Introduction to Noise, Measurement and Amplification

A.A. Clerk,1 M.H. Devoret,2 S.M. Girvin,3 Florian Marquardt,4 and R.J. Schoelkopf2 1Department of Physics, McGill University, 3600 rue University Montr´eal,QC Canada H3A 2T8∗ 2Department of Applied Physics, Yale University PO Box 208284, New Haven, CT 06520-8284 3Department of Physics, Yale University PO Box 208120, New Haven, CT 06520-8120 4Department of Physics, Center for NanoScience, and Center for Theoretical Physics, Ludwig-Maximilians-Universit¨atM¨unchen Theresienstr. 37, D-80333 M¨unchen,Germany (Dated: April 15, 2010)

The topic of has become extremely timely due to the rise of quantum informa- tion physics and the resulting interchange of ideas between the condensed matter and atomic, molecular, opticalquantum optics communities. This review gives a pedagogical introduction to the physics of quantum noise and its connections to quantum measurement and quantum ampli- fication. After introducing quantum noise spectra and methods for their detection, the basics of weak continuous measurements are described. Particular attention is given to the treatment of the standard quantum limit on linear amplifiers and position detectors within a general linear- response framework. This approach is shown how it relates to the standard Haus-Caves quantum limit for a bosonic amplifier known in quantum optics and its application to the case of electrical circuits is illustrated, including mesoscopic detectors and resonant cavity detectors.

Contents D. Scattering versus op-amp modes of operation 31 E. Linear response description of a position detector 33 I. Introduction 2 1. Detector back-action 33 2. Total output noise 34 II. Quantum Noise Spectra 6 3. Detector power gain 34 A. Introduction to quantum noise 6 4. Simplifications for a quantum-ideal detector 36 B. Quantum spectrum analyzers 7 5. Quantum limit on added noise and noise temperature 36 III. Quantum Measurements 9 F. Quantum limit on the noise temperature of a voltage A. Weak continuous measurements 10 amplifier 38 B. Measurement with a parametrically coupled resonant 1. Classical description of a voltage amplifier 39 cavity 11 2. Linear response description 40 1. QND measurement of the state of a using a 3. Role of noise cross-correlations 41 resonant cavity 13 G. Near quantum-limited mesoscopic detectors 42 2. Quantum limit relation for QND qubit state 1. dc SQUID amplifiers 42 detection 15 2. Quantum point contact detectors 42 3. Measurement of oscillator position using a 3. Single-electron transistors and resonant-level resonant cavity 16 detectors 42 H. Back-action evasion and noise-free amplification 43 IV. General Linear Response Theory 19 1. Degenerate parametric amplifier 44 2. Double-sideband cavity detector 45 A. Quantum constraints on noise 19 3. Stroboscopic measurements 46 1. Heuristic weak-measurement noise constraints 20 2. Generic linear-response detector 21 VI. Bosonic Scattering Description of a Two-Port 3. Quantum constraint on noise 22 Amplifier 46 4. Evading the detector quantum noise inequality 24 A. Scattering versus op-amp representations 46 B. Quantum limit on QND detection of a qubit 24 1. Scattering representation 47 2. Op-amp representation 47 V. Quantum Limit on Linear Amplifiers and 3. Converting between representations 48 Position Detectors 25 B. Minimal two-port scattering amplifier 49 arXiv:0810.4729v2 [cond-mat.mes-hall] 15 Apr 2010 A. Preliminaries on amplification 26 1. Scattering versus op-amp quantum limit 49 B. Standard Haus-Caves derivation of the quantum 2. Why is the op-amp quantum limit not achieved? 51 limit on a bosonic amplifier 27 C. Non-degenerate parametric amplifier 28 VII. Reaching the Quantum Limit in Practice 52 1. Gain and added noise 28 A. Importance of QND measurements 52 2. Bandwidth-gain tradeoff 29 B. Power matching versus noise matching 52 3. Effective temperature 30 VIII. Conclusions 53

Acknowledgements 54

∗Electronic address: [email protected] A. Basics of Classical and Quantum Noise 54 2

1. Classical noise correlators 54 motivation is that such systems can operate simultane- 2. The Wiener-Khinchin Theorem 55 ously at high frequencies and at low temperatures, en- 3. Square law detectors and classical spectrum tering the regime where ~ω > kBT . As such, quantum analyzers 56 zero-point fluctuations will play a more dominant role B. Quantum Spectrum Analyzers: Further Details 56 in determining their behaviour than the more familiar 1. Two-level system as a spectrum analyzer 56 thermal fluctuations. A second motivation comes from 2. Harmonic oscillator as a spectrum analyzer 58 the relation between quantum noise and quantum mea- 3. Practical quantum spectrum analyzers 59 surement. There exists an ever-increasing number of ex- a. Filter plus diode 59 periments in mesoscopic electronics where one is forced b. Filter plus photomultiplier 60 to think about the of the detection c. Double sideband heterodyne power spectrum 60 process, and about fundamental quantum limits which C. Modes, Transmission Lines and Classical constrain the performance of the detector or amplifier Input/Output Theory 60 used. 1. Transmission lines and classical input-output theory 61 Given the above, we will focus in this review on dis- 2. Lagrangian, Hamiltonian, and wave modes for a transmission line 63 cussing what is known as the “standard quantum limit” 3. Classical statistical mechanics of a transmission line 64 (SQL) on both displacement detection and amplification. 4. Amplification with a transmission line and a negative To preclude any possible confusion, it is worthwhile to resistance 66 state explicitly from the start that there is no limit to how well one may resolve the position of a particle in D. Quantum Modes and Noise of a Transmission Line 67 an instantaneous measurement. Indeed, in the typical 1. Quantization of a transmission line 67 Heisenberg microscope setup, one would scatter photons 2. Modes and the windowed Fourier transform 68 off an electron, thereby detecting its position to an ac- 3. Quantum noise from a resistor 70 curacy set by the wavelength of photons used. The fact that its momentum will suffer a large uncontrolled per- E. Back Action and Input-Output Theory for turbation, affecting its future motion, is of no concern Driven Damped Cavities 71 1. Photon shot noise inside a cavity and back action 71 here. Only as one tries to further increase the resolution 2. Input-output theory for a driven cavity 73 will one finally encounter relativistic effects (pair produc- 3. Quantum limited position measurement using a tion) that set a limit given by the cavity detector 77 of the electron. The situation is obviously very differ- 4. Back-action free single-quadrature detection 80 ent if one attempts to observe the whole trajectory of the particle. As this effectively amounts to measuring F. Information Theory and Measurement Rate 81 1. Method I 82 both position and momentum, there has to be a trade- 2. Method II 82 off between the accuracies of both, set by the Heisenberg uncertainty relation. The way this is enforced in practice G. Number Phase Uncertainty 83 is by the uncontrolled perturbation of the momentum during one position measurement adding to the noise in H. Using feedback to reach the quantum limit 84 1. Feedback using mirrors 84 later measurements, a phenomenon known as ”measure- 2. Explicit examples 85 ment back-action”. 3. Op-amp with negative voltage feedback 86 Just such a situation is encountered in “weak mea- surements” (Braginsky and Khalili, 1992), where one in- I. Additional Technical Details 87 1. Proof of quantum noise constraint 87 tegrates the signal over time, gradually learning more 2. Proof that a noiseless detector does not amplify 88 about the system being measured; this review will focus 3. Simplifications for a quantum-limited detector 89 on such measurements. There are many good reasons 4. Derivation of non-equilibrium Langevin equation 89 why one may be interested in doing a , 5. Linear-response formulas for a two-port bosonic rather than an instantaneous, strong projective measure- amplifier 90 ment. On a practical level, there may be limitations to a. Input and output impedances 90 the strength of the coupling between the system and the b. Voltage gain and reverse current gain 91 6. Details for the two-port bosonic voltage amplifier detector, which have to be compensated by integrating with feedback 92 the signal over time. One may also deliberately opt not to disturb the system too strongly, e.g. to be able to apply References 93 quantum feedback techniques for state control. More- over, as one reads out an oscillatory signal over time, one effectively filters away noise (e.g. of a technical na- I. INTRODUCTION ture) at other frequencies. Finally, consider an example like detecting the collective coordinate of motion of a mi- Recently, several advances have led to a renewed inter- cromechanical beam. Its zero-point uncertainty (ground est in the quantum mechanical aspects of noise in meso- state position fluctuation) is typically on the order of the scopic electrical circuits, detectors and amplifiers. One diameter of a proton. It is out of the question to reach 3 this accuracy in an instantaneous measurement by scat- added noise that is just the one expected for the product tering photons of such a small wavelength off the struc- of their respective gains. Therefore, as one approaches ture, since they would instead resolve the much larger the classical level, i.e. large overall gains, the SQL in its position fluctuations of the individual atoms comprising simplified form of half a photon added always applies. the beam (and induce all kinds of unwanted damage), in- Unlike traditional discussions of the amplifier SQL, we stead of reading out the center-of-mass coordinate. The will devote considerable attention to a general linear re- same holds true for other collective degrees of freedom. sponse approach based on the quantum relation between The prototypical example we will discuss several times susceptibilities and noise. This approach treats the am- is that of a weak measurement detecting the motion of a plifier or detector as a black box with an input port cou- harmonic oscillator (such as the mechanical beam). The pling to the signal source and an output port to access the measurement then actually follows the slow evolution of amplified signal. It is more suited for mesoscopic systems amplitude and phase of the oscillations (or, equivalently, than the quantum optics scattering-type approach, and the two quadrature components), and the SQL derives it leads us to the quantum noise inequality: a relation be- from the fact that these two do not com- tween the noise added to the output and the back-action mute. It essentially says that the measurement accuracy noise feeding back to the signal source. In the ideal case will be limited to resolving both quadratures down to the (what we term a “quantum-limited detector”), the prod- scale of the position fluctuations, within one uct of these two contributions reaches the minimum value mechanical damping time. Note that, in special appli- allowed by quantum mechanics. We will show that opti- cations, one might be interested only in one particular mizing this inequality on noise is a necessary pre-requisite quadrature of motion. Then the Heisenberg uncertainty for having a detector achieve the quantum-limit on a spe- relation does not enforce any SQL and one may again cific measurement task, such as linear ampification. obtain unlimited accuracy, at the expense of renouncing There are several motivations for understanding in all knowledge of the other quadrature. principle, and realizing in practice, amplifiers whose noise reaches this minimum quantum limit. Reaching the Position detection by weak measurement essentially quantum limit on continuous position detection has been amounts to amplifying the quantum signal up to a clas- one of the goals of many recent experiments on quan- sically accessible level. Therefore, the theory of quantum tum electro-mechanical (Cleland et al., 2002; Etaki et al., limits on displacement detection is intimately connected 2008; Flowers-Jacobs et al., 2007; Knobel and Cleland, to limits on how well an amplifier can work. If an ampli- 2003; LaHaye et al., 2004; Naik et al., 2006; Poggio et al., fier does not have any preference for any particular phase 2008; Regal et al., 2008) and opto-mechanical systems of the oscillatory signal, it is called “phase-preserving”, (Arcizet et al., 2006; Gigan et al., 2006; Marquardt and which is the case relevant for amplifying and thereby de- Girvin, 2009; Schliesser et al., 2008; Thompson et al., tecting both quadratures equally well1. We will derive 2008). As we will show, having a near-quantum lim- and discuss in great detail the SQL for phase-preserving ited detector would allow one to continuously monitor linear amplifiers (Caves, 1982; Haus and Mullen, 1962). the quantum zero-point fluctuations of a mechanical res- Quantum mechanics demands that such an amplifier onator. Having a quantum limited detector is also nec- adds noise that corresponds to half a photon added to essary for such tasks as single- NMR detection (Ru- each mode of the input signal, in the limit of high photon- gar et al., 2004), as well as gravitational wave detection number gain G. In contrast, for small gain, the minimum (Abramovici et al., 1992). The topic of quantum-limited number of added noise quanta, (1 − 1/G))/2, can be- detection is also directly relevant to recent activity ex- come arbitrarily small as the gain is reduced down to 1 ploring feedback control of quantum systems (Doherty (no amplification). One might ask, therefore, whether et al., 2000; Korotkov, 2001b; Ruskov and Korotkov, it shouldn’t be possible to evade the SQL by being con- 2002; Wiseman and Milburn, 1993, 1994); such schemes tent with small gains? The answer is no, since high gains necessarily need a close-to-quantum-limited detector. G  1 are needed to amplify the signal to a level where it can be read out (or further amplified) using classical de- This review is organized as follows. We start in Sec. II vices without their noise having any further appreciable by providing a short review of the basic statistical proper- effect, converting 1 input photon into G  1 output pho- ties of quantum noise, including its detection. In Sec. III tons. In the words of Caves, it is necessary to generate we turn to quantum measurements, and give a basic in- an output that “we can lay our grubby, classical hands troduction to weak, continuous measurements. To make on” (Caves, 1982). It is a simple exercise to show that things concrete, we discuss heuristically measurements feeding the input of a first, potentially low-gain amplifier of both a qubit and an oscillator using a simple resonant into a second amplifier results in an overall bound on the cavity detector, giving an idea of the origin of the quan- tum limit in each case. Sec. IV is devoted to a more rigorous treatment of quantum constraints on noise aris- ing from general quantum linear response theory. The 1 In the literature this is often referred to as a ’phase insensitive’ heart of the review is contained in Sec. V, where we give amplifier. We prefer the term ’phase-preserving’ to avoid any a thorough discussion of quantum limits on amplification ambiguity. and continuous position detection. We also briefly dis- 4 cuss various methods for beating the usual quantum lim- pecially relevant given recent experiments making use of its on added noise using back-action evasion techniques. microwave stripline resonators. These appendices appear We are careful to distinguish two very distinct modes as a separate on-line only supplement to the published ar- of amplifier operation (the “scattering” versus “op amp” ticle (Clerk et al., 2009), but are included in this arXiv modes); we expand on this in Sec. VI, where we discuss version of the article. In Table II, we list the contents of both modes of operation in a simple two-port bosonic these appendices. Note that while some aspects of the amplifier. Importantly, we show that an amplifier can topics discussed in this review have been studied in the be quantum limited in one mode of operation, but fail quantum optics and quantum dissipative systems com- to be quantum limited in the other mode of operation. munities and are the subject of several comprehensive Finally, in Sec. VII, we highlight a number of practical books (Braginsky and Khalili, 1992; Gardiner and Zoller, considerations that one must keep in mind when trying 2000; Haus, 2000; Weiss, 1999), they are somewhat newer to perform a quantum limited measurement. Table I pro- to the condensed matter physics community; moreover, vides a synopsis of the main results discussed in the text, some of the technical machinery developed in these fields as well as definitions of symbols used. is not directly applicable to the study of quantum noise In addition to the above, we have supplemented the in quantum electronic systems. Finally, note that while main text with several pedagogical appendices which this article is a review, there is considerable new mate- cover basic background topics. Particular attention is rial presented, especially in our discussion of quantum given to the quantum mechanics of transmission lines amplification (cf. Secs. V.D,VI). and driven electromagnetic cavities, topics which are es-

TABLE I: Table of symbols and main results.

Symbol Definition / Result General Definitions R ∞ iωt f[ω] Fourier transform of the function or f(t), defined via f[ω] = −∞ dtf(t)e † ˆ† R ∞ ˆ† iωt ˆ†  ˆ  (Note that for operators, we use the convention f [ω] = −∞ dtf (t)e , implying f [ω] = f[−ω] ) R +∞ iωt SFF [ω] Classical noise spectral density or power spectrum: SFF [ω] = −∞ dt e hF (t)F (0)i R +∞ iωt ˆ ˆ SFF [ω] Quantum noise spectral density: SFF [ω] = −∞ dt e hF (t)F (0)i ¯ ¯ 1 1 R +∞ iωt ˆ ˆ SFF [ω] Symmetrized quantum noise spectral density SFF [ω] = 2 (SFF [ω] + SFF [−ω]) = 2 −∞ dt e h{F (t), F (0)}i χAB (t) General linear response susceptibility describing the response of A to a perturbation which couples to B; i in the quantum case, given by the Kubo formula χAB (t) = − θ(t)h[Aˆ(t), Bˆ(0)]i [Eq. (2.14)] ~ A Coupling constant (dimensionless) between measured system and detector/amplifier, ˆ ˆ ˆ ˆ † e.g. V = AF (t)ˆσx, V = AxˆF, or V = A~ωcσˆzaˆ aˆ M, Ω Mass and angular frequency of a mechanical harmonic oscillator. q ~ xZPF Zero point uncertainty of a mechanical oscillator, xZPF = 2MΩ . γ0 Intrinsic damping rate of a mechanical oscillator due to coupling to a bath via Vˆ = AxˆFˆ : A2 γ0 = (SFF [Ω] − SFF [−Ω]) [Eq. (2.12)] 2M~Ω ωc Resonant frequency of a cavity κ, Qc Damping, quality factor of a cavity: Qc = ωc/κ

Sec. II Quantum noise spectra Teff [ω] Effective temperature at a frequency ω for a given quantum noise spectrum, defined via   SFF [ω] = exp ~ω [Eq. (2.8)] SFF [−ω] kBTeff [ω] Fluctuation-dissipation theorem relating the symmetrized noise spectrum to the dissipative part 1 ω for an equilibrium bath: S¯FF [ω] = coth( ~ )(SFF [ω] − SFF [−ω]) [Eq. (2.16)] 2 2kBT

Sec. III Quantum Measurements Number-phase uncertainty relation for a coherent state: 1 ∆N∆θ ≥ 2 [Eq. (3.6), (G12)] N˙ Photon number flux of a coherent beam δθ Imprecision noise in the measurement of the phase of a coherent beam Fundamental noise constraint for an ideal coherent beam: 1 SN˙ N˙ Sθθ = 4 [Eq. (3.8), (G21)] ¯0 Sxx(ω) symmetrized spectral density of zero-point position fluctuations of a damped harmonic oscillator S¯xx,tot(ω) total output noise spectral density (symmetrized) of a linear position detector, referred back to the oscillator S¯xx,add(ω) added noise spectral density (symmetrized) of a linear position detector, referred back to the oscillator

Sec. IV: General linear response theory 5

TABLE I: Table of symbols and main results.

Symbol Definition / Result xˆ Input signal Fˆ Fluctuating force from the detector, coupling tox ˆ via Vˆ = AxˆFˆ Iˆ Detector output signal General quantum constraint on the detector output noise, backaction noise and gain: 2 2  h ¯ i ¯ ¯ ¯ ~χ˜IF [ω] SIF [ω] SII [ω]SFF [ω] − SIF [ω] ≥ 1 + ∆ ˜ [Eq. (4.11)] 2 ~λ[ω]/2 ∗ 2 2 whereχ ˜IF [ω] ≡ χIF [ω] − [χFI [ω]] and ∆[z] = ( 1 + z − 1 + |z| )/2. [Note: 1 + ∆[z] ≥ 0 and ∆ = 0 in most cases of relevance, see discussion around Eq. (4.17)] α Complex proportionality constant characterizing a quantum-ideal detector: |α|2 = S¯ /S¯ and sin (arg α[ω]) = √ ~|λ[ω]|/2 [Eqs. (4.18,I17)] II FF ¯ ¯ SII [ω]SFF [ω]

Γmeas Measurement rate (for a QND qubit measurement) [Eq. 4.24] Γϕ Dephasing rate (due to measurement back-action) [Eqs. (3.27),(4.19)] Constraint on weak, continuous QND qubit state detection : η = Γmeas ≤ 1 [Eq. (4.25)] Γϕ

Sec. V: Quantum Limit on Linear Amplifiers and Position Detectors G Photon number (power) gain, e.g. in Eq. (5.7) √ Input-output relation for a bosonic scattering amplifier: ˆb† = Gaˆ† + Fˆ† [Eq.(5.7)] (∆a)2 Symmetrized field operator uncertainty for the scattering description of a bosonic amplifier: 2 1 † 2 (∆a) ≡ 2 {a,ˆ aˆ } − |hai| Standard quantum limit for the noise added by a phase-preserving bosonic scattering amplifier 2 1 in the high-gain limit, G  1, where h(∆a) iZPF = 2 : (∆b)2 2 1 G ≥ (∆a) + 2 [Eq. (5.10)]

GP [ω] Dimensionless power gain of a linear position detector or voltage amplifier (maximum ratio of the power delivered by the detector output to a load, vs. the power fed into signal source): 2 |χIF [ω]| GP [ω] = [Eq. (5.52)] 4Im χFF [ω]·Im χII [ω]] h i2 Im α 4kBTeff For a quantum-ideal detector, in the high-gain limit: GP ' [Eq. (5.57)] |α| ~ω  ω  S¯xx,eq[ω, T ] Intrinsic equilibrium noise S¯xx,eq[ω, T ] = coth ~ [−Im χxx[ω]] [Eq. (5.59)] ~ 2kBT Aopt Optimal coupling strength of a linear position detector which minimizes the added noise at frequency ω: ¯ 4 SII [ω] Aopt[ω] = 2 ¯ [Eq. (5.64)] |λ[ω]χxx[ω]| SFF [ω] γ[Aopt] Detector-induced damping of a quantum-limited linear position detector at optimal coupling, fulfills γ[Aopt] Im α √ 1 ~Ω γ +γ[A ] = α = 4k T  1 [Eq. (5.69)] 0 opt GP [Ω] B eff Standard quantum limit for the added noise spectral density of a linear position detector (valid at each frequency ω): Sxx,add[ω] ≥ limT →0 Sxx,eq[w, T ] [Eq. (5.62)] Effective increase in oscillator temperature due to coupling to the detector backaction, for an ideal detector, with ~Ω/kB  Tbath  Teff : γ·Teff +γ0·Tbath Ω Tosc ≡ → ~ + Tbath [Eq. (5.70)] γ+γ0 4kB

Zin,Zout Input and output impedances of a linear voltage amplifier Zs Impedance of signal source attached to input of a voltage amplifier λV Voltage gain of a linear voltage amplifier V˜ (t) Voltage noise of a linear voltage amplifier (Proportional to the intrinsic output noise of the generic linear-response detector [Eq. (5.81)] ) I˜(t) Current noise of a linear voltage amplifier (Related to the back-action force noise of the generic linear-response detector [Eqs. (5.80)] ) TN Noise temperature of an amplifier [defined in Eq. (5.74)] ZN Noise impedance of a linear voltage amplifier [Eq. 5.77)] Standard quantum limit on the noise temperature of a linear voltage amplifier: ~ω kBTN[ω] ≥ 2 [Eq.(5.89)]

Sec. VI: Bosonic Scattering Description of a Two-Port Amplifier

Vˆa(Vˆb) Voltage at the input (output) of the amplifier Relation to bosonic mode operators: Eq. (6.2a)

Iˆa(Iˆb) Current drawn at the input (leaving the output) of the amplifier 6

TABLE I: Table of symbols and main results.

Symbol Definition / Result Relation to bosonic mode operators: Eq. (6.2b) 0 λI Reverse current gain of the amplifier s[ω] Input-output 2 × 2 scattering matrix of the amplifier [Eq. (6.3)] 0 Relation to op-amp parameters λV , λI ,Zin,Zout: Eqs. (6.7) V˜ˆ (I˜ˆ) Voltage (current) noise operators of the amplifier

Fˆa[ω], Fˆb[ω] Input (output) port noise operators in the scattering description [Eq. (6.3)] Relation to op-amp noise operators V,˜ˆ I˜ˆ: Eq. (6.9)

II. QUANTUM NOISE SPECTRA ‘classical’) correlator vanishes in thermal equilibrium, just as it does classically: hxˆpˆ +p ˆxˆi = 0. Note how- A. Introduction to quantum noise ever that in the quantum case, the canonical commu- tation relation between position and momentum implies In classical physics, the study of a noisy time- there must be some correlations between the two, namely dependent quantity invariably involves its spectral den- hxˆ(0)ˆp(0)i−hpˆ(0)ˆx(0)i = i~. These correlations are easily sity S[ω]. The spectral density tells us the intensity of evaluated by writingx ˆ andp ˆ in terms of harmonic oscil- the noise at a given frequency, and is directly related to lator ladder operators. We find that in thermal equilib- the auto-correlation function of the noise.2 In a similar ~ ~ rium: hpˆ(0)ˆx(0)i = −i 2 and hxˆ(0)ˆp(0)i = +i 2 . Not only fashion, the study of quantum noise involves quantum are the position and momentum correlated, but their cor- noise spectral densities. These are defined in a manner relator is imaginary!3 This means that, despite the fact which mimics the classical case: that the position is an hermitian with real Z +∞ eigenvalues, its autocorrelation function is complex and iωt Sxx[ω] = dt e hxˆ(t)ˆx(0)i. (2.1) given from Eq. (2.2) by: −∞ 2  +iΩt −iΩt Gxx(t) = xZPF nB(~Ω)e + [nB(~Ω) + 1]e , Herex ˆ is a quantum operator (in the Heisenberg rep- (2.3) resentation) whose noise we are interested in, and the 2 where xZPF ≡ ~/2MΩ is the RMS zero-point uncer- angular brackets indicate the quantum statistical aver- tainty of x in the quantum ground state, and nB is the age evaluated using the quantum . Note Bose-Einstein occupation factor. The complex nature of that we will use S[ω] throughout this review to denote the autocorrelation function follows from the fact that the spectral density of a classical noise, while S[ω] will the operatorx ˆ does not commute with itself at different denote a quantum noise spectral density. times. As a simple introductory example illustrating impor- Because the correlator is complex it follows that the tant differences from the classical limit, consider the po- spectral density is not symmetric in frequency: sition noise of a simple harmonic oscillator having mass 2 M and frequency Ω. The oscillator is maintained in equi- Sxx[ω] = 2πxZPF (2.4) librium with a large heat bath at temperature T via some × {nB(~Ω)δ(ω + Ω) + [nB(~Ω) + 1]δ(ω − Ω)} infinitesimal coupling which we will ignore in considering the dynamics. The solutions of the Heisenberg equations In contrast, a classical autocorrelation function is always of motion are the same as for the classical case but with real, and hence a classical noise spectral density is always the initial position x and momentum p replaced by the symmetric in frequency. Note that in the high tempera- corresponding quantum operators. It follows that the ture limit kBT  ~Ω we have nB(~Ω) ∼ nB(~Ω) + 1 ∼ kBT position autocorrelation function is Ω . Thus, in this limit the Sxx[ω] becomes symmetric in~ frequency as expected classically, and coincides with Gxx(t) = hxˆ(t)ˆx(0)i (2.2) the classical expression for the position spectral density 1 (cf. Eq. (A12)). = hxˆ(0)ˆx(0)i cos(Ωt) + hpˆ(0)ˆx(0)i sin(Ωt). MΩ The Bose-Einstein factors suggest a way to understand the frequency-asymmetry of Eq. (2.4): the positive fre- Classically the second term on the RHS vanishes be- quency part of the spectral density has to do with stimu- cause in thermal equilibrium x and p are uncorrelated lated emission of energy into the oscillator and the nega- random variables. As we will see shortly below for the tive frequency part of the spectral density has to do with quantum case, the symmetrized (sometimes called the

3 Notice that this occurs because the product of two non- 2 For readers unfamiliar with the basics of classical noise, a com- commuting hermitian operators is not itself an hermitian op- pact review is given in Appendix A. erator. 7

TABLE II Contents of online appendix material. Page numbers refer to the supplementary material. Section Page A. Basics of Classical and Quantum Noise 1 B. Quantum Spectrum Analyzers: Further Details 4 C. Modes, Transmission Lines and Classical Input-Output Theory 8 D. Quantum Modes and Noise of a Transmission Line 15 E. backaction and Input-Output Theory for Driven Damped Cavities 18 F. Information Theory and Measurement Rate 29 G. Number Phase Uncertainty 30 H. Using Feedback to Reach the Quantum Limit 31 I. Additional Technical Details 34

10 Schoelkopf et al., 2003). Describing the TLS as a ficti- 8 tious spin-1/2 particle with spin down (spin up) repre- senting the ground state (excited state), its Hamiltonian 6 is Hˆ = ~ω01 σˆ , where ω is the energy splitting be- S [ ]/2Rk T 0 2 z ~ 01 VV B 4 tween the two states. The TLS is then coupled to an 2 external noise source via an additional term in the Hamil- tonian 0 Vˆ = AFˆσˆx, (2.5) -10 -5 0 5 10 where A is a coupling constant, and the operator Fˆ rep- h /kBT < 0 > 0 resents the external noise source. The coupling Hamil- emission absorption tonian Vˆ can lead to the exchange of energy between by reservoir by reservoir the two-level system and noise source, and hence tran- FIG. 1 Quantum noise spectral density of voltage fluctuations sitions between its two eigenstates. The corresponding across a resistor (resistance R) as a function of frequency at Fermi Golden Rule transition rates can be compactly ex- zero temperature (dashed line) and finite temperature (solid pressed in terms of the quantum noise spectral density of line). Fˆ, SFF [ω]: A2 Γ↑ = SFF [−ω01] (2.6a) ~2 emission of energy by the oscillator. That is, the posi- A2 Γ↓ = SFF [+ω01]. (2.6b) tive frequency part of the spectral density is a measure ~2 of the ability of the oscillator to absorb energy, while the Here, Γ↑ is the rate at which the qubit is excited from negative frequency part is a measure of the ability of the its ground to excited state; Γ↓ is the corresponding rate oscillator to emit energy. As we will see, this is gener- for the opposite, relaxation process. As expected, posi- ally true, even for non-thermal states. Fig. 1 illustrates tive (negative) frequency noise corresponds to absorption this idea for the case of the voltage noise spectral density (emission) of energy by the noise source. Note that if the of a resistor (see Appendix D.3 for more details). Note noise source is in thermal equilibrium at temperature T , that the result Eq. (2.4) can be extended to the case of the transition rates of the TLS must satisfy the detailed a bath of many harmonic oscillators. As described in −β ω01 balance relation Γ↑/Γ↓ = e ~ , where β = 1/kBT . Appendix D a resistor can be modeled as an infinite set This in turn implies that in thermal equilibrium, the of harmonic oscillators and from this model the John- quantum noise spectral density must satisfy: son/Nyquist noise of a resistor can be derived. β ω01 SFF [+ω01] = e ~ SFF [−ω01]. (2.7) The more general situation is where the noise source is B. Quantum spectrum analyzers not in thermal equilibrium; in this case, no general de- tailed balance relation holds. However, if we are con- The qualitative picture described in the previous sub- cerned only with a single particular frequency, then it is section can be confirmed by considering simple systems always possible to define an ‘effective temperature’ Teff which act as effective spectrum analyzers of quantum for the noise using Eq. (2.7), i.e. noise. The simplest such example is a quantum-two ~ω level system (TLS) coupled to a quantum noise source kBTeff [ω] ≡ h i (2.8) log SFF [ω] (Aguado and Kouwenhoven, 2000; Gavish et al., 2000; SFF [−ω] 8

Note that for a non-equilibrium system, Teff will in gen- spectrum is responsible for the damping. This also has a eral be frequency-dependent. In NMR language, Teff will simple heuristic interpretation: damping is caused by the simply be the ‘spin temperature’ of our TLS spectrom- net tendency of the noise source to absorb, rather than eter once it reaches steady state after being coupled to emit, energy from the oscillator. the noise source. The damping induced by the noise source may equiv- Another simple quantum noise spectrometer is a har- alently be attributed to the oscillator’s motion inducing monic oscillator (frequency Ω, mass M, position x) cou- an average value to hF i which is out-of-phase with x, pled to a noise source (see e.g. Dykman (1978); Schwinger i.e. δhA · F (t)i = −Mγx˙(t). Standard quantum linear (1961)). The coupling Hamiltonian is now: response theory yields: Z ˆ ˆ  †  ˆ 2 0 0 0 V = AxˆF = A xZPF(ˆa +a ˆ ) F (2.9) δhA · Fˆ(t)i = A dt χFF (t − t )hxˆ(t )i (2.13) wherea ˆ is the oscillator annihilation operator, Fˆ is the where we have introduced the susceptibility operator describing the fluctuating noise, and A is again ˆ −i D E a coupling constant. We see that (−AF ) plays the role of χFF (t) = θ(t) [Fˆ(t), Fˆ(0)] (2.14) a fluctuating force acting on the oscillator. In complete ~ analogy to the previous subsection, noise in Fˆ at the os- Using the fact that the oscillator’s motion only involves cillator frequency Ω can cause transitions between the the frequency Ω, we thus have: oscillator energy eigenstates. The corresponding Fermi 2 2 2A xZPF Golden Rule transition rates are again simply related to γ = [−ImχFF [Ω]] (2.15) the noise spectrum SFF [ω]. Incorporating these rates ~ into a simple master equation describing the probabil- A straightforward manipulation of Eq. (2.14) for χFF ity to find the oscillator in a particular energy state, one shows that this expression for γ is exactly equivalent to finds that the stationary state of the oscillator is a Bose- our previous expression, Eq. (2.12). Einstein distribution evaluated at the effective tempera- In addition to giving insight on the meaning of the sym- ture Teff [Ω] defined in Eq. (2.8). Further, one can use the metric and asymmetric parts of a quantum noise spec- master equation to derive a very classical-looking equa- tral density, the above example also directly yields the tion for the average energy hEi of the oscillator (see Ap- quantum version of the fluctuation-dissipation theorem pendix B.2): (Callen and Welton, 1951). As we saw earlier, if our noise source is in thermal equilibrium, the positive and nega- d hEi = P − γhEi (2.10) tive frequency parts of the noise spectrum are strictly dt related to one another by the condition of detailed bal- where ance (cf. Eq. (2.7)). This in turn lets us link the classical, symmetric-in-frequency part of the noise to the damping 2 2 A A S¯FF [Ω] (i.e. the asymmetric-in-frequency part of the noise). Let- P = (SFF [Ω] + SFF [−Ω]) ≡ (2.11) 4M 2M ting β = 1/(kBT ) and making use of Eq. (2.7), we have: A2x2 γ = ZPF (S [Ω] − S [−Ω]) (2.12) SFF [Ω] + SFF [−Ω] 2 FF FF S¯ [Ω] ≡ ~ FF 2 The two terms in Eq. (2.10) describe, respectively, heat- 1 = coth(β~Ω/2) (SFF [Ω] − SFF [−Ω]) ing and damping of the oscillator by the noise source. 2 ΩM The heating effect of the noise is completely analogous to = coth(β Ω/2)~ γ[Ω] (2.16) what happens classically: a random force causes the os- ~ A2 cillator’s momentum to diffuse, which in turn causes hEi Thus, in equilibrium, the condition that noise-induced to grow linearly in time at rate proportional to the force transitions obey detailed balance immediately implies noise spectral density. In the quantum case, Eq. (2.11) that noise and damping are related to one another via the indicates that it is the symmetric-in-frequency part of the temperature. For T  Ω, we recover the more familiar ¯ ~ noise spectrum , SFF [Ω], which is responsible for this ef- classical version of the fluctuation dissipation theorem: fect, and which thus plays the role of a classical noise 2 ¯ source. This is another reason why S¯FF [ω] is often re- A SFF [Ω] = 2kBT Mγ (2.17) ferred to as the “classical” part of the noise.4 In contrast, Further insight into the fluctuation dissipation theorem we see that the asymmetric-in-frequency part of the noise is provided in Appendix C.3, where we discuss it in the simple but instructive context of a transmission line ter- minated by an impedance Z[ω].

4 ˆ2 R ∞ ¯ We have thus considered two simple examples of how Note that with our definition, hF i = −∞(dω/2π)SFF [ω]. It is common in engineering contexts to define so-called “one-sided” one can measure quantum noise spectral densities. Fur- classical spectral densities, which are equal to two times our def- ther details, as well as examples of other quantum noise inition. spectrum analyzers, are given in Appendix B. 9

III. QUANTUM MEASUREMENTS +d

Having introduced both quantum noise and quantum z z spectrum analyzers, we are now in a position to intro- duce the general topic of quantum measurements. All FIG. 2 (Color online) Schematic illustration of position dis- practical measurements are affected by noise. Certain tributions of an atom in the detector plane of a Stern-Gerlach quantum measurements remain limited by quantum noise apparatus whose field gradient is in the z direction. For small even though they use completely ideal apparatus. As we values of the displacement d (described in the text), there is significant overlap of the distributions and the spin cannot be will see, the limiting noise here is associated with the fact unambiguously inferred from the position. For large values of that canonically conjugate variables are incompatible ob- d the spin is perfectly entangled with position and can be in- servables in quantum mechanics. ferred from the position. This is the limit of strong projective The simplest, idealized description of a quantum mea- measurement. surement, introduced by von Neumann (Bohm, 1989; Haroche and Raimond, 2006; von Neumann, 1932; Wheeler and Zurek, 1984), postulates that the mea- surement process instantaneously collapses the system’s where h~r|ξ±i = ψ1(~r ± dzˆ) are spatial orbitals peaked in onto one of the eigenstates of the observ- the plane of the detector. The deflection d is determined able to be measured. As a consequence, any initial super- by the device geometry and the magnetic field gradient. position of these eigenstates is destroyed and the values The z-direction position distribution of the particle for of observables conjugate to the measured observable are each spin component is shown in Fig. 2. If d is suffi- perturbed. This perturbation is an intrinsic feature of ciently large compared to the wave packet spread then, quantum mechanics and cannot be avoided in any mea- given the position of the particle, one can unambiguously surement scheme, be it of the “projection-type” described determine the distribution from which it came and hence by von Neumann or rather a weak, continuous measure- the value of the spin projection of the atom. This is the ment to be analyzed further below. limit of a strong ‘projective’ measurement. To form a more concrete picture of quantum measure- In the initial state one has hΨ0|σˆx|Ψ0i = +1, but in ment, we begin by noting that every quantum measure- the final state one has ment apparatus consists of a macroscopic ‘pointer’ cou- 1 hΨ |σˆ |Ψ i = {hξ |ξ i + hξ |ξ i} (3.3) pled to the microscopic system to be measured. (A spe- 1 x 1 2 − + + − cific model is discussed in Allahverdyan et al. (2001).) This pointer is sufficiently macroscopic that its position For sufficiently large d the states ξ± are orthogonal and can be read out ‘classically’. The interaction between the thus the act ofσ ˆz measurement destroys the spin coher- microscopic system and the pointer is arranged so that ence the two become strongly correlated. One of the simplest possible examples of a quantum measurement is that of hΨ1|σˆx|Ψ1i → 0. (3.4) the Stern-Gerlach apparatus which measures the projec- tion of the spin of an S = 1/2 atom along some chosen This is what we mean by projection or ‘col- direction. What is really measured in the experiment is lapse’. The result of measurement of the atom position will yield a random and unpredictable value of ± 1 for the the final position of the atom on the detector plate. How- 2 ever, the magnetic field gradient in the magnet causes z projection of the spin. This destruction of the coher- this position to be perfectly correlated (‘entangled’) with ence in the transverse spin components by a strong mea- the spin projection so that the latter can be inferred from surement of the longitudinal spin component is the first of the former. Suppose for example that the initial state of many examples we will see of the Heisenberg in action. Measurement of one variable destroys the atom is a product of a spatial wave function ξ0(~r) centered on the entrance to the magnet, and a spin state information about its conjugate variable. We will study which is the superposition of up and down spins corre- several examples in which we understand microscopically sponding to the eigenstate ofσ ˆ : how it is that the coupling to the measurement appara- x tus causes the ‘backaction’ quantum noise which destroys 1 our knowledge of the conjugate variable. |Ψ0i = √ {| ↑i + | ↓i} |ξ0i. (3.1) 2 In the special case where the eigenstates of the observ- able we are measuring are also stationary states (i.e. en- After passing through a magnet with field gradient in the ergy eigenstates), measuring the observable a second time z direction, an atom with spin up is deflected upwards would reproduce exactly the same measurement result, and an atom with spin down is deflected downwards. By thus providing a way to confirm the accuracy of the mea- the linearity of quantum mechanics, an atom in a spin surement scheme. These optimal kinds of repeatable mea- superposition state thus ends up in a superposition of surements are called “Quantum Non-Demolition” (QND) the form measurements (Braginsky and Khalili, 1992, 1996; Bra- 1 ginsky et al., 1980; Peres, 1993). A simple example would |Ψ1i = √ {| ↑i|ξ+i + | ↓i|ξ−i} , (3.2) 2 be a sequential pair of Stern-Gerlach devices oriented in 10 the same direction. In the absence of stray magnetic A. Weak continuous measurements perturbations, the second apparatus would always yield the same answer as the first. In practice, one terms a In discussing “real” quantum measurements, another measurement QND if the observable being measured is key notion to introduce is that of weak, continuous mea- an eigenstate of the ideal Hamiltonian of the measured surements (Braginsky and Khalili, 1992). Many mea- system (i.e. one ignores any couplings between this sys- surements in practice take an extended time-interval to tem and sources of dissipation). This is reasonable if complete, which is much longer than the “microscopic” such couplings give rise to dynamics on timescales longer time scales (oscillation periods etc.) of the system. The than what is needed to complete the measurement. This reason may be quite simply that the coupling strength point is elaborated in Sec. VII, where we discuss prac- between the detector and the system cannot be made tical considerations related to the quantum limit. We arbitrarily large, and one has to wait for the effect of also discuss in that section the fact that the repeatabil- the system on the detector to accumulate. For example, ity of QND measurements is of fundamental practical im- in our Stern-Gerlach measurement suppose that we are portance in overcoming detector inefficiencies (Gambetta only able to achieve small magnetic field gradients and et al., 2007). that consequently, the displacement d cannot be made large compared to the wave packet spread (see Fig. 2). In this case the states ξ± would have non-zero overlap A common confusion is to think that a QND measure- and it would not be possible to reliably distinguish them: ment has no effect on the state of the system being mea- we thus would only have a “weak” measurement. How- sured. While this is true if the initial state is an eigen- ever, by cascading together a series of such measurements state of the observable, it is not true in general. Consider and taking advantage of the fact that they are QND, we again our example of a spin oriented in the x direction. can eventually achieve an unambiguous strong projective The result of the firstσ ˆz measurement will be that the measurement: at the end of the cascade, we are certain state randomly and completely unpredictably collapses of whichσ ˆz eigenstate the spin is in. During this process, onto one of the twoσ ˆz eigenstates: the state is indeed al- the overlap of ξ± would gradually fall to zero correspond- tered by the measurement. However all subsequent mea- ing to a smooth continuous loss of phase in the surements using the same orientation for the detectors transverse spin components. At the end of the process, will always agree with the result of the first measure- the QND nature of the measurement ensures that the ment. Thus QND measurements may affect the state of probability of measuring σz =↑ or ↓ will accurately re- the system, but never the value of the observable (once it flect the initial wavefunction. Note that it is only in this is determined). Other examples of QND measurements case of weak continuous measurements that makes sense include: (i) measuring the electromagnetic field energy to define a measurement rate in terms of a rate of gain stored inside a cavity by determining the radiation pres- of information about the variable being measured, and sure exerted on a moving piston (Braginsky and Khalili, a corresponding dephasing rate, the rate at which infor- 1992), (ii) detecting the presence of a photon in a cav- mation about the conjugate variable is being lost. We ity by its effect on the phase of an atom’s superposition will see that these rates are intimately related via the state (Haroche and Raimond, 2006; Nogues et al., 1999), Heisenberg uncertainty principle. and (iii) the “dispersive” measurement of a qubit state While strong projective measurements may seem to by its effect on the frequency of a driven microwave res- be the ideal, there are many cases where one may in- onator (Blais et al., 2004; Lupa¸scu et al., 2007; Wallraff tentionally desire a weak continuous measurement; this et al., 2004), which is the first canonical example we will was already discussed in the introduction. There are describe below. many practical examples of weak, continuous measure- ment schemes. These include: (i) charge measurements, where the current through a device (e.g. quantum point In contrast to the above, in non-QND measurements, contact or single-electron transistor) is modulated by the the back-action of the measurement will affect the ob- presence/absence of a nearby charge, and where it is nec- servable being studied. The canonical example we will essary to wait for a sufficiently long time to overcome the consider below is the position measurement of a harmonic shot noise and distinguish between the two current val- oscillator. Since the position operator does not commute ues, (ii) the weak dispersive qubit measurement discussed with the Hamiltonian, the QND criterion is not fulfilled. below, (iii) displacement detection of a nano-mechanical Other examples of non-QND measurements include: (i) beam (e.g. optically or by capacitive coupling to a charge photon counting via photo-detectors that absorb the pho- sensor), where one looks at the two quadrature ampli- tons, (ii) continuous measurements where the observable tudes of the signal produced at the beam’s resonance does not commute with the Hamiltonian, thus inducing frequency. a time-dependence of the measurement result, (iii) mea- Not surprisingly, quantum noise plays a crucial role in surements that can be repeated only after a time longer determining the properties of a weak, continuous quan- than the energy relaxation time of the system (e.g. for a tum measurement. For such measurements, noise both qubit, T1). determines the back-action effect of the measurement on 11 the measured system, as well as how quickly information phase shift can be converted (via homodyne interferome- is acquired in the measurement process. Previously we try) into a changing intensity; this can then be detected saw that a crucial feature of quantum noise is the asym- using diodes or photomultipliers. metry between positive and negative frequencies; we fur- In this subsection, we will analyze weak, continuous ther saw that this corresponds to the difference between measurements made using such a parametric cavity de- absorption and emission events. For measurements, an- tector; this will serve as a good introduction to the other key aspect of quantum noise will be important: more general approaches presented in later sections. We as we will discuss extensively, quantum mechanics places will show that this detector is capable of reaching the constraints on the noise of any system capable of act- ‘quantum-limit’, meaning that it can be used to make a ing as a detector or amplifier. These constraints in turn weak, continuous measurement as optimally as is allowed place restrictions on any weak, continuous measurement, by quantum mechanics. This is true for both the (QND) and lead directly to quantum limits on how well one can measurement of the state of a qubit, and the (non-QND) make such a measurement. measurement of the position of a harmonic oscillator. In the rest of this section, we give an introduction to Complementary analyses of weak, continuous qubit mea- how one describes a weak, continuous quantum measure- surement are given in Makhlin et al. (2000, 2001) (using ment, considering the specific examples of using para- a single-electron transistor) and in Clerk et al. (2003); metric coupling to a resonant cavity for QND detection Gurvitz (1997); Korotkov (2001b); Korotkov and Averin of the state of a qubit and the (necessarily non-QND) (2001); Pilgram and B¨uttiker (2002) (using a quantum detection of the position of a harmonic oscillator. In the point contact). We will focus here on a high-Q cavity following section (Sec. IV), we give a derivation of a very detector; weak qubit measurement with a low-Q cavity general quantum mechanical constraint on the noise of was studied in (Johansson et al., 2006). any system capable of acting as a detector, and show It is worth noting the widespread usage of cavity de- how this constraint directly leads to the quantum limit tectors in experiment. One important current realization on qubit detection. Finally, in Sec. V, we will turn to the is a microwave cavity used to read out the state of a su- important but slightly more involved case of a quantum perconducting qubit (Blais et al., 2004; Lupa¸scu et al., linear amplifier or position detector. We will show that 2004; Duty et al., 2005; Il’ichev et al., 2003; Izmalkov the basic quantum noise constraint derived Sec. IV again et al., 2004; Lupa¸scu et al., 2005; Schuster et al., 2005; leads to a quantum limit; here, this limit is on how small Sillanp¨a¨a et al., 2005; Wallraff et al., 2004). Another one can make the added noise of a linear amplifier. class of examples are optical cavities used to measure Before leaving this introductory section, it is worth mechanical degree of freedom. Examples of such systems pointing out that the theory of weak continuous measure- include those where one of the cavity mirrors is mounted ments is sometimes described in terms of some set of aux- on a cantilever (Arcizet et al., 2006; Gigan et al., 2006; iliary systems which are sequentially and momentarily Kleckner and Bouwmeester, 2006). Related systems in- weakly coupled to the system being measured. (See Ap- volve a freely suspended mass (Abramovici et al., 1992; pendix E.) One then envisions a sequence of projective Corbitt et al., 2007), an optical cavity with a thin trans- von Neumann measurements on the auxiliary variables. parent membrane in the middle (Thompson et al., 2008) The weak entanglement between the system of interest and, more generally, an elastically deformable whispering and one of the auxiliary variables leads to a kind of par- gallery mode resonator (Schliesser et al., 2006). Systems tial collapse of the system wave function (more precisely where a microwave cavity is coupled to a mechanical el- the density matrix) which is described in mathematical ement are also under active study (Blencowe and Buks, terms not by projection operators, but rather by POVMs 2007; Regal et al., 2008; Teufel et al., 2008). (positive operator valued measures). We will not use this We start our discussion with a general observation. and the related ‘quantum trajectory’ language here, but The cavity uses interference and the wave nature of light direct the reader to the literature for more information to convert the input signal to a phase shifted wave. For on this important approach. (Brun, 2002; Haroche and small phase shifts we have a weak continuous measure- Raimond, 2006; Jordan and Korotkov, 2006; Peres, 1993) ment. Interestingly, it is the complementary particle na- ture of light which turns out to limit the measurement. As we will see, it both limits the rate at which we can B. Measurement with a parametrically coupled resonant make a measurement (via photon shot noise in the out- cavity put beam) and also controls the backaction disturbance of the system being measured (due to photon shot noise A simple yet experimentally practical example of a inside the cavity acting on the system being measured). quantum detector consists of a resonant optical or RF These two dual aspects are an important part of any cavity parametrically coupled to the system being mea- weak, continuous quantum measurement; hence, under- sured. Changes in the variable being measured (e.g. the standing both the output noise (i.e. the measurement state of a qubit or the position of an oscillator) shift the imprecision) and back-action noise of detectors will be cavity frequency and produce a varying phase shift in the crucial. carrier signal reflected from the cavity. This changing All of our discussion of noise in the cavity system will 12 be framed in terms of the number-phase uncertainty re- again be represented in terms of a noise spectral density I 2 lation for coherent states. A coherent photon state con- Sxx = Sθθ/4k . Here the superscript I refers to the fact tains a Poisson distribution of the number of photons, that this is noise representing imprecision in the mea- implying that the fluctuations in photon number obey surement, not actual fluctuations in the position. We (∆N)2 = N¯, where N¯ is the mean number of photons. also need to worry about backaction: each photon hit- Further, coherent states are over-complete and states of ting the mirror transfers a momentum 2~k to the mirror, different phase are not orthogonal to each other; this di- so photon shot noise corresponds to a random backac- 2 2 rectly implies (see Appendix G) that there is an uncer- tion force noise spectral density SFF = 4~ k SN˙ N˙ Mul- tainty in any measurement of the phase. For large N¯, tiplying these together we have the central result for the this is given by: product of the backaction force noise and the imprecision

