] P. Hadley, E. Delvigne, E. H. Visscher, S. L?ahteenm?aki, and J. E. Mooij. ?3e Tunneling Processes In a Superconducting Single-Electron Tunneling Transis- tor,? Physical Review B 58, no. 23 (1998), 15317-15320. [149] R. J. Fitzgerald, S. L. Pohlen, and M. Tinkham. ?Observation of Andreev Reflection In All-Superconducting Single-Electron Transistors,? Physical Review B 57, no. 18 (1998), 11073-11076. [150] M. Thalakulam, Z. Ji, and A.J. Rimberg. ?Sensitivity and linearity of Super- conducting Radio-Frequency Single-Electron Transistors: E?ects of Quantum Charge Fluctuations,? Physical Review Letters 93, no. 6 (2004), 066804/1-4. 254 [151] I would like to thank Rob Schoelkopf and Per Delsing for bringing this fact to our attention. [152] R. L. Kautz, G. Zimmerli, and J.M. Martinis. ?Self-Heating In the Coulomb- Blockade Electrometer,? Journal of Applied Physics 73, no. 5 (1993), 2386-2396. [153] Wahlgren, P. ?The Radio Frequency Single-Electron Transistor and the Hori- zon Picture for Tunneling.? Ph.D. dissertation, Chalmers University of Technol- ogy, G?oteborg University, 1998. 255