1 2 (∆θ)2 = . (3.5) I 2 ~ ¯ SFF S = S ˙ ˙ Sθθ = (3.9) 4N xx ~ NN 4 ¯ Thus, large-N coherent states obey the number-phase or in analogy with Eq. (3.6) uncertainty relation q 1 S SI = ~. (3.10) ∆N∆θ = (3.6) FF xx 2 2 Not surprisingly, the situation considered here is as ideal analogous to the position-momentum uncertainty rela- as possible. Thus, the RHS above is actually a lower tion. bound on the product of imprecision and back-action Eq. (3.6) can also be usefully formulated in terms of noise for any detector capable of significant amplification; noise spectral densities associated with the measurement. we will prove this rigorously in Sec. IV.A. Eq. (3.10) thus Consider a continuous photon beam carrying an average represents the quantum-limit on the noise of our detector. ˙ photon flux N. The variance in the number of photons As we will see shortly, having a detector with quantum- detected grows linearly in time and can be represented as limited noise is a prerequisite for reaching the quantum 2 (∆N) = SN˙ N˙ t, where SN˙ N˙ is the white-noise spectral limit on various different measurement tasks (e.g. contin- density of photon-flux fluctuations. On a physical level, uous position detection of an oscillator and QND qubit ˙ it describes photon shot noise, and is given by SN˙ N˙ = N. state detection). Note that in general, a given detector Consider now the phase of the beam. Any homodyne will have more noise than the quantum-limited value; we measurement of this phase will be subject to the same will devote considerable effort in later sections to deter- photon shot noise fluctuations discussed above (see Ap- mining the conditions needed to achieve the lower bound pendix G for more details). Thus, if the phase of the of Eq. (3.10). beam has some nominal small value θ0, the output sig- We now turn to the story of measurement using a high nal from the homodyne detector integrated up to time Q cavity; it will be similar to the above discussion, except R t t will be of the form I = θ0t + 0 dτ δθ(τ), where δθ is that we have to account for the filtering of the noise by a noise representing the imprecision in our measurement the cavity response. We relegate relevant calculational of θ0 due to the photon shot noise in the output of the details related to Appendix E. The cavity is simply de- homodyne detector. An unbiased estimate of the phase scribed as a single bosonic mode coupled weakly to elec- obtained from I is θ = I/t, which obeys hθi = θ0. Fur- tromagnetic modes outside the cavity. The Hamiltonian 2 ther, one has (∆θ) = Sθθ/t, where Sθθ is the spectral of the system is given by: density of the δθ white noise. Comparison with Eq. (3.5) ˆ † ˆ yields H = H0 + ~ωc (1 + Azˆ)a ˆ aˆ + Henvt. (3.11)

1 Here, H0 is the unperturbed Hamiltonian of the system Sθθ = . (3.7) 4N˙ whose variablez ˆ (which is not necessarily a position) is being measured,a ˆ is the annihilation operator for the The results above lead us to the fundamental cavity mode, and ωc is the cavity resonance frequency wave/particle relation for ideal coherent beams in the absence of the coupling A. We will take both A andz ˆ to be dimensionless. The term Hˆ describes the p 1 envt SN˙ N˙ Sθθ = (3.8) electromagnetic modes outside the cavity, and their cou- 2 pling to the cavity; it is responsible for both driving and Before we study the role that these uncertainty rela- damping the cavity mode. The damping is parameterized tions play in measurements with high Q cavities, let us by rate κ, which tells us how quickly energy leaks out of consider the simplest case of reflecting light from a mir- the cavity; we consider the case of a high quality-factor ror without a cavity. The phase shift of the beam (hav- cavity, where Qc ≡ ωc/κ  1. ing wave vector k) when the mirror moves a distance Turning to the interaction term in Eq. (3.11), we see x is 2kx. Thus, the uncertainty in the phase measure- that the parametric coupling strength A determines the ment corresponds to a position imprecision which can change in frequency of the cavity as the system variable 13 zˆ changes. We will assume for simplicity that the dy- I(t) namics ofz ˆ is slow compared to κ. In this limit the ΔI reflected phase shift simply varies slowly in time adia- batically following the instantaneous value ofz ˆ. We will also assume that the coupling A is small enough that the phase shifts are always very small and hence the measure- t ment is weak. Many photons will have to pass through the cavity before much information is gained about the value of the phase shift and hence the value ofz ˆ. We first consider the case of a ‘one-sided’ cavity where only one of the mirrors is semi-transparent, the other FIG. 3 (Color online) Distribution of the integrated output being perfectly reflecting. In this case, a wave incident for the cavity detector, I(t), for the two different qubit states. The separation of the means of the distributions grows linearly on the cavity (say, in a one-dimensional waveguide) will √in time, while the width of the distributions only grow as the be perfectly reflected, but with a phase shift θ determined t. by the cavity and the value ofz ˆ. The reflection coefficient at the bare cavity frequency ωc is simply given by (Walls and Milburn, 1994) The mean photon number in the cavity is found to be 1 + 2iAQczˆ r = − . (3.12) n¯ = Nt˙ WD, where again N˙ the mean photon flux incident 1 − 2iAQczˆ on the cavity. From this it follows that

Note that r has unit magnitude because all photons S = (A ω t )2S . (3.19) which are incident are reflected if the cavity is lossless. Fz Fz ~ c WD N˙ N˙ For weak coupling we can write the reflection phase shift Combining this with Eq. (3.15) again yields the same iθ as r = −e , where result as Eq. (3.10) obtained without the cavity. The parametric cavity detector (used in this way) is thus a θ ≈ 4QcAzˆ = (Aωczˆ)tWD (3.13) quantum-limited detector, meaning that the product of its noise spectral densities achieves the ideal minimum We see that the scattering phase shift is simply the fre- value. quency shift caused by the parametric coupling multi- We will now examine how the quantum limit on the plied by the Wigner delay time (Wigner, 1955) noise of our detector directly leads to quantum limits on different measurement tasks. In particular, we will ∂ ln r tWD = Im = 4/κ. (3.14) consider the cases of continuous position detection and ∂ω QND qubit state measurement. Thus the measurement imprecision noise power for a given photon flux N˙ incident on the cavity is given by 1. QND measurement of the state of a qubit using a resonant cavity I 1 Szz = 2 Sθθ. (3.15) (AωctWD) Here we specialize to the case where the system oper- atorz ˆ =σ ˆ represents the state of a spin-1/2 quantum The random part of the generalized backaction force con- z bit. Eq. (3.11) becomes jugate toz ˆ is from Eq. (3.11)

ˆ 1 † ˆ ∂Hˆ H = ~ω01σˆz + ~ωc (1 + Aσˆz)a ˆ aˆ + Henvt (3.20) Fˆ ≡ − = −A ω δnˆ (3.16) 2 z ∂zˆ ~ c We see thatσ ˆz commutes with all terms in the Hamilto- where, sincez ˆ is dimensionless, Fˆz has units of energy. nian and is thus a constant of the motion (assuming that † † Here δnˆ =n ˆ − n¯ =a ˆ aˆ − haˆ aˆi represents the photon Hˆenvt contains no qubit decay terms so that T1 = ∞) and number fluctuations around the meann ¯ inside the cavity. hence the measurement will be QND. From Eq. (3.13) we The backaction force noise spectral density is thus see that the two states of the qubit produce phase shifts ±θ0 where 2 SFz Fz = (A~ωc) Snn (3.17) θ0 = AωctWD. (3.21) As shown in Appendix E, the cavity filters the photon shot noise so that at low frequencies ω  κ the number As θ0  1, it will take many reflected photons before fluctuation spectral density is simply we are able to determine the state of the qubit. This is a direct consequence of the unavoidable photon shot Snn =nt ¯ WD. (3.18) noise in the output of the detector, and is a basic feature 14 of weak measurements– information on the input is only is a sum of a large number of statistically independent acquired gradually in time. terms and thus we can take the accumulated phase to be Let I(t) be the homodyne signal for the wave reflected gaussian distributed. Using the cumulant expansion we from the cavity integrated up to time t. Depending on the then obtain state of the qubit the mean value of I will be hIi = ±θ0t, *Z t 2+! and the RMS gaussian fluctuations about the mean will −iϕ 1 √ e = exp − dτ ∆ω01(τ) be ∆I = Sθθt. As illustrated in Fig. 3 and discussed 2 0 extensively in Makhlin et al. (2001), the integrated signal   is drawn from one of two gaussian distributions which 2 = exp − SFz Fz t . (3.26) are better and better resolved with increasing time (as ~2 long as the measurement is QND). The state of the qubit thus becomes ever more reliably determined. The signal Note also that the noise correlator above is naturally energy to noise energy ratio becomes symmetrized– the quantum asymmetry of the noise plays no role for this type of coupling. Eq. (3.26) yields the de- 2 2 hIi θ0 phasing rate SNR = 2 = t (3.22) (∆I) Sθθ 2 2 Γϕ = SFz Fz = 2θ0SN˙ N˙ . (3.27) which can be used to define the measurement rate via ~2 2 SNR θ0 1 Using Eqs. (3.23) and (3.27), we find the interest- Γmeas ≡ = = I . (3.23) 2 2Sθθ 2Szz ing conclusion that the dephasing rate and measurement rates coincide: There is a certain arbitrariness in the scale factor of 2 appearing in the definition of the measurement rate; this Γϕ 4 I = 2 SzzSFz Fz = 4SN˙ N˙ Sθθ = 1. (3.28) particular choice is motivated by precise information the- Γmeas ~ oretic grounds (as defined, Γmeas is the rate at which the ‘accessible information’ grows, c.f Appendix F). As we will see and prove rigorously, this represents the While Eq. (3.20) makes it clear that the state of the ideal, quantum-limited case for QND qubit detection: the qubit modulates the cavity frequency, we can easily re- best one can do is measure as quickly as one dephases. In write this equation to show that this same interaction keeping with our earlier discussions, it represents the en- term is also responsible for the back-action of the mea- forcement of the Heisenberg uncertainty principle. The surement (i.e. the disturbance of the qubit state by the faster you gain information about one variable, the faster measurement process): you lose information about the conjugate variable. Note that in general, the ratio Γϕ/Γmeas will be larger than one, as an arbitrary detector will not reach the quantum Hˆ = ~ ω + 2Aω aˆ†aˆ σˆ + ω aˆ†aˆ + Hˆ (3.24) 2 01 c z ~ c envt limit on its noise spectral densities. Such a non-ideal detector produces excess back-action beyond what is re- We now see that the interaction can also be viewed as quired quantum mechanically. providing a ‘light shift’ (i.e. ac Stark shift) of the qubit In addition to the quantum noise point of view pre- splitting frequency (Blais et al., 2004; Schuster et al., sented above, there is a second complementary way in 2005) which contains a constant part 2AnAω¯ c plus a ran- ˆ which to understand the origin of measurement induced domly fluctuating part ∆ω01 = 2Fz/~ which depends on † dephasing (Stern et al., 1990) which is analogous to our nˆ =a ˆ aˆ, the number of photons in the cavity. During a description of loss of transverse spin coherence in the measurement,n ˆ will fluctuate around its mean and act Stern-Gerlach experiment in Eq. (3.3). The measurement as a fluctuating back-action ‘force’ on the qubit. In the † takes the incident wave, described by a coherent state present QND case, noise inn ˆ =a ˆ aˆ cannot cause transi- |αi, to a reflected wave described by a (phase shifted) tions between the two qubit eigenstates. This is the op- coherent state |r↑ · αi or |r↓ · αi, where r↑/↓ is the qubit- posite of the situation considered in Sec. II.B, where we dependent reflection amplitude given in Eq. (3.12). Con- wanted to use the qubit as a spectrometer. Despite the sidering now the full state of the qubit plus detector, lack of any noise-induced transitions, there still is a back- measurement results in a state change: action here, as noise inn ˆ causes the effective splitting frequency of the qubit to fluctuate in time. For weak cou- 1   1  +iω01t/2 √ | ↑i + | ↓i ⊗ |αi → √ e | ↑i ⊗ |r↑ · αi pling, the resulting phase diffusion leads to measurement- 2 2 induced dephasing of superpositions in the qubit (Blais  −iω01t/2 et al., 2004; Schuster et al., 2005) according to + e | ↓i ⊗ |r↓ · αi (3.29)

D R t E −iϕ −i dτ ∆ω01(τ) e = e 0 . (3.25) As |r↑ · αi= 6 |r↓ · αi, the qubit has become entangled with the detector: the state above cannot be written as a For weak coupling the dephasing rate is slow and thus we product of a qubit state times a detector state. To assess are interested in long times t. In this limit the integral the coherence of the final qubit state (i.e. the relative 15 phase between ↑ and ↓), one looks at the off-diagonal not true. The two states |r↑αi and |r↓αi could in princi- matrix element of the qubit’s reduced density matrix: ple be orthogonal without them corresponding to differ- ent values of the detector output (i.e. θ). For example, ρ↓↑ ≡ Tr detectorh↓ |ψihψ| ↑i (3.30) the qubit may have become entangled with extraneous e+iω01t microscopic degrees of freedom in the detector. Thus, on = hr · α|r · αi (3.31) 2 ↑ ↓ a heuristic level, the origin of Eq. (3.35) is clear. +iω01t Returning to our one-sided cavity system, we see from e  2 ∗  = exp −|α| 1 − r↑r↓ (3.32) Eq. (3.28) that the one-sided cavity detector reaches 2 the quantum limit. It is natural to now ask why this In Eq. (3.31) we have used the usual expression for the is the case: is there a general principle in action here overlap of two coherent states. We see that the measure- which allows the one-sided cavity to reach the quantum limit? The answer is yes: reaching the quantum limit ment reduces the magnitude of ρ↑↓: this is dephasing. The amount of dephasing is directly related to the over- requires that there is no ‘wasted’ information in the de- lap between the different detector states that result when tector (Clerk et al., 2003). There should not exist any the qubit is up or down; this overlap can be straightfor- unmeasured quantity in the detector which could have been probed to learn more about the state of the qubit. wardly found using Eq. (3.32) and |α|2 = N¯ = Nt˙ , where ¯ In the single-sided cavity detector, information on the N is the mean number of photons that have reflected from state of the qubit is only available in (that is, is entirely the cavity after time t. We have encoded in) the phase shift of the reflected beam; thus, there is no ‘wasted’ information, and the detector does exp −|α|2 1 − r∗r  = exp −2Nθ¯ 2 ≡ exp [−Γ t] ↑ ↓ 0 ϕ indeed reach the quantum limit. (3.33) To make this idea of ‘no wasted information’ more con- with the dephasing rate Γ being given by: ϕ crete, we now consider a simple detector which fails to reach the quantum limit precisely due to ‘wasted’ infor- 2 ˙ Γϕ = 2θ0N (3.34) mation. Consider again a 1D cavity system where now both mirrors are slightly transparent. Now, a wave in- in complete agreement with the previous result in cident at frequency ωR on one end of the cavity will be Eq.(3.27). partially reflected and partially transmitted. If the ini- tial incident wave is described by a coherent state |αi, the scattered state can be described by a tensor product 2. Quantum limit relation for QND qubit state detection of the reflected wave’s state and the transmitted wave’s state: We now return to the ideal quantum limit relation of Eq. (3.28). As previously stated, this is a lower bound: |αi → |rσ · αi|tσ · αi (3.36) quantum mechanics enforces the constraint that in a where the qubit-dependent reflection and transmission QND qubit measurement the best you can possibly do is amplitudes rσ and tσ are given by (Walls and Milburn, measure as quickly as you dephase (Averin, 2003, 2000b; 1994): Clerk et al., 2003; Devoret and Schoelkopf, 2000; Ko- rotkov and Averin, 2001; Makhlin et al., 2001): 1 t↓ = (3.37) 1 + 2iAQc Γmeas ≤ Γϕ (3.35) 2iQ A r = c (3.38) ↓ 1 + 2iAQ While a detector with quantum limited noise has an c ∗ ∗ equality above, most detectors will be very far from this with t↑ = (t↓) and r↑ = (r↓) . Note that the in- 2 ideal limit, and will dephase the qubit faster than they cident beam is almost perfectly transmitted: |tσ| = 2 acquire information about its state. We provide a proof of 1 − O(AQc) . Eq. (3.35) in Sec. IV.B; for now, we note that its heuris- Similar to the one-sided case, the two-sided cavity tic origin rests on the fact that both measurement and could be used to make a measurement by monitoring the dephasing rely on the qubit becoming entangled with the phase of the transmitted wave. Using the expression for detector. Consider again Eq. (3.29), describing the evo- tσ above, we find that the qubit-dependent transmission lution of the qubit-detector system when the qubit is ini- phase shift is given by: tially in a superposition of ↑ and ↓. To say that we have ˜ ˜ θ↑/↓ = ±θ0 = ±2AQc (3.39) truly measured the qubit, the two detector states |r↑αi and |r↓αi need to correspond to different values of the where again the two signs correspond to the two different detector output (i.e. phase shift θ in our example); this qubit eigenstates. The phase shift for transmission is necessarily implies they are orthogonal. This in turn im- only half as large as in reflection so the Wigner delay plies that the qubit is completely dephased: ρ↑↓ = 0, time associated with transmission is just as we saw in Eq. (3.4) in the Stern-Gerlach example. 2 t˜ = . (3.40) Thus, measurement implies dephasing. The opposite is WD κ 16

Upon making the substitution of t˜WD for tWD, the one- Returning to our cavity detector, we note in closing sided cavity Eqs. (3.15) and (3.17) remain valid. However that it is often technically easier to work with the trans- the internal cavity photon number shot noise remains mission of light through a two-sided cavity, rather than fixed so that Eq. (3.18) becomes reflection from a one-sided cavity. One can still reach the quantum limit in the two-sided cavity case if on uses an ˜ Snn = 2¯ntWD. (3.41) asymmetric cavity in which the input mirror has much less transmission than the output mirror. Most photons which means that are reflected at the input, but those that enter the cav- 2 2 ity will almost certainly be transmitted. The price to be Snn = 2N˙ t˜ = 2S ˙ ˙ t˜ (3.42) WD NN WD paid is that the input carrier power must be increased. and

S = 2 2A2ω2t˜2 S . (3.43) Fz Fz ~ c WD N˙ N˙ 3. Measurement of oscillator position using a resonant cavity As a result the backaction dephasing doubles relative to the measurement rate and we have The qubit measurement discussed in the previous sub- section was an example of a QND measurement: the Γmeas 1 back-action did not affect the observable being measured. = 2SN˙ N˙ Sθθ = . (3.44) Γϕ 2 We now consider the simplest example of a non-QND measurement, namely the weak continuous measurement Thus the two-sided cavity fails to reach the quantum of the position of a harmonic oscillator. The detector limit by a factor of 2. will again be a parametrically-coupled resonant cavity, Using the entanglement picture, we may again alterna- where the position of the oscillator x changes the fre- tively calculate the amount of dephasing from the overlap quency of the cavity as per Eq. (3.11) (see, e.g., Tittonen between the detector states corresponding to the qubit et al. (1999)). Similar to the qubit case, for a sufficiently states ↑ and ↓ (cf. Eq. (3.31)). We find: weak coupling the phase shift of the reflected beam from the cavity will depend linearly on the position x of the −Γϕt e = ht↑α|t↓αihr↑α|r↓αi (3.45) oscillator (cf. Eq. (3.13)); by reading out this phase, we = exp −|α|2 (1 − (t )∗t − (r )∗r ) (3.46) may thus measure x. The origin of backaction noise is ↑ ↓ ↑ ↓ the same as before, namely photon shot noise in the cav- Note that both the change in the transmission and reflec- ity. Now however this represents a random force which tion amplitudes contribute to the dephasing of the qubit. changes the momentum of the oscillator. During the Using the expressions above, we find: subsequent time evolution these random force perturba- tions will reappear as random fluctuations in the position. ˜ 2 2 ˜ 2 ˜ 2 Γϕt = 4(θ0) |α| = 4(θ0) N¯ = 4(θ0) Nt˙ = 2Γmeast. Thus the measurement is not QND. This will mean that (3.47) the minimum uncertainty of even an ideal measurement Thus, in agreement with the quantum noise result, the is larger (by exactly a factor of 2) than the ‘true’ quantum two-sided cavity misses the quantum limit by a factor of uncertainty of the position (i.e. the ground state uncer- two. tainty). This is known as the standard quantum limit on Why does the two-sided cavity fail to reach the quan- weak continuous position detection. It is also an exam- tum limit? The answer is clear from Eq. (3.46): even ple of a general principle that a linear ‘phase-preserving’ though we are not monitoring it, there is information on amplifier necessarily adds noise, and that the minimum the state of the qubit available in the phase of the re- added noise exactly doubles the output noise for the case flected wave. Note from Eq. (3.38) that the magnitude where the input is vacuum (i.e. zero-point) noise. A more of the reflected wave is weak (∝ A2), but (unlike the general discussion of the quantum limit on amplifiers and transmitted wave) the difference in the reflection phase position detectors will be presented in Sec. V. associated with the two qubit states is large (±π/2). The We start by emphasizing that we are speaking here of a ‘missing information’ in the reflected beam makes a di- weak continuous measurement of the oscillator position. rect contribution to the dephasing rate (i.e. the second The measurement is sufficiently weak that the position term in Eq. (3.46)), making it larger than the measure- undergoes many cycles of oscillation before significant in- ment rate associated with measurement of the transmis- formation is acquired. Thus we are not talking about the sion phase shift. In fact, there is an equal amount of instantaneous position but rather the overall amplitude information in the reflected beam as in the transmitted and phase, or more precisely the two quadrature ampli- beam, so the dephasing rate is doubled. We thus have tudes describing the smooth envelope of the motion, a concrete example of the general principle connecting a failure to reach the quantum limit to the presence of xˆ(t) = Xˆ(t) cos(Ωt) + Yˆ (t) sin(Ωt). (3.48) ‘wasted information’. Note that the application of this principle to generalized quantum point contact detectors One can easily show that for an oscillator, the two is found in Clerk et al. (2003). quadrature amplitudes Xˆ and Yˆ are canonically conju- 17 gate and hence do not commute with each other i 12 [X,ˆ Yˆ ] = ~ = 2ix2 . (3.49) MΩ ZPF 10 As the measurement is both weak and continuous, it will yield information on both Xˆ and Yˆ . As such, one is effec- (arb. units) tively trying to simultaneously measure two incompatible 8 area prop. to T observables. This basic fact is intimately related to the ( ω ) property mentioned above, that even a completely ideal 6 measurement S weak continuous position measurement will have a total imprecision uncertainty which is twice the zero-point uncertainty. 4 We are now ready to start our heuristic analysis of po- sition detection using a cavity detector; relevant calcula- 2 tional details presented in Appendix E.3. Consider first the mechanical oscillator we wish to measure. We take it 0

to be a simple harmonic oscillator of natural frequency Ω Output Noise and mechanical damping rate γ0. For weak damping, and 0.0 0.5 1.0 1.5 2.0 at zero coupling to the detector, the spectral density of Frequency ω / Ω the oscillator’s position fluctuations is given by Eq. (2.4) with the delta function replaced by a Lorentzian5 FIG. 4 (Color online) Spectral density of the symmetrized ¯  output noise Sθθ[ω] of a linear position detector. The oscilla- 2 γ0 tor’s noise appears as a Lorentzian sitting above a noise floor Sxx[ω] = xZPF nB(~Ω) 2 2 (ω + Ω) + (γ0/2) (i.e. the measurement imprecision). As discussed in the text,  the width of the peak is proportional to the oscillator damp- γ0 ing rate γ0, while the area under the peak is proportional to + [nB(~Ω) + 1] 2 2 . (3.50) (ω − Ω) + (γ0/2) temperature. This latter fact can be used to calibrate the response of the detector. When we now weakly couple the oscillator to the cavity (as per Eq. (3.11), withz ˆ =x/x ˆ ZPF) and drive the cavity on resonance, the phase shift θ of the reflected beam will be proportional to x (i.e. δθ(t) = [dθ/dx]·x(t)). As such, FWHM of the curve will give the damping parameter γ0, the oscillator’s position fluctuations will cause additional and the area under the experimental curve fluctuations of the phase θ, over and above the intrinsic shot-noise induced phase fluctuations S . We consider Z ∞ dω k T θθ S¯ [ω] = B (3.52) the usual case where the noise spectrometer being used xx 2 −∞ 2π MΩ to measure the noise in θ (i.e. the noise in the homo- dyne current) measures the symmetric-in-frequency noise measures the temperature. What the experimentalist spectral density; as such, it is the symmetric-in-frequency actually plots in making such a curve is the output of position noise that we will detect. In the classical limit the entire detector-plus-following-amplifier chain. Impor- kBT  ~Ω, this is just given by: tantly, if the temperature is known, then the area of the measured curve can be used to calibrate the coupling 1 S¯xx[ω] ≡ (Sxx[ω] + Sxx[−ω]) of the detector and the gain of the total overall amplifier 2 chain (see, e.g., Flowers-Jacobs et al., 2007; LaHaye et al., kBT γ0 2004). One can thus make a calibrated plot where the ≈ 2 2 2 (3.51) 2MΩ (|ω| − Ω) + (γ0/2) measured output noise is referred back to the oscillator position. If we ignore back-action effects, we expect to see this Consider now the case where the oscillator is at zero Lorentzian profile riding on top of the background im- temperature. Eq. (3.50) then yields for the symmetrized precision noise floor; this is illustrated in Fig. 4. noise spectral density Note that additional stages of amplification would also add noise, and would thus further augment this back- ¯0 2 γ0/2 ground noise floor. If we subtract off this noise floor, the Sxx[ω] = xZPF 2 2 . (3.53) (|ω| − Ω) + (γ0/2) One might expect that one could see this Lorentzian directly in the output noise of the detector (i.e. the θ 5 This form is valid only for weak damping because we are as- noise), sitting above the measurement-imprecision noise suming that the oscillator frequency is still sharply defined. We floor. However, this neglects the effects of measurement have evaluated the Bose-Einstein factor exactly at frequency Ω and we have assumed that the Lorentzian centered at positive backaction. From the classical equation of motion we ex- (negative) frequency has negligible weight at negative (positive) pect the response of the oscillator to the backaction force frequencies. F = Fz/xZPF (cf. Eq. (3.16)) at frequency ω to produce 18 an additional displacement δx[ω] = χxx[ω]F [ω], where χxx[ω] is the mechanical susceptibility 10 ]

1 1 0 Total added position noise χ [ω] ≡ . (3.54) ! Measurement imprecision xx 2 2 8 M Ω − ω − iγ0ω / Contribution from backaction 2 These extra oscillator fluctuations will show up as ad- ) 6 ditional fluctuations in the output of the detector. For ZPT x simplicity, let us focus on this noise at the oscillator’s res- ( onance frequency Ω. As a result of the detector’s back-

[ 4 action, the total measured position noise (i.e. inferred spectral density) at the frequency Ω is given by: add

xx, 2

2 S |χ [Ω]| S¯ [Ω] = S¯0 [Ω] + xx [S [+Ω] + S [−Ω]] xx,tot xx 2 FF FF 1 + SI [+Ω] + SI [−Ω] (3.55) 0.1 1 10 2 xx xx ¯0 ¯ Coupling SFF / SFF,opt = Sxx[Ω] + Sxx,add[Ω] (3.56)

The first term here is just the intrinsic zero-point noise FIG. 5 (Color online) Noise power of the added position noise of the oscillator: of a linear position detector, evaluated at the oscillator’s reso- 2 nance frequency (S¯xx,add[Ω]), as a function of the magnitude ¯0 2xZPF Sxx[Ω] = = ~|χxx[Ω]|. (3.57) of the back-action noise spectral density SFF . SFF is pro- γ0 portional to the oscillator-detector coupling, and in the case The second term S¯ is the total noise added by the of the cavity detector, is also proportional to the power in- xx,add cident on the cavity. The optimal value of S is given by measurement, and includes both the measurement impre- FF I I 2 SFF,opt = ~MΩγ0/2 (cf. Eq. (3.59)). We have assumed that cision Sxx ≡ SzzxZPF and the extra fluctuations caused ¯ there are no correlations between measurement imprecision by the backaction. We stress that Sxx,tot corresponds to noise and back-action noise, as is appropriate for the cavity a position noise spectral density inferred from the output detector. of the detector: one simply scales the spectral density of 2 total output fluctuations S¯θθ,tot[Ω] by (dθ/dx) . Implicit in Eq. (3.57) is the assumption that the back action noise and the imprecision noise are uncorrelated added position uncertainty noise is plotted in Fig. 5 as and thus add in quadrature. It is not obvious that this is a function of SFF . We see that for high drive intensity, correct, since in the cavity detector the backaction noise the backaction noise dominates the position uncertainty, and output shot noise are both caused by the vacuum while for low drive intensity, the output shot noise (the noise in the beam incident on the cavity. It turns out last term in the equation above) dominates. ¯ there are indeed correlations, however the symmetrized The added noise (and hence the total noise Sxx,tot[Ω]) (i.e. ‘classical’) correlator S¯θF does vanish for our choice is minimized when the drive intensity is tuned so that of a resonant cavity drive. Further, Eq. (3.55) assumes SFF is equal to SFF,opt, with: that the measurement does not change the damping rate of the oscillator. Again, while this will not be true for ~ ~ SFF,opt = = MΩγ0. (3.59) an arbitrary detector, it is the case here for the cav- 2|χxx[Ω]| 2 ity detector when (as we have assumed) it is driven on resonance. Details justifying both these statements are The more heavily damped is the oscillator, the less sus- given in Appendix E; the more general case with non-zero ceptible it is to backaction noise and hence the higher is noise correlations and back-action damping is discussed the optimal coupling. At the optimal coupling strength, in Sec. V.E. the measurement imprecision noise and back-action noise Assuming we have a quantum-limited detector that each make equal contributions to the added noise, yield- I 2 ing: obeys Eq. (3.9) (i.e. SxxSFF = ~ /4) and that the shot noise is symmetric in frequency, the added position ¯ ~ ¯0 noise spectral density at resonance (i.e. second term in Sxx,add[Ω] = = Sxx[Ω] (3.60) Eq. (3.56)) becomes: MΩγ0  2 1  Thus, the spectral density of the added position noise is S¯ [Ω] = |χ[Ω]|2S + ~ . (3.58) xx,add FF 4 S exactly equal to the noise power associated with the os- FF cillator’s zero-point fluctuations. This represents a mini- Recall from Eq. (3.19) that the backaction noise is pro- mum value for the added noise of any linear position de- portional to the coupling of the oscillator to the detec- tector, and is referred to as the standard quantum limit tor and to the intensity of the drive on the cavity. The on position detection. Note that this limit only involves 19 the added noise of the detector, and thus has nothing to one tries to continuously monitor coherent qubit oscilla- do with the initial temperature of the oscillator. tions with a linear detector which is transversely coupled We emphasize that to reach the above quantum limit to the qubit (Korotkov and Averin, 2001); this is also on weak continuous position detection, one needs the a non-QND situation. Finally, if one only wants to de- detector itself to be quantum limited, i.e. the product tect the noise peak (as opposed to making a continuous I SFF Sxx must be as small as is allowed by quantum me- quantum-limited measurement), one could use two in- chanics, namely ~2/4. Having a quantum-limited detec- dependent detectors coupled to the oscillator and look tor however is not enough: in addition, one must be able at the cross-corelation between the two output noises: in to achieve sufficiently strong coupling to reach the opti- this case, there need not be any noise floor (Doiron et al., mum given in Eq. (3.59). Further, the measured output 2007; Jordan and B¨uttiker, 2005a). noise must be dominated by the output noise of the cav- In Table III, we give a summary of recent experiments ity, not by the added noise of following amplifier stages. which approach the quantum limit on weak, continu- A related, stronger quantum limit refers to the total ous position detection of a mechanical resonator. Note ¯ inferred position noise from the measurement, Sxx,tot[ω]. that in many of these experiments, the effects of detec- It follows from Eqs. (3.60),(3.56) that at resonance, the tor back-action were not seen. This could either be the smallest this can be is twice the oscillator’s zero point result of too low of a detector-oscillator coupling, or due noise: to the presence of excessive thermal noise. As we have shown, the back-action force noise serves to slightly heat ¯ ¯0 Sxx,tot[Ω] = 2Sxx[Ω]. (3.61) the oscillator. If it is already at an elevated temperature due to thermal noise, this additional heating can be very Half the noise here is from the oscillator itself, half is from hard to resolve. the added noise of the detector. Reaching this quantum In closing, we stress that this subsection has given only limit is even more challenging: one needs both to reach a very rudimentary introduction to the quantum limit on the quantum limit on the added noise and cool the oscil- position detection. A complete discussion which treats lator to its ground state. the important topics of back-action damping, effective Finally, we emphasize that the optimal value of the temperature, noise cross-correlation and power gain is coupling derived above was specific to the choice of min- given in Sec. V.E. imizing the total position noise power at the resonance frequency. If a different frequency had been chosen, the optimal coupling would have been different; one again finds that the minimum possible added noise corresponds IV. GENERAL LINEAR RESPONSE THEORY to the ground state noise at that frequency. It is inter- esting to ask what the total position noise would be as A. Quantum constraints on noise a function of frequency, assuming that the coupling has been optimized to minimize the noise at the resonance In this section, we will further develop the connection frequency, and that the oscillator is initially in the ground between quantum limits and noise discussed in previ- state. From our results above we have ous sections, focusing now on a more general approach. S¯ [ω] As before, we will emphasize the idea that reaching xx,tot the quantum limit requires a detector having “quantum-  2  2 γ0/2 ~ |χxx[ω]| ideal” noise properties. The approach here is different = xZPF 2 2 + + |χxx[Ω]| (|ω| − Ω) + (γ0/2) 2 |χxx[Ω]| from typical treatments in the quantum optics litera- 2  2  ture (Gardiner and Zoller, 2000; Haus, 2000), and uses xZPF (γ0/2) ≈ 1 + 3 2 2 (3.62) nothing more than features of quantum linear response. γ0 (|ω| − Ω) + (γ0/2) Our discussion here will expand upon Clerk (2004); Clerk which is plotted in Fig. 6. Assuming that the detector is et al. (2003); somewhat similar approaches to quantum quantum limited, one sees that the Lorentzian peak rises measurement are also discussed in Braginsky and Khalili above the constant background by a factor of three when (1992) and Averin (2003). the coupling is optimized to minimize the total noise In this subsection, we will start by heuristically sketch- power at resonance. This represents the best one can do ing how constraints on noise (similar to Eq. (3.9) for the when continuously monitoring zero-point position fluctu- cavity detector) can emerge directly from the Heisenberg ations. Note that the value of this peak-to-floor ratio is uncertainty principle. We then present a rigorous and a direct consequence of two simple facts which hold for general quantum constraint on noise. We introduce both an optimal coupling, at the quantum limit: i) the total the notion of a generic linear response detector, and the added noise at resonance (back-action plus measurement basic quantum constraint on detector noise. In the next imprecision) is equal to the zero-point noise, and ii) back- subsection (C), we will discuss how this noise constraint action and measurement imprecision make equal contri- leads to the quantum limit on QND state detection of a butions to the total added noise. Somewhat surprisingly, qubit. The quantum limit on a linear amplifier (or a po- the same maximum peak-to-floor ratio is obtained when sition detector) is discussed in detail in the next section. 20

TABLE III Synopsis of recent experiments approaching the quantum limit on continuous position detection of a mechanical ¯I resonator. The second column corresponds to the best measurement imprecision noise spectral density Sxx achieved in the ¯0 experiment. This value is compared against the zero-point position noise spectral density Sxx, calculated using the total measured resonator damping (which may include a back-action contribution). All spectral densities are at the oscillator’s ¯I resonance frequency Ω. As discussed in the text, there is no quantum limit on how small one can make Sxx; for an ideal ¯I ¯0 detector, one needs to tune the detector-resonator coupling so that Sxx = Sxx/2 in order to reach the quantum limit on position detection. The third column presents the product of the measured imprecision noise (unless otherwise noted) and measured back-action noises, divided by ~/2; this quantity must be one to achieve the quantum limit on the added noise. Experiment Mechanical Imprecision noise Detector Noise Product a p ¯I ¯ frequency [Hz] vs. zero-point noise SxxSFF /(~/2) p ¯I ¯0 Ω/(2π) Sxx/Sxx[Ω] Cleland et al. (2002) (quantum point contact) 1.5 × 106 4.2 × 104 Knobel and Cleland (2003) (single-eletron transistor) 1.2 × 108 1.8 × 102 LaHaye et al. (2004) (single-electron transistor) 2.0 × 107 5.4 Naik et al. (2006) (single-electron transistor) 2.2 × 107 5.3 b 8.1 × 102 ¯I 1 (if Sxx had been limited by SET shot noise) 1.5 × 10 Arcizet et al. (2006) (optical cavity) 8.1 × 105 0.87 Flowers-Jacobs et al. (2007) (atomic point-contact) 4.3 × 107 29 1.7 × 103 Regal et al. (2008) (microwave cavity) 2.4 × 105 21 Schliesser et al. (2008) (optical cavity) 4.1 × 107 0.50 Poggio et al. (2008) (quantum point contact) 5.2 × 103 63 Etaki et al. (2008) (d.c. SQUID) 2.0 × 106 47 Groblacher et al. (2009) (optical cavity) 9.5 × 102 0.57 Schliesser et al. (2009) (optical cavity) 6.5 × 107 5.5 1.0 × 102 Teufel et al. (2009) (microwave cavity) 1.0 × 106 0.63 aA blank value in this column indicates that back-action was not measured in the experiment. bNote that back-action effects dominated the mechanical Q in this measurement, lowering it from 1.2 × 105 to ∼ 4.2 × 102. If one compares the imprecision against the zero-point noise of the p ¯I ¯0 uncoupled mechanical resonator, one finds Sxx/Sxx[Ω] ∼ 0.33.

1. Heuristic weak-measurement noise constraints when trying to obtain a good estimate of the position by averaging several such measurements, it is not optimal As we already stressed in the introduction, there is no to make ∆x too small, because otherwise this additional fundamental quantum limit on the accuracy with which a perturbation, called the “back-action” of the measure- given observable can be measured, at least not within the ment device, will become large. The back-action can be framework of non-relativistic quantum mechanics. For described as a random force ∆F = ∆p/δt. A meaningful 2 ¯ example, one can, in principle, measure the position of a limit δt → 0 is obtained by keeping both ∆x δt ≡ Sxx 2 ¯ particle to arbitrary accuracy in the course of a projec- and ∆p /δt ≡ SFF fixed. In this limit, the deviations tion measurement. However, the situation is different δx(t) describing the finite measurement accuracy and when we specialize to continuous, non-QND measure- the fluctuations of the back-action force F can be de- scribed as white noise processes, hδx(t)δx(0)i = S¯xxδ(t) ments. Such a measurement can be envisaged as a series ¯ of instantaneous measurements, in the limit where the and hF (t)F (0)i = SFF δ(t). The Heisenberg uncertainty ¯ ¯ 2 spacing between the measurements δt is taken to zero. relation ∆p∆x ≥ ~/2 then implies SxxSFF ≥ ~ /4 (Bra- Each measurement in the series has a limited resolution ginsky and Khalili, 1992). Note this is completely analo- and perturbs the conjugate variables, thereby affecting gous to the relation Eq. (3.9) we derived for the resonant the subsequent dynamics and measurement results. Let cavity detector using the fundamental number-phase un- us discuss this briefly for the example of a series of posi- certainty relation. In this section, we will derive rigor- tion measurements of a free particle. ously more general quantum limit relations on noise spec- tral densities of this form. After initially measuring the position with an accuracy ∆x, the momentum suffers a random perturbation of size ∆p ≥ ~/(2∆x). Consequently, a second position mea- surement taking place a time δt later will have an addi- tional uncertainty of size δt(∆p/m) ∼ ~δt/(m∆x). Thus, 21

2.0 total noise

] 1.5

Ω eq. + back [ xx 0 action noise S signal input output "load" / ] 1.0 source operator ω

[ equilibrium

tot noise (T=0) detector xx, S 0.5 FIG. 7 (Color online) Schematic of a generic linear response detector.

0.0 interaction Hamiltonian: ˆ ˆ 0.0 0.4 0.8 1.2 Hint = AxˆF (4.1) Frequency ω/Ω Here, the operatorx ˆ (which is not necessarily a position FIG. 6 (Color online) Spectral density of measured position operator) carries the input signal. Note that becausex ˆ fluctuations of a harmonic oscillator, S¯xx,tot[ω], as a function belongs to the signal source, it necessarily commutes with of frequency ω, for a detector which reaches the quantum the detector variables I,ˆ Fˆ. limit at the oscillator frequency Ω. We have assumed that We will always assume the coupling strength A to be without the coupling to the detector, the oscillator would be small enough that we can accurately describe the output in its ground state. The y-axis has been normalized by the of the detector using linear response.6 We thus have: ¯0 zero-point position-noise spectral density Sxx[ω], evaluated at Z ω = Ω. One clearly sees that the total noise at Ω is twice the ˆ ˆ 0 0 0 zero-point value, and that the peak of the Lorentzian rises hI(t)i = hIi0 + A dt χIF (t − t )hxˆ(t )i, (4.2) a factor of three above the background. This background represents the measurement-imprecision, and is equal to 1/2 ˆ ¯0 where hIi0 is the input-independent value of the detector of Sxx(Ω). output at zero-coupling, and χIF (t) is the linear-response susceptibility or gain of our detector. Note that in Clerk et al. (2003) and Clerk (2004), this gain coefficient is denoted λ. Using standard time-dependent perturba- 2. Generic linear-response detector tion theory in the coupling Hˆint, one can easily derive Eq. (4.2), with χIF (t) given by a Kubo-like formula: To rigorously discuss the quantum limit, we would like i Dh iE to start with a description of a detector which is as gen- χIF (t) = − θ(t) Iˆ(t), Fˆ(0) (4.3) eral as possible. To that end, we will think of a detector ~ 0 as some physical system (described by some unspecified Here (and in what follows), the operators Iˆ and Fˆ are ˆ Hamiltonian Hdet and some unspecified density matrix Heisenberg operators with respect to the detector Hamil- ρˆ0) which is time-independent in the absence of coupling tonian, and the subscript 0 indicates an expectation value to the signal source. The detector has both an input with respect to the density matrix of the uncoupled de- port, characterized by an operator Fˆ, and an output tector. port, characterized by an operator Iˆ (see Fig. 7). The As we have already discussed, there will be unavoidable output operator Iˆ is simply the quantity which is read- noise in both the input and output ports of our detector. out at the output of the detector (e.g., the current in a single-electron transistor, or the phase shift in the cavity detector of the previous section). The input operator Fˆ is the detector quantity which directly couples to the input 6 The precise conditions for when linear response breaks down will signal (e.g., the qubit), and which causes a back-action depend on specific details of the detector. For example, in the disturbance of the signal source; in the cavity example cavity detector discussed in Sec. III.B, one would need the di- of the previous section, we had Fˆ =n ˆ, the cavity pho- mensionless coupling A to satisfy A  1/(Qchzˆi) to ensure that the non-linear dependence of the phase shift θ on the signal hzˆi ton number. As we are interested in weak couplings, we is negligible. This translates to the signal modulating the cavity will assume a simple bilinear form for the detector-signal frequency by an amount much smaller than its linewidth κ. 22

This noise is subject to quantum mechanical constraints, did nothing but amplify vacuum noise entering the out- and its presence is what limits our ability to make a mea- put port, this would heat up the system being measured surement or amplify a signal. We thus need to quantita- at the input port and hence produce excess backaction. tively characterize the noise in both these ports. Recall Thus, the ideal situation is to have χFI = 0, implying from the discussion in Sec. II.B that it is the symmetric- a high asymmetry between the input and output of the in-frequency part of a quantum noise spectral density detector, and requiring the detector to be in a state far which plays a role akin to classical noise. We will thus from thermodynamic equilibrium. We note that almost want to characterize the symmetrized noise correlators of all mesoscopic detectors that have been studied in detail our detector (denoted as always with a bar). Redefining (e.g. single electron transistors and generalized quantum these operators so that their average value is zero at zero point contacts) have been found to have a vanishing re- coupling (i.e. Fˆ → Fˆ − hFˆi0, Iˆ → Iˆ− hIˆi0), we have: verse gain: χFI = 0 (Clerk et al., 2003). For this reason, Z ∞ we will often focus on the ideal (but experimentally rel- 1 iωt S¯FF [ω] ≡ dte h{Fˆ(t), Fˆ(0)}i0 (4.4a) evant) situation where χFI = 0 in what follows. 2 −∞ Before proceeding, it is worth emphasizing that there Z ∞ 1 iωt is a relation between the detector gains χIF and χFI S¯II [ω] ≡ dte h{Iˆ(t), Iˆ(0)}i0 (4.4b) 2 −∞ and the unsymmetrized I-F quantum noise correlator, Z ∞ SIF [ω]. This spectral density, which need not be sym- 1 iωt S¯IF [ω] ≡ dte h{Iˆ(t), Fˆ(0)}i0 (4.4c) metric in frequency, is defined as: 2 −∞ Z ∞ ¯ iωt where {, } indicates the anti-commutator, SII represents SIF [ω] = dte hIˆ(t)Fˆ(0)i0 (4.7) the intrinsic noise in the output of the detector, and S¯FF −∞ describes the back-action noise seen by the source of the input signal. In general, there will be some correlation Using the definitions, one can easily show that: between these two kinds of noise; this is described by the ¯ 1 ∗ cross-correlator SIF . S¯IF [ω] = [SIF [ω] + SIF [−ω] ] (4.8a) Finally, we must also allow for the possibility that our 2 detector could operate in reverse (i.e. with input and ∗ i ∗ χIF [ω] − χFI [ω] = − [SIF [ω] − SIF [−ω] ](4.8b) output ports playing opposite roles). We thus intro- ~ duce the reverse gain χ of our detector. This is the FI ¯ response coefficient describing an experiment where we Thus, while SIF represents the classical part of the I-F couple our input signal to the output port of the detector quantum noise spectral density, the gains χIF , χFI are determined by the quantum part of this spectral density. (i.e. Hˆint = AxˆIˆ), and attempt to observe something at This also demonstrates that though the gains have an the input port (i.e. in hFˆ(t)i). In complete analogy to explicit factor of 1/ in their definitions, they have a Eq. (4.2), one would then have: ~ well defined ~ → 0 limit, as the asymmetric-in-frequency Z 0 0 0 part of SIF [ω] vanishes in this limit. hFˆ(t)i = hFˆi0 + A dt χFI (t − t )hxˆ(t )i (4.5) with: 3. Quantum constraint on noise i h i χFI (t) = − θ(t)h Fˆ(t), Iˆ(0) i0 (4.6) ~ Despite having said nothing about the detec- Note that if our detector is in a time-reversal symmet- tor’s Hamiltonian or state (except that it is time- ric, thermal equilibrium state, then Onsager reciprocity independent), we can nonetheless derive a very general ∗ quantum constraint on its noise properties. Note first relations would imply either χIF = χFI (if I and F have ∗ that for purely classical noise spectral densities, one al- the same parity under time-reversal) or χIF = −χFI (if I and F have the opposite parity under time-reversal) (see, ways has the inequality: e.g. Pathria (1996)). Thus, if the detector is in equilib- ¯ ¯ ¯ 2 rium, the presence of gain necessarily implies the pres- SII [ω]SFF [ω] − SIF [ω] ≥ 0 (4.9) ence of reverse gain. Non-zero reverse gain is also found in many standard classical electrical amplifiers such as This simply expresses the fact that the correlation be- op amps (Boylestad and Nashelsky, 2006). tween two different noisy quantities cannot be arbitrarily The reverse gain is something that we must worry large; it follows immediately from the Schwartz inequal- about even if we are not interested in operating our detec- ity. In the quantum case, this simple constraint becomes tor in reverse. To see why, note that to make a measure- modified whenever there is an asymmetry between the ment of the output operator Iˆ, we must necessarily couple detector’s gain and reverse gain. This asymmetry is pa- to it in some manner. If χFI 6= 0, the noise associated rameterized by the quantityχ ˜IF [ω]: with this coupling could in turn lead to additional back- ∗ action noise in the operator Fˆ. Even if the reverse gain χ˜IF [ω] ≡ χIF [ω] − (χFI [ω]) (4.10) 23

We will show below that the following quantum noise in- find, at any finite frequency ω 6= 0: equality (involving symmetrized noise correlators) is al- 2 ¯ 2 ¯ ways valid (see also Eq. (6.36) in Braginsky and Khalili (∆A) = SII [ω], (∆B) = SFF [ω] (4.15a) (1996)): Dn oE iδ A,ˆ Bˆ = 2Re e S¯IF [ω] (4.15b) ¯ ¯ ¯ 2 Dh iE Z +∞ Dh iE SII [ω]SFF [ω] − SIF [ω] ≥ A,ˆ Bˆ = dt cos(ωt + δ) Iˆ(t), Fˆ(0) χ˜ [ω] 2   S¯ [ω]  −∞ ~ IF 1 + ∆ IF (4.11)  iδ ∗ = i~Re e χIF [ω] − (χFI [ω]) (4.15c) 2 ~χ˜IF [ω]/2 In the last line, we have simply made use of the where Kubo formula definitions of the gain and reverse gain

2 2 (cf. Eqs. (4.3) and (4.6)). As a consequence of 1 + z − 1 + |z| ∆[z] = , (4.12) Eqs. (4.15a)-(4.15c), the Heisenberg uncertainty relation 2 (4.13) directly yields:

 iδ 2 To interpret the quantum noise inequality Eq. (4.11), S¯II [ω]S¯FF [ω] ≥ Re e S¯IF [ω] + (4.16) note that 1+∆[z] ≥ 0. Eq. (4.11) thus implies that if our 2 ~  iδ ∗2 detector has gain and does not have a perfect symmetry Re e χIF [ω] − (χFI [ω]) ∗ 4 between input and output (i.e. χIF 6= χFI ), then it must in general have a minimum amount of back-action and Maximizing the RHS of this inequality over δ then yields output noise; moreover, these two noises cannot be per- the general quantum noise constraint of Eq. (4.11). fectly anti-correlated. As we will show in the following With this derivation, we can now interpret the quan- sections, this constraint on the noise of a detector directly tum noise constraint Eq. (4.11) as stating that the noise leads to quantum limits on various different measurement at a given frequency in two observables, Iˆ and Fˆ, is tasks. Note that in the zero-frequency limit,χ ˜IF and ¯ bounded by the value of their commutator at that fre- SIF are both real, implying that the term involving ∆ in quency. The fact that Iˆ and Fˆ do not commute is neces- Eq. (4.11) vanishes. The result is a simpler looking in- sary for the existence of linear response (gain) from the equality found elsewhere in the literature (Averin, 2003; detector, but also means that the noise in both Iˆ and Clerk, 2004; Clerk et al., 2003). Fˆ cannot be arbitrarily small. A more detailed deriva- While Eq. (4.11) may appear vaguely reminiscent to tion, yielding additional important insights, is described the standard fluctuation dissipation theorem, its ori- in Appendix I.1. gin is quite different: in particular, the quantum noise Given the quantum noise constraint of Eq. (4.11), we constraint applies irrespective of whether the detector can now very naturally define a “quantum-ideal” detector is in equilibrium. Eq. (4.11) instead follows directly (at a given frequency ω) as one which minimizes the LHS from Heisenberg’s uncertainty relation applied to the fre- ˆ ˆ of Eq. (4.11)– a quantum-ideal detector has a minimal quency representation of the operators I and F . In its amount of noise at frequency ω. We will often be inter- most general form, the Heisenberg uncertainty relation ested in the ideal case where there is no reverse gain (i.e. gives a lower bound for the uncertainties of two observ- measuring Iˆ does not result in additional back-action ables in terms of their commutator and their noise cor- noise in Fˆ); the condition to have a quantum limited relator (Gottfried, 1966): detector thus becomes: 1 Dn oE2 1 Dh iE 2 ¯ ¯ ¯ 2 2 2 ˆ ˆ ˆ ˆ SII [ω]SFF [ω] − SIF [ω] = (4.17) (∆A) (∆B) ≥ A, B + A, B . (4.13) 4 4 2   ¯  ~χIF [ω] SIF [ω] 1 + ∆ Here we have assumed hAˆi = hBˆi = 0. Let us now choose 2 ~χIF [ω]/2 the Hermitian operators Aˆ and Bˆ to be given by the where ∆[z] is given in Eq. (4.12). Again, as we will cosine-transforms of Iˆ and Fˆ, respectively, over a finite discuss below, in most cases of interest (e.g. zero fre- time-interval T : quency and/or large amplifier power gain), the last term on the RHS will vanish. In the following sections, we r 2 Z T/2 Aˆ ≡ dt cos(ωt + δ) Iˆ(t) (4.14a) will demonstrate that the “ideal noise” requirement of T −T/2 Eq. (4.17) is necessary in order to achieve the quantum r 2 Z T/2 limit on QND detection of a qubit, or on the added noise Bˆ ≡ dt cos(ωt) Fˆ(t) (4.14b) of a linear amplifier. T −T/2 Before leaving our general discussion of the quantum noise constraint, it is worth emphasizing that achieving Note that we have phase shifted the transform of Iˆ rela- Eq. (4.17) places a very strong constraint on the prop- tive to that of Fˆ by a phase δ. In the limit T → ∞ we erties of the detector. In particular, there must exist a 24 tight connection between the input and output ports of to just consider the noise of the detector: one must nec- the detector– in a certain restricted sense, the operators Iˆ essarily also consider the noise associated with whatever and Fˆ must be proportional to one another (see Eq. (I13) system is coupled to Iˆ to readout the detector output, in Appendix I.1). As is discussed in Appendix I.1, this as this noise will be fed back to the detector input port, proportionality immediately tells us that a quantum-ideal causing additional back-action; we give an explicit ex- detector cannot be in equilibrium. The proportionality ample of this in the next sub-section, when we discuss exhibited by a quantum-ideal detector is parameterized QND qubit detection. Even more problematically, when ∗ by a single complex valued number α[ω], whose magni- χFI = χIF , there is never any amplification by the detec- tude is given by: tor. As we will discuss in Sec. V.E.3, the proper metric of the detector’s ability to amplify is its dimensionless 2 |α[ω]| = S¯II [ω]/S¯FF [ω] (4.18) power gain: what is the power supplied at the output of the detector versus the amount of power drawn at the ∗ While this proportionality requirement may seem purely input from the signal source? When χFI = χIF , one has formal, it does have a simple heuristic interpretation; negative feedback, with the result that the power gain as is discussed in Clerk et al. (2003), it may be viewed cannot be larger than 1 (cf. Eq. 5.53). There is thus no ∗ as a formal expression of the principle that a quantum- amplification when χFI = χIF . Further, if one also in- limited detector must not contain any ‘wasted informa- sists that the noise constraint of Eq. (4.11) is optimized, tion’ (cf. Sec. III.B.2). then one finds the power gain must be exactly 1; this is explicitly demonstrated in Appendix I.2. The detector thus will simply act as a transducer, reproducing the in- put signal at the output without any increase in energy. 4. Evading the detector quantum noise inequality We have here a specific example of a more general idea that will be discussed extensively in Sec. V: if a detector We now turn to situations where the RHS of Eq. (4.11) only acts as a transducer, it need not add any noise. vanishes, implying that there is no additional quantum constraint on the noise of our detector beyond what ex- At finite frequencies, there is a second way to make the RHS of the quantum noise constraint of Eq. (4.11) ists classically. In such situations, one could have a de- ¯ tector with perfectly correlated back-action and output vanish: one needs the quantity SIF [ω]/χ˜IF to be purely noises (i.e. S¯ S¯ = |S¯ |2), or even with a vanishing imaginary, and larger in magnitude than ~/2. In this FF II IF case, it would again be possible to have the LHS of the back-action S¯FF = 0. Perhaps not surprisingly, these sit- uations are not of much utility. As we will now show, in noise constraint of Eq. (4.11) equal to zero. However, cases where the RHS of Eq. (4.11) vanishes, the detector one again finds that in such a case, the dimensionless may be low noise, but will necessarily be deficient in an- power gain of the detector is at most equal to one; it thus other important regard: it will not be good enough that does not amplify. This is shown explicitly in Appendix we can ignore the extra noise associated with the mea- I.2. An important related statement is that a quantum- ˆ limited detector with a large power gain must have the surement of the detector output I. As we have already ¯ discussed, reading-out Iˆ will invariably involve coupling quantity SIF /χIF be real. Thus, at the quantum limit, the detector output to some other physical system. In the correlations between the back-action force and the intrin- ideal case, this coupling will not generate any additional sic output noise fluctuations must have the same phase back-action on the system coupled to the detector’s in- as the gain χIF . As we discuss further in Sec. V.F, this put port. In addition, the signal at the detector output requirement can be interpreted in terms of the principle should be large enough that any noise introduced in mea- of ‘no wasted information’ introduced in Sec. III.B.2. suring Iˆ is negligible; we already came across this idea in our discussion of the resonant cavity detector (see com- ments following Eq. (3.60)). This means that we need B. Quantum limit on QND detection of a qubit our detector to truly amplify the input signal, not sim- ply reproduce it at the output with no gain in energy. In Sec. III.B, we discussed the quantum limit on QND As we now show, a detector which evades the quantum qubit detection in the specific context of a resonant cavity constraint of Eq. (4.11) by making the RHS of the in- detector. We will now show how the full quantum noise equality zero will necessarily fail in one or both of the constraint of Eq. (4.11) directly leads to this quantum above requirements. limit for an arbitrary weakly-coupled detector. Similar The most obvious case where the quantum noise con- to Sec. III.B, we couple the input operator of our generic straint vanishes is for a detector which has equal forward linear-response detector to theσ ˆz operator of the qubit ∗ and reverse gains, χFI = χIF . As we have already men- we wish to measure (i.e. we takex ˆ =σ ˆz in Eq. (4.1)); tioned, this relation will necessarily hold if the the detec- we also consider the QND regime, whereσ ˆz commutes tor is time-reversal symmetric and in equilibrium, and with the qubit Hamiltonian. As we saw in Sec. III.B, Iˆ and Fˆ have the same parity under time-reversal. In the quantum limit in this case involves the inequality this case, the relatively large reverse gain implies that in Γmeas ≤ Γϕ, where Γmeas is the measurement rate, and analyzing a given measurement task, it is not sufficient Γϕ is the back-action dephasing rate. For the latter quan- 25 tity, we can directly use the results of our calculation for This yields: the cavity system, where we found the dephasing rate 2 2 was set by the zero-frequency noise in the cavity pho- A (χIF ) Γmeas = (4.24) ton number (cf. Eq. 3.27). In complete analogy, the 2S¯II back-action dephasing rate here will be determined by the zero-frequency noise in the input operator Fˆ of our Putting this all together, we find that the “efficiency” detector: ratio η = Γmeas/Γϕ is given by:

2 2A 2 2 Γ = S¯ (4.19) Γmeas ~ (χIF ) ϕ 2 FF η ≡ = (4.25) ~ Γϕ 4S¯II S¯FF We omit frequency arguments in this subsection, as it is always the zero-frequency susceptibilities and spectral In the case where our detector has a vanishing reverse densities which appear. gain (i.e. χFI = 0), the quantum-limit bound η ≤ 1 fol- lows immediately from the quantum noise constraint of The measurement rate (the rate at which information Eq. (4.11). We thus see that achieving the quantum- on the state of qubit is acquired) is also defined in com- limit for QND qubit detection requires both a detec- plete analogy to what was done for the cavity detector. tor with quantum-ideal noise properties, as defined by We imagine we turn the measurement on at t = 0 and Eq. (4.17), as well as a detector with a vanishing noise start to integrate up the output I(t) of our detector: cross-correlator: S¯IF = 0. Z t If in contrast χFI 6= 0, it would seem that it is possi- mˆ (t) = dt0Iˆ(t0) (4.20) ble to have η ≥ 1. This is of course an invalid inference: 0 as discussed, χFI 6= 0 implies that we must necessarily consider the effects of extra noise injected into detector’s The probability distribution of the integrated output ˆ mˆ (t) will depend on the state of the qubit; for long times, output port when one measures I, as the reverse gain we may approximate the distribution corresponding to will bring this noise back to the qubit, causing extra de- each qubit state as being gaussian. Noting that we have phasing. The result is that one can do no better than chosen Iˆ so that its expectation vanishes at zero coupling, η = 1. To see this explicitly, consider the extreme case χIF = χFI and S¯II = S¯FF = 0, and suppose we use the average value of hmˆ (t)i corresponding to each qubit ˆ state is (in the long time limit of interest): a second detector to read-out the output I of the first detector. This second detector has input and output op- erators Fˆ , Iˆ ; we also take it to have a vanishing reverse hmˆ (t)i↑ = AχIF t, hmˆ (t)i↓ = −AχIF t (4.21) 2 2 gain, so that we do not have to also worry about how its The variance of both distributions is, to leading order, output is read-out. Coupling the detectors linearly in the ˆˆ independent of the qubit state: standard way (i.e. Hint,2 = IF2), the overall gain of the

two detectors in series is χI2F2 ·χIF , while the back-action 2 2 2 ¯ driving the qubit dephasing is described by the spectral hmˆ (t)i↑/↓ − hmˆ (t)i↑/↓ ≡ hhmˆ (t)ii↑/↓ = SII t (4.22) 2 density (χFI ) SF2F2 . Using the fact that our second de- For the last equality above, we have taken the long-time tector must itself satisfy the quantum noise inequality, limit, which results in the variance ofm ˆ being determined we have: ¯ completely by the zero-frequency output noise SII [ω = 0] 2 of the detector. The assumption here is that due to h 2 i ~ 2 (χFI ) S¯F F S¯I I ≥ (χI F · χIF ) (4.26) the weakness of the measurement, the measurement time 2 2 2 2 4 2 2 (i.e. 1/Γ ) will be much longer than the autocorrela- meas Thus, the overall chain of detectors satisfies the usual, tion time of the detector’s noise. zero-reverse gain quantum noise inequality, implying that We can now define the measurement rate, in complete we will still have η ≤ 1. analogy to the cavity detector of the previous section (cf. Eq. (3.23)), by how quickly the resolving power of the measurement grows:7 V. QUANTUM LIMIT ON LINEAR AMPLIFIERS AND 2 POSITION DETECTORS 1 [hmˆ (t)i↑ − hmˆ (t)i↓] 2 2 ≡ Γmeast. (4.23) 4 hhmˆ (t)ii↑ + hhmˆ (t)ii↓ In the previous section, we established the fundamen- tal quantum constraint on the noise of any system ca- pable of acting as a linear detector; we further showed that this quantum noise constraint directly leads to the 7 The strange looking factor of 1/4 here is purely chosen for con- “quantum limit” on non-demolition qubit detection us- venience; we are defining the measurement rate based on the information theoretic definition given in Appendix F. This fac- ing a weakly-coupled detector. In this section, we turn tor of four is consistent with the definition used in the cavity to the more general situation where our detector is a system. phase-preserving quantum linear amplifier: the input to 26 the detector is described by some time-dependent oper- For a “pump frequency” ωP equalling twice the “signal atorx ˆ(t) which we wish to have amplified at the output frequency”, ωP = 2ωS, the resulting dynamics will lead of our detector. As we will see, the quantum limit in to an amplification of the initial oscillator position, with this case is a limit on how small one can make the noise the energy provided by the external driving: added by the amplifier to the signal. The discussion in this section both furthers and generalizes the heuristic λt p(0) −λt discussion of position detection using a cavity detector x(t) = x(0)e cos(ωSt) + e sin(ωSt) (5.2) MωS presented in Sec. III.B. In this section, we will start by presenting a heuris- Thus, one of the quadratures is amplified exponen- tic discussion of quantum constraints on amplification. tially, at a rate λ = δω/2, while the other one decays. We will then demonstrate explicitly how the previously- In a quantum-mechanical description, this produces a discussed quantum noise constraint leads directly to the squeezed state out of an initial coherent state. Such quantum limit on the added noise of a phase-preserving a system is called a “degenerate parametric amplifier”, linear amplifier; we will examine both the cases of a and we discuss its quantum dynamics in more detail in generic linear position detector and a generic voltage am- Sec. V.H. We will see that such an amplifier, which only plifier, following the approach outlined in Clerk (2004). amplifies a single quadrature, is not required quantum We will also spend time explicitly connecting the linear mechanically to add any noise (Braginsky and Khalili, response approach we use here to the bosonic scattering 1992; Caves, 1982; Caves et al., 1980). formulation of the quantum limit favoured by the quan- Can we now change this parametric amplification tum optics community (Caves, 1982; Courty et al., 1999; scheme slightly in order to make both signal quadratures Grassia, 1998; Haus and Mullen, 1962), paying particular grow with time? It turns out this is impossible, as long attention to the case of two-port scattering amplifier. We as we restrict ourselves to a driven system with a sin- will see that there are some important subtleties involved gle degree of freedom. The reason in in converting between the two approaches. In particular, is that Liouville’s theorem requires phase space volume there exists a crucial difference between the case where to be conserved during motion. More formally, this is the input signal is tightly coupled to the input of the related to the conservation of Poisson brackets, or, in amplifier (the case usually considered in the quantum quantum mechanics, to the conservation of commutation optics community), versus the case where, similar to an relations. Nevertheless, it is certainly desirable to have ideal op-amp, the input signal is only weakly coupled to an amplifier that acts equally on both quadratures (a the input of the amplifier (the case usually considered in so-called “phase-preserving” or “phase-insensitive” am- the solid state community). plifier), since the signal’s phase is often not known be- forehand. The way around the restriction created by Liouville’s theorem is to add more degrees of freedom, A. Preliminaries on amplification such that the phase space volume can expand in both quadratures (i.e. position and momentum) of the inter- What exactly does one mean by ‘amplification’? As esting signal degree of freedom, while being compressed we will see (cf. Sec. V.E.3), a precise definition requires in other directions. This is achieved most easily by cou- that the energy provided at the output of the amplifier pling the signal oscillator to another oscillator, the “idler be much larger than the energy drawn at the input of mode”. The external driving now modulates the cou- the amplifier– the “power gain” of the amplifier must pling between these oscillators, at a frequency that has to be larger than one. For the moment, however, let us equal the sum of the oscillators’ frequencies. The result- work with the cruder definition that amplification in- ing scheme is called a phase-preserving non-degenerate volves making some time-dependent signal ‘larger’. To parametric amplifier (see Sec. V.C). set the stage, we will first consider an extremely sim- ple classical analogue of a linear amplifier. Imagine the Crucially, there is a price to pay for the introduction of “signal” we wish to amplify is the coordinate x(t) of a an extra degree of freedom: there will be noise associated harmonic oscillator; we can write this signal as: with the “idler” oscillator, and this noise will contribute to the noise in the output of the amplifier. Classically, p(0) one could make the noise associated with the “idler” os- x(t) = x(0) cos(ωSt) + sin(ωSt) (5.1) cillator arbitrarily small by simply cooling it to zero tem- MωS perature. This is not possible quantum-mechanically; Our signal has two quadrature amplitudes, i.e. the am- there are always zero-point fluctuations of the idler oscil- plitude of the cosine and sine components of x(t). To lator to contend with. It is this noise which sets a funda- “amplify” this signal, we start at t = 0 to parametrically mental quantum limit for the operation of the amplifier. drive the oscillator by changing its frequency ωS periodi- We thus have a heuristic accounting for why there is a cally in time: ωS(t) = ω0 +δω sin(ωP t), where we assume quantum-limit on the added noise of a phase-preserving δω  ω0. The well-known physical example is a swing linear amplifier: one needs extra degrees of freedom to whose motion is being excited by effectively changing the amplify both signal quadratures, and such extra degrees length of the pendulum at the right frequency and phase. of freedom invariably have noise associated with them. 27

B. Standard Haus-Caves derivation of the quantum limit and the output operators must satisfy the usual commu- on a bosonic amplifier tation relations:

 † h †i We now make the ideas of the previous subsection a,ˆ aˆ = 1, ˆb, ˆb = 1 (5.5) more precise by quickly sketching the standard deriva- tion of the quantum limit on the noise added by a phase- Second, the linearity of the amplifier and the fact that it preserving amplifier. This derivation is originally due is phase preserving (i.e. both signal quadratures are am- to Haus and Mullen (1962), and was both clarified and plified the same way) implies a simple relation between extended by Caves (1982); the amplifier quantum limit the output operator ˆb and the input operatora ˆ: was also motivated in a slightly different manner by √ √ Heffner (1962).8 While extremely compact, the Haus- ˆb = Ga,ˆ ˆb† = Gaˆ† (5.6) Caves derivation can lead to confusion when improperly applied; we will discuss this in Sec. V.D, as well as in where G is the dimensionless “photon number gain” of the Sec. VI, where we apply this argument carefully to the amplifier. It is immediately clear however this expression important case of a two-port quantum voltage amplifier, cannot possibly be correct as written because it violates and discuss the connection to the general linear-response the fundamental bosonic commutation relation [ˆb, ˆb†] = formulation of Sec. IV. 1. We are therefore forced to write The starting assumption of this derivation is that both √ √ ˆ ˆ ˆ† † ˆ† the input and output ports of the amplifier can be de- b = Gaˆ + F, b = Gaˆ + F (5.7) scribed by sets of bosonic modes. If we focus on a nar- ˆ row bandwidth centered on frequency ω, we can describe where F is an operator representing additional noise a classical signal E(t) in terms of a complex number a added by the amplifier. Based on the discussion of the ˆ defining the amplitude and phase of the signal (or equiva- previous subsection, we can anticipate what F repre- lently the two quadrature amplitudes) (Haus, 2000; Haus sents: it is noise associated with the additional degrees and Mullen, 1962) of freedom which must invariably be present in a phase- preserving amplifier. ˆ E(t) ∝ i[ae−iωt − a∗e+iωt]. (5.3) As F represents noise, it has a vanishing expectation value; in addition, one also assumes that this noise is uncorrelated with the input signal, implying [Fˆ, aˆ] = In the quantum case, the two signal quadratures of E(t) ˆ † ˆ ˆ † (i.e. the real and imaginary parts of a(t)) cannot be mea- [F, aˆ ] = 0 and hFaˆi = hFaˆ i = 0. Insisting that ˆ ˆ† sured simultaneously because they are canonically conju- [b, b ] = 1 thus yields: gate; this is in complete analogy to a harmonic oscillator h i (cf. Eq. (3.48)). As a result a, a∗ must be elevated to the Fˆ, Fˆ† = 1 − G (5.8) status of photon ladder operators: a → a,ˆ a∗ → a† . Consider the simplest case, where there is only a single The question now becomes how small can we make the mode at both the input and output, with corresponding noise described by Fˆ? Using Eqs. (5.7), the noise at the operatorsa ˆ and ˆb.9 It follows that the input signal into amplifier output ∆b is given by: the amplifier is described by the expectation value haˆi, 2 2 1 D E while the output signal is described by hˆbi. Correspond- (∆b) = G (∆a) + {Fˆ, Fˆ†} 2 ingly, the symmetrized noise in both these quantities is 1 D E described by: ≥ G (∆a)2 + [Fˆ, Fˆ†] 2 1 2 |G − 1| (∆a)2 ≡ {a,ˆ aˆ†} − |haˆi|2 (5.4) ≥ G (∆a) + (5.9) 2 2 We have used here a standard inequality to bound the 2 with an analogous definition for (∆b) . expectation of {Fˆ, Fˆ†}. The first term here is simply To derive a quantum limit on the added noise of the the amplified noise of the input, while the second term amplifier, one uses two simple facts. First, both the input represents the noise added by the amplifier. Note that if there is no amplification (i.e. G = 1), there need not be any added noise. However, in the more relevant case of large amplification (G  1), the added noise cannot 8 Note that Caves (1982) provides a thorough discussion of why vanish. It is useful to express the noise at the output as the derivation of the amplifier quantum limit given in Heffner an equivalent noise at (“referred to”) the input by simply (1962) is not rigorously correct 9 To relate this to the linear response detector of Sec. IV.A, one dividing out the photon gain G. Taking the large-G limit, could naively writex ˆ, the operator carrying the input signal, as, we have: e.g.,x ˆ =a ˆ +a ˆ†, and the output operator Iˆ as, e.g., Iˆ = ˆb + ˆb† 2 (we will discuss how to make this correspondence in more detail (∆b) 1 ≥ (∆a)2 + (5.10) in Sec. VI) G 2 28

Thus, we have a very simple demonstration that an am- C. Non-degenerate parametric amplifier plifier with a large photon gain must add at least half a quantum of noise to the input signal. Equivalently, the 1. Gain and added noise minimum value of the added noise is simply equal to the zero-point noise associated with the input mode; the to- Before we start our discussion of how the Haus-Caves tal output noise (referred to the input) is at least twice formulation of the quantum limit connects to the general the zero point input noise. Note that both these conclu- linear-response approach of Sec. IV, it is useful to con- sions are identical to what we found in our analysis of sider a specific example. To that end, we analyze here the resonant cavity position detector in Sec. III.B.3. We a non-degenerate parametric amplifier, a linear, phase- will discuss in later sections how this conclusion can also preserving amplifier which reaches the quantum limit on be reached using the general linear-response language of its added noise (Gordon et al., 1963; Louisell et al., 1961; Sec. IV (cf. Sec. V.E, V.F). Mollow and Glauber, 1967a,b) and directly realizes the ideas of the previous sub-section. One possible realiza- As already discussed, the added noise operator F is tion (Yurke et al., 1989) is a cavity with three internal associated with additional degrees of freedom (beyond resonances that are coupled together by a non-linear el- input and output modes) necessary for phase-preserving ement (such as a Josephson junction) whose symmetry amplification. To see this more concretely, note that ev- permits three-wave mixing. The three modes are called ery linear amplifier is inevitably a non-linear system con- the pump, idler and signal and their struc- sisting of an energy source and a ‘spigot’ controlled by ture (illustrated in Fig. 8) obeys ωP = ωI + ωS. The the input signal which redirects the energy source partly system Hamiltonian is then to the output channel and partly to some other chan- nel(s). Hence there are inevitably other degrees of free- ˆ  † † †  Hsys = ~ ωPaˆPaˆP + ωIaˆI aˆI + ωSaˆSaˆS dom involved in the amplification process beyond the in-   put and output channels. An explicit example is the † † † + i~η aˆSaˆI aˆP − aˆSaˆIaˆP . (5.13) quantum parametric amplifier, to be discussed in the next subsection. Further insights into amplifier added We have made the rotating wave approximation in the noise and its connection to the fluctuation-dissipation three-wave mixing term, and without loss of generality, theorem can be obtained by considering a simple model we take the non-linear susceptibility η to be real and posi- where a transmission line is terminated by an effective tive. The system is driven at the pump frequency and the negative impedance; we discuss this model in Appendix three wave mixing term permits a single pump photon to C.4. split into an idler photon and a signal photon. This pro- cess is stimulated by signal photons already present and To see explicitly the role of the additional degrees of leads to gain. A typical mode of operation would be the freedom, note first that for G > 1 the RHS of Eq. (5.8) is negative resistance reflection mode in which the input negative. Hence the simplest possible form for the added signal is reflected from a non-linear cavity and the re- noise is flected beam extracted using a circulator (Bergeal et al., 2008; Yurke et al., 1989). √ √ The non-linear EOMs become tractable if we assume ˆ ˆ† ˆ† ˆ F = G − 1d , F = G − 1d (5.11) the pump has large amplitude and can be treated classi- cally by making the substitution

−iωPt −i(ωI+ωS)t where dˆ and dˆ† represent a single additional mode of aˆP = ψPe = ψPe , (5.14) the system. This is the minimum number of additional where without loss of generality we take ψ to be real degrees of freedom that must inevitably be involved in P and positive. We note here the important point that if the amplification process. Note that for this case, the this approximation is not valid, then our amplifier would inequality in Eq. (5.9) is satisfied as an equality, and in any case not be the linear amplifier which we seek. the added noise takes on its minimum possible value. With this approximation we can hereafter forget about If instead we had, say, two additional modes (coupled the dynamics of the pump degree of freedom and deal inequivalently): with the reduced system Hamiltonian ˆ  † †  √ Hsys = ~ ωIaˆI aˆI + ωSaˆSaˆS + i~λ Fˆ = G − 1(cosh θdˆ† + sinh θdˆ ) (5.12) 1 2   † † −i(ωI+ωS)t +i(ωI+ωS)t aˆSaˆI e − aˆSaˆIe (5.15) it is straightforward to show that the added noise is in- where λ ≡ ηψP. Transforming to the interaction repre- evitably larger than the minimum. This again can be sentation we are left with the following time-independent interpreted in terms of wasted information, as the extra quadratic Hamiltonian for the system degrees of freedom are not being monitored as part of the ˆ  † †  measurement process and so information is being lost. Vsys = i~λ aˆSaˆI − aˆSaˆI (5.16) 29

To get some intuitive understanding of the physics, let damping rate of the cavity modes). It then suffices to us temporarily ignore the damping of the cavity modes find the steady state solution of these EOMs that would result from their coupling to modes outside the cavity. We now have a pair of coupled EOMs for the 2λ † 2 ˆ aˆS = aˆI − √ bS,in (5.20) two modes κS κS 2λ 2 ˙ † aˆ† = aˆ − ˆb† . (5.21) aˆS = +λaˆ I S √ I,in I κI κI † aˆ˙ = +λaˆ (5.17) I S The output signal of the non-degnerate paramp is for which the solutions are the signal leaving the cavity signal mode and entering the external bath modes; it is described by an oper- ˆ † ator bS,out. The standard input-output theory treat- aˆS(t) = cosh(λt)ˆaS(0) + sinh(λt)ˆaI (0) † † ment of the extra-cavity modes (Walls and Milburn, aˆI (t) = sinh(λt)a ˆS(0) + cosh(λt)a ˆI (0) (5.18) 1994), presented in Appendix E, yields the simple re- lation (cf. Eq. (E37)): We see that the amplitude in the signal channel grows ex- ponentially in time and that the effect of the time evolu- ˆ ˆ √ bS,out = bS,in + κSaˆS. (5.22) tion is to perform a simple unitary transformation which † mixesa ˆS witha ˆI in such a way as to preserve the commu- The first term corresponds to the reflection of the signal tation relations. Note the close connection with the form and noise incident on the cavity from the bath, while found from very general arguments in Eqs. (5.7)-(5.11). the second term corresponds to radiation from the cavity We may now tackle the full system which includes the mode into the bath. Using this, we find that the output coupling between the cavity modes and modes external signal from the cavity is given by to the cavity. Such a coupling is of course necessary in Q2 + 1 2Q order to feed the input signal into the cavity, as well as ex- ˆb = ˆb + ˆb† , (5.23) tract the amplified output signal. It will also result in the S,out Q2 − 1 S,in Q2 − 1 I,in damping of the cavity modes, which will cut off the expo- where Q ≡ √2λ is proportional to the pump amplitude nential growth found above and yield a fixed amplitude κIκS gain. We will present and motivate the main results in and inversely proportional to the cavity decay rates. We this subsection, relegating details to how one treats the have to require Q2 < 1 to make sure that the parametric bath modes (so-called input-output theory (Walls and amplifier does not settle into self-sustained oscillations, Milburn, 1994)) to Appendix E. Working in the stan- i.e. it works below threshold. Under that condition, we dard Markovian limit, we obtain the following EOMs in can define the photon-number gain G0 via the interaction representation: p − G = (Q2 + 1)/(Q2 − 1), (5.24) κ √ 0 aˆ˙ = − S aˆ + λaˆ† − κ ˆb , S 2 S I S S,in such that ˙ † κI † √ ˆ† aˆI = − aˆI + λ aˆS − κI bI,in. (5.19) ˆ p ˆ p ˆ† 2 bS,out = − G0 bS,in − G0 − 1 bI,in. (5.25)

Here κS and κI are the respective damping rates of the ˆ ˆ In the ideal case, the noise associated with bI,in, bS,in cavity signal mode and the idler mode. The coupling to is simply vacuum noise. As a result, the input-output extra-cavity modes also lets signals and noise enter the relation Eq. (5.25) is precisely of the Haus-Caves form cavity from the baths: this is described by the bosonic Eq. (5.11) for an ideal, quantum-limited amplifier. It ˆ ˆ operators bS,in and bI,in which drive (respectively) the demonstrates that the non-degenerate parametric ampli- ˆ signal and idler modes. bS,in describes both the input fier reaches the quantum limit for minimum added noise. signal to be amplified plus vacuum noise entering from In the limit of large gain the output noise (referred to the ˆ the bath coupled to the signal mode, whereas bI,in simply input) for a vacuum input signal is precisely doubled. describes vacuum noise.10 Let us fix our attention on signals inside a frequency window δω centered on ωS (hence zero frequency in the 2. Bandwidth-gain tradeoff interaction representation). For simplicity, we will first consider the case where the signal bandwidth δω is al- The above results neglected the finite bandwidth δω most infinitely narrow (i.e. much much smaller than the of the input signal to the amplifier. The gain G0 given in Eq. (5.24) is only the gain at precisely the mean sig- nal frequency ωS; for a finite bandwidth, we also need to understand how the power gain varies as a function of 10 Note that the ˆb operators are not dimensionless, as ˆb†ˆb represents frequency over the entire signal bandwidth. As we will a photon flux. This is discussed fully in Appendix E. see, a parametric amplifier suffers from the fact that as 30

3. Effective temperature ω I Recall that in Sec. II.B, we introduced the concept of an effective temperature of a non-equilibrium system, ω Eq. (2.8). As we will discuss, this concept plays an impor- P tant role in quantum-limited amplifiers; the degenerate paramp gives us a first example of this. Returning to the ωS behaviour of the paramp at the signal frequency, we note that Eq. (5.25) implies that even for vacuum input to both the signal and idler ports, the output will contain a real photon flux. To quantify this in a simple way, it FIG. 8 (Color online) Energy level scheme of the non- is useful to introduce temporal modes which describe the degenerate (phase preserving) parametric oscillator. input and output fields during a particular time interval [j∆t, (j + 1)∆t] (where j is an integer):

Z (j+1)∆t one increases the overall magnitude of the gain at the 1 ˆ BˆS,in,j = √ dτbS,in(τ), (5.29) center frequency ωS (e.g. by increasing the pump ampli- ∆t j∆t tude), one simultaneously narrows the frequency range ˆ ˆ over which the gain is appreciable. Heuristically, this with the temporal modes BS,out,j and BI,in,j being de- is because parametric amplification involves using the fined analogously. These temporal modes are discussed pump energy to decrease the damping and hence increase further in Appendix D.2, where we discuss the windowed the quality factor of the signal mode resonance. This in- Fourier transform (cf. Eq. (D18)). crease in quality factor leads to amplification, but it also With the above definition, we find that the output reduces the bandwidth over whicha ˆS can respond to the mode will have a real occupancy even if the input mode ˆ input signal bS,in. is empty: To deal with a finite signal bandwidth, one simply ˆ† ˆ Fourier transforms Eqs. (5.19). The resulting equations n¯S,out = h0|BS,out,jBS,out,j|0i are easily solved and substituted into Eq. (5.22), result- † = G0h0|Bˆ BˆS,in,j|0i + ing in a frequency-dependent generalization of the input- S,in,j ˆ ˆ† output relation given in Eq. (5.25): (G0 − 1)h0|BI,in,jBI,in,j|0i ˆ ˆ 0 ˆ† = G0 − 1. (5.30) bS,out[ω] = −g[ω]bS,in[ω] − g [ω]bI,in[ω]

(5.26) The dimensionless mode occupancyn ¯S,out is best thought of as a photon flux per unit bandwidth (cf. Eq.(D26)). Here, g[ω] is the frequency-dependent gain of the ampli- This photon flux is equivalent to the photon flux that fier, and g0[ω] satisfies |g0[ω]|2 = |g[ω]|2 − 1. In the rele- would appear in equilibrium at the very high effective vant limit where G = |g[0]|2  1 (i.e. large gain at the 0 temperature (assuming large gain G ) signal frequency), one has to an excellent approximation: 0 √   T ≈ ω G . (5.31) κS−κI eff ~ S 0 G0 − i κ +κ (ω/D) g[ω] = I S , (5.27) 1 − i(ω/D) This is an example of a more general principle, to be discussed in Sec. V.E.4: a high gain amplifier must have with associated with it a large effective temperature scale. Re- 1 κSκI ferring this total output noise back to the input, we have D = √ . (5.28) G0 κS + κI (in the limit G0  1): As always, we work in an where the T ω ω ω eff = ~ S + ~ S = ~ S + T . (5.32) signal frequency has been shifted to zero. D represents G 2 2 2 N the effective operating bandwidth of the amplifier. Com- 0 ponents of the signal with frequencies (in the rotating This corresponds to the half photon of vacuum noise as- frame) |ω|  D are strongly amplified, while components sociated with the signal source, plus the added noise of with frequencies |ω|  D are not amplified at all, but a half photon of our phase preserving amplifier (i.e. the can in fact be slightly attenuated. As we already antici- noise temperature TN is equal to its quantum limited pated, the amplification bandwidth D becomes progres- value). Here, the added noise is simply the vacuum noise sively smaller as the√ pump power and G0 are increased, associated with the idler port. with the product G0D remaining constant. In a para- The above argument is merely suggestive that the out- metric amplifier increasing the gain via increasing the put noise looks like an effective temperature. In fact, pump strength comes with a price: the effective operat- it is possible to show that the photon number distribu- ing bandwidth is reduced. tion of the output is precisely that of a Bose-Einstein 31 distribution at temperature Teff . From Eq. (5.13) we see aout bin that the action of the paramp is to destroy a pump pho- ton and create a pair of new photons, one in the signal amplifier channel and one in the idler channel. Using the SU(1,1) symmetry of the quadratic hamiltonian in Eq. (5.16) it ain bout is possible to show that, for vacuum input, the output input line output line of the paramp is a so-called ‘two-mode squeezed state’ of the form (Caves and Schumaker, 1985; Gerry, 1985; FIG. 9 (Color online) Schematic of a two-port bosonic ampli- Knight and Buzek, 2004) fier. Both the input and outputs of the amplifier are attached to transmission lines. The incoming and outgoing wave am- −1/2 αb† b† |Ψouti = Z e S I |0i (5.33) plitudes in the input (output) transmission line are labelled aˆin, aˆout (ˆbin, ˆbout) respectively. The voltages at the end of where α is a constant related to the gain and, to simplify the two lines (Vˆa, Vˆb) are linear combinations of incoming and the notation, we have dropped the ‘out’ labels on the outgoing wave amplitudes. operators. The normalization constant Z can be worked out by expanding the exponential and using √ † n (bS) |0, 0i = n!|n, 0i (5.34) TABLE IV Two different amplifier modes of operation. to obtain Mode Input Signal Output Signal ∞ s(t) o(t) −1/2 X n |Ψouti = Z α |n, ni (5.35) Scattering s(t) = ain(t) o(t) = bout(t)) n=0 (ain indep. of aout) (bout indep. of bin) and hence Op-amp s(t) = Va(t) o(t) = Vb(t) (ain depends on aout) (bout depends on bin) 1 Z = (5.36) 1 − |α|2 so the state is normalizable only for |α|2 < 1. This appearance of finite entropy in a subsystem even Because this output is obtained by unitary evolution when the full system is in a pure state is a purely quan- from the vacuum input state, the output state is a pure tum effect. Classically the entropy of a composite system state with zero entropy. In light of this, it is interesting is at least as large as the entropy of any of its compo- to consider the reduced density matrix obtain by tracing nents. Entanglement among the components allows this over the idler mode. The pure state density matrix is: lower bound on the entropy to be violated in a quantum system.11 In this case the two-mode squeezed state has ρ = |Ψouti hΨout| strong entanglement between the signal and idler chan- ∞ nels (since their photon numbers are fluctuating identi- X αnα∗m = |n, ni hm, m| (5.37) cally). Z m,n=0 If we now trace over the idler mode we are left with the reduced density matrix for the signal channel D. Scattering versus op-amp modes of operation

∞ X |α|2nS We now begin to address the question of how the stan- ρ˜ = Tr {ρ} = |n i hn | S Idler S Z S dard Haus-Caves derivation of the amplifier quantum nS=0 limit presented in Sec. V.B relates to the general linear 1 −β ω a† a response approach of Sec. IV. Recall that in Sec. III.B.3, ≡ e ~ S S S (5.38) Z we already used this latter approach to discuss position detection with a cavity detector, reaching similar conclu- which is a pure thermal equilibrium distribution with ef- sions (i.e. at best, the detector adds noise equal to the fective Boltzmann factor zero-point noise). In that linear-response-based discus- e−β~ωS = |α|2 < 1. (5.39) sion, we saw that a crucial aspect of the quantum limit was the trade-off between back-action noise and measure- The effective temperature can be obtained from the re- quirement that the signal mode occupancy is G0 − 1

1 11 This paradox has prompted Charles Bennett to remark that a = G0 − 1 (5.40) eβ~ωS − 1 classical house is at least as dirty as its dirtiest room, but a quantum house can be dirty in every room and still perfectly which in the limit of large gain reduces to Eq. (5.31). clean over all. 32

cold the amplifier’s input port, is precisely equal to the signal load to be amplified (i.e. the input signal), irrespective of the amplitude of the wave leaving the input port (i.e. a ). vacuum noise out uin uout Further, the output signal is taken to be the amplitude aout bin of the outgoing wave exiting the output of the amplifier (i.e. b ), again, irrespective of whatever might be en- amplifier out input line output line tering the output port (see Table IV). In this situation, ain bout the Haus-Caves description of the quantum limit in the circulator previous subsection is almost directly applicable; we will make this precise in Sec. VI. Back-action is indeed ir- FIG. 10 Illustration of a bosonic two-port amplifier used in relevant, as the prescribed experimental conditions mean the scattering mode of operation. The “signal” is an incoming that it plays no role. We will call this mode of oper- wave in the input port of the amplifier, and does not depend on what is coming out of the amplifier. This is achieved by ation the “scattering mode”, as it is most relevant to connecting the input line to a circulator and a “cold load” time-dependent experiments where the experimentalist (i.e. a zero temperature resistor): all that goes back towards launches a signal pulse at the input of the amplifier and the source of the input signal is vacuum noise. looks at what exits the output port. One is usually only interested in the scattering mode of operation in cases where the source producing the input signal is matched to the input of the amplifier: only in this case is the input ment imprecision noise. We saw that reaching the quan- wave ain perfectly transmitted into the amplifier. As we tum limit required both a detector with “ideal” noise, as will see in Sec. VI, such a perfect matching requires a rel- well as an optimization of the detector-oscillator coupling atively strong coupling between the signal source and the strength. Somewhat disturbingly, none of these ideas ap- input of the amplifier; as such, the amplifier will strongly peared explicitly in the Haus-Caves derivation; this can enhance the damping of the signal source. give the misleading impression that the quantum limit The second mode of linear amplifier operation is what never has anything to do with back-action. A further we call the “op-amp” mode; this is the mode one usu- confusion comes from the fact that many detectors have ally has in mind when thinking of an amplifier which input and outputs that cannot be described by a set of is weakly coupled to the signal source, and will be the bosonic modes. How does one apply the above arguments focus of the next two subsections. The key difference to such systems? from the “scattering” mode is that here, the input sig- The first step in resolving these seeming inconsisten- nal is not simply the amplitude of a wave incident on cies is to realize that there are really two different ways the input port of the amplifier; similarly, the output sig- in which one can use a given amplifier or detector. In de- nal is not the amplitude of a wave exiting the output ciding how to couple the input signal (i.e. the signal to be port. As such, the Haus-Caves derivation of the quan- amplified) to the amplifier, and in choosing what quan- tum limit does not directly apply. For the bosonic ampli- tity to measure, the experimentalist essentially enforces fier discussed here, the op-amp mode would correspond boundary conditions; as we will now show, there are in to using the amplifier as a voltage op-amp. The input general two distinct ways in which to do this. For con- signal would thus be the voltage at the end of the input creteness, consider the situation depicted in Fig. 9: a two- transmission line. Recall that the voltage at the end of a port voltage amplifier where the input and output ports transmission line involves the amplitude of both left and of the amplifier are attached to one-dimensional trans- right moving waves, i.e. Va(t) ∝ Re [ain(t) + aout(t)]. At mission lines (see App. C for a quick review of quantum first, this might seem quite confusing: if the signal source transmission lines). Similar to the previous subsection, determines Va(t), does this mean it sets the value of both we focus on a narrow bandwidth signal centered about a ain(t) and aout(t)? Doesn’t this violate causality? These frequency ω. At this frequency, there exists both a right- fears are of course unfounded. The signal source enforces moving and a left-moving wave in each transmission line. the value of Va(t) by simply changing ain(t) in response We label the corresponding amplitudes in the input (out- to the value of aout(t). While there is no violation of put) line with ain, aout (bin, bout), as per Fig. 9. Quantum causality, the fact that the signal source is dynamically mechanically, these amplitudes become operators, much responding to what comes out of the amplifier’s input in the same way that we treated the mode amplitude a port implies that back-action is indeed relevant. as an operator in the previous subsection. We will ana- The op-amp mode of operation is relevant to the typ- lyze this two-port bosonic amplifier in detail in Sec. VI; ical situation of “weak coupling” between the signal here, we will only sketch its operation to introduce the source and amplifier input. By “weak coupling”, we mean two different amplifier operation modes. This will then here something stronger than just requiring that the am- allow us to understand the subtleties of the Haus-Caves plifier be linear: we require additionally that the ampli- quantum limit derivation. fier does not appreciably change the dissipation of the sig- In the first kind of setup, the experimentalist arranges nal source. This is analogous to the situation in an ideal things so that ain, the amplitude of the wave incident on voltage op-amp, where the amplifier input impedance is 33

much larger than the impedance of the signal source. We noise of a voltage amplifier (to be discussed in Sec. V.F). stress the op-amp mode and this limit of weak coupling is the relevant situation in most electrical measurements. Thus, we see that the Haus-Caves formulation of the 1. Detector back-action quantum limit is not directly relevant to amplifiers or de- tectors operated in the usual op-amp mode of operation. We first consider the consequence of noise in the detec- We clearly need some other way to describe quantum tor input port. As we have already seen in Sec. II.B, the ˆ amplifiers used in this regime. As we will demonstrate in fluctuating back-action force F acting on our oscillator the remainder of this section, the general linear-response will lead to both damping and heating of the oscillator. approach of Sec. IV is exactly what is needed. To see To model the intrinsic (i.e. detector-independent) heating this, we will have to expand the discussion of Sec. IV to and damping of the oscillator, we will also assume that include the concepts of input and output impedance, as our oscillator is coupled to an equilibrium heat bath. In well as power gain. The linear response approach will the weak-coupling limit that we are interested in, one allow us to see (similar to Sec. III.B.3) that reaching the can use lowest-order perturbation theory in the coupling ˆ quantum limit in the op-amp mode does indeed require a A to describe the effects of the back-action force F on trade-off between back-action and measurement impreci- the oscillator. A full quantum treatment (see Appendix sion, and requires use of an amplifier with ideal quantum I.4) shows that the oscillator is described by an effective 14 noise properties (cf. Eq. 4.11). This approach also has classical Langevin equation: the added benefit of being directly applicable to systems 2 Mx¨(t) = −MΩ x(t) − Mγ0x˙(t) + F0(t) where the input and output of the amplifier are not de- Z scribed by bosonic modes.12 In the next section (Sec. VI), −MA2 dt0γ(t − t0)x ˙(t0) − A · F (t) (5.41) we will return to the scattering description of a two-port voltage amplifier (Fig. 10), and show explicitly how an The position x(t) in the above equation is not an op- amplifier can be quantum-limited when used in the scat- erator, but is simply a classical variable whose fluctua- tering mode of operation, but miss the quantum limit tions are driven by the fluctuating forces F (t) and F0(t). when used in the op-amp mode of operation. Nonetheless, the noise in x calculated from Eq. (5.41) corresponds precisely to S¯xx[ω], the symmetrized quan- tum mechanical noise in the operatorx ˆ. The fluctuating E. Linear response description of a position detector force exerted by the detector (which represents the heat- ing part of the back-action) is described by A · F (t) in In this subsection, we will examine the amplifier quan- Eq. (5.41); it has zero mean, and a spectral density given 2 tum limit for a two-port linear amplifier in the usual weak by A S¯FF [ω] in Eq. (4.4a). The kernel γ(t) describes coupling, “op-amp” regime of operation. Our discussion the damping effect of the detector. It is given by the here will make use of the results we have obtained for the asymmetric part of the detector’s quantum noise, as was noise properties of a generic linear response detector in derived in Sec. II.B (cf. Eq. (2.12)). Sec. IV, including the fundamental quantum noise con- Eq. (5.41) also describes the effects of an equilibrium straint of Eq. (4.11). For simplicity, we will start with heat bath at temperature T0 which models the intrin- the concrete problem of continuous position detection of sic (i.e. detector-independent) damping and heating of a harmonic oscillator. Our discussion will thus generalize the oscillator. The parameter γ0 is the damping aris- the discussion of position detection using a cavity detec- ing from this bath, and F0 is the corresponding fluc- tor given in Sec. III.B.3. We start with a generic detector tuating force. The spectral density of the F0 noise is (as introduced in Sec. IV.A) coupled at its input to the determined by γ0 and T0 via the fluctuation-dissipation 13 positionx ˆ of a harmonic oscillator (cf. Eq. (4.1)). We theorem (cf. Eq. (2.16)). T0 and γ0 have a simple phys- would like to understand the total output noise of our ical significance: they are the temperature and damping amplifier in the presence of the oscillator, and, more im- of the oscillator when the coupling to the detector A is portantly, how small we can make the amplifier’s contri- set to zero. bution to this noise. The resulting lower bound is known To make further progress, we recall from Sec. II.B that as the standard quantum limit (SQL) on position detec- even though our detector will in general not be in equi- tion, and is analogous to the quantum limit on the added librium, we may nonetheless assign it an effective tem- perature Teff [ω] at each frequency (cf. Eq. (2.8)). The

12 Note that the Haus-Caves derivation for the quantum limit of a scattering amplifier has recently been generalized to the case of 14 Note that we have omitted a back-action term in this equation fermionic operators (Gavish et al., 2004). which leads to small renormalizations of the oscillator frequency 13 For consistency with previous sections, our coupling Hamiltonian and mass. These terms are not important for the following dis- does not have a minus sign. This is different from the conven- cussion, so we have omitted them for clarity; one can consider tion of Clerk (2004), where the coupling Hamiltonian is written M and Ω in this equation to be renormalized quantities. See Hint = −Axˆ · Fˆ. Appendix I.4 for more details. 34 effective temperature of an out-of-equilibrium detector is simply a measure of the asymmetry of the detector’s g Aλ Bgy quantum noise. We are often interested in the limit where the internal detector timescales are much faster than the −1 −1 −1 timescales relevant to the oscillator (i.e. Ω , γ , γ0 ). We may then take the ω → 0 limit in the expression for FD x F I y Teff , yielding: A B S¯ (0) γ + γ γ + γ 2k T ≡ FF (5.42) 0 Ag out ld B eff Mγ(0) In this limit, the oscillator position noise calculated from Eq. (5.41) is given by: FIG. 11 (Color online) Schematic of a generic linear-response position detector, where an auxiliary oscillator y is driven by ¯ 1 2(γ0 + γ)kB γ0T + γTeff the detector output. Sxx[ω] = 2 2 2 2 2 M (ω − Ω ) + ω (γ + γ0) γ0 + γ (5.43)

This is exactly what would be expected if the oscillator Here, SII , SFF and SIF are the (classical) detector noise were only attached to an equilibrium Ohmic bath with a correlators calculated in the absence of any coupling to ¯ damping coefficient γΣ = γ0 + γ and temperature T = the oscillator. Note importantly that we have included (γ0T + γTeff )/γΣ. the fact that the two kinds of detector noise (in Iˆ and in Fˆ) may be correlated with one another. To apply the classically-derived Eq. (5.46) to our 2. Total output noise quantum detector-plus-oscillator system, we recall from Sec. II.B that symmetrized quantum noise spectral densi- The next step in our analysis is to link fluctuations ties play the role of classical noise. The LHS of Eq. (5.46) in the position of the oscillator (as determined from thus becomes S , the total symmetrized quantum- Eq. (5.41)) to noise in the output of the detector. As II,tot mechanical output noise of the detector, while the RHS discussed in Sec. III.B.3, the output noise consists of the will now contain the symmetrized quantum-mechanical intrinsic output noise of the detector (i.e. “measurement detector noise correlators S¯ , S¯ and S¯ , defined as in imprecision noise”) plus the amplified position fluctua- FF II IF Eq. (4.4a). Though this may seem rather ad-hoc, one can tions in the position of the oscillator. The latter contains easily demonstrate that Eq. (5.46) thus interpreted would both an intrinsic part and a term due to the response of be quantum-mechanically rigorous if the detector corre- the oscillator to the backaction. lation functions obeyed Wick’s theorem. Thus, quantum To start, imagine that we can treat both the oscillator corrections to Eq. (5.46) will arise solely from the non- position x(t) and the detector output I(t) as classically Gaussian nature of the detector noise correlators. We fluctuating quantities. Using the linearity of the detec- expect from the central limit theorem that such correc- tor’s response, we can then write δI , the fluctuating total tions will be small in the relevant limit where ω is much part of the detector’s output, as: smaller that the typical detector frequency ∼ kBTeff /~, δItotal[ω] = δI0[ω] + AχIF [ω] · δx[ω] (5.44) and neglect these corrections in what follows. Note that the validity of Eq. (5.46) for a specific model of a tunnel The first term (δI0) describes the intrinsic (oscillator- independent) fluctuations in the detector output, and junction position detector has been explicitly verified in 2 Clerk and Girvin (2004). has a spectral density S¯II [ω]. If we scale this by |χIF | , we have the measurement imprecision noise discussed in Sec. III.B.3. The second term corresponds to the ampli- 3. Detector power gain fied fluctuations of the oscillator, which are in turn given by solving Eq. (5.41): Before proceeding, we need to consider our detector  1/M  once again in isolation, and return to the fundamental δx[ω] = − (F [ω] − A · F [ω]) (ω2 − Ω2) + iωΩ/Q[ω] 0 question of what we mean by amplification. To be able to say that our detector truly amplifies the motion of the ≡ χxx[ω](F0[ω] − A · F [ω]) (5.45) oscillator, it is not sufficient to simply say the response where Q[ω] = Ω/(γ0 + γ[ω]) is the oscillator quality fac- function χIF must be large (note that χIF is not dimen- tor. It follows that the spectral density of the total noise sionless!). Instead, true amplification requires that the in the detector output is given classically by: power delivered by the detector to a following amplifier be much larger than the power drawn by the detector SII,tot[ω] = SII [ω] 2 4 2  at its input– i.e., the detector must have a dimension- + |χxx[ω]χIF [ω]| A SFF [ω] + A SF F [ω] 0 0 less power gain GP [ω] much larger than one. As we have 2 + 2A Re [χxx[ω]χIF [ω]SIF [ω]] (5.46) already discussed, if the power gain was not large, we 35 would need to worry about the next stage in the ampli- As the oscillator y is being driven on resonance, the rela- fication of our signal, and how much noise is added in tion between y and I is given by y[ω] = χyy[ω]I[ω] with −1 that process. Having a large power gain means that by χyy[ω] = −i[ωMγout[ω]] . From conservation of energy, the time our signal reaches the following amplifier, it is we have that the net power flow into the output oscillator so large that the added noise of this following amplifier from the detector is equal to the power dissipated out of is unimportant. The power gain is analogous to the di- the oscillator through the intrinsic damping γld. We thus mensionless photon number gain G that appears in the have: standard Haus-Caves description of a bosonic linear am- 2 plifier (cf. Eq. (5.7)). Pout ≡ Mγld · y˙ 2 2 2 To make the above more precise, we start with the idea = Mγldω |χyy[ω]| · |BAχIF χxx[ω]FD| case of no reverse gain, χFI = 0. We will define the power 1 γld 2 gain G [ω] of our generic position detector in a way that = · |BAχIF χxx[ω]FD| P M 2 is analogous to the power gain of a voltage amplifier. (γld + γout[ω]) Imagine we drive the oscillator we are trying to measure (5.50) with a force 2FD cos ωt; this will cause the output of our detector hIˆ(t)i to also oscillate at frequency ω. To Using the above definitions, we find that the ratio be- optimally detect this signal in the detector output, we tween Pout and Pin is independent of γ0, but depends on further couple the detector output I to a second oscillator γld: with natural frequency ω, mass M, and position y: there 2 2 2 0 ˆ Pout 1 A B |χIF [ω]| γld/γout[ω] is a new coupling term in our Hamiltonian, Hint = BI ·yˆ, = 2 2 2 where B is a coupling strength. The oscillations in hI(t)i Pin M ω γout[ω]γ[ω] (1 + γld/γout[ω]) will now act as a driving force on the auxiliary oscillator (5.51) y (see Fig 11). We can consider the auxiliary oscillator y as a “load” we are trying to drive with the output of our We now define the detector power gain GP [ω] as the value detector. of this ratio maximized over the choice of γld . The max- To find the power gain, we need to consider both Pout, imum occurs for γld = γout[ω] (i.e. the load oscillator is the power supplied to the output oscillator y from the de- “matched” to the output of the detector), resulting in: tector, and P , the power fed into the input of the ampli- in   fier. Consider first Pin. This is simply the time-averaged Pout GP [ω] ≡ max power dissipation of the input oscillator x caused by the Pin back-action damping γ[ω]. Using a bar to denote a time 2 2 2 1 A B |χIF | average, we have: = 2 2 4M ω γoutγ 2 2 2 2 P ≡ Mγ[ω] · x˙ 2 = Mγ[ω]ω |χ [ω]| F (5.47) |χIF [ω]| in xx D = (5.52) 4Im χFF [ω] · Im χII [ω] Note that the oscillator susceptibility χxx[ω] depends on both the back-action damping γ[ω] as well as the intrinsic In the last line, we have used the relation between the oscillator damping γ0 (cf. Eq. (5.45)). damping rates γ[ω] and γout[ω] and the linear-response Next, we need to consider the power supplied to the susceptibilities χFF [ω] and χII [ω] (cf. Eqs. (2.15) and “load” oscillator y at the detector output. This oscilla- (5.48)). We thus find that the power gain is a simple di- tor will have some intrinsic, detector-independent damp- mensionless ratio formed by the three different response ing γld, as well as a back-action damping γout. In the coefficients characterizing the detector, and is indepen- same way that the back-action damping γ of the input dent of the coupling constants A and B. As we will see oscillator x is determined by the quantum noise in Fˆ in subsection V.F, it is completely analogous to the power (cf. Eq. (2.14)-(2.12)), the back-action damping of the gain of a voltage amplifier, which is also determined by load oscillator y is determined by the quantum noise in three parameters: the voltage gain, the input impedance the output operator Iˆ: and the output impedance. Note that there are other important measures of power gain commonly in use in B2 γ [ω] = [−Im χ [ω]] the engineering community: we will comment on these out Mω II in Sec. VII.B. B2 S [ω] − S [−ω] Finally, the above results are easily generalized to = II II (5.48) M~ω 2 the case where the detector’s reverse gain χFI is non- vanishing. For simplicity, we present results for the case 2 where χII is the linear-response susceptibility which de- where β = Re (χIF χFI ) /|χIF | ≥ 0, implying that there termines how hIˆi responds to a perturbation coupling to is no positive feedback. Maximizing the ratio of Pout/Pin Iˆ: over choices of γld now yields Z ∞ i Dh iE iωt 2GP χII [ω] = − dt Iˆ(t), Iˆ(0) e (5.49) GP,rev = √ ≤ 1/β. (5.53) ~ 0 1 + 2βGP + 1 + 4βGP 36

Here, GP,rev is the power gain in the presence of reverse (cf. Eq. (I13)), one finds: gain, while GP is the zero-reverse gain power gain given   2 2 ~ω 2 by Eq. (5.52). One can confirm that GP,rev is a mono- (Im α) coth 2k T + (Re α) G [ω] = B eff (5.56) tonic increasing function of GP , and is bounded by 1/β. P 2 ∗ |α| As we have previously noted in Sec.IV.A.4, if χFI = χIF , there is no additional quantum noise constraint on our where α[ω] is the parameter characterizing a quantum detector beyond what exists classically (i.e. the RHS of limited detector in Eq. (4.18); recall that |α[ω]|2 deter- Eq. (4.11) vanishes). We now see explicitly that when mines the ratio of SII and SFF . It follows immediately ∗ that for a detector with ‘ideal noise’ to also have a large χFI = χIF , the power gain of our detector can be at most 1, as β = 1. Thus, while there is no minimum power gain (GP  1), one absolutely needs kBTeff  ~ω: back-action noise required by quantum mechanics in this a large power gain implies a large effective detector tem- case, there is also no amplification: at best, our detec- perature. In the large GP limit, we have tor would act as a transducer. Note further that if the  2 ∗ Im α kBTeff detector has χFI = χ and optimizes the inequality of G ' (5.57) IF P |α| ω/2 Eq. (4.11), then one can show GP,rev must be one (cf. Ap- ~ pendix I.2): the detector is simply a transducer. This Thus, the effective temperature of a quantum-ideal detec- is in keeping with the results obtained using the Haus- tor does more than just characterize the detector back- Caves approach, which also yields the conclusion that a action– it also determines the power gain. noiseless detector is a transducer. Finally, an additional consequence of the large GP [ω], large Teff limit is that the gain χIF and noise cross- correlator S¯IF are in phase: S¯IF /χIF is purely real, up to 4. Simplifications for a quantum-ideal detector corrections which are small as ω/Teff . This is shown ex- plicitly in Appendix I.3. Thus, we find that a large power We now consider the important case where our detec- gain detector with ideal quantum noise cannot have sig- tor has no reverse gain (allowing it to have a large power nificant out-of-phase correlations between its output and gain), and also has “ideal” quantum noise (i.e. it satisfies input noises. This last point may be understood in terms the ideal noise condition of Eq. (4.17)). Fulfilling this of the idea of wasted information: if there were signifi- ˆ ˆ condition immediately places some powerful constraints cant out-of-phase correlations between I and F , it would on our detector. be possible to improve the performance of the amplifier by using feedback. We will discuss this point more fully First, note that we have defined in Eq. (5.42) the effec- ¯ tive temperature of our detector based on what happens in Sec. VI. Note that as SIF /χIF is real, the last term at the input port; this is the effective temperature seen in the quantum noise constraint of Eq. (4.11) vanishes. by the oscillator we are trying to measure. We could also consider the effective temperature of the detector as seen 5. Quantum limit on added noise and noise temperature at the output (i.e. by the oscillator y used in defining the power gain). This “output” effective temperature is We now turn to calculating the noise added to our determined by the quantum noise in the output operator ˆ signal (i.e. hxˆ(t)i) by our generic position detector. To I: characterize this added noise, it is useful to take the total (symmetrized) noise in the output of the detector, and ~ω kBTeff,out[ω] ≡ (5.54) refer it back to the input by dividing out the gain of the log (S [+ω]/S [−ω]) II II detector: ¯ For a general out-of-equilibrium amplifier, Teff,out does ¯ SII,tot[ω] Sxx,tot[ω] ≡ 2 2 (5.58) not have to be equal to the input effective temperature A |χIF [ω]| Teff defined by Eq. (2.8). However, for a quantum-ideal ¯ detector, the effective proportionality between input and Sxx,tot[ω] is simply the frequency-dependent spectral output operators (cf. Eq. (I13)) immediately yields: density of position fluctuations inferred from the output of the detector. It is this quantity which will directly determine the sensitivity of the detector– given a certain T [ω] = T [ω] (5.55) eff,out eff detection bandwidth, what is the smallest variation of x that can be resolved? The quantity S¯ [ω] will have Thus, a detector with quantum-ideal noise necessarily xx,tot contributions both from the intrinsic fluctuations of the has the same effective temperature at its input and its input signal, as well as a contribution due to the detector. output. This is all the more remarkable given that a We first define S¯ [ω, T ] to be the symmetrized equi- quantum-ideal detector cannot be in equilibrium, and xx,eq librium position noise of our damped oscillator (whose thus T cannot represent a real physical temperature. eff damping is γ + γ) at temperature T : Another important simplification for a quantum-ideal 0   detector is the expression for the power gain. Using ¯ ~ω Sxx,eq[ω, T ] = ~ coth [−Im χxx[ω]] (5.59) the proportionality between input and output operators 2kBT 37 where the oscillator susceptibility χxx[ω] is defined in terms in Eq. (5.61)). We would thus expect S¯xx,add[ω] to Eq. (5.45). The total inferred position noise may then attain a minimum value at an optimal choice of coupling be written: A = Aopt where both these terms make equal contribu- tions (see Fig. 5). Defining φ[ω] = arg χ [ω], we thus  γ  xx S¯ [ω] ≡ 0 · S¯ [ω, T ] + S¯ [ω] (5.60) have the bound: xx,tot γ + γ xx,eq 0 xx,add 0 "s # S¯ S¯ Re χ∗ e−iφ[ω]S¯  In the usual case where the detector noise can be approx- ¯ II FF IF IF Sxx,add[ω] ≥ 2|χxx[ω]| 2 + 2 imated as being white, this spectral density will consist of |χIF | |χIF | a Lorentzian sitting atop a constant noise floor (cf. Fig. 6) (5.63) The first term in Eq. (5.60) represents position noise aris- where the minimum value at frequency ω is achieved ing from the fluctuating force δF0(t) associated with the when: intrinsic (detector-independent) dissipation of the oscil- s S¯ [ω] lator (cf. Eq. (5.41)). The prefactor of this term arises A2 = II (5.64) opt 2 ¯ because the strength of the intrinsic Langevin force act- |χIF [ω]χxx[ω]| SFF [ω] ing on the oscillator is proportional to γ0, not to γ0 + γ. 2 2 The second term in Eq. (5.60) represents the added Using the inequality X + Y ≥ 2|XY | we see that this ¯ position noise due to the detector. It has contributions value serves as a lower bound on Sxx,add even in the pres- from both from the detector’s intrinsic output noise S¯II ence of detector-dependent damping. In the case where as well as from the detector’s back-action noise S¯FF , and the detector-dependent damping is negligible, the RHS of may be written: Eq. (5.63) is independent of A, and thus Eq. (5.64) can be satisfied by simply tuning the coupling strength A; in ¯ ¯ SII 2 2 ¯ the more general case where there is detector-dependent Sxx,add[ω] = 2 2 + A |χxx| SFF + |χIF | A damping, the RHS is also a function of A (through the re- ∗ sponse function χ [ω]), and it may no longer be possible 2Re χ∗ (χ ) S¯  xx IF xx IF to achieve Eq. (5.64) by simply tuning A.15 2 (5.61) |χIF | While Eq. (5.63) is certainly a bound on the added ¯ For clarity, we have omitted writing the explicit fre- displacement noise Sxx,add[ω], it does not in itself rep- resent the quantum limit. Reaching the quantum limit quency dependence of the gain χIF , susceptibility χxx, and noise correlators; they should all be evaluated at the requires more than simply balancing the detector back- frequency ω. Note that the first term on the RHS cor- action and intrinsic output noises (i.e. the first two terms responds to the “measurement imprecision” noise of our in Eq. (5.61)); one also needs a detector with “quantum- detector, S¯I (ω). ideal” noise properties, that is a detector which satis- xx fies Eq. (4.17). Using the quantum noise constraint of We can now finally address the quantum limit on the ¯ added noise in this setup. As discussed in Sec. V.D, Eq. (4.11) to further bound Sxx,add[ω], we obtain: the Haus-Caves derivation of the quantum limit (cf. ¯ χxx[ω] Sec. V.B) is not directly applicable to the position de- Sxx,add[ω] ≥ 2 × χ tector we are describing here; nonetheless, we may use IF s its result to guess what form the quantum limit will take "  2   ¯  ~|χIF | 2SIF ¯ 2 1 + ∆ + SIF here. The Haus-Caves argument told us that the added 2 χ noise of a phase-preserving linear amplifier must be at ~ IF least as large as the zero point noise. We thus anticipate  ∗ −iφ[ω] ¯ # Re χIF e SIF that if our detector has a large power gain, the spectral + (5.65) |χIF | density of the noise added by the detector (i.e. S¯xx,add[ω]) must be at least as big as the zero point noise of our where the function ∆[z] is defined in Eq. (4.12). The damped oscillator: minimum value of S¯xx,add[ω] in Eq. (5.65) is now achieved ¯ ¯ when one has both an optimal coupling (i.e. Eq. (5.64)) Sxx,add[ω] ≥ lim Sxx,eq[w, T ] = |~Im χxx[ω]| (5.62) T →0 and a quantum limited detector, that is one which satis- fies Eq. (4.11) as an equality. We will now show that the bound above is rigorously Next, we consider the relevant case where our detector correct at each frequency ω. is a good amplifier and has a power gain G [ω]  1 over The first step is to examine the dependence of the P ¯ the width of the oscillator resonance. As we have dis- added noise Sxx,add[ω] (as given by Eq. (5.61)) on the cussed, this implies that the ratio S¯ /χ is purely real, coupling strength A. If we ignore for a moment the IF IF detector-dependent damping of the oscillator, the sit- uation is the same as the cavity position detector of Sec. III.B.3: there is an optimal value of the coupling 15 Note that in the heuristic discussion of position detection using strength A which corresponds to a trade-off between a resonant cavity detector in Sec. III.B.3, these concerns did not imprecision noise and back-action (i.e. first and second arise as there was no back-action damping. 38

¯ up to small ~ω/(kBTeff ) corrections (see Sec. (IV.A.4) imaginary, Eq. (5.67) tells us that SIF should be zero. and Appendix I.3 for more details). This in turn implies Assuming we have a quantum-limited detector with a ¯ that ∆[2SIF /~χIF ] = 0; we thus have: large power gain (kBTeff  ~Ω), the remaining condition on the coupling A (Eq. (5.64)) may be written as: S¯xx,add[ω] ≥ s  γ[A ] Im α 1 Ω  2  ¯ 2 ¯ opt ~ ~ SIF cos [φ[ω]] SIF = p = (5.69) 2|χxx[ω]|  + + (5.66) γ0 + γ[Aopt] α 2 GP [Ω] 4kBTeff 2 χIF χIF As γ[A] ∝ A2 is the detector-dependent damping of ¯ Finally, as there is no further constraint on SIF /χIF (be- the oscillator, we thus have that to achieve the quantum- yond the fact that it is real), we can minimize the expres- limited value of S¯ [Ω] with a large power gain, ¯ xx,add sion over its value. The minimum Sxx,add[ω] is achieved one needs the intrinsic damping of the oscillator to be for a detector whose cross-correlator satisfies: much larger than the detector-dependent damping. The detector-dependent damping must be small enough to S¯ [ω] IF = −~ cot φ[ω], (5.67) compensate the large effective temperature of the detec- χIF optimal 2 tor; if the bath temperature satisfies ~Ω/kB  Tbath  with the minimum value being given by: Teff , Eq. (5.69) implies that at the quantum limit, the temperature of the oscillator will be given by: ¯ ¯ Sxx,add[ω] = ~|Im χxx[ω]| = lim Sxx,eq[ω, T ] min T →0 γ · Teff + γ0 · Tbath ~Ω (5.68) Tosc ≡ → + Tbath (5.70) γ + γ0 4kB where S¯xx,eq[ω, T ] is the equilibrium contribution to ¯ Sxx,tot[ω] defined in Eq. (5.59). Thus, in the limit of Thus, at the quantum limit and for large Teff , the detec- 16 a large power gain, we have that at each frequency, the tor raises the oscillator’s temperature by ~Ω/4kB. As minimum displacement noise added by the detector is pre- expected, this additional heating is only half the zero- cisely equal to the noise arising from a zero temperature point energy; in contrast, the quantum-limited value of bath. This conclusion is irrespective of the strength of S¯xx,add[ω] corresponds to the full zero-point result, as the intrinsic (detector-independent) oscillator damping. it also includes the contribution of the intrinsic output We have thus derived the amplifier quantum limit (in noise of the detector. the context of position detection) for a two-port ampli- Finally, we return to Eq. (5.65); this is the constraint fier used in the “op-amp” mode of operation. Though we on the added noise S¯xx,add[ω] before we assumed our de- reached a conclusion similar to that given by the Haus- tector to have a large power gain, and consequently a Caves approach, the linear-response, quantum noise ap- large Teff . Note crucially that if we did not require a proach used is quite different. This approach makes ex- large power gain, then there need not be any added noise. plicitly clear what is needed to reach the quantum limit. Without the assumption of a large power gain, the ratio We find that to reach the quantum-limit on the added S¯IF /χIF can be made imaginary with a large magni- ¯ displacement noise Sxx,add[ω] with a large power gain, tude. In this limit, 1 + ∆[2S¯IF /χIF ] → 0: the quan- one needs: tum constraint on the amplifier noises (e.g. the RHS of Eq. (4.11)) vanishes. One can then easily use Eq. (5.65) 1. A quantum limited detector, that is a detec- to show that the added noise S¯ [ω] can be zero. This tor which satisfies the “ideal noise” condition of xx,add confirms a general conclusion that we have seen several Eq. (4.17), and hence the proportionality condition times now (cf. Secs. IV.A.4, V.B): if a detector does not of Eq. (I13). amplify (i.e. the power gain is unity), it need not produce 2. A coupling A which satisfies Eq. (5.64). any added noise.

3. A detector cross-correlator S¯IF which satisfies Eq. (5.67). F. Quantum limit on the noise temperature of a voltage amplifier Recall that condition (1) is identical to what is required for quantum-limited detection of a qubit; it is rather de- We now turn attention to the quantum limit on the manding, and requires that there is no “wasted” informa- added noise of a generic linear voltage amplifier used in tion about the input signal in the detector which is not the “op-amp” mode of operation (see, e.g. , Devoret and revealed in the output (Clerk et al., 2003). Also note that cot φ changes quickly as a function of frequency across the oscillator resonance, whereas S¯IF will be roughly constant; condition (2) thus implies that it will not be 16 If in contrast our oscillator was initially at zero temperature (i.e. possible to achieve a minimal S¯xx,add[ω] across the en- Tbath = 0), one finds that the effect of the back-action (at the tire oscillator resonance. A more reasonable goal is to quantum limit and for G  1) is to heat the oscillator to a ¯ P optimize Sxx,add[ω] at resonance, ω = Ω. As χxx[Ω] is temperature ~Ω/(kB ln 5). 39

~ at the amplifier input, and is amplified at the output; as V λV such, it is completely analogous to the intrinsic detector Zin vin Zout output noise SII of our linear response detector. In con- input output trast, the current noise source of the voltage amplifier ~ λI IL Vout ZS I represents back-action: this fluctuating current (spectral ZL density SI˜I˜[ω]) flows back across the parallel combination of the source impedance and amplifier input impedance, FIG. 12 Schematic of a linear voltage amplifier, including a producing an additional fluctuating voltage at its input. reverse gain λ . V˜ and I˜ represent the standard voltage and I The current noise is thus analogous to the back-action current noises of the amplifier, as discussed further in the text. The case with reverse gain is discussed in detail in Sec. VI. noise SFF of our generic linear response detector. Putting the above together, the total voltage at the input terminals of the amplifier is:

Zin h i ZsZin Schoelkopf (2000)). For such amplifiers, the added noise vin,tot(t) = vin(t) + V˜ (t) − I˜(t) is usually expressed in terms of the “noise temperature” Zin + Zs Zs + Zin of the amplifier; we will define this concept, and demon- ' vin(t) + V˜ (t) − ZsI˜(t) (5.71) strate that, when appropriately defined, this noise tem- perature must be bigger than ~ω/(2kB), where ω is the In the second line, we have taken the usual limit of signal frequency. Though the voltage amplifier is closely an ideal voltage amplifier which has an infinite input analogous to the position detector treated in the previ- impedance (i.e. the amplifier draws zero current). The ous section, its importance makes it worthy of a separate spectral density of the total input voltage fluctuations is discussion. Similar to the last subsection, our discussion thus: here will use the general linear-response approach. In contrast, in Sec. VI, we will present the bosonic scatter- SVV,tot[ω] = Svinvin [ω] + SVV,add[ω]. (5.72) ing description of a two-port voltage amplifier, a descrip- Here S is the spectral density of the voltage fluc- tion similar to that used in formulating the Haus-Caves vinvin proof of the amplifier quantum limit. We will then be tuations of the input signal vin(t), and SVV,add is the in a good position to contrast the linear-response and amplifier’s contribution to the total noise at the input: scattering approaches, and will clearly see there that the 2 S [ω] = S + |Z | S − 2Re [Z∗S ] scattering mode of operation and the “op-amp” mode of VV,add V˜ V˜ s I˜I˜ s V˜ I˜ operation discussed here are not equivalent. We stress (5.73) that the general treatment presented here can also be applied directly to the system discussed in Sec. VI. For clarity, we have dropped the frequency index for the spectral densities appearing on the RHS of this equation. It is useful to now consider a narrow bandwidth input signal at a frequency ω, and ask the following question: if 1. Classical description of a voltage amplifier the signal source was simply an equilibrium resistor at a temperature T , how much hotter would it have to be to Let us begin by recalling the standard schematic de- 0 produce a voltage noise equal to S [ω]? The result- scription of a voltage amplifier, see Fig. 12. The in- VV,tot ing increase in the source temperature is defined as the put voltage to be amplified v (t) is produced by a cir- in noise temperature T [ω] of the amplifier and is a conve- cuit which has a Thevenin-equivalent impedance Z , the N s nient measure of the amplifier’s added noise. It is stan- source impedance. We stress that we are considering the dard among engineers to define the noise temperature as- “op-amp” mode of amplifier operation, and thus the in- suming the initial temperature of the resistor T  ω. put signal does not correspond to the amplitude of a 0 ~ One may then use the classical expression for the thermal wave incident upon the amplifier (see Sec. V.D). The noise of a resistor, which yields the definition: amplifier itself has an input impedance Zin and an out- put impedance Zout, as well as a voltage gain coefficient 2Re Zs · kBTN[ω] ≡ SVV,tot[ω] (5.74) λV : assuming no current is drawn at the output (i.e. Z → ∞ in Fig. 12), the output voltage V (t) is sim- iφ load out Writing Zs = |Zs|e , we have: ply λV times the voltage across the input terminals of the amplifier.   1 SV˜ V˜ −iφ  The added noise of the amplifier is standardly repre- 2kBTN = + |Zs|SI˜I˜ − 2Re e SV˜ I˜ cos φ |Zs| sented by two noise sources placed at the amplifier input. (5.75) ˜ There is both a voltage noise source V (t) in series with It is clear from this expression that TN will have a mini- ˜ the input voltage source, and a current noise source I(t) mum as a function of |Zs|. For |Zs| too large, the back- in parallel with input voltage source (Fig. 12). The volt- action current noise of the amplifier will dominate TN, age noise produces a fluctuating voltage V˜ (t) (spectral while for |Zs| too small, the voltage noise of the ampli- density SV˜ V˜ [ω]) which simply adds to the signal voltage fier (i.e. its intrinsic output noise will dominate). The 40

situation is completely analogous to the position detec- in the classical description. First, as the operator Qˆ rep- tor of the last section; there, we needed to optimize the resents a back-action force, its fluctuations correspond to coupling strength A to balance back-action and intrinsic the amplifier’s current noise I˜(t): output noise contributions, and thus minimize the to- 2 ¯ tal added noise. Optimizing the source impedance thus SI˜I˜[ω] ↔ ω SQQ[ω]. (5.80) yields a completely classical minimum bound on TN: Similarly, the fluctuations in the operator Vˆ , when re- q out 2 ferred back to the amplifier input, will correspond to the kBTN ≥ S ˜ ˜ S ˜˜ − [Im S ˜ ˜] − Re S ˜ ˜ (5.76) V V II V I V I voltage noise V˜ (t) discussed above: where the minimum is achieved for an optimal source S¯ [ω] impedance which satisfies: VoutVout SV˜ V˜ [ω] ↔ 2 (5.81) |λV | s SV˜ V˜ [ω] |Zs[ω]|opt = ≡ ZN (5.77) A similar correspondence holds for the cross-correlator of SI˜I˜[ω] these noise sources:

Im SV˜ I˜[ω] ¯ sin φ[ω] = − (5.78) SV Q[ω] opt p out S ˜ ˜ [ω]S ˜˜[ω] SV˜ I˜[ω] ↔ +iω (5.82) V V II λV The above equations define the so-called noise impedance To proceed, we need to identify the input and output ZN. We stress again that the discussion so far in this impedances of the amplifier, and then define its power subsection has been completely classical. gain. The first step in this direction is to assume that the output of the amplifier (Vˆout) is connected to an external circuit via a term: 2. Linear response description ˆ 0 ˆ Hint = qout(t) · Vout (5.83) It is easy to connect the classical description of a volt-

age amplifier to the quantum mechanical description of a where ˜iout = dqout/dt is the current in the external generic linear response detector; in fact, all that is needed circuit. We may now identify the input and output is a “relabeling” of the concepts and quantities we in- impedances of the amplifier in terms of the damping at troduced in Sec. V.E when discussing a linear position the input and output. Using the Kubo formulae for detector. Thus, the quantum voltage amplifier will be conductance and resistance yields (cf. Eq. (2.12) and ˆ characterized by both an input operator Q and an out- Eq. (5.48), with the substitutions Fˆ → Qˆ and Iˆ → Vˆ ): put operator Vˆout; these play the role, respectively, of ˆ ˆ ˆ F and I in the position detector. Vout represents the 1/Zin[ω] = iωχQQ[ω] (5.84) output voltage of the amplifier, while Qˆ is the operator χVV [ω] which couples to the input signal vin(t) via a coupling Zout[ω] = (5.85) Hamiltonian: −iω 1 i.e. hI˜iniω = vin[ω] and hV iω = Zout[ω]˜iout[ω], Hˆint = vin(t) · Qˆ (5.79) Zin[ω] where the subscript ω indicates the Fourier transform of a time-dependent expectation value. In more familiar terms, I˜ˆ − dQ/dtˆ represents the cur- in We will consider throughout this section the case of no rent flowing into the amplifier.17 The voltage gain of our reverse gain, χ = 0. We can define the power gain amplifier λ will again be given by the Kubo formula of QVout V G exactly as we did in Sec. V.E.3 for a linear position Eq. (4.3), with the substitutions Fˆ → Q,ˆ Iˆ → Vˆ (we P out detector. G is defined as the ratio of the power delivered will assume these substitutions throughout this section). P to a load attached to the amplifier output divided by We can now easily relate the fluctuations of the in- the power drawn by the amplifier, maximized over the put and output operators to the noise sources used to impedance of the load. One finds: describe the classical voltage amplifier. As usual, sym- metrized quantum noise spectral densities S¯[ω] will play 2 |λV | the role of the classical spectral densities S[ω] appearing GP = (5.86) 4Re (Zout)Re (1/Zin) Expressing this in terms of the linear response coefficients χVV and χQQ, we obtain an expression which is com- 17 Note that one could have instead written the coupling Hamil- ˆ pletely analogous to Eq. (5.52) for the power gain for a tonian in the more traditional form Hˆint(t) = φ(t) · I˜in, where R 0 0 position detector: φ = dt vin(t ) is the flux associated with the input voltage. The linear response results we obtain are exactly the same. We 2 prefer to work with the charge Qˆ in order to be consistent with |λV | GP = (5.87) the rest of the text. 4Im χQQ · Im χVV 41

Finally, we may again define the effective tempera- coupling (i.e. tuning the source impedance Zs to match ture Teff [ω] of the amplifier via Eq. (2.8), and define a the noise impedance, cf. Eq. (5.77) - (5.78)); one also quantum-limited voltage amplifier as one which satisfies needs to have an amplifier with “ideal” quantum noise, the ideal noise condition of Eq. (4.17). For such an am- that is an amplifier satisfying Eq. (4.17). plifier, the power gain will again be determined by the Finally, we need to test our initial assumption that effective temperature via Eq. (5.56). |Zs|  |Zin|, taking |Zs| to be equal to its optimal value Turning to the noise, we can again calculate the to- ZN. Using the proportionality condition of Eq. (I13) tal symmetrized noise at the output port of the amplifier and the fact that we are in the large power gain limit following the same argument used to the get the out- (GP [ω]  1), we find: put noise of the position detector (cf. Eq. (5.46)). As Z [ω] α ω 1 we did in the classical approach, we will again assume N ~ = = p  1 (5.90) that the input impedance of the amplifier is much larger Re Zin[ω] Im α 4kBTeff 2 GP [ω] that source impedance: Zin  Zs; we will test this as- sumption for consistency at the end of the calculation. It follows that |ZN|  |Zin| in the large power gain, large Focusing only on the amplifier contribution to this noise effective temperature regime of interest, thus justifying (as opposed to the intrinsic noise of the input signal), and the form of Eq. (5.73). Eq. (5.90) is analogous to the referring this noise back to the amplifier input, we find case of the displacement detector, where we found that that the symmetrized quantum noise spectral density de- reaching the quantum limit on resonance required the detector-dependent damping to be much weaker than the scribing the added noise of the amplifier, S¯VV,add[ω], sat- isfies the same equation we found for a classical voltage intrinsic damping of the oscillator (cf. Eq. (5.69)). amplifier, Eq. (5.73), with each classical spectral density Thus, similar to the situation of the displacement de- S[ω] being replaced by the corresponding symmetrized tector, the linear response approach allows us both to quantum spectral density S¯[ω] as per Eqs. (5.80) - (5.82). derive rigorously the quantum limit on the noise tempera- It follows that the amplifier noise temperature will ture TN of an amplifier, and to state conditions that must again be given by Eq. (5.75), and that the optimal be met to reach this limit. To reach the quantum-limited noise temperature (after optimizing over the source value of TN with a large power gain, one needs both a impedance) will be given by Eq. (5.76). Whereas classi- tuned source impedance Zs, and an amplifier which pos- cally nothing more could be said, quantum mechanically, sesses ideal noise properties (cf. Eq. (4.17) and Eq. (I13)). we now get a further bound from the quantum noise con- straint of Eq. (4.11) and the requirement of a large power gain. The latter requirement tells us that the voltage gain 3. Role of noise cross-correlations ¯ λV [ω] and the cross-correlator SVoutQ must be in phase (cf. Sec. (IV.A.4) and Appendix I.3). This in turn means Before leaving the topic of a linear voltage amplifier, that S¯ must be purely imaginary. In this case, the we pause to note the role of cross-correlations in current V˜ I˜ and voltage noise in reaching the quantum limit. First, quantum noise constraint may be re-written as: note from Eq. (5.78) that in both the classical and quan- 2 tum treatments, the noise impedance ZN of the amplifier 2  ω  S¯ [ω]S¯ [ω] − Im S¯  ≥ ~ (5.88) will have a reactive part (i.e. Im Z 6= 0) if there are V˜ V˜ I˜I˜ V˜ I˜ 2 N out-of-phase correlations between the amplifier’s current Using these results in Eq. (5.76), we find the ultimate and voltage noises (i.e. if Im SVI 6= 0). Thus, if such quantum limit on the noise temperature:18 correlations exist, it will not be possible to minimize the noise temperature (and hence, reach the quantum limit), ω if one uses a purely real source impedance Z . k T [ω] ≥ ~ (5.89) s B N 2 More significantly, note that the final classical expres- sion for the noise temperature TN explicitly involves the Similar to the case of the position detector, reaching the real part of the SVI correlator (cf. Eq. (5.76)). In con- quantum limit here is not simply a matter of tuning the trast, we have shown that in the quantum case, Re S¯VI must be zero if one wishes to reach the quantum limit while having a large power gain (cf. Appendix I.3); as such, this quantity does not appear in the final expres- 18 Note that our definition of the noise temperature T conforms N sion for the minimal TN. It also follows that to reach the with that of Devoret and Schoelkopf (2000) and most electrical quantum limit while having a large power gain, an ampli- engineering texts, but is slightly different than that of Caves (1982). Caves assumes the source is initially at zero temperature fier cannot have significant in-phase correlations between (i.e. T0 = 0), and consequently uses the full quantum expression its current and voltage noise. for its equilibrium noise. In contrast, we have assumed that This last statement can be given a heuristic explana- kBT0  ~ω. The different definition of the noise temperature tion. If there are out-of-phase correlations between cur- used by Caves leads to the result kBTN ≥ ~ω/(ln 3) as opposed to our Eq. (5.89). We stress that the difference between these rent and voltage noise, we can easily make use of these by results has nothing to do with physics, but only with how one appropriately choosing our source impedance. However, defines the noise temperature. if there are in-phase correlations between current and 42 voltage noise, we cannot use these simply by tuning the et al., 2008). In practice, it can be difficult to achieve source impedance. We could however have used them by the theoretically-predicted quantum-limited performance implementing feedback in our amplifier. The fact that we due to spurious heating caused by the dissipation in the have not done this means that these correlations repre- shunt resistances used in the SQUID. This effect can be sent a kind of missing information; as a result, we must significantly ameliorated, however, by adding cooling fins necessarily miss the quantum limit. In Sec. VI.B, we to the shunts (Wellstood et al., 1994). explicitly give an example of a voltage amplifier which misses the quantum limit due to the presence of in-phase current and voltage fluctuations; we show how this am- 2. Quantum point contact detectors plifier can be made to reach the quantum limit by adding feedback in Appendix H. A quantum point contact (QPC) is a narrow conduct- ing channel formed in a two-dimensional gas. The cur- rent through the constriction is very sensitive to nearby G. Near quantum-limited mesoscopic detectors charges, and thus the QPC acts as a charge-to-current amplifier. It has been shown theoretically that the QPC Having discussed the origin and precise definition of can achieve the amplifier quantum limit, both in the the quantum limit on the added noise of a linear, phase- regime where transport is due to tunneling (Gurvitz, preserving amplifier, we now provide a brief review of 1997), as well as in regimes where the transmission is not work examining whether particular detectors are able (in small (Aleiner et al., 1997; Clerk et al., 2003; Korotkov principle) to achieve this ideal limit. We will focus on and Averin, 2001; Levinson, 1997; Pilgram and B¨uttiker, the “op-amp” mode of operation discussed in Sec. V.D, 2002). Experimentally, QPCs are in widespread use as where the detector is only weakly coupled to the system detectors of quantum dot . The back-action de- producing the signal to be amplified. As we have repeat- phasing of QPC detectors was studied in (Buks et al., edly stressed, reaching the quantum limit in this case 1998; Sprinzak et al., 2000); a good agreement was found requires the detector to have “quantum ideal noise”, as with the theoretical prediction, confirming that the QPC defined by Eq. (4.17). Heuristically, this corresponds to has quantum-limited back-action noise. the general requirement of no wasted information: there should be no other quantity besides the detector output that could be monitored to provide information on the in- 3. Single-electron transistors and resonant-level detectors put signal (Clerk et al., 2003). We have already given one simple but relevant example of a detector which reaches A metallic single-electron transistor (SET) consists of the amplifier quantum limit: the parametric cavity de- a small metallic island attached via tunnel junctions to tector, discussed extensively in Sec. III.B. Here, we turn larger source and drain electrodes. Because of Coulomb to other more complex detectors. blockade effects, the conductance of a SET is very sen- sitive to nearby charges, and hence it acts as a sensitive charge-to-current amplifier. Considerable work has inves- 1. dc SQUID amplifiers tigated whether metallic SETs can approach the quan- tum limit in various different operating regimes. The- The dc SQUID (superconducting quantum interference oretically, the performance of a normal-metal SET in device) is a detector based on a superconducting ring hav- the sequential tunneling regime was studied by Aassime ing two Josephson junctions. It can in principle be used et al. (2001); Devoret and Schoelkopf (2000); Johans- as a near quantum-limited voltage amplifier or flux-to- son et al. (2003, 2002); Makhlin et al. (2000); Shnirman voltage amplifier. Theoretically, this was investigated us- and Sch¨on(1998). In this regime, where transport is ing a quantum Langevin approach (Danilov et al., 1983; via a sequence of energy-conserving tunnel events, one Koch et al., 1980), as well as more rigorously by using is far from optimizing the quantum noise constraint of perturbative techniques (Averin, 2000b) and mappings Eq. (4.17), and hence one cannot reach the quantum to quantum impurity problems (Clerk, 2006). Experi- limit (Korotkov, 2001b; Shnirman and Sch¨on,1998). If ments on SQUIDS have also confirmed its potential for one instead chooses to work with a normal-metal SET in near quantum-limited operation. M¨uck et al. (2001) were the cotunneling regime (a higher-order tunneling process able to achieve a noise temperature TN approximately involving a virtual transition), then one can indeed ap- 1.9 times the quantum limited value at an operating fre- proach the quantum limit (Averin, 2000a; van den Brink, quency of ω = 2π × 519 MHz. Working at lower fre- 2002). However, by virtue of being a higher-order pro- quencies appropriate to gravitational wave detection ap- cess, the related currents and gain factors are small, im- plications, Vinante et al. (2001) were able to achieve a pinging on the practical utility of this regime of opera- TN approximately 200 times the quantum limited value tion. It is worth noting that while most theory on SETs at a frequency ω = 2π × 1.6 kHz; more recently, the assume a dc voltage bias, to enhance bandwidth, experi- same group achieved a TN approximately 10 times the ments are usually conducted using the rf-SET configura- quantum limit at a frequency ω = 2π × 1.6 kHz (Falferi tion (Schoelkopf et al., 1998), where the SET changes the 43 damping of a resonant LC circuit. Korotkov and Paala- resonant frequency Ω. Its motion may be written in nen (1999) have shown that this mode of operation for terms of quadrature operators defined as in Eq. (3.48): a sequential tunneling SET increases the measurement ˆ ˆ imprecision noise by approximately a factor of 2. The xˆ(t) = Xδ(t) cos (Ωt + δ) + Yδ(t) sin (Ωt + δ) measurement properties of a normal-metal, sequential- (5.91) tunneling rf-SET (including back-action) were studied experimentally in Turek et al. (2005). Here,x ˆ(t) is the Heisenberg-picture position operator of Measurement using superconducting SET’s has also the oscillator. The quadrature operators can be written been studied. Clerk et al. (2002) have shown that so- in terms of the (Schr¨odinger-picture)oscillator creation called incoherent Cooper-pair tunneling processes in a su- and destruction operators as: perconducting SET can have a noise temperature which ˆ  i(Ωt+δ) † −i(Ωt+δ) is approximately a factor of two larger than the quantum Xδ(t) = xZPF ceˆ +c ˆ e (5.92a) limited value. The measurement properties of a super- ˆ  i(Ωt+δ) † −i(Ωt+δ) conducting SET biased at a point of incoherent Cooper- Yδ(t) = −ixZPF ceˆ − cˆ e (5.92b) pair tunneling have been probed recently in experiment As previously discussed, the two quadrature ampli- (Naik et al., 2006; Thalakulam et al., 2004). tude operators Xˆ and Yˆ are canonically conjugate The quantum measurement properties of phase- δ δ (cf. Eq.(3.49)). Making a measurement of one quadra- coherent, non-interacting resonant level detectors have ture amplitude, say Xˆδ, will thus invariably lead to back- also been studied theoretically (Averin, 2000b; Clerk and ˆ Stone, 2004; Gavish et al., 2006; Mozyrsky et al., 2004). action disturbance of the other, conjugate quadrature Yδ. These systems are similar to metallic SET, except that However, due to the dynamics of a harmonic oscillator, the central island only has a single level (as opposed to this disturbance will not affect the measured quadrature a continuous density of states), and Coulomb-blockade at later times. One can already see this from the classical effects are typically neglected. These detectors can reach equations of motion. Suppose our oscillator is driven by the quantum limit in the regime where the voltage and a time-dependent force F (t) which only has appreciable temperature are smaller than the intrinsic energy broad- bandwidth near Ω. We may write this as: ening of the level due to tunneling. They can also reach F (t) = FX (t) cos(Ωt + δ) + FY (t) sin(Ωt + δ) (5.93) the quantum limit in a large-voltage regime that is analo- gous to the cotunneling regime in a metallic SET (Averin, where FX (t),FY (t) are slowly varying compared to Ω. 2000b; Clerk and Stone, 2004). The influence of dephas- Using the fact that the oscillator has a high-quality factor ing processes on such a detector was studied in Clerk and Q = Ω/γ, one can easily find the equations of motion: Stone (2004). d γ F (t) X (t) = − X (t) − Y , (5.94a) dt δ 2 δ 2mΩ d γ F (t) H. Back-action evasion and noise-free amplification Y (t) = − Y (t) + X . (5.94b) dt δ 2 δ 2mΩ

Having discussed in detail quantum limits on phase- Thus, as long as FY (t) and FX (t) are uncorrelated and preserving linear amplifiers (i.e. amplifiers which measure sufficiently slow, the dynamics of the two quadratures are both quadratures of a signal equally well), we now return completely independent; in particular, if Yδ is subject to to the situation discussed at the very start of Sec. V.A: a narrow-bandwidth, noisy force, it is of no consequence imagine we wish only to amplify a single quadrature of to the evolution of Xδ. An ideal measurement of Xδ will some time-dependent signal. For this case, there need result in a back-action force having the form in Eq. (5.93) not be any added noise from the measurement. Unlike with FY (t) = 0, implying that Xδ(t) will be completely the case of amplifying both quadratures, Liouville’s the- unaffected by the measurement. orem does not require the existence of any additional Not surprisingly, if one can measure and amplify Xδ degrees of freedom when amplifying a single quadrature: without any back-action, there need not be any added phase space volume can be conserved during amplifica- noise due to the amplification. In such a setup, the tion simply by contracting the unmeasured quadrature only added noise is the measurement-imprecision noise (cf. Eq. 5.2). As no extra degrees of freedom are needed, associated with intrinsic fluctuations of the amplifier out- there need not be any extra noise associated with the put. These may be reduced (in principle) to an arbitrar- amplification process. ily small value by simply increasing the amplifier gain Alternatively, single-quadrature detection can take a (e.g. by increasing the detector-system coupling): in an form similar to a QND measurement, where the back- ideal setup, there is no back-action penalty on the mea- action does not affect the dynamics of the quantity be- sured quadrature associated with this increase. ing measured (Bocko and Onofrio, 1996; Braginsky and The above conclusion can lead to what seems like a Khalili, 1992; Braginsky et al., 1980; Caves, 1982; Caves contradiction. Imagine we use a back-action evading am- et al., 1980; Thorne et al., 1978). For concreteness, con- plifier to make a “perfect” measurement of Xδ (i.e. neg- sider a high-Q harmonic oscillator with position x(t) and ligible added noise). We would then have no uncertainty 44

as to the value of this quadrature. Consequently, we 1. Degenerate parametric amplifier would expect the quantum state of our oscillator to be a squeezed state, where the uncertainty in Xδ is much Perhaps the simplest example of a phase non- smaller than xZPF. However, if there is no back-action preserving amplifier is the degenerate parametric ampli- acting on Xδ, how is the amplifier able to reduce its un- fier; the classical version of this system was described at certainty? This seeming paradox can be fully resolved the start of Sec. V.A (cf. Eq. 5.2). The setup is similar by considering the conditional aspects of an ideal single to the non-degenerate parametric amplifier discussed in quadrature measurement, where one considers the state Sec. V.C, except that the idler mode is eliminated, and of the oscillator given a particular measurement history the non-linearity converts a single pump photon into two (Clerk et al., 2008; Ruskov et al., 2005). signal photons at frequency ωS = ωP/2. As we now show, the resulting dynamics causes one signal quadrature to be amplified, while the other is attenuated, in such a way that it is not necessary to add extra noise to preserve the It is worth stressing that the possibility of amplifying a canonical commutation relations. single quadrature without back-action (and hence, with- The system Hamiltonian is out added noise) relies crucially on our oscillator resem-   bling a perfect harmonic oscillator: the oscillator Q must Hˆ = ω aˆ† aˆ + ω aˆ† aˆ be large, and non-linearities (which could couple the two sys ~ P P P S S S quadratures) must be small. In addition, the envelope of  † † †  + i~η aˆSaˆSaˆP − aˆSaˆSaˆP . (5.95) the non-vanishing back-action force FX (t) must have a narrow bandwidth. One should further note that a very Treating the pump classically as before, the analog of high precision measurement of Xδ will produce a very Eq. (5.16) is large backaction force FX . If the system is not nearly perfectly harmonic, then the large amplitude imparted λ   Vˆ = i aˆ† aˆ† − aˆ aˆ , (5.96) to the conjugate quadrature Yδ will inevitably leak back sys ~ 2 S S S S into Xδ. where λ/2 = ηψP , and the analog of Eq. (5.19) is:

˙ κS † √ ˆ aˆS = − aˆS + λaˆ − κS bS,in (5.97) Amplifiers or detectors which treat the two signal 2 S quadratures differently are known in the quantum op- The dimensionless quadrature operators corresponding tics literature as ‘phase sensitive’; we prefer the desig- to the signal mode are: nation ‘phase-non-preserving’ since they do not preserve the phase of the original signal. Such amplifiers invari- 1  †  i  †  xˆS = √ aˆ +a ˆS , yˆS = √ aˆ − aˆS (5.98) ably rely on some internal clock (i.e. an oscillator with a 2 S 2 S well defined phase) which breaks time-translation invari- ance and picks out the phase of the quadrature that will which obey [ˆxS, yˆS] = i. We can define quadrature oper- ˆ ˆ be amplified (i.e. the choice of δ used to define the two ators XS,in/out, YS,in/out corresponding to the input and quadratures in Eq. (5.91)); we will see this explicitly in output fields in a completely analogous manner. what follows. This leads to an important caveat: even The steady state solution of Eq. (5.97) for the output in a situation where the interesting information is in a fields becomes single signal quadrature, to benefit from using a phase √ 1 non-preserving amplifier, we must know in advance the XˆS,out = GXˆS,in, YˆS,out = √ YˆS,in, (5.99) precise phase of this quadrature. If we do not know this G phase, we will have either to revert to a phase-preserving where the number gain G is given by amplification scheme (and thus be susceptible to added noise) or we would have to develop a sophisticated and λ + κ /22 high speed quantum feedback scheme to dynamically G = S . (5.100) λ − κ /2 adapt the measurement to the correct quadrature in real S time (Armen et al., 2002). In what follows, we will make We thus see clearly that the amplifier treats the two the above ideas concrete by considering a few examples quadratures differently. One quadrature is amplified, the 19 of quantum, phase non-preserving amplifiers. other attenuated, with the result that the commutation relation can be preserved without the necessity of extra degrees of freedom and added noise. Note that the large- amplitude pump mode has played the role of a clock in 19 One could in principle generalize the linear response approach the degenerate paramp: it is the phase of the pump which of Sec. IV to deal with phase non-preserving detectors. How- ever, as such detectors are not time-translational invariant, such picks out which quadrature of the signal will be amplified. a description becomes rather cumbersome and is not particularly Before ending our discussion here, it is important to helpful. We prefer instead to present concrete examples. stress that while the degenerate parametric amplifier is 45 phase-sensitive and has no added noise, it is not an ex- at ωc. Making a single-quadrature measurement requires ample of back-action evasion (see Caves et al. (1980), that one drives the cavity equally at the two sideband fre- ¯ footnote on p. 342). This amplifier is operated in the quencies. The amplitude of the driving field bin entering scattering mode of amplifier operation, a mode where (as the cavity will be chosen to have the form: discussed extensively in Sec. V.D) back-action is not at all p relevant. Recall that in this mode of operation, the am- i N˙   ¯b (t) = − eiδe−i(ωc−Ω)t − e−iδe−i(ωc+Ω)t plifier input is perfectly impedance matched to the signal in 4 p source, and the input signal is simply the amplitude of an N˙ incident wave on the amplifier input. This mode of oper- = sin(Ωt + δ)e−iωct. (5.101) ation necessarily requires a strong coupling between the 2 ˆ input signal and the amplifier input (i.e. hbS,ini). If one Here, N˙ is the photon number flux associated with the instead tried to weakly couple the degenerate parametric cavity drive (see Appendix E for more details on how amplifier to a signal source, and operate it in the “op- to properly include a drive using input-output theory). amp” mode of operation (cf. Sec. V.D), one finds that Such a drive could be produced by taking a signal at the there is indeed a back-action disturbance of the measured cavity resonance frequency, and amplitude modulating it quadrature. We have yet another example which demon- at the mechanical frequency. strates that one must be very careful to distinguish the To understand the effect of this drive, note that it sends “op-amp” and “scattering” mode of amplifier operation. the cavity both photons with frequency (ωc − Ω) and photons with frequency (ωc + Ω). The first kind of drive photon can be converted to a cavity photon if a quanta is 2. Double-sideband cavity detector absorbed from the mechanical oscillator; the second kind of drive photon can be converted to a cavity photon if a We now turn to a simple but experimentally-relevant quanta is emitted to the mechanical oscillator. The result detector that is truly back-action evading. We will take is that we can create a cavity photon by either absorb- ing or emitting a mechanical oscillator quanta. Keeping as our input signal the positionx ˆ of a mechanical oscilla- i2δ tor. The amplifier setup we consider is almost identical track that there is a well-defined relative phase of e to the cavity position detector discussed in Sec. III.B.3: between the two kinds of drive photons, we would expect we again have a single-sided resonant cavity whose fre- the double-sideband drive to yield an effective cavity- quency depends linearly on the oscillator’s position, with oscillator interaction of the form: the Hamiltonian being given by Eq. (3.11) (withz ˆ = p  † iδ −iδ †  Veff ∝ N˙ aˆ e cˆ + e cˆ + h.c. (5.102a) x/xˆ ZPF). We showed in Sec. III.B.3 and Appendix E.3 p † that by driving the cavity on resonance, it could be used ∝ N˙ (ˆa +a ˆ )Xˆδ (5.102b) to make a quantum limited position measurement: one can operate it as a phase-preserving amplifier of the me- This is exactly what is found in a full calculation (see Ap- chanical’s oscillators position, and achieve the minimum pendix E.4). Note that we have written the interaction possible amount of added noise. To use the same system in an interaction picture in which the fast oscillations of to make a back-action free measurement of one oscillator the cavity and oscillator operators have been removed. quadrature only, one simply uses a different cavity drive. In the second line, we have made use of Eqs. (5.92) to ˆ Instead of driving at the cavity resonance frequency ωc, show that the effective interaction only involves the Xδ one drives at the two sidebands associated with the me- oscillator quadrature. chanical motion (i.e. at frequencies ωc ± Ω, where Ω is We thus see from Eq. (5.102b) that the cavity is only as always the frequency of the mechanical resonator). As coupled to the oscillator Xδ quadrature. As shown rig- we will see, such a drive results in an effective interaction orously in Appendix E.3, the result is that the system which only couples the cavity to one quadrature of the only measures and amplifies this quadrature: the light ˆ ˆ oscillator’s motion. This setup was first proposed as a leaving the cavity has a signature of Xδ, but not of Yδ. means of back-action evasion in Braginsky et al. (1980); Further, Eq. (5.102b) implies that the cavity operator p † further discussion can be found in Braginsky and Khalili N˙ aˆ +a ˆ will act as a noisy force on the Yδ quadra- (1992); Caves et al. (1980), as well as in Clerk et al. ture. While this will cause a back-action heating of Yδ, (2008), which gives a fully quantum treatment and con- it will not affect the measured quadrature Xδ. We thus siders conditional aspects of the measurement. In what have a true back-action evading amplifier: the cavity out- follows, we sketch the operation of this system following put light lets one measure Xδ free from any back-action Clerk et al. (2008); details are provided in Appendix E.4. effect. Note that in deriving Eq. (5.102a), we have used We will start by requiring that our system be in the the fact that the cavity operators have fluctuations in a “good-cavity” limit, where ωc  Ω  κ (κ is the damp- narrow bandwidth ∼ κ  Ω: the back-action force noise ing of the cavity mode); we will also require the mechan- is slow compared to the oscillator frequency. If this were ical oscillator to have a high Q-factor. In this regime, the not the case, we could still have a back-action heating of two sidebands associated with the mechanical motion at the measured Xδ quadrature. Such effects, arising from ωc ±Ω are well-separated from the main cavity resonance a non-zero ratio κ/Ω, are treated in Clerk et al. (2008). 46

Finally, as there is no back-action on the measured discuss the physics in terms of the bosonic voltage ampli- Xδ quadrature, the only added noise of the amplification fier first introduced in Sec. V.D. Recall that in that sub- scheme is measurement-imprecision noise (e.g. shot noise section, we demonstrated that the standard Haus-Caves in the light leaving the cavity). This added noise can derivation of the quantum limit was not directly relevant be made arbitrarily small by increasing the gain of the to the usual weak-coupling “op-amp” mode of amplifier detector by, for example, increasing the strength of the operation, a mode where the input signal is not simply cavity drive N˙ . In a real system where κ/ωM is non-zero, the amplitude of a wave incident on the amplifier. In this the finite-bandwidth of the cavity number fluctuations section, we will expand upon that discussion, giving an leads to a small back-action on the Xδ. As a result, one explicit discussion of the differences between the op-amp cannot make the added noise arbitrarily small, as too description of an amplifier presented in Sec. V.E, and the large a cavity drive will heat the measured quadrature. scattering description often used in the quantum optics Nonetheless, for a sufficiently small ratio κ/ωM , one can literature (Courty et al., 1999; Grassia, 1998). We will still beat the standard quantum limit on the added noise see that what one means by “back-action” and “added (Clerk et al., 2008). noise” are not the same in the two descriptions! Further, even though an amplifier may reach the quantum limit when used in the scattering mode (i.e. its added noise 3. Stroboscopic measurements is as small as allowed by commutation relations), it can nonetheless fail to achieve the quantum limit when used With sufficiently high bandwidth it should be able to in the op-amp mode. Finally, the discussion here will also do stroboscopic measurements in sync with the oscillator allow us to highlight important aspects of the quantum motion which could allow one to go below the standard limit not easily discussed in the more general context of quantum limit in one of the quadratures of motion (Bra- Sec. IV. ginsky and Khalili, 1992; Caves et al., 1980). To under- stand this idea, imagine an extreme form of phase sen- sitive detection in which a Heisenberg microscope makes A. Scattering versus op-amp representations a strong high-resolution measurement which projects the In the bosonic scattering approach, a generic linear oscillator onto a state of well defined position X0 at time t = 0: amplifier is modeled as a set of coupled bosonic modes. To make matters concrete, we will consider the specific ∞ case of a voltage amplifier with distinct input and output X −i(n+1/2)Ωt ΨX0 (t) = ane |ni , (5.103) ports, where each port is a semi-infinite transmission line n=0 (see Fig. 9). We start by recalling that a quantum trans- mission line can be described as a set of non-interacting where the coefficients obey an = hn|X0i. Because the bosonic modes (see Appendix D for a quick review). De- position is well-defined the momentum is extremely un- noting the input transmission line with an a and the out- certain. (Equivalently the momentum kick delivered by put transmission line with a b, the current and voltage the backaction of the microscope makes the oscillator operators in these lines may be written: momentum uncertain.) Thus the wave packet quickly spreads out and the position uncertainty becomes large. Z ∞ ˆ dω  ˆ −iωt  However because of the special feature that the har- Vq(t) = Vq[ω]e + h.c. (6.1a) 0 2π monic oscillator levels are evenly spaced, we can see from Z ∞ dω  −iωt  Eq. (5.103) that the wave packet reassembles itself pre- Iˆq(t) = σq Iˆq[ω]e + h.c. (6.1b) cisely once each period of oscillation because einΩt = 1 0 2π for every integer n. (At half periods, the packet re- with assembles at position −X0.) Hence stroboscopic mea- r surements made once (or twice) per period will be back- ω Vˆ [ω] = ~ Z (ˆq [ω] +q ˆ [ω]) (6.2a) action evading and can go below the standard quantum q 2 q in out limit. The only limitations will be the finite anharmonic- s ity and damping of the oscillator. Note that the possi- ~ω Iˆq[ω] = (ˆqin[ω] − qˆout[ω]) (6.2b) bility of using mesoscopic electron detectors to perform 2Zq stroboscopic measurements has recently received atten- tion (Jordan and B¨uttiker, 2005b; Ruskov et al., 2005). Here, q can be equal to a or b, and we have σa = 1, σb = −1. The operatorsa ˆin[ω], aˆout[ω] are bosonic annihila- tion operators;a ˆin[ω] describes an incoming wave in the VI. BOSONIC SCATTERING DESCRIPTION OF A input transmission line (i.e. incident on the amplifier) TWO-PORT AMPLIFIER having frequency ω, whilea ˆout[ω] describes an outgoing ˆ ˆ wave with frequency ω. The operators bin[ω] and bout[ω] In this section, we return again to the topic of Sec. V.F, describe analogous waves in the output transmission line. quantum limits on a quantum voltage amplifier. We now We can think of Vˆa as the input voltage to our amplifier, 47 and Vˆb as the output voltage. Similarly, Iˆa is the current Sec. 5.10, where we presented the Haus-Caves derivation drawn by the amplifier at the input, and Iˆb the current of the quantum limit (cf. Eq. (5.7)). It thus follows that drawn at the output of the amplifier. Finally, Za (Zb) is in the scattering mode of operation, the matrix element the characteristic impedance of the input (output) trans- s21[ω] represents the gain of our amplifier at frequency ω, 2 mission line. |s21[ω]| the corresponding “photon number gain”, and As we have seen, amplification invariably requires ad- Fˆb the added noise operator of the amplifier. The opera- ditional degrees of freedom. Thus, to amplify a signal at tor Fˆa represents the back-action noise in the scattering a particular frequency ω, there will be 2N bosonic modes mode of operation; this back-action has no effect on the involved, where the integer N is necessarily larger than added noise of the amplifier in the scattering mode. 2. Four of these modes are simply the frequency-ω modes Similar to Sec. V.B, one can now apply the standard ˆ in the input and output lines (i.e.a ˆin[ω],ˆaout[ω],bin[ω] argument of Haus and Mullen (1962) and Caves (1982) ˆ and bout[ω]). The remaining 2(N − 2) modes describe to our amplifier. This argument tells us that since the auxiliary degrees of freedom involved in the amplifica- “out” operators must have the same commutation rela- tion process; these additional modes could correspond to tions as the “in” operators, the added noise Fˆb cannot be frequencies different from the signal frequency ω. The arbitrarily small in the large gain limit (i.e. |s21|  1). auxiliary modes can also be divided into incoming and Note that this version of the quantum limit on the added outgoing modes. It is thus convenient to represent them noise has nothing to do with back-action. As already as additional transmission lines attached to the amplifier; discussed, this is perfectly appropriate for the scatter- these additional lines could be semi-infinite, or could be ing mode of operation, as in this mode, the experimen- terminated by active elements. talist ensures that the signal going into the amplifier is completely independent of whatever is coming out of the amplifier. This mode of operation could be realized in 1. Scattering representation time-dependent experiments, where a pulse is launched at the amplifier. This mode is not realized in most weak- In general, our generic two-port bosonic amplifier will coupling amplification experiments, where the signal to be described by a N × N scattering matrix which deter- be amplified is not identical to an incident wave ampli- mines the relation between the outgoing mode operators tude. and incoming mode operators. The form of this matrix is constrained by the requirement that the output modes obey the usual canonical bosonic commutation relations. 2. Op-amp representation It is convenient to express the scattering matrix in a form which only involves the input and output lines explicitly: In the usual op-amp amplifier mode of operation (de- ! ! ! scribed extensively in Sec. IV), the input and output sig- aˆ [ω] s [ω] s [ω] aˆ [ω] out = 11 12 in nals are not simply incoming/outgoing wave amplitudes; ˆ ˆ bout[ω] s21[ω] s22[ω] bin[ω] thus, the scattering representation is not an optimal de- ! scription of our amplifier. The system we are describing Fˆ [ω] + a (6.3) here is a voltage amplifier: thus, in the op-amp mode, ˆ Fb[ω] the experimentalist would ensure that the voltage at the end of the input line (Vˆa) is equal to the signal to be Here Fˆa[ω] and Fˆb[ω] are each an unspecified linear amplified, and would read out the voltage at the end of combination of the incoming auxiliary modes introduced the output transmission line (Vˆb) as the output of the above. They thus describe noise in the outgoing modes of amplifier. From Eq. (6.1a), we see that this implies that the input and output transmission lines which arises from the amplitude of the wave going into the amplifier, ain, the auxiliary modes involved in the amplification pro- will depend on the amplitude of the wave exiting the am- cess. Note the similarity between Eq. (6.3) and Eq. (5.7) plifier, aout. for the simple one-port bosonic amplifier considered in Thus, if we want to use our amplifier as a voltage am- Sec. V.B. plifier, we would like to find a description which is more In the quantum optics literature, one typically views tailored to our needs than the scattering representation Eq. (6.3) as the defining equation of the amplifier; we will of Eq. (6.3). This can be found by simply re-expressing call this the scattering representation of our amplifier. the scattering matrix relation of Eq. (6.3) in terms of The representation is best suited to the scattering mode voltages and currents. The result will be what we term of amplifier operation described in Sec. V.D. In this mode the “op amp” representation of our amplifier, a repre- of operation, the experimentalist ensures that haˆin[ω]i is sentation which is standard in the discussion of classical precisely equal to the signal to be amplified, irrespective amplifiers (see, e.g., Boylestad and Nashelsky (2006)). of what is coming out of the amplifier. Similarly, the In this representation, one views Vˆa and Iˆb as inputs output signal from the amplifier is the amplitude of the to the amplifier: Vˆ is set by whatever we connect to ˆ a outgoing wave in the output line, hbout[ω]i. If we focus on the amplifier input, while Iˆb is set by whatever we con- ˆ bout, we have precisely the same situation as described in nect to the amplifier output. In contrast, the outputs 48 of our amplifier are the voltage in the output line, Vˆb, load impedance to be Zload, some simple algebra yields: and the current drawn by the amplifier at the input, Iˆa. 1 1 λ0 λ Note that this interpretation of voltages and currents is = + I V (6.6) identical to how we viewed the voltage amplifier in the Zin,loaded Zin Zload + Zout linear-response/quantum noise treatment of Sec. V.F. Using Eqs. (6.1a) and (6.1b), and suppressing fre- It is of course undesirable to have an input impedance quency labels for clarity, Eq. (6.3) may be written ex- which depends on the load. Thus, we see yet again that plicitly in terms of the voltages and current in the input it is undesirable to have appreciable reverse gain in our amplifier (cf. Sec. IV.A.2). (Vˆa, Iˆa) and output (Vˆb, Iˆb) transmission lines:

! ! ! ! Vˆ λ −Z Vˆ λ · V˜ˆ b = V out a + V (6.4) 3. Converting between representations Iˆ 1 λ0 Iˆ ˜ˆ a Zin I b I Some straightforward algebra now lets us express the The coefficients in the above matrix are completely deter- op-amp parameters appearing in Eq. (6.4) in terms of the mined by the scattering matrix of Eq. (6.3) (see Eqs. (6.7) scattering matrix appearing in Eq. (6.3): below); moreover, they are familiar from the discussion of r Z s a voltage amplifier in Sec. V.F. λV [ω] is the voltage gain λ = 2 b 21 (6.7a) 0 V of the amplifier, λI [ω] is the reverse current gain of the Za D r amplifier, Zout is the output impedance, and Zin is the Z s λ0 = 2 b 12 (6.7b) input impedance. The last term on the RHS of Eq. (6.4) I Z D ˆ a describes the two familiar kinds of amplifier noise. V˜ is (1 + s )(1 + s ) − s s Z = Z 11 22 12 21 (6.7c) the usual voltage noise of the amplifier (referred back to out b D ˜ˆ the amplifier input), while I is the usual current noise of 1 1 (1 − s11)(1 − s22) − s12s21 the amplifier. Recall that in this standard description of = (6.7d) Zin Za D a voltage amplifier (cf. Sec. V.F), I˜ represents the back- action of the amplifier: the system producing the input where all quantities are evaluated at the same frequency signal responds to these current fluctuations, resulting in ω, and D is defined as: an additional fluctuation in the input signal going into the amplifier. Similarly, λV · V˜ represents the intrinsic D = (1 + s11)(1 − s22) + s12s21 (6.8) output noise of the amplifier: this contribution to the total output noise does not depend on properties of the Further, the voltage and current noises in the op-amp input signal. Note that we are using a sign convention representation are simple linear combinations of the noises Fˆ and Fˆ appearing in the scattering representa- where a positive hIˆai indicates a current flowing into the a b tion: amplifier at its input, while a positive hIˆbi indicates a current flowing out of the amplifier at its output. Also ! ! ! ˜ˆ 1 1+s11 ˆ ˆ ˆ V p − 2 2s Fa note that the operators Va and Ib on the RHS of Eq. (6.4) = 2~ωZa 21 ˜ˆ s22−1 s12 ˆ will have noise; this noise is entirely due to the systems Za · I D − D Fb attached to the input and output of the amplifier, and as (6.9) such, should not be included in what we call the added noise of the amplifier. Again, all quantities above are evaluated at frequency ω. Additional important properties of our amplifier follow Eq. (6.9) immediately leads to an important conclusion immediately from quantities in the op-amp representa- and caveat: what one calls the “back-action” and “added tion. As discussed in Sec. V.E, the most important mea- noise” in the scattering representation (i.e. Fa and Fb ) sure of gain in our amplifier is the dimensionless power are not the same as the “back-action” and “added noise” gain. This is the ratio between power dissipated at the defined in the usual op-amp representation. For example, output to that dissipated at the input, taking the output the op-amp back-action I˜ˆ does not in general coincide current I to be V /Z : B B out with the Fˆa, the back-action in the scattering picture. If we are indeed interested in using our amplifier as a 2  0 −1 (λV ) Zin λV λI Zin voltage amplifier, we are interested in the total added GP ≡ · 1 + (6.5) 4 Zout 2 Zout noise of our amplifier as defined in the op-amp repre- sentation. As we saw in Sec. V.F (cf. Eq. (5.71)), this Another important quantity is the loaded input quantity involves both the noises I˜ˆ and V˜ˆ . We thus see impedance: what is the input impedance of the ampli- explicitly something already discussed in Sec. V.D: it is fier in the presence of a load attached to the output? In very dangerous to make conclusions about how an ampli- 0 the presence of reverse current gain λI 6= 0, the input fier behaves in the op-amp mode of operation based on impedance will depend on the output load. Taking the its properties in the scattering mode of operation. As we 49 will see, even though an amplifier is “ideal” in the scatter- by the impedances of the input and output transmission ing mode (i.e. Fa as small as possible), it can nonetheless lines. From Eqs. (6.7), we have: fail to reach the quantum limit in the op-amp mode of r operation. Zb ˆ λV = 2 G (6.11a) In what follows, we will calculate the op-amp noises V˜ Za ˆ 0 and I˜ in a minimal bosonic voltage amplifier, and show λI = 0 (6.11b) explicitly how this description is connected to the more Zout = Zb (6.11c) general linear-response treatment of Sec. V.F. However, 1 1 before proceeding, it is worth noting that Eqs. (6.7a)- = (6.11d) Zin Za (6.7d) are themselves completely consistent with linear- response theory. Using linear-response, one would calcu- We immediately see that our amplifier looks less ideal 0 late the op-amp parameters λV , λI ,Zin and Zout using as an op-amp. The input and output impedances are Kubo formulas (cf. Eqs. (5.84), (5.85) and the discussion the same as those of the input and output transmission following Eq. (5.79)). These in turn would involve corre- line. However, for an ideal op-amp, we would have liked lation functions of Iˆa and Vˆb evaluated at zero coupling Zin → ∞ and Zout → 0. to the amplifier input and output. Zero coupling means Also of interest are the expressions for the amplifier that there is no input voltage to the amplifier (i.e. a short noises in the op-amp representation: circuit at the amplifier input, Vˆa = 0) and there is noth- ˆ ! 1 1 ! ! ing at the amplifier output drawing current (i.e. an open V˜ p − √ Fˆ = − 2 ωZ 2 2 G a ˆ ˆ ~ a ˆ circuit at the amplifier output, Ib = 0). Eq. (6.4) tells Za · I˜ 1 0 Fb ˆ ˆ us that in this case, Vb and Ia reduce to (respectively) (6.12) ˆ ˆ the noise operators λV V˜ and I˜. Using the fact that the commutators of Fˆa and Fˆb are completely determined As s12 = 0, the back-action noise is the same in both the by the scattering matrix (cf. Eq. (6.3)), we verify explic- op-amp and scattering representations: it is determined ˆ itly in Appendix I.5 that the Kubo formulas yield the completely by the noise operator Fa. However, the volt- ˆ same results for the op-amp gains and impedances as age noise (i.e. the intrinsic output noise) involves both Fa Eqs. (6.7a)-(6.7d) above. and Fˆb. We thus have the unavoidable consequence that there will be correlations in I˜ˆ and V˜ˆ . Note that from basic linear response theory, we know that there must B. Minimal two-port scattering amplifier be some correlations between I˜ˆ and V˜ˆ if there is to be gain (i.e. λV is given by a Kubo formula involving these 1. Scattering versus op-amp quantum limit operators, cf. Eq. (4.3)). To make further progress, we note again that commu- In this subsection we demonstrate that an amplifier tators of the noise operators Fˆa and Fˆb are completely which is “ideal” and minimally complex when used in the determined by Eq. (6.10) and the requirement that the scattering operation mode fails, when used as a voltage output operators obey canonical commutation relations. op-amp, to have a quantum limited noise temperature. We thus have: The system we look at is very similar to the amplifier h i considered by Grassia (1998), though our conclusions are Fˆ , Fˆ† = 1 (6.13a) somewhat different than those found there. a a In the scattering representation, one might guess that h ˆ ˆ†i Fb, Fb = 1 − |G| (6.13b) an “ideal” amplifier would be one where there are no h i h i reflections of signals at the input and output, and no ˆ ˆ ˆ ˆ† Fa, Fb = Fa, Fb = 0 (6.13c) way for incident signals at the output port to reach the input. In this case, Eq. (6.3) takes the form: We will be interested in the limit of a large power ! ! ! ! gain, which requires |G|  1. A minimal solution to the aˆ 0 0 aˆ Fˆ above equations would be to have the noise operators de- out = √ in + a (6.10) ˆ ˆ bout G 0 bin Fˆb termined by two independent (i.e. mutually commuting) auxiliary input mode operators u and v† : √ in in where we have defined G ≡ s . All quantities above 21 ˆ should be evaluated at the same frequency ω; for clarity, Fa =u ˆin (6.14) we will omit writing the explicit ω dependence of quan- ˆ p † Fb = |G| − 1ˆvin (6.15) tities throughout this section. Turning to the op-amp representation, the above equa- Further, to minimize the noise of the amplifier, we take tion implies that our amplifier has no reverse gain, and the operating state of the amplifier to be the vacuum for that the input and output impedances are simply given both these modes. With these choices, our amplifier is in 50

cold Thus, we need to calculate the symmetrized spectral den- load sities of the current and voltage noises, and their cross- correlation. It is easy to confirm from the definitions of vacuum noise uin uout Eq. (6.1a) and (6.1b) that these quantities take the form: Dn oE aout bin V˜ˆ [ω], V˜ˆ †(ω0) ¯ SVV [ω] = 0 (6.17a) input line output line 4πδ(ω − ω ) Dn oE ain bout I˜ˆ[ω], I˜ˆ†(ω0) 1/2 1/2 † S¯ [ω] = (6.17b) cin cout=G cin+(G-1) v in II 4πδ(ω − ω0) Dn oE ˜ˆ ˜ˆ† 0 ideal V [ω], I (ω ) S¯ [ω] = (6.17c) 1-port VI 4πδ(ω − ω0) amp. The expectation values here are over the operating state vout vin vacuum noise of the amplifier; we have chosen this state to be the vac- uum for the auxiliary mode operatorsu ˆin andv ˆin to min- imize the noise. FIG. 13 Schematic of a “minimal” two-port amplifier which Taking |s |  1, and using Eqs. (6.14) and (6.15), we reaches the quantum limit in the scattering mode of operation, 21 but misses the quantum limit when used as a weakly-coupled have op-amp. See text for further description ωZ ωZ S¯ [ω] = ~ a (σ + σ ) = ~ a (6.18a) VV 4 uu vv 2 ~ω ~ω S¯II [ω] = σuu = (6.18b) exactly the minimal form described by Grassia (1998): an Za Za input and output line coupled to a negative resistance box ω ω S¯ [ω] = ~ σ = ~ (6.18c) and an auxiliary “cold load” via a four-port circulator VI 2 uu 2 (see Fig. 13). The negative resistance box is nothing but the single-mode bosonic amplifier discussed in Sec. V.B; where we have defined: an explicit realization of this element would be the para- D ˆ† ˆ† E metric amplifier discussed in Sec. V.C. The “cold load” σab ≡ aˆb + b aˆ (6.19) is a semi-infinite transmission line which models dissipa- tion due to a resistor at zero-temperature (i.e. its noise and have used the fact that there cannot be any correla- is vacuum noise, cf. Appendix D). tions between the operators u and v in the vacuum state Note that within the scattering picture, one would con- (i.e. huˆvˆ†i = 0). clude that our amplifier is ideal: in the large gain limit, It follows immediately from the above equations that ˆ the noise added by the amplifier to bout corresponds to a our minimal amplifier does not optimize the quantum single quantum at the input: noise constraint of Eq. (5.88): Dn oE ˆ ˆ†  ω 2 Fb , Fb |G| − 1 Dn oE ¯ ¯  ¯ 2 ~ = vˆ† , vˆ → 1 (6.16) SVV [ω]SII [ω] − Im SVI = 2 × . (6.20) |G| |G| in in 2

This however is not the quantity which interests us: as The noise product S¯VV S¯II is precisely twice the we want to use this system as a voltage op-amp, we would quantum-limited value. As a result, the general the- like to know if the noise temperature defined in the op- ory of Sec. V.F tells us if one couples an input signal amp picture is as small as possible. We are also usually weakly to this amplifier (i.e. Zs  Zin), it is impossible interested in the case of a signal which is weakly cou- to reach the quantum limit on the added noise. Thus, pled to our amplifier; here, weak-coupling means that while our amplifier is ideal in the scattering mode of op- the input impedance of the amplifier is much larger than eration (cf. Eq. (6.16)), it fails to reach the quantum the impedance of the signal source (i.e. Zin  Zs). In limit when used in the weak-coupling, op-amp mode of this limit, the amplifier only slightly increases the total operation. Our amplifier’s failure to have “ideal” quan- damping of the signal source. tum noise also means that if we tried to use it to do To address whether we can reach the op-amp quan- QND qubit detection, the resulting back-action dephas- tum limit in the weak-coupling regime, we can make use ing would be twice as large as the minimum required by of the results of the general theory presented in Sec. V.F. quantum mechanics (cf. Sec. IV.B). In particular, we need to check whether the quantum One might object to the above conclusions based on noise constraint of Eq. (5.88) is satisfied, as this is a pre- the classical expression for the minimal noise tempera- requisite for reaching the (weak-coupling) quantum limit. ture, Eq. (5.76). Unlike the quantum noise constraint 51 of Eq. (5.88), this equation also involves the real part 2. Why is the op-amp quantum limit not achieved? of S¯VI , and is optimized by our “minimal” amplifier. However, this does not mean that one can achieve a Returning to the more interesting case of a weak noise temperature of ~ω/2 at weak coupling! Recall amplifier-signal coupling, one might still be puzzled as from Sec. V.F that in the usual process of optimizing to why our seemingly ideal amplifier misses the quantum the noise temperature, one starts by assuming the weak limit. While the mathematics behind Eq. (6.20) is fairly coupling condition that the source impedance Zs is much transparent, it is also possible to understand this result smaller than the amplifier input impedance Zin. One heuristically. To that end, note again that the amplifier ¯ then finds that to minimize the noise temperature, |Zs| noise cross-correlation SIV does not vanish in the large- should be tuned to match the noise impedance of the gain limit (cf. Eq. (6.18c)). Correlations between the two p amplifier ZN ≡ S¯VV /S¯II . However, in our mini- amplifier noises represent a kind of information, as by making use of them, we can improve the performance of mal bosonic√ amplifier, it follows from Eqs. (6.18) that Z = Z / 2 ∼ Z : the noise impedance is on the the amplifier. It is easy to take advantage of out-of-phase N in in ˜ ˜ ¯ order of the input impedance. Thus, it is impossible to correlations between I and V (i.e. Im SVI ) by simply match the source impedance to the noise impedance while tuning the phase of the source impedance (cf. Eq. (5.75)). at the same time satisfying the weak coupling condition However, one cannot take advantage of in-phase noise ¯ Zs  Zin. correlations (i.e. Re SVI ) as easily. To take advantage of the information here, one needs to modify the ampli- Despite its failings, our amplifier can indeed yield a fier itself. By feeding back some of the output voltage quantum-limited noise temperature in the op-amp mode to the input, one could effectively cancel out some of the of operation if we no longer insist on a weak coupling back-action current noise I˜ and thus reduce the over- to the input signal. To see this explicitly, imagine we all magnitude of S¯II . Hence, the unused information in connected our amplifier to a signal source with source the cross-correlator Re S¯VI represents a kind of wasted impedance Zs. The total output noise of the amplifier, information: had we made use of these correlations via referred back to the signal source, will now have the form: a feedback loop, we could have reduced the noise tem- perature and increased the information provided by our   amplifier. The presence of a non-zero Re S¯VI thus corre- ˆ ZsZa ˆ ˆ V˜ = − I˜+ V˜ (6.21) sponds to wasted information, implying that we cannot tot Z + Z s a reach the quantum limit. Recall that within the linear- response approach, we were able to prove rigorously that Note that this classical-looking equation can be rigor- a large-gain amplifier with ideal quantum noise must have ¯ ously justified within the full quantum theory if one starts Re SVI = 0 (cf. the discussion following Eq. (5.57)); thus, ¯ with a full description of the amplifier and the signal a non-vanishing Re SVI rigorously implies that one can- source (e.g. a parallel LC oscillator attached in parallel not be at the quantum limit. In Appendix H, we give an explicit demonstration of how feedback may be used to the amplifier input). Plugging in the expressions for I˜ˆ to utilize these cross-correlations to reach the quantum ˆ and V˜ , we find: limit. Finally, yet another way of seeing that our amplifier does not reach the quantum limit (in the weak coupling r "     ˆ ~ω ZsZa 2 regime) is to realize that this system does not have a V˜tot = √ uˆin − 2 Zs + Za Za well defined effective temperature. Recall from Sec.V.F # that a system with “ideal” quantum noise (i.e. one that p † Z (ˆu − vˆ ) satisfies Eq. (5.88) as an equality) necessarily has the a in in same effective temperature at its input and output ports r    (cf. Eq. (5.55)). Here, that implies the requirement: ~ωZa Zs − Za † = uˆin − vˆin (6.22) 2 ¯ 2 Zs + Za |λV | · SVV = ZinS¯II ≡ 2kBTeff (6.23) Zout

Thus, if one tunes Zs to Za = Zin, the modeu ˆin does not In contrast, our minimal bosonic amplifier has very dif- contribute to the total added noise, and one reaches the ferent input and output effective temperatures: quantum limit. Physically speaking, by matching the signal source to the input line, the “back-action” noise ~ω 2kBTeff,in = ZinS¯II = (6.24) described by Fˆa =u ˆin does not feed back into the input 2 2 of the amplifier. Note that achieving this matching ex- |λV | · S¯VV plicitly requires one to be far from weak coupling! Having 2kBTeff,out = = 2|G|~ω (6.25) Zout Zs = Za means that when we attach the amplifier to the signal source, we will dramatically increase the damping This huge difference in effective temperatures means that of the signal source. it is impossible for the system to possess “ideal” quan- 52 tum noise, and thus it cannot reach the weak-coupling B. Power matching versus noise matching quantum limit. While it implies that one is not at the quantum limit, In Sec. V, we saw that an important part of reach- the fact that Teff,in  Teff,out can nonetheless be viewed ing the quantum limit on the added noise of an amplifier as an asset. From a practical point of view, a large Teff,in (when used in the op-amp mode of operation) is to op- can be dangerous. Even though the direct effect of the timize the coupling strength to the amplifier. For a po- large Teff,in is offset by an appropriately weak coupling sition detector, this condition corresponds to tuning the to the amplifier (see Eq. (5.70) and following discussion), strength of the back-action damping γ to be much smaller this large Teff,in can also heat up other degrees of freedom than the intrinsic oscillator damping (cf. Eq. (5.69)). For if they couple strongly to the back-action noise of the am- a voltage amplifier, this condition corresponds to tuning plifier. This can in turn lead to unwanted heating of the the impedance of the signal source to be equal to the input system. As Teff,in is usually constant over a broad noise impedance (cf. Eq. (5.77)), an impedance which is range of frequencies, this unwanted heating effect can be much smaller than the amplifier’s input impedance (cf. quite bad. In the minimal amplifier discussed here, this Eq. (5.90)). problem is circumvented by having a small Teff,in. The√ In this subsection, we make the simple point that opti- only price that is payed is that the added noise will be 2 mizing the coupling (i.e. source impedance) to reach the the quantum limit value. We discuss this issue further in quantum limit is not the same as what one would do to Sec. VII. optimize the power gain. To understand this, we need to introduce another measure of power gain commonly used in the engineering community, the available power gain VII. REACHING THE QUANTUM LIMIT IN PRACTICE GP,avail. For simplicity, we will discuss this quantity in the context of a linear voltage amplifier, using the nota- A. Importance of QND measurements tions of Sec. V.F; it can be analogously defined for the position detector of Sec.V.E. GP,avail tells us how much The fact that QND measurements are repeatable is of power we are providing to an optimally matched output fundamental practical importance in overcoming detec- load relative to the maximum power we could in principle tor inefficiencies (Gambetta et al., 2007). A prototypical have extracted from the source. This is in marked con- example is the electron shelving technique (Nagourney trast to the power gain GP , which was calculated using et al., 1986; Sauter et al., 1986) used to measure trapped the actual power drawn at the amplifier input. ions. A related technique is used in present implementa- For the available power gain, we first consider Pin,avail. tions of ion-trap based quantum computation. Here the This is the maximum possible power that could be deliv- (extremely long-lived) hyperfine state of an ion is read ered to the input of the amplifier, assuming we optimized out via state-dependent optical fluorescence. With prop- both the value of the input impedance Zin and the load erly chosen circular polarization of the exciting laser, only impedance Zload while keeping Zs fixed. For simplicity, one hyperfine state fluoresces and the transition is cy- we will take all impedances to be real in our discussion. cling; that is, after fluorescence the ion almost always re- In general, the power drawn at the input of the amplifier 2 2 turns to the same state it was in prior to absorbing the ex- is given by Pin = vinZin/ (Zs + Zin) . Maximizing this citing photon. Hence the measurement is QND. Typical over Zin, we obtain the available input power Pin,avail: experimental parameters (Wineland et al., 1998) allow 6 2 the cycling transition to produce N ∼ 10 fluorescence vin Pin,avail = (7.1) photons. Given the photomultiplier quantum efficiency 4Zs and typically small solid angle coverage, only a very small numbern ¯d will be detected on average. The probability The maximum occurs for Zin = Zs. of getting zero detections (ignoring dark counts for sim- The output power supplied to the load Pout = 2 plicity) and hence misidentifying the hyperfine state is vload/Zload is calculated as before, keeping Zin and Zs P (0) = e−n¯d . Even for a very poor overall detection ef- distinct. One has: −5 ficiency of only 10 , we still haven ¯d = 10 and nearly 2  2 perfect fidelity F = 1 − P (0) ≈ 0.999955. It is important vout Zload Pout = to note that the total time available for measurement Zload Zout + Zload is not limited by the phase coherence time (T ) of the 2 λ2v2  Z 2 Z /Z qubit or by the measurement-induced dephasing (Gam- in in load out = 2 (7.2) betta et al., 2006; Korotkov, 2001a; Makhlin et al., 2001; Zout Zin + Zs (1 + Zload/Zout) Schuster et al., 2005), but rather only by the rate at which the qubit makes real transitions between measure- As a function of Zload, Pout is maximized when Zload = ment (ˆσz) eigenstates. In a perfect QND measurement Zout: there is no measurement-induced state mixing (Makhlin 2 2  2 et al., 2001) and the relaxation rate 1/T1 is unaffected λ vin Zin Pout,max = (7.3) by the measurement process. 4Zout Zin + Zs 53

The available power gain is now defined as: calls for the application of non-equilibrium cooling tech- niques which exploit back-action to reduce the effective 2  2 Pout,max λ Zs Zin temperature of the mechanical device, a technique that GP,avail ≡ = Pin,avail Zout Zin + Zs has been demonstrated recently both in superconducting circuits and optomechanical setups (see Marquardt and 4Zs/Zin = GP (7.4) Girvin (2009) for a review). (1 + Z /Z )2 s in The ability to do quantum limited position detection

We see that GP,avail is strictly less than or equal to the will in turn open up many new interesting avenues of re- power gain GP ; equality is only achieved when Zs = Zin search. Among the most significant, it will allow the pos- (i.e. when the source impedance is “power matched” to sibility of doing quantum feedback control (Wiseman and the input of the amplifier). The general situation where Milburn, 1993, 1994), where one uses the continuously- GP,avail < GP indicates that we are not drawing as much obtained measurement output to tailor the state of the power from the source as we could, and hence the actual mechanical resonator. The relevant theoretical frame- power supplied to the load is not as large as it could be. work is that of quantum conditional evolution and quan- Consider now a situation where we have achieved the tum trajectories (see e.g. Brun, 2002; Jacobs and Steck, quantum limit on the added noise. This necessarily 2006), where one tracks the state of a measured quan- means that we have “noise matched”, i.e. taken Zs to be tum system in a particular run of the experiment. The equal to the noise impedance ZN. The available power application of these ideas has only recently been explored gain in this case is: in condensed matter contexts (Bern´ad et al., 2008; Goan and Milburn, 2001; Goan et al., 2001; Korotkov, 1999, 2 ZN p 2001b; Oxtoby et al., 2008, 2006). Fully understanding GP,avail ' λ ' 2 GP  GP (7.5) Zout the potential of these techniques, as well as differences that occur in condensed matter versus atomic physics We have used Eq. (5.90), which tells us that the noise contexts, remains an active area of research. Other im- impedance is smaller than the input impedance by a large √ portant directions in nanomechanics include the possi- factor 1/(2 G ). Thus, as reaching the quantum limit P bility of detecting quantum jumps in the state of a me- requires the use of a source impedance much smaller than chanical resonator via QND measurement of its energy Z , it results in a dramatic drop in the available power in (Jayich et al., 2008; Santamore et al., 2004a,b; Thomp- gain compared to the case where we “power match” (i.e. son et al., 2008), as well as the possibility of making take Z = Z ). In practice, one must decide whether it s in backaction evading measurements (cf. Sec. V.H). Back- is more important to minimize the added noise, or max- action evasion using a microwave cavity detector cou- imize the power provided at the output of the amplifier: pled to a nanomechanical resonator was recently reported one cannot do both at the same time. (Hertzberg et al., 2010). Another area distinct from nanomechanics where rapid VIII. CONCLUSIONS progress is being made is the readout of solid state qubits using microwave signals sent through cavities whose In this review, we have given an introduction to quan- transmission properties are controlled by the qubit. At tum limits for position detection and amplification, lim- the moment, one is close to achieving good fidelity single- its which are tied to fundamental constraints on quantum shot QND readout, which is a prerequisite for a large noise correlators. We wish to end by briefly emphasiz- number of applications in quantum information process- ing notable current developments and pointing out future ing. The gradually growing information about the qubit perspectives in the field. state is extracted from the measured noisy microwave sig- As we have repeatedly emphasized, much of our dis- nal trace, leading to a corresponding collapse of the qubit cussion has been directly relevant to the measurement state. This process can also be described by conditional of mechanical nanoresonators, a topic of considerable re- quantum evolution and quantum trajectories. cent attention. These nanoresonators are typically stud- A promising method for superconducting qubit read- ied by coupling them either to electrical (often super- out currently employed is a so-called “latching mea- conducting) circuits or to optical cavities. A key goal is surement”, where the hysteretic behaviour of a strongly to achieve quantum-limited continuous position detection driven anharmonic system (e.g. a Josephson junction) is (cf. Sec. V.E); current experiments are coming tantalyz- exploited to toggle between two states depending on the ingly close to this limit (cf. Table III). Although being qubit state (Lupa¸scu et al., 2006; Siddiqi et al., 2004). able to follow the nanoresonator’s motion with a precision Although this is then no longer a linear measurement set by the quantum limit is in principle independent from scheme and therefore distinct from what we have dis- being at low temperatures, it becomes interesting only cussed in this review, it can be turned into a linear am- when the systems are near their ground state; one could plifier for a sufficiently weak input signal. An interesting then, e.g., monitor the oscillator’s zero-point fluctuations and important open question is whether such a setup can (cf. Sec. III.B.3). Given the comparatively small values reach the quantum limit on linear amplification. of mechanical frequencies (mostly less than GHz), this Both qubit detection and mechanical measurements 54 in electrical circuits would benefit from quantum-limited define its ‘windowed’ Fourier transform as follows: on-chip amplifiers. Such amplifiers are now being devel- Z +T/2 oped using the tools of circuit quantum electrodynamics, 1 iωt VT [ω] = √ dt e V (t), (A2) employing Josephson junctions or SQUIDs coupled to mi- T −T/2 crowave transmission line cavities (Bergeal et al., 2008; Castellanos-Beltran et al., 2008). Such an amplifier has where T is the sampling time. In the limit T  τc the integral is a sum of a large number N ≈ T of random already been used to perform continuous position detec- τc tion with a measurement imprecision below the SQL level uncorrelated terms. We can think of the value of the (Teufel et al., 2009). integral as the end point of a random walk in the complex plane which starts at the√ origin. Because the distance traveled will scale with T , our choice of normalization makes the statistical properties of V [ω] independent of Acknowledgements the sampling time T (for sufficiently large√ T ). Notice that VT [ω] has the peculiar√ units of volts secs which is This work was supported in part by NSERC, the Cana- usually denoted volts/ Hz. dian Institute for Advanced Research (CIFAR), NSA The spectral density (or ‘power spectrum’) of the noise under ARO contract number W911NF-05-1-0365, the is defined to be the ensemble averaged quantity NSF under grants DMR-0653377 and DMR-0603369, the 2 David and Lucile Packard Foundation, the W.M. Keck SVV [ω] ≡ lim h|VT [ω]| i = lim hVT [ω]VT [−ω]i (A3) Foundation and the Alfred P. Sloan Foundation. We ac- T →∞ T →∞ knowledge the support and hospitality of the Aspen Cen- The second equality follows from the fact that V (t) is ter for Physics where part of this work was carried out. real valued. The Wiener-Khinchin theorem (derived in F.M. acknowledges support via DIP, GIF and through Appendix A.2) tells us that the spectral density is equal SFB 631, SFB/TR 12, the Nanosystems Initiative Mu- to the Fourier transform of the autocorrelation function nich, and the Emmy-Noether program. S. M. G. acknowl- Z +∞ edges the support of the Centre for Advanced Study at iωt the Norwegian Academy of Science and Letters where SVV [ω] = dt e GVV (t). (A4) −∞ part of this work was carried out. The inverse transform relates the autocorrelation func- tion to the power spectrum

Appendix A: Basics of Classical and Quantum Noise Z +∞ dω G (t) = e−iωtS [ω]. (A5) VV 2π VV 1. Classical noise correlators −∞ We thus see that a short auto-correlation time implies Consider a classical random voltage signal V (t). The a spectral density which is non-zero over a wide range of signal is characterized by zero mean hV (t)i = 0, and frequencies. In the limit of ‘white noise’ autocorrelation function 2 GVV (t) = σ δ(t) (A6) 0 0 GVV (t − t ) = hV (t)V (t )i (A1) the spectrum is flat (independent of frequency) whose sign and magnitude tells us whether the voltage 2 fluctuations at time t and time t0 are correlated, anti- SVV [ω] = σ (A7) correlated or statistically independent. We assume that In the opposite limit of a long autocorrelation time, the the noise process is stationary (i.e., the statistical proper- signal is changing slowly so it can only be made up out ties are time translation invariant) so that GVV depends of a narrow range of frequencies (not necessarily centered only on the time difference. If V (t) is Gaussian dis- on zero). tributed, then the mean and autocorrelation completely Because V (t) is a real-valued classical variable, it natu- specify the statistical properties and the probability dis- rally follows that GVV (t) is always real. Since V (t) is not tribution. We will assume here that the noise is due to a quantum operator, it commutes with its value at other the sum of a very large number of fluctuating charges times and thus, hV (t)V (t0)i = hV (t0)V (t)i. From this it so that by the central limit theorem, it is Gaussian dis- follows that GVV (t) is always symmetric in time and the tributed. We also assume that GVV decays (sufficiently power spectrum is always symmetric in frequency rapidly) to zero on some characteristic correlation time scale τ which is finite. c SVV [ω] = SVV [−ω]. (A8) The spectral density of the noise as measured by a spectrum analyzer is a measure of the intensity of the As a prototypical example of these ideas, let us con- signal at different frequencies. In order to understand sider a simple harmonic oscillator of mass M and fre- the spectral density of a random signal, it is useful to quency Ω. The oscillator is maintained in equilibrium 55 with a large heat bath at temperature T via some in- This proves the Wiener-Khinchin theorem stated in finitesimal coupling which we will ignore in considering Eq. (A4). the dynamics. The solution of Hamilton’s equations of A useful application of these ideas is the following. motion are Suppose that we have a noisy signal V (t) = V¯ + η(t) which we begin monitoring at time t = 0. The integrated 1 x(t) = x(0) cos(Ωt) + p(0) sin(Ωt) signal up to time t is given by MΩ p(t) = p(0) cos(Ωt) − x(0)MΩ sin(Ωt), (A9) Z T I(T ) = dt V (t) (A15) where x(0) and p(0) are the (random) values of the po- 0 sition and momentum at time t = 0. It follows that the and has mean position autocorrelation function is hI(T )i = VT.¯ (A16) Gxx(t) = hx(t)x(0)i (A10) Provided that the integration time greatly exceeds the 1 = hx(0)x(0)i cos(Ωt) + hp(0)x(0)i sin(Ωt). autocorrelation time of the noise, I(T ) is a sum of a large MΩ number of uncorrelated random variables. The central limit theorem tells us in this case that I(t) is gaussian Classically in equilibrium there are no correlations be- distributed even if the signal itself is not. Hence the tween position and momentum. Hence the second term 1 2 2 probability distribution for I is fully specified by its mean vanishes. Using the equipartition theorem 2 MΩ hx i = 1 and its variance kBT , we arrive at 2 Z T h(∆I)2i = dtdt0 hη(t)η(t0)i. (A17) kBT G (t) = cos(Ωt) (A11) 0 xx MΩ2 From the definition of spectral density above we have the which leads to the spectral density simple result that the variance of the integrated signal grows linearly in time with proportionality constant given k T S [ω] = π B [δ(ω − Ω) + δ(ω + Ω)] (A12) by the noise spectral density at zero frequency xx MΩ2 h(∆I)2i = S [0] T. (A18) which is indeed symmetric in frequency. VV As a simple application, consider the photon shot noise of a coherent laser beam. The total number of photons 2. The Wiener-Khinchin Theorem detected in time T is Z T From the definition of the spectral density in Eqs.(A2- N(T ) = dt N˙ (t). (A19) A3) we have 0

Z T Z T The photo-detection signal N˙ (t) is not gaussian, but 1 0 iω(t−t0) 0 SVV [ω] = dt dt e hV (t)V (t )i rather is a point process, that is, a sequence of delta func- T 0 0 tions with random Poisson distributed arrival times and T +2B(t) 1 Z Z mean photon arrival rate N˙ . Nevertheless at long times = dt dτ eiωτ hV (t + τ/2)V (t − τ/2)i T 0 −2B(t) the mean number of detected photons (A13) hN(T )i = NT˙ (A20) where will be large and the photon number distribution will be gaussian with variance B (t) = t if t < T/2 2 = T − t if t > T/2. h(∆N) i = SN˙ N˙ T. (A21) Since we know that for a Poisson process the variance is If T greatly exceeds the noise autocorrelation time τc then it is a good approximation to extend the bound B(t) equal to the mean in the second integral to infinity, since the dominant con- h(∆N)2i = hN(T )i, (A22) tribution is from small τ. Using time translation invari- ance gives it follows that the shot noise power spectral density is

1 Z T Z +∞ S (0) = N.˙ (A23) S [ω] = dt dτ eiωτ hV (τ)V (0)i N˙ N˙ VV T 0 −∞ Since the noise is white this result happens to be valid at Z +∞ all frequencies, but the noise is gaussian distributed only = dτ eiωτ hV (τ)V (0)i . (A14) −∞ at low frequencies. 56

3. Square law detectors and classical spectrum analyzers we obtain

Now that we understand the basics of classical noise, hIi = SVV (−ω0) + SVV (ω0) (A27) we can consider how one experimentally measures a clas- sical noise spectral density. With modern high speed showing as expected that the classical square law detector digital sampling techniques it is perfectly feasible to di- measures the symmetrized noise power. rectly measure the random noise signal as a function of We thus have two very different basic approaches for time and then directly compute the autocorrelation func- the measurement of classical noise spectral densities: for tion in Eq. (A1). This is typically done by first per- low RF frequencies, one can directly measure the noise forming an analog-to-digital conversion of the noise sig- autocorrelation, whereas for high microwave frequencies, nal, and then numerically computing the autocorrelation one uses a filter and a square law detector. For noise function. One can then use Eq. (A4) to calculate the signals in intermediate frequency ranges, a combination noise spectral density via a numerical Fourier transform. of different methods is generally used. The whole story Note that while Eq. (A4) seems to require an ensemble becomes even more complicated, as at very high frequen- average, in practice this is not explicitly done. Instead, cies (e.g. in the far infrared), devices such as the so- one uses a sufficiently long averaging time T (i.e. much called ‘Fourier Transform spectrometer’ are in fact based longer than the correlation time of the noise) such that on a direct measurement of the equivalent of an auto- a single time-average is equivalent to an ensemble aver- correlation function of the signal. In the infrared, visible age. This approach of measuring a noise spectral density and ultraviolet, noise spectrometers use gratings followed directly from its autocorrelation function is most appro- by a slit acting as a filter. priate for signals at RF frequencies well below 1 MHz. For microwave signals with frequencies well above 1 GHz, a very different approach is usually taken. Here, the Appendix B: Quantum Spectrum Analyzers: Further Details standard route to obtain a noise spectral density involves first shifting the signal to a lower intermediate frequency 1. Two-level system as a spectrum analyzer via a technique known as heterodyning (we discuss this more in Sec. B.3.c). This intermediate-frequency signal In this sub-appendix, we derive the Golden Rule tran- is then sent to a filter which selects a narrow frequency sition rates Eqs. (2.6) describing a quantum two-level sys- range of interest, the so-called ‘resolution bandwidth’. tem coupled to a noise source (cf. Sec. II.B). Our deriva- Finally, this filtered signal is sent to a square-law detector tion is somewhat unusual, in that the role of the contin- (e.g. a diode), and the resulting output is averaged over uum as a noise source is emphasized from the outset. We a certain time-interval (the inverse of the so-called ‘video start by treating the noise F (t) in Eq. (2.5) as being a bandwidth’). It is this final output which is then taken classically noisy variable. We assume that the coupling to be a measure of the noise spectral density. A is under our control and can be made small enough It helps to put the above into equations. Ignoring for that the noise can be treated in lowest order perturba- simplicity the initial heterodyning step, let tion theory. We take the state of the two-level system to be Vf [ω] = f[ω]V [ω] (A24) ! α (t) be the voltage at the output of the filter and the input |ψ(t)i = g . (B1) of the square law detector. Here, f[ω] is the (ampli- αe(t) tude) transmission coefficient of the filter and V [ω] is the Fourier transform of the noisy signal we are measuring. In the interaction representation, first-order time- From Eq. (A5) it follows that the output of the square dependent perturbation theory gives law detector is proportional to i Z t Z +∞ ˆ dω 2 |ψI(t)i = |ψ(0)i − dτ V (τ)|ψ(0)i. (B2) hIi = |f[ω]| SVV [ω]. (A25) ~ 0 −∞ 2π Approximating the narrow band filter centered on fre- If we initially prepare the two-level system in its ground quency ±ω as20 state, the amplitude to find it in its excited state at time 0 t is from Eq. (B2) 2 |f[ω]| = δ(ω − ω0) + δ(ω + ω0) (A26) iA Z t αe = − dτ he|σˆx(τ)|giF (τ), ~ 0 iA Z t 20 iω01τ A linear passive filter performs a convolution Vout(t) = = − dτ e F (τ). (B3) R +∞ 0 0 0 ~ 0 −∞ dt F (t − t )Vin(t ) where F is a real-valued (and causal) function. Hence it follows that f[ω], which is the Fourier trans- form of F , obeys f[−ω] = f ∗[ω] and hence |f[ω]|2 is symmetric Since the integrand in Eq. (B3) is random, αe is a sum in frequency. of a large number of random terms; i.e. its value is the 57

endpoint of a random walk in the complex plane (as dis- We now see that our two-level system does indeed act as a cussed above in defining the spectral density of classical quantum spectrum analyzer for the noise. Operationally, noise). As a result, for times exceeding the autocorre- we prepare the system either in its ground state or in its lation time τc of the noise, the integral will not grow excited state, weakly couple it to the noise source, and linearly with time but rather only as the square root of after an appropriate interval of time (satisfying the above time, as expected for a random walk. We can now com- inequalities) simply measure whether the system is now pute the probability in its excited state or ground state. Repeating this pro- tocol over and over again, we can find the probability A2 Z t Z t 2 −iω01(τ1−τ2) of making a transition, and thereby infer the rate and pe(t) ≡ |αe| = 2 dτ1dτ2 e F (τ1)F (τ2) ~ 0 0 hence the noise spectral density at positive and nega- (B4) tive frequencies. Naively one imagines that a spectrom- which we expect to grow quadratically for short times eters measures the noise spectrum by extracting a small t < τc, but linearly for long times t > τc. Ensemble amount of the signal energy from the noise source and averaging the probability over the random noise yields analyzes it. This is not the case however. There must be energy flowing in both directions if the noise is to be A2 Z t Z t −iω01(τ1−τ2) fully characterized. p¯e(t) = 2 dτ1dτ2 e hF (τ1)F (τ2)i ~ 0 0 We now rigorously treat the quantity Fˆ(τ) as a quan- (B5) tum Heisenberg operator which acts in the Hilbert space Introducing the noise spectral density of the noise source. The previous derivation is unchanged ˆ ˆ Z +∞ (the ordering of F (τ1)F (τ2) having been chosen cor- iωτ SFF (ω) = dτ e hF (τ)F (0)i, (B6) rectly in anticipation of the quantum treatment), and −∞ Eqs. (2.6a,2.6b) are still valid provided that we interpret the angular brackets in Eq. (B5,B6) as representing a and utilizing the Fourier transform defined in Eq. (A2) quantum expectation value (evaluated in the absence of and the Wiener-Khinchin theorem from Appendix A.2, the coupling to the spectrometer): we find that the probability to be in the excited state 21 indeed increases linearly with time at long times, +∞ Z X S (ω) = dτ eiωτ ρ hα|Fˆ(τ)|γihγ|Fˆ(0)|αi. 2 FF αα A −∞ α,γ p¯e(t) = t SFF (−ω01) (B7) ~2 (B10) Here, we have assumed a stationary situation, where The time derivative of the probability gives the transition the density matrix ρ of the noise source is diagonal in rate from ground to excited states the energy eigenbasis (in the absence of the coupling to A2 the spectrometer). However, we do not necessarily as- Γ↑ = 2 SFF (−ω01) (B8) sume that it is given by the equilibrium expression. This ~ yields the standard quantum mechanical expression for Note that we are taking in this last expression the spec- the spectral density: tral density on the negative frequency side. If F were a +∞ strictly classical noise source, hF (τ)F (0)i would be real, Z i iωτ X (α−γ )τ ˆ 2 SFF (ω) = dτ e ραα e ~ |hα|F |γi| and SFF (−ω01) = SFF (+ω01). However, because as we −∞ discuss below F is actually an operator acting on the en- α,γ h i X ˆ 2 vironmental degrees of freedom, Fˆ(τ), Fˆ(0) 6= 0 and = 2π~ ραα |hα|F |γi| δ(γ − α − ~ω)(B11). α,γ SFF (−ω01) 6= SFF (+ω01). Another possible experiment is to prepare the two-level Substituting this expression into Eqs. (2.6a,2.6b), we de- system in its excited state and look at the rate of decay rive the familiar Fermi Golden Rule expressions for the into the ground state. The algebra is identical to that two transition rates. above except that the sign of the frequency is reversed: In standard courses, one is not normally taught that A2 the transition rate of a discrete state into a continuum Γ↓ = 2 SFF (+ω01). (B9) as described by Fermi’s Golden Rule can (and indeed ~ should!) be viewed as resulting from the continuum act- ing as a quantum noise source which causes the am- plitudes of the different components of the wave func- 21 Note that for very long times, where there is a significant de- tion to undergo random walks. The derivation presented pletion of the probability of being in the initial state, first-order here hopefully provides a motivation for this interpreta- perturbation theory becomes invalid. However, for sufficiently tion. In particular, thinking of the perturbation (i.e. the small A, there is a wide range of times τc  t  1/Γ for which Eq. B7 is valid. Eqs. (2.6a) and (2.6b) then yield well-defined coupling to the continuum) as quantum noise with a rates which can be used in a master equation to describe the full small but finite autocorrelation time (inversely related dynamics including long times. to the bandwidth of the continuum) neatly explains why 58 the transition probability increases quadratically for very Given these transition rates, we may immediately write short times, but linearly for very long times. a simple master equation for the probability pn(t) that It it is important to keep in mind that our expressions there are n quanta in the oscillator: for the transition rates are only valid if the autocorrela- tion time of our noise is much shorter that the typical d p = [nΓ p + (n + 1)Γ p ] time we are interested in; this typical time is simply the dt n ↑ n−1 ↓ n+1 inverse of the transition rate. The requirement of a short − [nΓ↓ + (n + 1)Γ↑] pn (B14) autocorrelation time in turn implies that our noise source must have a large bandwidth (i.e. there must be large The first two terms describe transitions into the state |ni number of available photon frequencies in the vacuum) from the states |n + 1i and |n − 1i, and hence increase and must not be coupled too strongly to our system. This pn. In contrast, the last two terms describe transitions is true despite the fact that our final expressions for the out of the state |ni to the states |n + 1i and |n − 1i, and transition rates only depend on the spectral density at hence decrease pn. The stationary state of the oscillator the transition frequency (a consequence of energy con- is given by solving Eq. (B14) for d p = 0, yielding: servation). dt n One standard model for the continuum is an infinite   −n Ω/(kBTeff ) − Ω/(kBTeff ) collection of harmonic oscillators. The electromagnetic pn = e ~ 1 − e ~ (B15) continuum in the hydrogen atom case mentioned above is a prototypical example. The vacuum electric field noise where the effective temperature T [Ω] is defined in coupling to the hydrogen atom has an extremely short eff Eq. (2.8). Eq. (B15) describes a thermal equilibrium autocorrelation time because the range of mode frequen- distribution of the oscillator, with an effective oscillator cies ωα (over which the dipole matrix element coupling temperature Teff [Ω] determined by the quantum noise the atom to the mode electric field E~ is significant) is α spectrum of Fˆ. This is the same effective temperature extremely large, ranging from many times smaller than that emerged in our discussion of the TLS spectrum an- the transition frequency to many times larger. Thus, the alyzer. As we have seen, if the noise source is in thermal autocorrelation time of the vacuum electric field noise is equilibrium at a temperature T , then T [Ω] = T . considerably less than 10−15s, whereas the decay time of eq eff eq In the more general case where the noise source is not the hydrogen 2p state is about 10−9s. Hence the inequal- in thermal equilibrium, T only serves to characterize ities needed for the validity of our expressions are very eff the asymmetry of the quantum noise, and will vary with easily satisfied. frequency 22. We can learn more about the quantum noise spectrum ˆ 2. Harmonic oscillator as a spectrum analyzer of F by also looking at the dynamics of the oscillator. In particular, as the average energy hEi of the oscillator P∞ 1  We now provide more details on the system described is just given by hE(t)i = n=0 ~Ω n + 2 pn(t), we can in Sec. II.B, where a harmonic oscillator acts as a spec- use the master equation Eq. (B14) to derive an equa- trometer of quantum noise. We start with the coupling tion for its time dependence. One thus finds Eq. (2.10). Hamiltonian givein in Eq. (2.9). In analogy to the TLS By demanding dhEi/dt = 0 in this equation, we find spectrometer, noise in Fˆ at the oscillator frequency Ω that the combination of damping and heating effects can cause transitions between its eigenstates. We as- causes the energy to reach a steady state mean value sume both that A is small, and that our noise source of hEi = P/γ. This implies that the finite ground state has a short autocorrelation time, so we may again use energy hEi = ~Ω/2 of the oscillator is determined via the perturbation theory to derive rates for these transitions. balance between the ‘heating’ by the zero-point fluctua- There is a rate for increasing the number of quanta in tions of the environment (described by the symmetrized the oscillator by one, taking a state |ni to |n + 1i: correlator at T = 0) and the dissipation. It is possible to take an alternative but equally correct viewpoint, where 2 only the deviation hδEi = hEi − Ω/2 from the ground A  2  ~ Γn→n+1 = (n + 1)xZPF SFF [−Ω] ≡ (n + 1)Γ↑ state energy is considered. Its evolution equation ~2 (B12) As expected, this rate involves the noise at −Ω, as en- d hδEi = hδEi(Γ − Γ ) + Γ Ω (B16) ergy is being absorbed from the noise source. Similarly, dt ↑ ↓ ↑~ there is a rate for decreasing the number of quanta in the oscillator by one: only contains a decay term at T = 0, leading to hδEi → 0.

2 A 2  Γn→n−1 = nxZPF SFF [Ω] ≡ nΓ↓ (B13) ~2 22 Note that the effective temperature can become negative if the This rate involves the noise at +Ω, as energy is being noise source prefers emitting energy versus absorbing it; in the emitted to the noise source. present case, that would lead to an instability. 59

3. Practical quantum spectrum analyzers average energy of the resonant filter circuit, taking this quantity as a proxy for the output of the diode. Sure As we have seen, a ‘quantum spectrum analyzer’ can enough, if we subject the filter circuit to purely classical in principle be constructed from a two level system (or noise, it would cause the average energy of the circuit a harmonic oscillator) in which we can separately mea- hEi to increase an amount directly proportional to the sure the up and down transition rates between states classical spectrum SFF [Ω]. We now consider hEi in the differing by some precise energy ~ω > 0 given by the case of a quantum noise source, and ask how it relates to frequency of interest. The down transition rate tells us the quantum noise spectral density SFF [Ω]. the noise spectral density at frequency +ω and the up The quantum case is straightforward to analyze using transition rate tells us the noise spectral density at −ω. the approach of Sec. B.2. Unlike the classical case, the While we have already discussed experimental implemen- noise will both lead to additional fluctuations of the filter tation of these ideas using two-level systems and oscilla- circuit and increase its damping rate by an amount γ tors, similar schemes have been implemented in other sys- (c.f. Eq. (2.12)). To make things quantitative, we let neq tems. A number of recent experiments have made use of denote the average number of quanta in the filter circuit superconductor-insulator-superconductor junctions (Bil- prior to coupling to Fˆ(t), i.e. langeon et al., 2006; Deblock et al., 2003; Onac et al., 2006) to measure quantum noise, as the current-voltage 1 neq =   , (B17) characteristics of such junctions are very sensitive to ~Ω exp k T − 1 the absorption or emission of energy (so-called photon- B eq assisted transport processes). It has also been suggested and let neff represent the Bose-Einstein factor associated that tunneling of flux in a SQUID can be used to measure with the effective temperature Teff [Ω] of the noise source quantum noise (Amin and Averin, 2008). Fˆ(t), In this subsection, we discuss additional methods for the detection of quantum noise. Recall from Sec. A.3 that 1 neff =   . (B18) one of the most basic classical noise spectrum analyzers exp ~Ω − 1 consists of a linear narrow band filter and a square law kB Teff [Ω] detector such as a diode. In what follows, we will consider One then finds (Gavish et al., 2000; Lesovik and Loosen, a simplified quantum treatment of such a device where we 1997): do not explicitly model a diode, but instead focus on the energy of the filter circuit. We then turn to various noise γ ∆hEi = ~Ω · (neff − neq) (B19) detection schemes making use of a photomultiplier. We γ0 + γ will show that depending on the detection scheme used, one can measure either the symmetrized quantum noise This equation has an extremely simple interpretation: spectral density S¯[ω], or the non-symmetrized spectral the first term results from the expected heating effect density S[ω]. of the noise, while the second term results from the noise source having increased the circuit’s damping by an amount γ. Re-expressing this result in terms of the symmetric and anti-symmetric in frequency parts of the a. Filter plus diode quantum noise spectral density SFF [Ω], we have: Using the simple treatment we gave of a harmonic os- S¯ (Ω) − n + 1  (S [Ω] − S [−Ω]) ∆hEi = FF eq 2 FF FF cillator as a quantum spectrum analyzer in Sec. B.2, one 2m (γ + γ) can attempt to provide a quantum treatment of the clas- 0 sical ‘filter plus diode’ spectrum analyzer discussed in (B20) Sec. A.3. This approach is due to Lesovik and Loosen We see that ∆hEi is in general not simply proportional (1997) and Gavish et al. (2000). The analysis starts by to the symmetrized noise S¯FF [Ω]. Thus, the ‘filter plus modeling the spectrum analyzer’s resonant filter circuit diode’ spectrum analyzer does not simply measure the as a harmonic oscillator of frequency Ω weakly coupled symmetrized quantum noise spectral density. We stress to some equilibrium dissipative bath. The oscillator thus that there is nothing particularly quantum about this re- has an intrinsic damping rate γ0  Ω, and is initially at sult. The extra term on the RHS of Eq. (B20) simply a finite temperature Teq. One then drives this damped reflects the fact that coupling the noise source to the fil- oscillator (i.e. the filter circuit) with the noisy quantum ter circuit could change the damping of this circuit; this ˆ force F (t) whose spectrum at frequency Ω is to be mea- could easily happen in a completely classical setting. As sured. long as this additional damping effect is minimal, the In the classical ‘filter plus diode’ spectrum analyzer, second term in Eq. (B20) will be minimal, and our spec- the output of the filter circuit was sent to a square law trum analyzer will (to a good approximation) measure detector, whose time-averaged output was then taken as the symmetrized noise. Quantitatively, this requires: the measured spectral density. To simplify the analy- sis, we can instead consider how the noise changes the neff  neq. (B21) 60

We now see where quantum mechanics enters: if the noise c. Double sideband heterodyne power spectrum to be measured is close to being zero point noise (i.e. neff → 0), the above condition can never be satisfied, and At RF and microwave frequencies, practical spectrome- thus it is impossible to ignore the damping effect of the ters often contain heterodyne stages which mix the initial noise source on the filter circuit. In the zero point limit, frequency down to a lower frequency ωIF (possibly in the this damping effect (i.e. second term in Eq. (B20)) will classical regime). Consider a system with a mixer and lo- always be greater than or equal to the expected heating cal oscillator at frequency ωLO that mixes both the upper effect of the noise (i.e. first term in Eq. (B20)). sideband input at ωu = ωLO +ωIF and the lower sideband input at ωl = ωLO − ωIF down to frequency ωIF. This can be achieved by having a Hamiltonian with a 3-wave mixing term which (in the rotating wave approximation) b. Filter plus photomultiplier is given by

† † † † † † We now turn to quantum spectrum analyzers involving V = λ[ˆaIFaˆlaˆLO +a ˆIFaˆl aˆLO] + λ[ˆaIFaˆuaˆLO +a ˆIFaˆuaˆLO] a square law detector we can accurately model– a photo- (B24) multiplier. As a first example of such a system, consider a The interpretation of this term is that of a Raman pro- photomultiplier with a narrow band filter placed in front cess. Notice that there are two energy conserving pro- of it. The mean photocurrent is then given by cesses that can create an IF photon which could then activate the photodetector. First, one can absorb an LO Z +∞ photon and emit two photons, one at the IF and one at 2 hIi = dω |f[ω]| r[ω]SVV[ω], (B22) the lower sideband. The second possibility is to absorb −∞ an upper sideband photon and create IF and LO photons. Thus we expect from this that the power in the IF chan- where f is the filter (amplitude) transmission function nel detected by a photomultiplier would be proportional defined previously and r[ω] is the response of the pho- to the noise power in the following way todetector at frequency ω, and SVV represents the elec- tric field spectral density incident upon the photodetec- I ∝ S[+ωl] + S[−ωu] (B25) tor. Naively one thinks of the photomultiplier as a square law detector with the square of the electric field repre- since creation of an IF photon involves the signal source senting the optical power. However, according to the either absorbing a lower sideband photon from the mixer Glauber theory of (ideal) photo-detection (Gardiner and or the signal source emitting an upper sideband photon Zoller, 2000; Glauber, 2006; Walls and Milburn, 1994), into the mixer. In the limit of small IF frequency this photocurrent is produced if, and only if, a photon is expression would reduce to the symmetrized noise power absorbed by the detector, liberating the initial photo- ¯ electron. Glauber describes this in terms of normal or- I ∝ S[+ωLO] + S[−ωLO] = 2S[ωLO] (B26) dering of the photon operators in the electric field auto- correlation function. In our language of noise power at which is the same as for a ‘classical’ spectrum analyzer positive and negative frequencies, this requirement be- with a square law detector (c.f. Appendix A.3). For equi- comes simply that r[ω] vanishes for ω > 0. Approximat- librium noise spectral density from a resistance R0 de- rived in Appendix D we would then have ing the narrow band filter centered on frequency ±ω0 as in Eq. (A26), we obtain SVV[ω] + SVV[−ω] = 2R0~|ω|[2nB(~|ω|) + 1], (B27)

hIi = r[−ω0]SVV[−ω0] (B23) Assuming our spectrum analyzer has high input impedance so that it does not load the noise source, this which shows that this particular realization of a quantum voltage spectrum will determine the output signal of the spectrometer only measures electric field spectral density analyzer. This symmetrized quantity does not vanish at at negative frequencies since the photomultiplier never zero temperature and the output contains the vacuum emits energy into the noise source. Also one does not noise from the input. This vacuum noise has been seen see in the output any ‘vacuum noise’ and so the output in experiment. (Schoelkopf et al., 1997) (ideally) vanishes as it should at zero temperature. Of course real photomultipliers suffer from imperfect quan- tum efficiencies and have non-zero dark current. Note Appendix C: Modes, Transmission Lines and Classical that we have assumed here that there are no additional Input/Output Theory fluctuations associated with the filter circuit. Our re- sult thus coincides with what we found in the previous In this appendix we introduce a number of important subsection for the ‘filter plus diode’ spectrum analyzer classical concepts about electromagnetic signals which (c.f. Eq. (B20), in the limit where the filter circuit is ini- are essential to understand before moving on to the study tially at zero temperature (i.e. neq = 0). of their quantum analogs. A signal at carrier frequency 61

p ω can be described in terms of its amplitude and phase where Zc ≡ `/c is called the characteristic impedance or equivalently in terms of its two quadrature amplitudes of the line. In terms of the left and right propagating modes, Eqs. (C4) and C5 become s(t) = X cos(ωt) + Y sin(ωt). (C1) → → vp∂xV + ∂tV = 0 (C8) We will see in the following that the physical oscillations v ∂ V ← − ∂ V ← = 0 (C9) of this signal in a transmission line are precisely the sinu- p x t soidal oscillations of a simple harmonic oscillator. Com- √ where vp ≡ 1/ `c is the wave phase velocity. These parison of Eq. (C1) with x(t) = x0 cos ωt+(p0/Mω) sin ωt shows that we can identify the quadrature amplitude X equations have solutions which propagate by uniform with the coordinate of this oscillator and thus the quadra- translation without changing shape since the line is dis- ture amplitude Y is proportional to the momentum con- persionless jugate to X. Quantum mechanically, X and Y become → x operators Xˆ and Yˆ which do not commute. Thus their V (x, t) = Vout(t − ) (C10) vp quantum fluctuations obey the Heisenberg uncertainty ← x relation. V (x, t) = Vin(t + ), (C11) v Ordinarily (e.g., in the absence of squeezing), the phase p choice defining the two quadratures is arbitrary and so where V and V are arbitrary functions of their argu- their vacuum (i.e. zero-point) fluctuations are equal in out ments. For an infinite transmission line, Vout and Vin are completely independent. However for the case of a semi- X = Y . (C2) ZPF ZPF infinite line terminated at x = 0 (say) by some system S, Thus the canonical commutation relation becomes these two solutions are not independent, but rather re- lated by the boundary condition imposed by the system. ˆ ˆ 2 We have [X, Y ] = iXZPF. (C3)

We will see that the fact that X and Y are canoni- V (x = 0, t) = [Vout(t) + Vin(t)] (C12) cally conjugate has profound implications both classically 1 I(x = 0, t) = [Vout(t) − Vin(t)], (C13) and quantum mechanically. In particular, the action of Zc any circuit element (beam splitter, attenuator, amplifier, etc.) must preserve the Poisson bracket (or in the quan- from which we may derive tum case, the commutator) between the signal quadra- tures. This places strong constraints on the properties Vout(t) = Vin(t) + ZcI(x = 0, t). (C14) of these circuit elements and in particular, forces every amplifier to add noise to the signal. 0 d x 1. Transmission lines and classical input-output theory V (x,t) a) I(x,t) We begin by considering a coaxial transmission line Zc , vp V(x,t) modeled as a perfectly conducting wire with inductance per unit length of ` and capacitance to ground per unit V (x,t) length c as shown in Fig. 14. If the voltage at position x L LL at time t is V (x, t), then the charge density is q(x, t) = b) cV (x, t). By charge conservation the current I and the CCC charge density are related by the continuity equation a

∂tq + ∂xI = 0. (C4) I Vout The constitutive relation (essentially Newton’s law) gives c) the acceleration of the charges V

`∂tI = −∂xV. (C5) Vin

We can decouple Eqs. (C4) and (C5) by introducing left FIG. 14 a) Coaxial transmission line, indicating voltages and and right propagating modes currents as defined in the main text. b) Lumped element representation of a transmission line with capacitance per unit V (x, t) = [V → + V ←] (C6) length c = C/a and inductance per unit length ` = L/a. c) 1 Discrete LC resonator terminating a transmission line. I(x, t) = [V → − V ←] (C7) Zc 62 √ 2 If the system under study is just an open circuit so where the resonant frequency is Ω0 ≡ 1/ L0C0. Note that I(x = 0, t) = 0, then Vout = Vin, meaning that the that the term ZcI(x = 0, t) in Eq. (C17) results in the outgoing wave is simply the result of the incoming wave linear viscous damping rate γ ≡ Zc/L0. reflecting from the open circuit termination. In general If we solve the equation of motion of the oscillator, we however, there is an additional outgoing wave radiated can predict the outgoing field. In the present instance of by the current I that is injected by the system dynamics a simple oscillator we have a particular example of the into the line. In the absence of an incoming wave we have general case where the system responds linearly to the input field. We can characterize any such system by a V (x = 0, t) = ZcI(x = 0, t), (C15) complex, frequency dependent impedance Z[ω] defined by indicating that the transmission line acts as a simple re- V (x = 0, ω) sistor which, instead of dissipating energy by Joule heat- Z[ω] = − . (C20) ing, carries the energy away from the system as propa- I(x = 0, ω) gating waves. The fact that the line can dissipate energy Note the peculiar minus sign which results from our def- despite containing only purely reactive elements is a con- inition of positive current flowing to the right (out of the sequence of its infinite extent. One must be careful with system and into the transmission line). Using Eqs. (C12, the order of limits, taking the length to infinity before C13) and Eq. (C20) we have allowing time to go to infinity. In this way the outgoing waves never reach the far end of the transmission line and Vout[ω] = r[ω]Vin[ω], (C21) reflect back. Since this is a conservative Hamiltonian sys- tem, we will be able to quantize these waves and make a where the reflection coefficient r is determined by the quantum theory of resistors (Caldeira and Leggett, 1983) impedance mismatch between the system and the line in Appendix D. The net power flow carried to the right and is given by the well known result by the line is Z[ω] − Z r[ω] = c . (C22) 1 2 2 Z[ω] + Zc P = [Vout(t) − Vin(t)]. (C16) Zc If the system is constructed from purely reactive (i.e. The fact that the transmission line presents a dissipa- lossless) components, then Z[ω] is purely imaginary and tive impedance to the system means that it causes damp- the reflection coefficient obeys |r| = 1 which is consistent ing of the system. It also however opens up the possibility with Eq. (C16) and the energy conservation requirement of controlling the system via the input field which par- of no net power flow into the lossless system. For exam- tially determines the voltage driving the system. From ple, for the series LC oscillator we have been considering, this point of view it is convenient to eliminate the output we have field by writing the voltage as 1 Z[ω] = + jωL0, (C23) jωC0 V (x = 0, t) = 2Vin(t) + ZcI(x = 0, t). (C17) where, to make contact with the usual electrical engi- As we will discuss in more detail below, the first term neering sign conventions, we have used j = −i. If the drives the system and the second damps it. From damping γ of the oscillator induced by coupling it to the Eq. (C14) we see that measurement of the outgoing field transmission line is small, the quality factor of the reso- can be used to determine the current I(x = 0, t) injected nance will be high and we need only consider√ frequencies by the system into the line and hence to infer the system near the resonance frequency Ω0 ≡ 1/ L0C0 where the dynamics that results from the input drive field. impedance has a zero. In this case we may approximate As a simple example, consider the system consisting of 2 an LC resonator shown in Fig. (14 c). This can be viewed Z[ω] ≈ 2 [Ω0 − ω] = 2jL0(ω − Ω0) (C24) as a simple harmonic oscillator whose coordinate Q is the jC0Ω0 charge on the capacitor plate (on the side connected to which yields for the reflection coefficient L0). The current I(x = 0, t) = Q˙ plays the role of the ω − Ω + jγ/2 velocity of the oscillator. The equation of motion for the r[ω] = 0 (C25) oscillator is readily obtained from ω − Ω0 − jγ/2

Q = C [−V (x = 0+, t) − L I˙(x = 0+, t)]. (C18) showing that indeed |r| = 1 and that the phase of the 0 0 reflected signal winds by 2π upon passing through the 23 Using Eq. (C17) we obtain a harmonic oscillator damped resonance. by the transmission line and driven by the incoming waves

23 ¨ 2 ˙ 2 For the case of resonant transmission through a symmetric cav- Q = −Ω0Q − γQ − Vin(t), (C19) ity, the phase shift only winds by . L0 π 63

Turning to the more general case where the system The momentum conjugate to ϕ(x) is simply the charge also contains lossy elements, one finds that Z[ω] is no density longer purely imaginary, but has a real part satisfying Re Z[ω] > 0. This in turn implies via Eq. (C22) that δL q(x, t) ≡ = c∂tϕ = cV (x, t) (C30) |r| < 1. In the special case of impedance matching δ∂tϕ Z[ω] = Zc, all the incident power is dissipated in the system and none is reflected. The other two limits of in- and so the Hamiltonian is given by terest are open circuit termination with Z = ∞ for which Z  1 1  r = +1 and short circuit termination Z = 0 for which H = dx q2 + (∂ ϕ)2 . (C31) 2c 2` x r = −1. Finally, if the system also contains an active device We know from our previous results that the charge which has energy being pumped into it from a separate density consists of left and right moving solutions of ar- external source, it may under the right conditions be de- bitrary fixed shape. For example we might have for the scribed by an effective negative resistance Re Z[ω] < 0 right moving case over a certain frequency range. Eq. (C22) then gives |r| ≥ 1, implying |Vout| > |Vin|. Our system will thus act q(t−x/vp) = αk cos[k(x−vpt)]+βk sin[k(x−vpt)]. (C32) like the one-port amplifier discussed in Sec. V.D: it am- plifies signals incident upon it. We will discuss this idea A confusing point is that since q is real valued, we see of negative resistance further in Sec. C.4; a physical real- that it necessarily contains both eikx and e−ikx terms ization is provided by the two-port reflection parametric even if it is only right moving. Note however that for amplifier discussed in Appendix V.C. k > 0 and a right mover, the eikx is associated with the positive frequency term e−iωkt while the e−ikx term is associated with the negative frequency term e+iωkt where 2. Lagrangian, Hamiltonian, and wave modes for a ωk ≡ vp|k|. For left movers the opposite holds. We can transmission line appreciate this better if we define

Prior to moving on to the case of quantum noise it Z ( r 2 ) 1 −ikx 1 k is useful to review the classical statistical mechanics of Ak ≡ √ dx e √ q(x, t) − i ϕ(x, t) L 2c 2` transmission lines. To do this we need to write down the Lagrangian and then determine the canonical mo- (C33) menta and the Hamiltonian. Very conveniently, the sys- where for simplicity we have taken the fields to obey pe- tem is simply a large collection of harmonic oscillators riodic boundary conditions on a length L. Thus we have (the normal modes) and hence can be readily quantized. (in a form which anticipates the full quantum theory) This representation of a physical resistor is essentially the 1 X one used by Caldeira and Leggett (Caldeira and Leggett, H = (A∗A + A A∗) . (C34) 2 k k k k 1983) in their seminal studies of the effects of dissipation k on tunneling. The only difference between this model and the vacuum fluctuations in free space is that the rel- The classical equation of motion (C29) yields the simple ativistic bosons travel in one dimension and do not carry result a polarization label. This changes the density of states as ∂ A = −iω A . (C35) a function of frequency, but has no other essential effect. t k k k It is convenient to define a flux variable (Devoret, 1997) Thus Z t ϕ(x, t) ≡ dτ V (x, τ), (C26) q(x, t) −∞ r c X where V (x, t) = ∂ ϕ(x, t) is the local voltage on the trans- = eikx A (0)e−iωkt + A∗ (0)e+iωkt (C36) t 2L k −k mission line at position x and time t. Each segment of k r the line of length dx has inductance ` dx and the voltage c X h i = A (0)e+i(kx−ωkt) + A∗(0)e−i(kx−ωkt) . drop along it is −dx ∂x∂tϕ(x, t). The flux through this 2L k k k inductance is thus −dx ∂xϕ(x, t) and the local value of the current is given by the constitutive equation (C37) 1 I(x, t) = − ∂xϕ(x, t). (C27) We see that for k > 0 (k < 0) the wave is right (left) ` moving, and that for right movers the eikx term is asso- The Lagrangian for the system is ciated with positive frequency and the e−ikx term is as- Z ∞ Z ∞   sociated with negative frequency. We will return to this c 2 1 2 Lg ≡ dx L(x, t) = dx (∂tϕ) − (∂xϕ) , in the quantum case where positive (negative) frequency 0 0 2 2` (C28) will refer to the destruction (creation) of a photon. Note The Euler-Lagrange equation for this Lagrangian is sim- that the right and left moving voltages are given by ply the wave equation 2 2 2 vp∂xϕ − ∂t ϕ = 0. (C29) 64

r 1 X h i V → = A (0)e+i(kx−ωkt) + A∗(0)e−i(kx−ωkt) (C38) 2Lc k k k>0 r 1 X h i V ← = A (0)e+i(kx−ωkt) + A∗(0)e−i(kx−ωkt) (C39) 2Lc k k k<0 The voltage spectral density for the right moving waves is thus

2π X S→ [ω] = {hA A∗iδ(ω − ω ) + hA∗A iδ(ω + ω )} (C40) VV 2Lc k k k k k k k>0 The left moving spectral density has the same expression but k < 0. Using Eq. (C16), the above results lead to a net power flow (averaged over one cycle) within a frequency band defined by a pass filter G[ω] of

vp X P = P → − P ← = sgn(k)[G[ω ]hA A∗i + G[−ω ]hA∗A i] . (C41) 2L k k k k k k k

3. Classical statistical mechanics of a transmission line introduces into the circuit. Consider the example of a ter- minating resistor at the end of a transmission line. If the Now that we have the Hamiltonian, we can consider resistance R is matched to the characteristic impedance the classical statistical mechanics of a transmission line in Zc of a transmission line, the terminating resistor acts thermal equilibrium at temperature T . Since each mode as a black body because it absorbs 100% of the power k is a simple harmonic oscillator we have from Eq. (C34) incident upon it. If the resistor is held at temperature T and the equipartition theorem it will bring the transmission line modes into equilibrium at the same temperature (at least for the case where the ∗ hAkAki = kBT. (C42) transmission line has finite length). The rate at which the equilibrium is established will depend on the impedance Using this, we see from Eq. (C40) that the right moving mismatch between the resistor and the line, but the final voltage signal has a simple white noise power spectrum. temperature will not. Using Eq. (C41) we have for the right moving power in a bandwidth B (in Hz rather than radians/sec) the very A good way to understand the fluctuation-dissipation simple result theorem is to represent the resistor R which is terminat- ing the Zc line in terms of a second semi-infinite trans- → vp X ∗ ∗ mission line of impedance R as shown in Fig. (15). First P = hG[ωk]A Ak + G[−ωk]AkA i 2L k k consider the case when the R line is not yet connected to k>0 Z +∞ the Zc line. Then according to Eq. (C22), the open termi- kBT dω 2 = G[ω] nation at the end of the Zc line has reflectivity |r| = 1 2 −∞ 2π so that it does not dissipate any energy. Additionally = kBTB. (C43) of course, this termination does not transmit any sig- nals from the R line into the Zc. However when the where we have used the fact mentioned in connection two lines are connected the reflectivity becomes less than with Eq. (A26) and the discussion of square law detec- unity meaning that incoming signals on the Zc line see tors that all passive filter functions are symmetric in fre- a source of dissipation R which partially absorbs them. quency. The absorbed signals are not turned into heat as in a One of the basic laws of statistical mechanics is Kirch- true resistor but are partially transmitted into the R line hoff’s law stating that the ability of a hot object to emit which is entirely equivalent. Having opened up this port radiation is proportional to its ability to absorb. This for energy to escape from the Zc system, we have also follows from very general thermodynamic arguments con- allowed noise energy (thermal or quantum) from the R cerning the thermal equilibrium of an object with its ra- line to be transmitted into the Zc line. This is completely diation environment and it means that the best possible equivalent to the effective circuit shown in Fig. (16 a) in emitter is the black body. In electrical circuits this princi- which a real resistor has in parallel a random current ple is simply a form of the fluctuation dissipation theorem generator representing thermal noise fluctuations of the which states that the electrical thermal noise produced electrons in the resistor. This is the essence of the fluc- by a circuit element is proportional to the dissipation it tuation dissipation theorem. 65

In order to make a quantitative analysis in terms of R V V Z the power flowing in the two lines, voltage is not the best R c c variable to use since we are dealing with more than one VR V value of line impedance. Rather we define incoming and c outgoing fields via FIG. 15 (Color online) Semi-infinite transmission line of impedance Zc terminated by a resistor R which is represented 1 ← as a second semi-infinite transmission line. Ain = √ Vc (C44) Zc 1 → Aout = √ Vc (C45) Zc I 1 → N R Bin = √ V (C46) R R 1 ← Bout = √ V (C47) R R SII R VN SVV Normalizing by the square root of the impedance allows us to write the power flowing to the right in each line in the simple form

2 2 Pc = (Aout) − (Ain) (C48) a) b) 2 2 PR = (Bin) − (Bout) (C49) FIG. 16 Equivalent circuits for noisy resistors. The out fields are related to the in fields by the s matrix ! ! Aout Ain = s (C50) R, and the energy transmitted into it by thermal fluctua- Bout Bin tions in the R line are both proportional to the absorption coefficient Requiring continuity of the voltage and current at the interface between the two transmission lines, we can solve 2 2 4RZc for the scattering matrix s: A = 1 − |r| = |t| = 2 . (C55) (R + Zc) ! +r t The requirement that the transmission line Zc come to s = (C51) t −r equilibrium with the resistor allows us to readily compute the spectral density of current fluctuations of the random where current source shown in Fig. (16 a). The power dissipated in Zc by the current source attached to R is R − Zc r = (C52) Z +∞ 2 R + Zc dω R Zc √ P = SII [ω] 2 2 RZ −∞ 2π (R + Zc) t = c . (C53) R + Zc (C56)

Note that |r|2 + |t|2 = 1 as required by energy conserva- For the special case R = Zc we can equate this to the tion and that s is unitary with det (s) = −1. By moving right moving power P → in Eq. (C43) because left moving the point at which the phase of the Bin and Bout fields waves in the Zc line are not reflected and hence cannot are determined one-quarter wavelength to the left, we contribute to the right moving power. Requiring P = can put s into different standard form P → yields the classical Nyquist result for the current noise of a resistor ! 0 +r it 2 s = (C54) S [ω] = k T (C57) it +r II R B which has det (s0) = +1. or in the electrical engineering convention As mentioned above, the energy absorbed from the Zc 4 S [ω] + S [−ω] = k T. (C58) line by the resistor R is not turned into heat as in a II II R B true resistor but is is simply transmitted into the R line, which is entirely equivalent. Kirchhoff’s law is now easy We can derive the equivalent expression for the volt- to understand. The energy absorbed from the Zc line by age noise of a resistor (see Fig. 16 b) by considering the 66 voltage noise at the open termination of a semi-infinite fields transmission line with Zc = R. For an open termination  in   out  V → = V ← so that the voltage at the end is given by XA XA  Xin   Xout  ~q in =  B  , ~q out =  B . (C63) ← →  in   out  V = 2V = 2V (C59)  YB   YB  in out YA YA and thus using Eqs. (C40) and (C42) we find The Poisson brackets amongst the different quadrature → amplitudes is given by SVV = 4SVV = 2RkBT (C60) in in {qi , qj } ∝ Jij, (C64) which is equivalent to Eq. (C57). or equivalently the quantum commutators are

in in 2 [qi , qj ] = iXZPFJij, (C65) 4. Amplification with a transmission line and a negative resistance where   0 0 0 +1 We close our discussion of transmission lines by fur-   ther expanding upon the idea mentioned at the end of  0 0 +1 0  J ≡   . (C66) App. C.1 that one can view a one-port amplifier as a  0 −1 0 0  transmission line terminated by an effective negative re- −1 0 0 0 sistance. The discussion here will be very general: we will explore what can be learned about amplification by sim- In order for the transformation to be canonical, the same ply extending the results we have obtained on transmis- Poisson bracket or commutator relations must hold for sion lines to the case of an effective negative resistance. the outgoing field amplitudes Our general discussion will not address the important is- out out 2 [qi , qj ] = iXZPFJij. (C67) sues of how one achieves an effective negative resistance over some appreciable frequency range: for such ques- In the case of a non-linear device these relations would tions, one must focus on a specific physical realization, apply to the small fluctuations in the input and output such as the parametric amplifier discussed in Sec. V.C. fields around the steady state solution. Assuming a linear device (or linearization around the steady state solution) We start by noting that for the case −Zc < R < 0 the power gain G is given by we can define a 4 × 4 real-valued scattering matrixs ˜ in analogy to the 2 × 2 complex-valued scattering matrix s G = |r|2 > 1, (C61) in Eq. (C51) which relates the output fields to the input fields and the s0 matrix introduced in Eq. (C54) becomes out in qi =s ˜ijqj . (C68) √ √ ! G ± G − 1 Eq. (C67) puts a powerful constraint on on thes ˜ matrix, s0 = − √ √ (C62) namely that it must be symplectic. That is,s ˜ and its ± G − 1 G transpose must obey where the sign choice depends on the branch cut chosen sJ˜ s˜T = J. (C69) in the analytic continuation of the off-diagonal elements. From this it follows that This transformation is clearly no longer unitary (because there is no energy conservation since we are ignoring the dets ˜ = ±1. (C70) work done by the amplifier power supply). Note however that we still have det (s0) = +1. It turns out that this This in turn immediately implies Liouville’s theorem naive analytic continuation of the results from positive to that Hamiltonian evolution preserves phase space volume negative resistance is not strictly correct. As we will show (since dets ˜ is the Jacobian of the transformation which in the following, we must be more careful than we have propagates the amplitudes forward in time). been so far in order to insure that the transformation Let us further assume that the device is phase preserv- from the in fields to the out fields must be canonical. ing, that is that the gain or attenuation is the same for In order to understand the canonical nature of the both quadratures. One form for thes ˜ matrix consistent transformation between input and output modes, it is with all of the above requirements is necessary to delve more deeply into the fact that the   + cos θ sin θ 0 0 two quadrature amplitudes of a mode are canonically   conjugate. Following the complex amplitudes defined  sin θ − cos θ 0 0  s˜ =   . (C71) in Eqs. (C44-C47), let us define a vector of real-valued  0 0 − cos θ sin θ  quadrature amplitudes for the incoming and outgoing 0 0 sin θ + cos θ 67

This simply corresponds to a beam splitter and is the Appendix D: Quantum Modes and Noise of a Transmission equivalent of Eq. (C51) with r = cos θ. As mentioned in Line connection with Eq. (C51), the precise form of the scat- tering matrix depends on the choice of planes at which 1. Quantization of a transmission line the phases of the various input and output waves are measured. Recall from Eq. (C30) and the discussion in Appendix Another allowed form of the scattering matrix is: C that the momentum conjugate to the transmission line flux variable ϕ(x, t) is the local charge density q(x, t).   Hence in order to quantize the transmission line modes + cosh θ + sinh θ 0 0 we simply promote these two physical quantities to quan-  + sinh θ + cosh θ 0 0  s˜0 = −   . tum operators obeying the commutation relation  0 0 + cosh θ − sinh θ    [ˆq(x), ϕˆ(x0)] = −i δ(x − x0) (D1) 0 0 − sinh θ + cosh θ ~ √ (C72) from which it follows that the mode amplitudes defined If one takes cosh θ = G, this scattering matrix is in Eq. (C33) become quantum operators obeying essentially the canonically correct formulation of the ˆ ˆ† negative-resistance scattering matrix we tried to write in [Ak0 , Ak] = ~ωkδkk0 (D2) Eq. (C62). Note that the off-diagonal terms have changed and we may identify the usual raising and lowering oper- sign for the Y quadrature relative to the naive expression ators by in Eq. (C62) (corresponding to the other possible an- ˆ p ˆ alytic continuation choice). This is necessary to satisfy Ak = ~ωk bk (D3) the symplecticity condition and hence make the transfor- ˆ mation canonical. The scattering matrixs ˜0 can describe where bk destroys a photon in mode k. The quantum amplification. Unlike the beam splitter scattering ma- form of the Hamiltonian in Eq. (C34) is thus trixs ˜ above,s ˜0 is not unitary (even though dets ˜0 = 1). X  1 Unitarity would correspond to power conservation. Here, H = ω ˆb†ˆb + . (D4) ~ k k k 2 power is not conserved, as we are not explicitly tracking k the power source supplying our active system. For the quantum case the thermal equilibrium expression The form of the negative-resistance amplifier scattering then becomes matrixs ˜0 confirms many of the general statements we ˆ† ˆ made about phase-preserving amplification in Sec. V.B. hAkAki = ~ωknB(~ωk), (D5) First, note that the requirement of finite gain G > 1 and phase preservation makes all the diagonal elements ofs ˜0 which reduces to Eq. (C42) in the classical limit ~ωk  (i.e. cosh θ ) equal. We see that to amplify the A mode, kBT . it is impossible to avoid coupling to the B mode (via the We have seen previously in Eqs. (C6) that the volt- sinh θ term) because of the requirement of symplecticity. age fluctuations on a transmission line can be resolved We thus see that it is impossible classically or quantum into right and left moving waves which are functions of a mechanically to build a linear phase-preserving amplifier combined space-time argument whose only effect is to amplify the desired signal. The x x V (x, t) = V →(t − ) + V ←(t + ). (D6) presence of the sinh θ term above means that the output vp vp signal is always contaminated by amplified noise from → at least one other degree of freedom (in this case the B Thus in an infinite transmission line, specifying V ev- erywhere in space at t = 0 determines its value for all mode). If the thermal or quantum noise in A and B → are equal in magnitude (and uncorrelated), then in the times. Conversely specifying V at x = 0 for all times limit of large gain where cosh θ ≈ sinh θ, the output noise fully specifies the field at all spatial points. In prepa- (referred to the input) will be doubled. This is true for ration for our study of the quantum version of input- both classical thermal noise and quantum vacuum noise. output theory in Appendix E, it is convenient to extend Eqs. (C38-C39) to the quantum case (x = 0): The negative resistance model of an amplifier here r gives us another way to think about the noise added by 1 X p h i Vˆ →(t) = ω ˆb e−iωkt + h.c. an amplifier: crudely speaking, we can view it as being 2Lc ~ k k directly analogous to the fluctuation-dissipation theorem k>0 r simply continued to the case of negative dissipation. Just Z ∞ dω ωZ h i = ~ c ˆb→[ω]e−iωt + h.c. (D7) as dissipation can occur only when we open up a new 2π 2 channel and thus we bring in new fluctuations, so ampli- 0 fication can occur only when there is coupling to an ad- In the second line, we have defined: ditional channel. Without this it is impossible to satisfy r vp X the requirement that the amplifier perform a canonical ˆb→[ω] ≡ 2π ˆb δ(ω − ω ) (D8) L k k transformation. k>0 68

In a similar fashion, we have: We have already seen that using classical statistical r mechanics, the voltage noise in equilibrium is white. Z ∞ dω ωZ h i Vˆ ←(t) = ~ c ˆb←[ω]e−iωt + h.c. (D9) The corresponding analysis of the temporal modes using 0 2π 2 Eqs. (D13) shows that the quantum commutator obeys r vp X ˆb←[ω] ≡ 2π ˆb δ(ω − ω ) (D10) [ˆb→(t), ˆb†→(t0)] = δ(t − t0). (D16) L k k k<0 In deriving this result, we have converted summations One can easily verify that among the ˆb→[ω], ˆb←[ω] opera- over mode index to integrals over frequency. Further, tors and their conjugates, the only non-zero commutators because (for finite time resolution at least) the integral is are given by: dominated by frequencies near +Ω0 we can, within the Markov (Wigner Weisskopf) approximation, extend the   †   † ˆb→[ω], ˆb→[ω0] = ˆb←[ω], ˆb←[ω0] = 2πδ(ω − ω0) lower limit of frequency integration to minus infinity and thus arrive at a delta function in time. If we further take (D11) the right moving modes to be in thermal equilibrium, We have taken the continuum limit L → ∞ here, allowing then we may similarly approximate: us to change sums on k to integrals. We have thus ob- tained the description of a quantum transmission line in ˆ†→ 0 ˆ→ 0 hb (t )b (t)i = nB(~Ω0)δ(t − t ) (D17a) terms of left and right-moving frequency resolved modes, ˆ→ ˆ†→ 0 0 as used in our discussion of amplifiers in Sec. VI (see hb (t)b (t )i = [1 + nB(~Ω0)] δ(t − t ). (D17b) Eqs. 6.2). Note that if the right-moving modes are fur- Equations (D15) to (D17b) indicate that Vˆ →(t) can be ther taken to be in thermal equilibrium, one finds (again, treated as the quantum operator equivalent of white in the continuum limit): noise; a similar line of reasoning applies mutatis mutan-  †  dis to the left moving modes. We stress that these re- ˆ→  ˆ→ 0 0 b [ω] b [ω ] = 2πδ(ω − ω )nB(~ω) (D12a) sults rely crucially on our assumption that we are dealing

 † with a relatively narrow band of frequencies in the vicin- ˆ→ ˆ→ 0  0 b [ω] b [ω ] = 2πδ(ω − ω ) [1 + nB(~ω)] ity of Ω0; the resulting approximations we have made are known as the Markov approximation. As one can al- . (D12b) ready see from the form of Eqs. (D7,D9), and as will be discussed further, the actual spectral density of vacuum We are typically interested in a relatively narrow band noise on a transmission line is not white, but is linear in of frequencies centered on some characteristic drive or frequency. The approximation made in Eq. (D16) treats resonance frequency Ω . In this case, it is useful to work 0 it as a constant within the narrow band of frequencies in the time-domain, in a frame rotating at Ω . Fourier 0 of interest. If the range of frequencies of importance is transforming 24 Eqs. (D8) and (D10), one finds: large then the Markov approximation is not applicable. r vp X ˆb→(t) = e−i(ωk−Ω0)tˆb (0), (D13a) L k k>0 2. Modes and the windowed Fourier transform r vp X ˆb←(t) = e−i(ωk−Ω0)tˆb (0). (D13b) L k While delta function correlations can make the quan- k<0 tum noise relatively easy to deal with in both the time These represent temporal right and left moving modes. and frequency domain, it is sometimes the case that it Note that the normalization factor in Eqs. (D13) has been is easier to deal with a ‘smoothed’ noise variable. The chosen so that the right moving photon flux at x = 0 and introduction of an ultraviolet cutoff regulates the math- time t is given by ematical singularities in the noise operators evaluated at equal times and is physically sensible because every real hN˙ i = hˆb†→(t)ˆb→(t)i (D14) measurement apparatus has finite time resolution. A sec- ond motivation is that real spectrum analyzers output a In the same rotating frame, and within the approxima- time varying signal which represents the noise power in tion that all relevant frequencies are near Ω0, Eq. (D7) a certain frequency interval (the ‘resolution bandwidth’) becomes simply: averaged over a certain time interval (the inverse ‘video r bandwidth’). The mathematical tool of choice for dealing → ~Ω0Zc h → †→ i Vˆ (t) ≈ ˆb (t) + ˆb (t) (D15) with such situations in which time and frequency both 2 appear is the ‘windowed Fourier transform’. The win- dowed transform uses a kernel which is centered on some frequency window and some time interval. By summa- 24 As in the main text, we use in this appendix a convention which tion over all frequency and time windows it is possible differs from the one commonly used in quantum optics:a ˆ[ω] = to invert the transformation. The reader is directed to R +∞ +iωt † † R +∞ +iωt † −∞ dt e aˆ(t) anda ˆ [ω] = [ˆa[−ω]] = −∞ dt e aˆ (t). (Mallat, 1999) for the mathematical details. 69

Δt For our present purposes where we are interested in ω area = 2π just a single narrow frequency range centered on Ω0, a convenient windowed transform kernel for smoothing the quantum noise is simply a box of width ∆t representing m B the finite integration time of our detector. In the frame mp 2π/Δt rotating at Ω0 we can define

Z tj+1 ω=Ω t ˆ→ 1 ˆ→ 0 Bj = √ dτ b (τ) (D18) ∆t tj

-m B-mp where tj = j(∆t) denotes the time of arrival of the jth temporal mode at the point x = 0. Recall that ˆb→ has ˆ→ a photon flux normalization and so Bj is dimensionless. From Eq. (D16) we see that these smoothed operators p obey the usual bosonic commutation relations FIG. 17 (Color online) Schematic figure indicating how the → †→ various modes defined by the windowed Fourier transform tile [Bˆ , Bˆ ] = δjk. (D19) j k the time-frequency plane. Each individual cell corresponds to † a different mode, and has an area 2π. The state Bj |0i has a single photon occupying basis mode j, which is centered in frequency space at Ω0 and in time space on the interval j∆t < t < (j + 1)∆t (i.e. this temporal mode passes the point x = 0 during the jth (δΩ)(∆t) << 1) and spread out over a large time interval time interval.) This basis mode is much like a note in a T centered at time T0. This could be given for example musical score: it has a certain specified pitch and occurs by at a specified time for a specified duration. Just as we 2 (j∆t−T0) can play notes of different frequencies simultaneously, we − 2 −i(δΩ)(j∆t) ψj = N e 4T e (D24) can define other temporal modes on the same time in- terval and they will be mutually orthogonal provided the where N is the appropriate normalization constant. angular frequency spacing is a multiple of 2π/∆t. The The state having n photons in the mode is simply result is a set of modes Bm,p labeled by both a frequency 1 n index m and a time index p. p labels the time interval as √ Ψ† |0i. (D25) before, while m labels the angular frequency: n! 2π The concept of ‘wave function of the photon’ is fraught ω = Ω + m (D20) m 0 ∆t with dangers. In the very special case where we re- strict attention solely to the subspace of single photon The result is, as illustrated in Fig. (17), a complete lat- Fock states, we can usefully think of the amplitudes {ψj} tice of possible modes tiling the frequency-time phase as the ‘wave function of the photon’ (Cohen-Tannoudji space, each occupying area 2π corresponding to the time- et al., 1989) since it tells us about the spatial mode which frequency uncertainty principle. is excited. In the general case however it is essential to We can form other modes of arbitrary shapes centered keep in mind that the transmission line is a collection of on frequency Ω0 by means of linear superposition of our coupled LC oscillators with an infinite number of degrees basis modes (as long as they are smooth on the time scale of freedom. Let us simplify the argument by considering ∆t). Let us define a single LC oscillator. We can perfectly well write a wave function for the system as a function of the coordinate X ˆ→ Ψ = ψjBj . (D21) (say the charge q on the capacitor). The ground state j wave function χ0(q) is a gaussian function of the coor- dinate. The one photon state created by Ψ† has a wave This is also a canonical bosonic mode operator obeying function χ1(q) ∼ qχ0(q) proportional to the coordinate [Ψ, Ψ†] = 1 (D22) times the same gaussian. In the general case χ is a wave functional of the charge distribution q(x) over the entire provided that the coefficients obey the normalization con- transmission line. dition Using Eq. (D17a) we have

X 2 ˆ†→ ˆ→ |ψj| = 1. (D23) hBj Bk i = nB(~Ω0)δjk (D26) j independent of our choice of the coarse-graining time win- We might for example want to describe a mode which is dow ∆t. This result allows us to give meaning to the centered at a slightly higher frequency Ω0 + δΩ (obeying phrase one often hears bandied about in descriptions of 70

πn amplifiers that ‘the noise temperature corresponds to a where ϕn is the normal coordinate and kn ≡ L . Sub- mode occupancy of X photons’. This simply means that stitution of this form into the Lagrangian and carrying the photon flux per unit bandwidth is X. Equivalently out the spatial integration yields a set of independent the flux in bandwidth B is harmonic oscillators representing the normal modes.

X ∞ N˙ = (B∆t) = XB. (D27) X  c 1  ∆t L = ϕ˙ 2 − k2 ϕ2 . (D32) g 2 n 2` n n The interpretation of this is that X photons in a tempo- n=1 ral mode of duration ∆t pass the origin in time ∆t. Each From this we can find the momentum operatorp ˆ canon- mode has bandwidth ∼ 1 and so there are B∆t inde- n ∆t ically conjugate to the coordinate operatorϕ ˆ and quan- pendent temporal modes in bandwidth B all occupying n tize the system to obtain an expression for the operator the same time interval ∆t. The longer is ∆t the longer it representing the voltage at the end of the transmission takes a given mode to pass the origin, but the more such line in terms of the mode creation and destruction oper- modes fit into the frequency window. ators As an illustration of these ideas, consider the following ∞ r elementary question: What is the mode occupancy of a X Ωn Vˆ = ~ i(ˆb† − ˆb ). (D33) laser beam of power P and hence photon flux N˙ = P ? n n Ω0 Lc We cannot answer this without knowing the coherence~ n=1 time or equivalently the bandwidth. The output of a The spectral density of voltage fluctuations is then found good laser is like that of a radio frequency oscillator–it to be has essentially no amplitude fluctuations. The frequency ∞ 2π X Ωn is nominally set by the physical properties of the oscilla- S [ω] = ~ n ( Ω )δ(ω + Ω ) tor, but there is nothing to pin the phase which conse- VV L c B ~ n n n=1 quently undergoes slow diffusion due to unavoidable noise perturbations. This leads to a finite phase coherence time +[nB(~Ωn) + 1]δ(ω − Ωn) , (D34) τ and corresponding frequency spread 1/τ of the laser where n ( ω) is the Bose occupancy factor for a photon spectrum. (A laser beam differs from a thermal source B ~ with energy ~ω. Taking the limit L → ∞ and converting that has been filtered to have the same spectrum in that the summation to an integral yields it has smaller amplitude fluctuations.) Thus we expect  that the mode occupancy is X = Nτ˙ . A convenient ap- SVV(ω) = 2Zc~|ω| nB(~|ω|)Θ(−ω)+[nB(~|ω|)+1]Θ(ω) , proximate description in terms of temporal modes is to (D35) take the window interval to be ∆t = τ. Within the jth where Θ is the step function. We see immediately that interval we take the phase to be a (random) constant ϕj at zero temperature there is no noise at negative frequen- so that (up to an unimportant normalization constant) cies because energy can not be extracted from zero-point we have the coherent state motion. However there remains noise at positive frequen- √ Y Xeiϕj Bˆ†→ cies indicating that the vacuum is capable of absorbing e j |0i (D28) energy from another quantum system. The voltage spec- j tral density at both zero and non-zero temperature is which obeys plotted in Fig. (1). √ Eq. (D35) for this ‘two-sided’ spectral density of a re- ˆ→ iϕk hBk i = Xe (D29) sistor can be rewritten in a more compact form and 2Z ω S [ω] = c~ , (D36) ˆ†→ ˆ→ VV − ω/k T hBk Bk i = X. (D30) 1 − e ~ B which reduces to the more familiar expressions in various 3. Quantum noise from a resistor limits. For example, in the classical limit kBT  ~ω the spectral density is equal to the Johnson noise result25 Let us consider the quantum equivalent to Eq. (C60), SVV[ω] = 2ZckBT, (D37) SVV = 2RkBT , for the case of a semi-infinite transmis- sion line with open termination, representing a resistor. in agreement with Eq. (C60). In the quantum limit it From Eq. (C27) we see that the proper boundary con- reduces to dition for the ϕ field is ∂xϕ(0, t) = ∂xϕ(L, t) = 0. (We have temporarily made the transmission line have a large SVV[ω] = 2Zc~ωΘ(ω). (D38) but finite length L.) The normal mode expansion that satisfies these boundary conditions is r ∞ 2 X ϕ(x, t) = ϕ (t) cos(k x), (D31) 25 Note again that in the engineering convention this would be L n n n=1 SVV[ω] = 4ZckBT . 71

Again, the step function tells us that the resistor can only 2006; Harris et al., 2007; H¨ohberger-Metzger and Kar- absorb energy, not emit it, at zero temperature. rai, 2004; Marquardt et al., 2007, 2006; Meystre et al., If we use the engineering convention and add the noise 1985; Schliesser et al., 2006; Teufel et al., 2008; Thomp- at positive and negative frequencies we obtain son et al., 2008; Wilson-Rae et al., 2007). In the case of a microwave cavity containing a qubit, the state-dependent ~ω polarizability of the qubit acts as a source which shifts SVV[ω] + SVV[−ω] = 2Zc~ω coth (D39) 2kBT the frequency of the cavity (Blais et al., 2004; Schuster et al., 2005; Wallraff et al., 2004). for the symmetric part of the noise, which appears in the The dephasing of a qubit in a microwave cavity and quantum fluctuation-dissipation theorem (cf. Eq. (2.16)). the fluctuations in the radiation pressure in an optical The antisymmetric part of the noise is simply cavity both depend on the quantum noise in the number of photons inside the cavity. We here use a simple equa- SVV[ω] − SVV[−ω] = 2Zc~ω, (D40) tion of motion method to exactly solve for this quantum noise in the perturbative limit where the dynamics of the yielding qubit or mirror degree of freedom has only a weak back action effect on the cavity. SVV[ω] − SVV[−ω] ~ω = tanh . (D41) In the following, we first give a basic discussion of S [ω] + S [−ω] 2k T VV VV B the cavity field noise spectrum, deferring the detailed This quantum treatment can also be applied to any microscopic derivation to subsequent subsections. We arbitrary dissipative network (Burkhard et al., 2004; De- then provide a review of the input-output theory for voret, 1997). If we have a more complex circuit con- driven cavities, and employ this theory to analyze the taining capacitors and inductors, then in all of the above important example of a dispersive position measurement, where we demonstrate how the standard quantum limit expressions, Zc should be replaced by Re Z[ω] where Z[ω] is the complex impedance presented by the circuit. can be reached. Finally, we analyze an example where In the above we have explicitly quantized the stand- a modified dispersive scheme is used to detect only one ing wave modes of a finite length transmission line. We quadrature of a harmonic oscillator’s motion, such that could instead have used the running waves of an infinite this quadrature does not feel any back-action. line and recognized that, as the in classical treatment in Eq. (C59), the left and right movers are not independent. 1. Photon shot noise inside a cavity and back action The open boundary condition at the termination requires V ← = V → and hence b→ = b←. We then obtain Consider a degree of freedomz ˆ coupled parametrically → with strength A to the cavity oscillator SVV [ω] = 4SVV [ω] (D42) ˆ † † Hint = ~ωc(1 + Azˆ) [ˆa aˆ − haˆ aˆi] (E1) and from the quantum analog of Eq. (C40) we have where following Eq. (3.11), we have taken A to be dimen- 4~|ω| sionless, and usez ˆ to denote the dimensionless system SVV [ω] = {Θ(ω)(nB + 1) + Θ(−ω)nB} 2cvp variable that we wish to probe. For example,z ˆ could represent the dimensionless position of a mechanical os- = 2Zc~|ω| {Θ(ω)(nB + 1) + Θ(−ω)nB} cillator (D43) xˆ zˆ ≡ . (E2) in agreement with Eq. (D35). xZPF We have subtracted the haˆ†aˆi term so that the mean Appendix E: Back Action and Input-Output Theory for force on the degree of freedom is zero. To obtain the full Driven Damped Cavities Hamiltonian, we would have to add the cavity damping and driving terms, as well as the Hamiltonian governing A high Q cavity whose resonance frequency can be the intrinsic dynamics of the systemz ˆ. From Eq. (3.17) parametrically controlled by an external source can act we know that the back action noise force acting onz ˆ is as a very simple quantum amplifier, encoding informa- proportional to the quantum fluctuations in the number tion about the external source in the phase and ampli- of photonsn ˆ =a ˆ†aˆ in the cavity, tude of the output of the driven cavity. For example, S (t) = haˆ†(t)ˆa(t)ˆa†(0)ˆa(0)i − haˆ†(t)ˆa(t)i2. (E3) in an optical cavity, one of the mirrors could be move- nn able and the external source could be a force acting on For the case of continuous wave driving at frequency that mirror. This defines the very active field of optome- ωL = ωc + ∆ detuned by ∆ from the resonance, the chanics, which also deals with microwave cavities cou- cavity is in a coherent state |ψi obeying pled to nanomechanical systems and other related setups (Arcizet et al., 2006; Brown et al., 2007; Gigan et al., aˆ(t) = e−iωLt[¯a + dˆ(t)] (E4) 72 where the first term is the ‘classical part’ of the mode am- (a) (b) plitude ψ(t) =ae ¯ −iωLt determined by the strength of the drive field, the damping of the cavity and the detuning 4 4 ∆, and d is the quantum part. By definition, aˆ|ψi = ψ|ψi (E5) 2 0 so the coherent state is annihilated by dˆ: ˆ d|ψi = 0. (E6) -4 0 That is, in terms of the operator dˆ, the coherent state -6 06-6 0 6 looks like the undriven quantum ground state . The dis- Noise power Frequency placement transformation in Eq. (E4) is canonical since (c) 3 [ˆa, aˆ†] = 1 ⇒ [d,ˆ dˆ†] = 1. (E7) Substituting the displacement transformation into Eq. (E3) and using Eq. (E6) yields ˆ ˆ† 2 Snn(t) =n ¯hd(t)d (0)i, (E8) - wheren ¯ = |a¯|2 is the mean cavity photon number. If we set the cavity energy damping rate to be κ, such that the amplitude damping rate is κ/2, then the undriven state 1 obeys

† +i∆t − κ |t| hdˆ(t)dˆ (0)i = e e 2 . (E9) This expression will be justified formally in the subse-

Noise temperature 0 quent subsection, after introducing input-output theory. 0 1 2 3 4 5 6 We thus arrive at the very simple result Frequency i∆t− κ |t| S (t) =ne ¯ 2 . (E10) nn FIG. 18 (Color online) (a) Noise spectrum of the photon num- The power spectrum of the noise is, via the Wiener- ber in a driven cavity as a function of frequency when the Khinchin theorem (Appendix A.2), simply the Fourier cavity drive frequency is detuned from the cavity resonance transform of the autocorrelation function given in by ∆ = +3κ (left peak) and ∆ = −3κ (right peak). (b) Ef- Eq. (E10) fective temperature Teff of the low frequency noise, ω → 0, as a function of the detuning ∆ of the drive from the cavity Z +∞ resonance. (c) Frequency-dependence of the effective noise iωt κ Snn[ω] = dt e Snn(t) =n ¯ 2 2 . temperature, for different values of the detuning. −∞ (ω + ∆) + (κ/2) (E11) As can be seen in Fig. 18a, for positive detuning ∆ = ωL − ωc > 0, i.e. for a drive that is blue-detuned with or equivalently respect to the cavity, the noise peaks at negative ω. This Snn[ω] − Snn[−ω] means that the noise tends to pump energy into the de- = tanh(β~ω/2). (E13) gree of freedomz ˆ (i.e. it contributes negative damping). Snn[ω] + Snn[−ω] For negative detuning the noise peaks at positive ω cor- Ifz ˆ is the coordinate of a harmonic oscillator of frequency responding to the cavity absorbing energy fromz ˆ. Basi- ω (or some non-conserved observable of a qubit with level cally, the interaction withz ˆ (three wave mixing) tries to splitting ω), then that system will acquire a temperature Raman scatter the drive photons into the high density of Teff [ω] in the absence of coupling to any other environ- states at the cavity frequency. If this is uphill in energy, ment. In particular, if the characteristic oscillation fre- thenz ˆ is cooled. quency of the systemz ˆ is much smaller than κ, then we As discussed in Sec. B.2 (c.f. Eq. (2.8)), at each fre- have the simple result quency ω, we can use detailed balance to assign the noise an effective temperature T [ω]: 1 2 S [ω] − S [−ω] eff = lim nn nn kBTeff ω→0+ ω Snn[ω] + Snn[−ω] Snn[ω] ~ = e~ω/kB Teff [ω] ⇔ d ln Snn[ω] Snn[−ω] = 2 d ω ~ω ~ kBTeff [ω] ≡ (E12) 1 −4∆ h Snn[ω] i log = 2 2 . (E14) Snn[−ω] ~ ∆ + (κ/2) 73

As can be seen in Fig. 18, the asymmetry in the noise is the goal of keeping the bath modes instead of trac- changes sign with detuning, which causes the effective ing them out. This is obviously necessary for the situa- temperature to change sign. tions we have in mind, where the output field emanating First we discuss the case of a positive Teff , where this from the cavity contains the information acquired during mechanism can be used to laser cool an oscillating me- a measurement of the system coupled to the cavity. As chanical cantilever, provided Teff is lower than the in- we learned from the classical treatment, we can elimi- trinsic equilibrium temperature of the cantilever. (Ar- nate the outgoing waves in favor of a damping term for cizet et al., 2006; Brown et al., 2007; Gigan et al., 2006; the system. However we can recover the solution for the Harris et al., 2007; H¨ohberger-Metzger and Karrai, 2004; outgoing modes completely from the solution of the equa- Marquardt et al., 2007; Schliesser et al., 2006; Thompson tion of motion of the damped system being driven by the et al., 2008; Wilson-Rae et al., 2007). A simple classical incoming waves. argument helps us understand this cooling effect. Sup- In order to drive the cavity we must partially open one pose that the moveable mirror is at the right hand end of its ports which exposes the cavity both to the external of a cavity being driven below the resonance frequency. drive and to the vacuum noise outside which permits en- If the mirror moves to the right, the resonance frequency ergy in the cavity to leak out into the surrounding bath. will fall and the number of photons in the cavity will rise. We will formally separate the degrees of freedom into in- There will be a time delay however to fill the cavity and ternal cavity modes and external bath modes. Strictly so the extra radiation pressure will not be fully effective speaking, once the port is open, these modes are not dis- in doing work on the mirror. During the return part of tinct and we only have ‘the modes of the universe’ (Gea- the oscillation as the mirror moves back to the left, the Banacloche et al., 1990a,b; Lang et al., 1973). However time delay in emptying the cavity will cause the mirror to for high Q cavities, the distinction is well-defined and we have to do extra work against the radiation pressure. At can model the decay of the cavity in terms of a spon- the end of the cycle it ends up having done net positive taneous emission process in which an internal boson is work on the light field and hence is cooled. The effect can destroyed and an external bath boson is created. We therefore be understood as being due to the introduction assume a single-sided cavity. For a high Q cavity, this of some extra optomechanical damping. physics is accurately captured in the following Hamilto- The signs reverse (and Teff becomes negative) if the nian cavity is driven above resonance, and consequently the ˆ ˆ ˆ ˆ cantilever motion is heated up. In the absence of in- H = Hsys + Hbath + Hint. (E15) trinsic mechanical losses, negative values of the effective The bath Hamiltonian is temperature indicate a dynamical instability of the can- tilever (or population inversion in the case of a qubit), ˆ X ˆ†ˆ Hbath = ~ωqbqbq (E16) where the amplitude of motion grows until it is finally q stabilized by nonlinear effects. This can be interpreted as negative damping introduced by the optomechanical where q labels the quantum numbers of the independent coupling and can be used to create parametric amplifica- harmonic oscillator bath modes obeying tion of mechanical forces acting on the oscillator. ˆ ˆ† Finally, we mention that cooling towards the quantum [bq, bq0 ] = δq,q0 . (E17) ground state of a mechanical oscillator (where phonon numbers become much less than one), is only possible Note that since the bath terminates at the system, there (Marquardt et al., 2007; Wilson-Rae et al., 2007) in the is no translational invariance, the normal modes are “far-detuned regime”, where −∆ = ω  κ (in contrast standing not running waves, and the quantum numbers to the ω  κ regime discussed above). q are not necessarily wave vectors. The coupling Hamiltonian is (within the rotating wave approximation) 2. Input-output theory for a driven cavity ˆ X h †ˆ ∗ˆ† i Hint = −i~ fqaˆ bq − fq bqaˆ . (E18) The results from the previous section can be more for- q mally and rigorously derived in a full quantum theory For the moment we will leave the system (cavity) Hamil- of a cavity driven by an external coherent source. The tonian to be completely general, specifying only that it theory relating the drive, the cavity and the outgoing consists of a single degree of freedom (i.e. we concentrate waves radiated by the cavity is known as input-output on only a single resonance of the cavity with frequency theory and the classical description was presented in Ap- ωc) obeying the usual bosonic commutation relation pendix C. The present quantum discussion closely fol- lows standard references on the subject (Walls and Mil- [ˆa, aˆ†] = 1. (E19) burn, 1994; Yurke, 1984; Yurke and Denker, 1984). The crucial feature that distinguishes such an approach from (N.B. this does not imply that it is a harmonic oscilla- many other treatments of quantum-dissipative systems tor. We will consider both linear and non-linear cavities.) 74

Note that the most general linear coupling to the bath Using ˆ† † ˆ modes would include terms of the form b aˆ and bqa but q Z x0 these are neglected within the rotating wave approxima- 1 dx δ(x − x0) = (E27) tion because in the interaction representation they os- −∞ 2 cillate at high frequencies and have little effect on the dynamics. we obtain for the cavity EOM The Heisenberg equation of motion (EOM) for the i κ ˙ ˆ X −iωq(t−t0)ˆ bath variables is aˆ = [Hsys, aˆ] − aˆ − fqe bq(t0). (E28) 2 ~ q ˆ˙ i ˆ ˆ ˆ ∗ bq = [H, bq] = −iωqbq + fq aˆ (E20) ~ The second term came from the part of the bath motion representing the wave radiated by the cavity and, within We see that this is simply the EOM of a harmonic oscil- the Markov approximation, has become a simple linear lator driven by a forcing term due to the motion of the damping term for the cavity mode. Note the important cavity degree of freedom. Since this is a linear system, factor of 2. The amplitude decays at half the rate of the the EOM can be solved exactly. Let t < t be a time in 0 intensity (the energy decay rate κ). the distant past before any wave packet launched at the Within the spirit of the Markov approximation it is cavity has reached it. The solution of Eq. (E20) is p 2 further convenient to treat f ≡ |fq| as a constant and Z t define the density of states (also taken to be a constant) ˆ −iωq(t−t0)ˆ −iωq(t−τ) ∗ bq(t) = e bq(t0) + dτ e fq aˆ(τ). by t0 X (E21) ρ = δ(ω − ω ) (E29) The first term is simply the free evolution of the bath c q q while the second represents the waves radiated by the cavity into the bath. so that the Golden Rule rate becomes The EOM for the cavity mode is κ = 2πf 2ρ. (E30) i X aˆ˙ = [Hˆ , aˆ] − f ˆb . (E22) sys q q We can now define the so-called ‘input mode’ ~ q 1 X Substituting Eq. (E21) into the last term above yields ˆb (t) ≡ √ e−iωq(t−t0)ˆb (t ) . (E31) in 2πρ q 0 q X ˆ X −iωq(t−t0)ˆ fqbq = fqe bq(t0) q q For the case of a transmission line treated in Appendix ˆ→ Z t D, this coincides with the field b moving towards the X 2 −i(ωq−ωc)(t−τ) +iωc(τ−t) + |fq| dτ e [e aˆ(τ)], (E23) cavity [see Eq. (D13a)]. We finally have for the cavity q t0 EOM √ where the last term in square brackets is a slowly varying ˙ i κ ˆ aˆ = [Hˆsys, aˆ] − aˆ − κ bin(t). (E32) function of τ. To simplify our result, we note that if ~ 2 the cavity system were a simple harmonic oscillator of Note that when a wave packet is launched from the bath frequency ω then the decay rate from the n = 1 single c towards the cavity, causality prevents it from knowing photon excited state to the n = 0 ground state would be about the cavity’s presence until it reaches the cavity. given by the following Fermi Golden Rule expression Hence the input mode evolves freely as if the cavity were X not present until the time of the collision at which point κ(ω ) = 2π |f |2δ(ω − ω ). (E24) c q c q it begins to drive the cavity. Since ˆb (t) evolves under q in the free bath Hamiltonian and acts as the driving term in From this it follows that the cavity EOM, we interpret it physically as the input mode. Eq. (E32) is the quantum analog of the classi- Z +∞ dν X κ(ω + ν)e−iν(t−τ) = |f |2e−i(ωq−ωc)(t−τ). cal equation (C19), for our previous example of an LC- 2π c q oscillator driven by a transmission line. The latter would −∞ q (E25) also have been first order in time if as in Eq. (C35) we We now make the Markov approximation which assumes had worked with the complex amplitude A instead of the that κ(ν) = κ is a constant over the range of frequencies coordinate Q. relevant to the cavity so that Eq. (E25) may be repre- Eq. (E31) for the input mode contains a time label sented as just as in the interaction representation. However it is best interpreted as simply labeling the particular linear X 2 −i(ωq−ωc)(t−τ) |fq| e = κδ(t − τ). (E26) combination of the bath modes which is coupled to the q system at time t. Some authors even like to think of 75 the bath modes as non-propagating while the cavity flies Proceeding as before we obtain along the bath (taken to be 1D) at a velocity v. The √ system then only interacts briefly with the local mode ˙ i κ ˆ aˆ = [Hˆsys, aˆ] + aˆ − κ bout(t). (E36) positioned at x = vt before moving on and interacting ~ 2 with the next local bath mode. We will elaborate on this Subtracting Eq. (E36) from Eq. (E32) yields view further at the end of this subsection. The expression for the power P (energy per time) √ in ˆb (t) = ˆb (t) + κ aˆ(t) (E37) impinging on the cavity depends on the normalization out in ˆ chosen in our definition of bin. It can be obtained, for which is consistent with our interpretation of the out- ˆ example, by imagining the bath modes bq to live on a one- going field as the reflected incoming field plus the field dimensional waveguide with propagation velocity v and radiated by the cavity out through the partially reflecting length L (using periodic boundary conditions). In that mirror. case we have to sum over all photons to get the average The above results are valid for any general cavity power flowing through a cross-section of the waveguide, Hamiltonian. The general procedure is to solve Eq. (E32) D E P ˆ†ˆ fora ˆ(t) for a given input field, and then solve Eq. (E37) Pin = q ~ωq(vp/L) bqbq . Inserting the definition for ˆ to obtain the output field. For the case of an empty cav- bin, Eq. (E31), the expression for the input power carried ity we can make further progress because the cavity mode by a monochromatic beam at frequency ω is is a harmonic oscillator D E ˆ† ˆ ˆ † Pin(t) = ~ω bin(t)bin(t) (E33) Hsys = ~ωcaˆ a.ˆ (E38)

Note that this has the correct dimensions due to our In this simple case, the cavity EOM becomes ˆ √ choice of normalization for bin (with dimensions ω). In κ √ aˆ˙ = −iω aˆ − aˆ − κ ˆb (t). (E39) the general case, an integration over frequencies is needed c 2 in (as will be discussed further below). An analogous for- mula holds for the power radiated by the cavity, to be Eq. (E39) can be solved by Fourier transformation, yield- discussed now. ing ˆ The output mode bout(t) is radiated into the bath and √ evolves freely after the system interacts with ˆb (t). If κ ˆ in aˆ[ω] = − bin[ω] (E40) the cavity did not respond at all, then the output mode i(ωc − ω) + κ/2 √ would simply be the input mode reflected off the cav- ˆ = − κχc[ω − ωc]bin[ω] (E41) ity mirror. If the mirror is partially transparent then the output mode will also contain waves radiated by the and cavity (which is itself being driven by the input mode partially transmitted into the cavity through the mirror) ˆ ω − ωc − iκ/2 ˆ bout[ω] = bin[ω] (E42) and hence contains information about the internal dy- ω − ωc + iκ/2 namics of the cavity. To analyze this output field, let which is the result for the reflection coefficient quoted in t1 > t be a time in the distant future after the input field has interacted with the cavity. Then we can write Eq. (3.12). For brevity, here and in the following, we will an alternative solution to Eq. (E20) in terms of the final sometimes use the susceptibility of the cavity, defined as rather than the initial condition of the bath 1 χc[ω − ωc] ≡ (E43) Z t1 −i(ω − ωc) + κ/2 ˆ −iωq (t−t1)ˆ −iωq(t−τ) ∗ bq(t) = e bq(t1) − dτ e fq aˆ(τ). t For the case of steady driving on resonance where ω = ωc, (E34) the above equations yield Note the important minus sign in the second term as- √ sociated with the fact that the time t is now the lower κ ˆb [ω] = aˆ[ω]. (E44) limit of integration rather than the upper as it was in out 2 Eq. (E21). Defining In steady state, the incoming power equals the outgoing power, and both are related to the photon number inside 1 X the single-sided cavity by ˆb (t) ≡ √ e−iωq(t−t1)ˆb (t ), (E35) out 2πρ q 1 q D E κ P = ω ˆb† (t)ˆb (t) = ω aˆ†(t)ˆa(t) (E45) ~ out out ~ 4 we see that this is simply the free evolution of the bath modes from the distant future (after they have interacted Note that this does not coincide with the naive expecta- † with the cavity) back to the present, indicating that it is tion, which would be P = ~ωκ aˆ aˆ . The reason for this indeed appropriate to interpret this as the outgoing field. discrepancy is the the interference between the part of the 76 incoming wave which is promptly reflected from the cav- as is required for the cavity bosonic quantum degree of ity and the field radiated by the cavity. The naive expres- freedom. We can interpret this as saying that the cavity sion becomes correct after the drive has been switched off zero-point fluctuations arise from the vacuum noise that (where ignoring the effect of the incoming vacuum noise, enters through the open port. We also now have a simple ˆ √ we would have bout = κaˆ). We note in passing that for physical interpretation of the quantum noise in the num- a driven two-sided cavity with coupling constants κL and ber of photons in the driven cavity in Eqs. (E3,E8,E11). κR (where κ = κL + κR), the incoming power sent into It is due to the vacuum noise which enters the cavity the left port is related to the photon number by through the same ports that bring in the classical drive. The interference between the vacuum noise and the clas- 2 † P = ~ωκ /(4κL) aˆ aˆ . (E46) sical drive leads to the photon number fluctuations in the cavity. Here for κL = κR the interference effect completely elim- inates the reflected beam and we have in contrast to In thermal equilibrium, ξˆ also contains thermal radi- Eq. (E45) ation. If the bath is being probed only over a narrow range of frequencies centered on ωc (which we have as- κ † P = ~ω aˆ aˆ . (E47) sumed in making the Markov approximation) then we 2 have to a good approximation (consistent with the above Eq. (E39) can also be solved in the time domain to commutation relation) obtain ˆ† ˆ 0 0 −(iωc+κ/2)(t−t0) hξ (t)ξ(t )i = Nδ(t − t ) (E54) aˆ(t) = e aˆ(t0) ˆ ˆ† 0 0 √ Z t hξ(t)ξ (t )i = (N + 1)δ(t − t ) (E55) −(iωc+κ/2)(t−τ)ˆ − κ dτ e bin(τ). (E48) t 0 where N = nB(~ωc) is the thermal equilibrium occupa- If we take the input field to be a coherent drive at fre- tion number of the mode at the frequency of interest. We quency ωL = ωc + ∆ so that its amplitude has a classical can gain a better understanding of Eq. (E54) by Fourier and a quantum part transforming it to obtain the spectral density

ˆ −iωLt ¯ ˆ +∞ bin(t) = e [bin + ξ(t)] (E49) Z 0 S[ω] = dt hξˆ†(t)ξˆ(t0)ieiω(t−t ) = N. (E56) and if we take the limit t0 → ∞ so that the initial tran- −∞ sient in the cavity amplitude has damped out, then the solution of Eq. (E48) has the form postulated in Eq. (E4) As mentioned previously, this dimensionless quantity is with the spectral density that would be measured by a photo- √ multiplier: it represents the number of thermal photons κ ¯ passing a given point per unit time per unit bandwidth. a¯ = − bin (E50) −i∆ + κ/2 Equivalently the thermally radiated power in a narrow and (in the frame rotating at the drive frequency) bandwidth B is √ Z t P = ~ωNB. (E57) dˆ(t) = − κ dτ e+(i∆−κ/2)(t−τ)ξˆ(τ). (E51) −∞ One often hears the confusing statement that the noise Even in the absence of any classical drive, the input added by an amplifier is a certain number N of photons field delivers vacuum fluctuation noise to the cavity. No- (N = 20, say for a good cryogenic HEMT amplifier op- tice that from Eqs. (E31, E49) erating at 5 GHz). This means that the excess output ˆ ˆ† 0 ˆ ˆ† 0 noise (referred back to the input by dividing by the power [bin(t), bin(t )] = [ξ(t), ξ (t )] gain) produces a flux of N photons per second in a 1 Hz 1 X 0 6 = e−i(ωq−ωL)(t−t ) bandwidth, or 10 N photons per second in 1 MHz of 2πρ q bandwidth (see also Eq. (D27)). 0 We can gain further insight into input-output theory = δ(t − t ), (E52) ˆ by using the following picture. The operator bin(t) repre- which is similar to Eq. (D16) for a quantum transmission sents the classical drive plus vacuum fluctuations which line. This is the operator equivalent of white noise. Using are just about to arrive at the cavity. We will be able Eq. (E48) in the limit t0 → −∞ in Eqs. (E4,E51) yields to show that the output field is simply the input field a short while later after it has interacted with the cav- † ˆ ˆ† [ˆa(t), aˆ (t)] = [d(t), d (t)] ity. Let us consider the time evolution over a short time Z t Z t 0 −(−i∆+κ/2)(t−τ) period ∆t which is very long compared to the inverse = κ dτ dτ e bandwidth of the vacuum noise (i.e., the frequency scale −∞ −∞ beyond which the vacuum noise cannot be treated as con- −(+i∆+κ/2)(t−τ 0) 0 e δ(τ − τ ) stant due to some property of the environment) but very = 1 (E53) short compared to the cavity system’s slow dynamics. In 77 this circumstance it is useful to introduce the quantum of a dispersive position measurement, which employs a Wiener increment related to Eq. (D18) cavity whose resonance frequency shifts in response to the motion of a harmonic oscillator. This physical sys- Z t+∆t tem was considered heuristically in Sec. III.B.3. Here we dWc ≡ dτ ξˆ(τ) (E58) t will present a rigorous derivation using the (linearized) equations of motion for the coupled cavity and oscillator which obeys system. Let the dimensionless position operator [dWc , dWc†] = ∆t. (E59) 1 In the interaction picture (in a displaced frame in zˆ = xˆ = [ˆc† +c ˆ] (E64) x which the classical drive has been removed) the Hamilto- ZPF nian term that couples the cavity to the quantum noise be the coordinate of a harmonic oscillator whose energy of the environment is from Eq. (E18) is √ Vˆ = −i κ(ˆa†ξˆ− aˆξˆ†). (E60) † ~ HM = ~ωMcˆ cˆ (E65)

Thus the time evolution operator (in the interaction pic- and whose position uncertainty in the quantum ground ture) on the jth short time interval [tj, tj + ∆t] is p 2 state is xZPF = h0|xˆ |0i.

√ † † This Hamiltonian could be realized for example by ˆ κ(ˆa dWc −aˆ dWc) Uj = e (E61) mounting one of the cavity mirrors on a flexible cantilever (see the discussion above). Using this we can readily evolve the incoming temporal When the mirror moves, the cavity resonance fre- mode forward in time by a small step ∆t quency shifts, √ dWc0 = Uˆ †dWcUˆ ≈ dWc + κ∆t a.ˆ (E62) ω˜c = ωc[1 + Azˆ(t)] (E66) Recall that in input-output theory we formally defined the outgoing field as the bath field far in the future prop- where for a cavity of length L, A = −xZPF/L. agated back (using the free field time evolution) to the Assuming that the mirror moves slowly enough for the present, which yielded cavity to adiabatically follow its motion (i.e. Ω  κ), the outgoing light field suffers a phase shift which follows the √ ˆ ˆ bout = bin + κa.ˆ (E63) changes in the mirror position. This phase shift can be detected in the appropriate homodyne set up as discussed Eq. (E62) is completely equivalent to this. Thus we con- in Sec. III.B, and from this phase shift we can determine firm our understanding that the incoming field is the bath the position of the mechanical oscillator. In addition to temporal mode just before it interacts with the cavity and the actual zero-point fluctuations of the oscillator, our the outgoing field is the bath temporal mode just after it measurement will suffer from shot noise in the homodyne interacts with the cavity. signal and from additional uncertainty due to the back This leads to the following picture which is especially action noise of the measurement acting on the oscillator. useful in the quantum trajectory approach to conditional All of these effects will appear naturally in the derivation quantum evolution of a system subject to weak continu- below. ous measurement (Gardiner et al., 1992; Walls and Mil- We begin by considering the optical cavity equation ¯ burn, 1994). On top of the classical drive bin(t), the bath of motion based on Eq. (E32) and the optomechanical supplies to the system a continuous stream of “fresh” har- coupling Hamiltonian in Eq. (E1). These yield monic oscillators, each in their ground state (if T = 0). √ ˙ κ ˆ Each oscillator with its quantum fluctuation dWc inter- aˆ = −iωc(1 + Azˆ)ˆa − aˆ − κbin. (E67) acts briefly for a period ∆t with the system and then 2 is disconnected to propagate freely thereafter, never in- Let the cavity be driven by a laser at a frequency ωL = teracting with the system again. Within this picture it ωc +∆ detuned from the cavity by ∆. Moving to a frame is useful to think of the oscillators arrayed in an infinite rotating at ωL we have stationary line and the cavity flying over them at speed v and touching each one for a time ∆t. κ √ p aˆ˙ = +i(∆ − Aω zˆ)ˆa − aˆ − κˆb . (E68) c 2 in

3. Quantum limited position measurement using a cavity and we can write the incoming field as a constant plus detector white noise vacuum fluctuations (again, in the rotating frame) We will now apply the input-output formalism intro- ˆ ¯ ˆ duced in the previous section to the important example bin = bin + ξ (E69) 78 and similarly for the cavity field following Eq. (E4) The mechanical oscillator equation of motion which is identical in form to that of the optical cavity aˆ =a ¯ + d.ˆ (E70) γ0 √ i ∂tcˆ = −[ + iΩ]ˆc − γ0ηˆ(t) + [Hˆint, cˆ(t)], (E81) Substituting these expressions into the equation of mo- 2 ~ tion, we find that the constant classical fields obey ˆ √ where Hint is the Hamiltonian in Eq. (E1) andη ˆ is κ the mechanical vacuum noise from the (zero tempera- a¯ = − ¯b (E71) κ/2 − i∆ in ture) bath which is causing the mechanical damping at rate γ0. Using Eq. (E70) and expanding to first order and the new quantum equation of motion is, after ne- in small fluctuations yields the equation of motion lin- glecting a small term dˆzˆ: earized about the steady state solution

˙ κ √ γ0 √ g ˆ ˆ† dˆ= +i∆dˆ− iAω a¯zˆ − dˆ− κξ.ˆ (E72) ∂tcˆ = −[ + iΩ]ˆc − γ0ηˆ(t) + 2√ [ξ(t) − ξ (t)]. (E82) c 2 2 κ The quantum limit for position measurement will be It is useful to consider an equivalent formulation in reached only at zero detuning, so we specialize to the which we expand the Hamiltonian in Eq. (E1) to second case ∆ = 0. We also choose the incoming field amplitude order in the quantum fluctuations about the classical so- and phase to obey lution q ˆ ˆ† ˆ ˆ ¯ Hint ≈ ~ωcd d +x ˆF, (E83) bin = −i N,˙ (E73) where the force (including the coupling A) is (up to a so that sign) s N˙ ~g † a¯ = +2i , (E74) Fˆ = −i [dˆ− dˆ ]. (E84) κ xZPF Note that the radiation pressure fluctuations (photon where N˙ is the incoming photon number flux. The quan- shot noise) inside the cavity provide a forcing term. The tum equation of motion for the cavity then becomes state of the field inside the cavity in general depends on ˙ κ √ the past history of the cantilever position. However for dˆ= +gzˆ − dˆ− κξ,ˆ (E75) 2 this special case of driving the cavity on resonance, the dependence of the cavity field on the cantilever history is where the opto-mechanical coupling constant is propor- such that the latter drops out of the radiation pressure. tional to the laser drive amplitude To see this explicitly, consider the equation of motion for s the force obtained from Eq. (E75) N˙ √ g ≡ 2Aωc = Aωc n.¯ (E76) ˙ κ g √ κ Fˆ = − Fˆ + i ~ κ[ξˆ− ξˆ†]. (E85) 2 xZPF and Within our linearization approximation, the position of N˙ the mechanical oscillator has no effect on the radiation n¯ = |a¯|2 = 4 (E77) pressure (photon number in the cavity), but of course it κ does affect the phase of the cavity field (and hence the is the mean cavity photon number. Eq. (E75) is easily outgoing field) which is what we measure in the homo- solved by Fourier transformation dyne detection. Thus for this special casez ˆ does not appear on the 1 n √ o dˆ[ω] = gzˆ[ω] − κξˆ[ω] . (E78) RHS of either Eq. (E85) or Eq. (E82), which means that [κ/2 − iω] there is no optical renormalization of the cantilever fre- quency (‘optical spring’) or optical damping of the can- Let us assume that we are in the limit of low mechan- tilever. The lack of back-action damping in turn implies ical frequency relative to the cavity damping, Ω  κ, that the effective temperature Teff of the cavity detector so that the cavity state adiabatically follows the motion is infinite (cf. Eq. (2.8)). For this special case of zero of the mechanical oscillator. Then we obtain to a good detuning the back action force noise is controlled by a approximation single quadrature of the incoming vacuum noise (which 2 n √ o interferes with the classical drive to produce photon num- dˆ[ω] = gzˆ[ω] − κξˆ[ω] (E79) κ ber fluctuations). This is illustrated in the cavity ampli- 2 n √ o tude phasor diagram of Fig. (19). We see that the vac- dˆ†[ω] = gzˆ[ω] − κξˆ†[ω] (E80) ˆ ˆ† ˆ κ uum noise quadrature ξ + ξ conjugate to F controls the 79

where the equilibrium fluctuations in position are given by √  † zˆ0[ω] ≡ − γ0 χM[ω − Ω]ˆη[ω] + χM[ω + Ω]ˆη [ω] . (E90) We can now obtain the power spectrum Szz describing the total position fluctuations of the cantilever driven by the mechanical vacuum noise plus the radiation pressure shot noise. From Eqs. (E89, E90) we find

Sxx[ω] 2 = Szz[ω] xZPF 2 = γ0|χ[ω − Ω]| (E91) 2 xZPF 2 + |χM[ω − Ω] − χM[ω + Ω]| SFF . ~2

Note that (assuming high mechanical Q, i.e. γ0  Ω) the equilibrium part has support only at positive frequencies while the back action induced position noise is symmetric in frequency reflecting the effective infinite temperature of the back action noise. Symmetrizing this result with FIG. 19 (Color online) Phasor diagram for the cavity ampli- respect to frequency (and using γ0  Ω) we have tude showing that (for our choice of parameters) the imag- ˆ  ¯0  inary quadrature of the vacuum noise ξ interferes with the ¯ ¯0 Sxx[Ω] ¯ classical drive to produce photon number fluctuations while Sxx[ω] ≈ Sxx[ω] 1 + 2 SFF , (E92) ~ the real quadrature produces phase fluctuations which lead to measurement imprecision. The quantum fluctuations are il- ¯0 where Sxx[ω] is the symmetrized spectral density for posi- lustrated in the usual fashion, depicting the Gaussian Wigner tion fluctuations in the ground state given by Eq. (3.53). density of the coherent state in terms of color intensity. Now that we have obtained the effect of the back action noise on the position fluctuations, we must turn our at- tention to the imprecision of the measurement due to shot noise in the output. The appropriate homodyne quadra- phase noise which determines the measurement impreci- ture variable to monitor to be sensitive to the output sion (shot noise in the homodyne signal). This will be phase shift caused by position fluctuations is discussed further below. The solution for the cantilever position can again be ˆ ˆ† Iˆ = bout + b , (E93) obtained by Fourier transformation. For frequencies out small on the scale of κ the solution of Eq. (E85) is which, using the input-output results above, can be writ- ten ˆ 2i~g nˆ ˆ† o F [ω] = √ ξ[ω] − ξ [ω] (E86) Iˆ = −(ξˆ+ ξˆ†) + λx.ˆ (E94) xZPF κ and hence the back action force noise spectral density is We see that the cavity homodyne detector system acts at low frequencies as a position transducer with gain 4g 4 2g2 λ = √ . (E95) ~ x κ SFF [ω] = 2 (E87) ZPF xZPFκ The first term in Eq. (E94) represents the vacuum noise in agreement with Eq. (3.17). that mixes with the homodyne local oscillator to produce Introducing a quantity proportional to the cantilever the shot noise in the output. The resulting measurement (mechanical) susceptibility (within the rotating wave ap- imprecision (symmetrized) spectral density referred back proximation we are using) to the position of the oscillator is

1 ¯I 1 χM[ω − Ω] ≡ , (E88) Sxx = . (E96) γ0 λ2 −i(ω − Ω) + 2 Comparing this to Eq. (E87) we see that we reach the we find from Eq. (E82) quantum limit relating the imprecision noise to the back action noise i ˆ zˆ[ω] =z ˆ0[ω] − xZPF {χM[ω − Ω] − χM[ω + Ω]} F [ω], 2 ~ S¯I S¯ = ~ (E97) (E89) xx FF 4 80 in agreement with Eq. (3.9). frequency where we find that the optimal coupling leads Notice also from Eq. (E94) that the quadrature of the to an optimal back action noise vacuum noise which leads to the measurement impreci- sion is conjugate to the one which produces the back 2 2γ S¯ = ~ = ~ 0 . (E100) action force noise as illustrated previously in Fig. (19). FF,opt 2S¯0 [Ω] 4x2 Recall that the two quadratures of motion of a harmonic xx ZPF oscillator in its ground state have no classical (i.e., sym- This makes sense because the higher the damping the less metrized) correlation. Hence the symmetrized cross cor- susceptible the oscillator is to back action forces. At this relator optimal coupling the total output noise spectral density at frequency Ω referred to the position is simply twice S¯ [ω] = 0 (E98) IF the vacuum value vanishes. Because there is no correlation between the ¯ ¯0 output imprecision noise and the forces controlling the Sxx,tot[Ω] = 2Sxx[Ω], (E101) position fluctuations, the total output noise referred back to the position of the oscillator is simply in agreement with Eq. (3.61). Evaluation of Eq. (E100) at the optimal coupling yields the graph shown in ¯ ¯ ¯I Sxx,tot[ω] = Sxx[ω] + Sxx (E99) Fig. (6). The background noise floor is due to the  S¯0 [Ω]  2 frequency independent imprecision noise with value ¯0 xx ¯ ~ 1 0 = S [ω] 1 + SFF + . S¯ [Ω]. The peak value at ω = Ω rises a factor of three xx 2 4S¯ 2 xx ~ FF above this background. This expression again clearly illustrates the competition We derived the gain λ in Eq. (E95) by direct solution between the back action noise proportional to the drive of the equations of motion. With the results we have de- laser intensity and the measurement imprecision noise rived above, it is straightforward to show that the Kubo which is inversely proportional. We again emphasize that formula in Eq. (4.3) yields equivalent results. We have all of the above relations are particular to the case of zero already seen that the classical (i.e. symmetrized) correla- detuning of the cavity drive field from the cavity. tions between the output signal Iˆ and the force Fˆ which The total output noise at some particular frequency couples to the position vanishes. However the Kubo for- will be a minimum at some optimal drive intensity. The mula evaluates the quantum (i.e. antisymmetric) corre- precise optimal value depends on the frequency chosen. lations for the uncoupled system (A = g = 0). Hence we Typically this is taken to be the mechanical resonance have

  i ˆ ˆ†  2i~g ˆ ˆ†  χIF (t) = − θ(t) − ξ(t − δt) + ξ (t − δt) , √ ξ(0) − ξ (0) , (E102) ~ xZPF κ 0

where δt is a small (positive) time representing the delay all of the key results in Sec. V.E derived from completely between the time when the vacuum noise impinges on general considerations of linear response theory. the cavity and when the resulting outgoing wave reaches the homodyne detector. (More precisely it also compen- sates for certain small retardation effects neglected in the limit ω  κ used in several places in the above deriva- 4. Back-action free single-quadrature detection tions.) Using the fact that the commutator between the two quadratures of the vacuum noise is a delta function, We now provide details on the cavity single-quadrature Fourier transformation of the above yields (in the limit detection scheme discussed in Sec. V.H.2. We again con- ω δt  1 the desired result sider a high-Q cavity whose resonance frequency is mod- ulated by a high-Q mechanical oscillator with co-ordinate

χIF [ω] = λ. (E103) xˆ (cf. Eqs. (E1) and (E64)). To use this system for ampli- fication of a single quadrature, we will consider the typi- Similarly we readily find that the small retardation cal case of a fast cavity (ωc  Ω), and take the “good cav- causes the reverse gain to vanish. Hence all our results ity” limit, where Ω  κ. As explained in the main text, are consistent with the requirements needed to reach the the crucial ingredient for single-quadrature detection is standard quantum limit. to take an amplitude-modulated cavity drive described ¯ Thus with this study of the specific case of an oscillator by the classical input field bin given in Eq. (5.101). As parametrically coupled to a cavity, we have reproduced before (cf. Eq. (E4)), we may write the cavity annihila- 81 tion operatora ˆ as the sum of a classical piecea ¯(t) and a where f(t) is a complex function which is slowly varying quantum piece dˆ; only dˆ is influenced by the mechanical on the scale of an oscillator period. oscillator.a ¯(t) is easily found from the classical (noise- The equations of motion are easily solved upon Fourier free) equations of motion for the isolated cavity; making transformation, resulting in: use of the conditions ωc  Ω  κ, we have " ˆ ∗ p Xδ[ω] = −xZPF · χM [ω] i (f [−ω] − f[ω]) (E109a) Nκ˙ a¯(t) ' cos (Ωt + δ) e−iωct (E104) 2Ω # √ iδ −iδ †  To proceed with our analysis, we work in an interac- + γ0 e ηˆ(ω + Ω) + e ηˆ (ω − Ω) tion picture with respect to the uncoupled cavity and oscillator Hamiltonians. Making standard rotating-wave " ˆ ∗ approximations, the Hamiltonian in the interaction pic- Yδ[ω] = ixZPF · χM [ω] (−i)(f[ω] + f [−ω]) (E109b) ture takes the simple form corresponding to Eq. (5.102b): √ + γ eiδηˆ(ω + Ω) − e−iδηˆ†(ω − Ω) ˜  ˆ ˆ† iδ −iδ † 0 Hint = ~A d + d e cˆ + e cˆ # √ ˜ ˆ ˆ†  ˆ −2iAχc[ω] κ ξ(ω + ωc) + ξ (ω − ωc)  † Xδ = ~A˜ dˆ+ dˆ , (E105) xZPF where the cavity and mechanical susceptibilities χc, χM where are defined in Eqs. (E43) and (E88). p ˆ Nκ˙ As anticipated, the detected quadrature Xδ is com- A˜ = A · ω , (E106) c 4Ω pletely unaffected by the measurement: Eq. (E109a) is identical to what we would have if there were no coupling and in the second line, we have made use of the definition between the oscillator and the cavity. In contrast, the of the quadrature operators Xˆ , Yˆ given in Eqs. (5.92). δ δ conjugate quadrature Yˆ experiences an extra stochastic The form of H was discussed heuristically in the main δ int force due to the cavity: this is the measurement back- text in terms of Raman processes where photons are re- action. moved from the classical drive ¯b and either up or down in Turning now to the output field from the cavity ˆb , converted to the cavity frequency via absorption or emis- out we use the input-output relation Eq. (E37) to find in the sion of a mechanical phonon. Alternatively, we can think lab (i.e. non-rotating) frame: of the drive yielding a time-dependent cavity-oscillator   coupling which “follows” the Xδ quadrature. Note that ˆ ¯ −i(ω − ωc) − κ/2 ˆ bout[ω] = bout[ω] + ξ[ω] we made crucial of use of the good cavity limit (κ  Ω) −i(ω − ωc) + κ/2 ˆ √ to drop terms in Hint which oscillate at frequencies ±2Ω. A˜ κ These terms represent Raman sidebands which are away −i χ [ω − ω ] · Xˆ (ω − ω ) x c c δ c from the cavity resonance by a distance ±2Ω. In the good ZPF cavity limit, the density of photon states is negligible so (E110) far off resonance and these processes are suppressed. The first term on the RHS simply represents the output Similar to Eqs. (E39) and (E81), the Heisenberg equa- field from the cavity in the absence of the mechanical tions of motion (in the rotating frame) follow directly oscillator and any fluctuations. It will yield sharp peaks from Hint and the dissipative terms in the total Hamil- at the two sidebands associated with the drive, ω = ωc ± tonian: Ω. The second term on the RHS of Eq. (E110) represents κ √ ∂ dˆ = − dˆ− κξˆ(t)eiωct − iA˜ eiδcˆ + e−iδcˆ† (E107a) the reflected noise of the incident cavity drive. This noise t 2 will play the role of the “intrinsic ouput noise” of this γ0 √ iΩt −iδ h  ˆ ˆ† i amplifier. ∂tcˆ = − cˆ − γ0ηˆ(t)e − ie A˜ d + d − f(t) 2 Finally, the last term on the RHS of Eq. (E110) is the (E107b) amplified signal: it is simply the amplified quadrature ˆ ˆ Xδ of the oscillator. This term will result in a peak in As before, ξ(t) represents the unavoidable noise in the the output spectrum at the resonance frequency of the cavity drive, andη ˆ(t), γ0 are the noisy force and damp- cavity, ω . As there is no back-action on the measured ing resulting from an equilibrium bath coupled to the c Xˆδ quadrature, the added noise can be made arbitrarily mechanical oscillator. Note from Eq. (E107a) that as an- small by simply increasing the drive strength N˙ (and ticipated, the cavity is only driven by one quadrature of hence A˜). the oscillator’s motion. We have also included a driving force F (t) on the mechanical oscillator which has some narrow bandwidth centered on the oscillator frequency; Appendix F: Information Theory and Measurement Rate this force is parameterized as: 2 Suppose that we are measuring the state of a qubit F (t) = ~ Re f(t)e−iΩte−iδ (E108) xZPF via the phase shift ±θ0 from a one-sided cavity. Let I(t) 82

be the homodyne signal integrated up to time t as in Using Eq. (F1) the polarization is easily evaluated Sec. III.B. We would like to understand the relationship   Iθ0 between the signal-to-noise ratio defined in Eq. (3.22), mI = tanh . (F9) and the rate at which information about the state of the Sθθ qubit is being gained. The probability distribution for I The information gain is thus conditioned on the state of the qubit σ = ±1 is   2  2  2  1 2 Iθ0 I θ0 1 −(I − σθ0t) II = tanh ≈ (F10) p(I|σ) = √ exp . (F1) 2 Sθθ 2 Sθθ 2πSθθt 2Sθθt where the second equality is only valid for small |m|. Based on knowledge of this conditional distribution, we Ensemble averaging this over all possible measurement now present two distinct but equivalent approaches to results yields the mean information gain at short times giving an information theoretic basis for the definition of the measurement rate. 1 θ2 I ≈ 0 t (F11) 2 Sθθ 1. Method I which justifies the definition of the measurement rate given in Eq. (3.23). Suppose we start with an initial qubit density matrix ! 1 2. Method II 2 0 ρ0 = . (F2) 0 1 2 An alternative information theoretic derivation is to After measuring for a time t, the new density matrix consider the qubit plus measurement device to be a sig- conditioned on the results of the measurement is naling channel. The two possible inputs to the channel ! are the two states of the qubit. The output of the chan- p+ 0 nel is the result of the measurement of I. By toggling ρ1 = (F3) 0 p− the qubit state back and forth, one can send information through the signal channel to another party. The chan- where it will be convenient to parameterize the two prob- nel is noisy because even for a fixed state of the qubit, abilities by the polarization m ≡ Tr(σzρ1) by the measured values of the signal I have intrinsic fluctu- 1 ± m ations. Shannon’s noisy channel coding theorem (Cover p = . (F4) and Thomas, 1991) tells us the maximum rate at which ± 2 information can be reliably sent down the channel by tog- The information gained by the measurement is the en- gling the state of the qubit and making measurements of tropy loss26 of the qubit I. It is natural to take this rate as defining the measure- ment rate for our detector. I = Tr(ρ ln ρ − ρ ln ρ ). (F5) 1 1 0 0 The reliable information gain by the receiver on a noisy We are interested in the initial rate of gain of information channel is a quantity known as the ‘mutual information’ 2 of the communication channel (Clerk et al., 2003; Cover at short times θ0t  Sθθ where m will be small. In this limit we have and Thomas, 1991) m2 Z +∞ ( ) I ≈ . (F6) X 2 R = − dI p(I) ln p(I) − p(σ)[p(I|σ) ln p(I|σ)] −∞ σ We must now calculate m conditioned on the measure- (F12) ment result I The first term is the Shannon entropy in the signal I X when we do not know the input signal (the value of the m ≡ σp(σ|I). (F7) I qubit). The second term represents the entropy given σ that we do know the value of the qubit (averaged over From Bayes theorem we can express this in terms of the two possible input values). Thus the first term is p(I|σ), which is the quantity we know, signal plus noise, the second is just the noise. Subtracting the two gives the net information gain. Expanding this p(I|σ)p(σ) expression for short times yields p(σ|I) = P 0 0 . (F8) σ0 p(I|σ )p(σ ) 1 (hI(t)i − hI(t)i )2 R = + − 8 Sθθt 2 θ0 26 It is important to note that we use throughout here the physi- = t cist’s entropy with the natural logarithm rather than the log base 2Sθθ 2 which gives the information in units of bits. = Γmeast (F13) 83 exactly the same result as Eq. (F11). (Here hI(t)iσ is the and mean value of I given that the qubit is in state σ.) hYˆ i = 0. (G9) Appendix G: Number Phase Uncertainty If the phase of this wave undergoes a small modulation In this appendix, we briefly review the number-phase due for example to weak parametric coupling to a qubit uncertainty relation, and from it we derive the relation- then one can estimate the phase by ship between the spectral densities describing the photon number fluctuations and the phase fluctuations. Con- hYˆ i hθi = . (G10) sider a coherent state labeled by its classical amplitude ˆ α hXi  |α|2  |αi = exp − exp{αaˆ†}|0i. (G1) This result is of course only valid for small angles, θ  1. 2 For N¯  1, the uncertainty will be

This is an eigenstate of the destruction operator hYˆ 2i 1 h0|dˆdˆ†|0i 1 (∆θ)2 = = 2 = . (G11) aˆ|αi = α|αi. (G2) (hXˆi)2 2N¯ 4N¯ Thus using Eq. (G5) we arrive at the fundamental quan- It is convenient to make the unitary displacement trans- tum uncertainty relation formation which maps the coherent state onto a new vac- uum state and the destruction operator onto 1 ∆θ∆N = . (G12) 2 aˆ = α + dˆ (G3) where d annihilates the new vacuum. Then we have Using the input-output theory described in Ap- pendix E we can restate the results above in terms of N¯ = hNˆi = h0|(α∗ + dˆ†)(α + dˆ)|0i = |α|2, (G4) noise spectral densities. Let the amplitude of the field coming in to the homodyne detector be and ˆ ¯ ˆ (∆N)2 = h(Nˆ − N¯)2i = |α|2h0|dˆdˆ†|0i = N.¯ (G5) bin = bin + ξ(t) (G13)

Now define the two quadrature amplitudes where ξˆ(t) is the vacuum noise obeying 1 Xˆ = √ (ˆa +a ˆ†) (G6) 2 [ξˆ(t), ξˆ†(t0)] = δ(t − t0). (G14) i Yˆ = √ (ˆa† − aˆ). (G7) 2 We are using a flux normalization for the field operators so Each of these amplitudes can be measured in a homodyne experiment. For convenience, let us take α to be real and ˙ ˆ† ˆ ¯ 2 positive. Then N = hbinbini = |bin| (G15) √ hXˆi = 2α (G8) and

2 ˙ ˙ ˙ ¯∗ ˆ† ¯ ˆ ¯∗ ˆ† ¯ ˆ ¯ 4 ˙ hN(t)N(0)i − N = h0|(bin + ξ (t))(bin + ξ(t))(bin + ξ (0))(bin + ξ(0))|0i − bin = Nδ(t). (G16)

From this it follows that the shot noise spectral density ture operator is

S ˙ ˙ = N.˙ (G17) NN i(ˆb† − ˆb ) (ξˆ† − ξˆ) θˆ = in in = hθˆi + i (G18) ˆ† ˆ 2¯b Similarly the phase can be estimated from the quadra- hbin + bini in 84 which has noise correlator cold load 1 hδθˆ(t)δθˆ(0)i = δ(t) (G19) 4N˙ by,in by,out corresponding to the phase imprecision spectral density mirror Y 1 Sθθ = . (G20) ˙ 4N ay,out ay,in We thus arrive at the fundamental quantum limit relation input line output line bz,out az,in ax,out bx,in 1 pS S = . (G21) θθ N˙ N˙ 2 bz,in Z az,out ax,in X bx,out

1/2 1/2 † Appendix H: Using feedback to reach the quantum limit cin cout=G cin+(G-1) v in ideal In Sec. (VI.B), we demonstrated that any two port am- 1-port plifier whose scattering matrix has s11 = s22 = s12 = 0 amp. will fail to reach quantum limit when used as a weakly coupled op-amp; at best, it will miss optimizing the quan- v v tum noise constraint of Eq. (5.88) by a factor of two. in out Reaching the quantum limit thus requires at least one FIG. 20 Schematic of a modified minimal two-port amplifier, of s11, s22 and s12 to be non-zero. In this subsection, where partially reflecting mirrors have been inserted in the we demonstrate how this may be done. We show that input and output transmission lines, as well as in the line by introducing a form of negative feedback to the “min- leading to the cold load. By tuning the reflection coefficient imal” amplifier of the previous subsection, one can take of the mirror in the cold load arm (mirror Y ), we can in- advantage of noise correlations to reduce the back-action duce negative feedback which takes advantage of correlations current noise SII by a factor of two. As a result, one is between current and voltage noise. This then allows this sys- able to reach the weak-coupling (i.e. op-amp) quantum tem to reach the quantum limit as a weakly coupled voltage limit. Note that quantum amplifiers with feedback are op amp. See text for further description also treated in Courty et al. (1999); Grassia (1998). On a heuristic level, we can understand the need for either reflections or reverse gain to reach the quantum impedance matched to the circulator. Each mirror will limit. A problem with the “minimal” amplifier of the be described by a 2 × 2 unitary scattering matrix: last subsection was that its input impedance was too low in comparison to its noise impedance Z ∼ Z . From ! ˆ ! N a aˆj,out bj,in = Uj · (H1) general expression for the input impedance, Eq. (6.7d), ˆb aˆ we see that having non-zero reverse gain (i.e. s 6= 0) j,out j,in 12 ! and/or non-zero reflections (i.e. s11 6= 0 and/or s22 6= 0) cos θj − sin θj Uj = (H2) could lead to Zin  Za. This is exactly what occurs sin θ cos θ when feedback is used to reach the quantum limit. Keep j j in mind that having non-vanishing reverse gain is danger- Here, the index j can take on three values: j = z for the 0 ous: as we discussed earlier, an appreciable non-zero λI mirror in the input line, j = y for the mirror in the arm can lead to the highly undesirable consequence that the going to the cold load, and j = x for the mirror in the amplifier’s input impedance depends on the impedance output line. The mode aj describes the “internal” mode of the load connected to its output (cf. Eq. (6.6)). which exists between the mirror and circulator, while the mode bj describes the “external” mode on the other side of the mirror. We have taken the Uj to be real for con- 1. Feedback using mirrors venience. Note that θj = 0 corresponds to the case of no mirror (i.e. perfect transmission). To introduce reverse gain and reflections into the “min- It is now a straightforward though tedious exercise to imal” two-port bosonic amplifier of the previous subsec- construct the scattering matrix for the entire system. tion, we will insert mirrors in three of the four arms lead- From this, one can identify the reduced scattering matrix ing from the circulator: the arm going to the input line, s appearing in Eq. (6.3), as well as the noise operators the arm going to the output line, and the arm going to Fj. These may then in turn be used to obtain the op-amp the auxiliary “cold load” (Fig. 20). Equivalently, one description of the amplifier, as well as the commutators could imagine that each of these lines is not perfectly of the added noise operators. These latter commutators 85 determine the usual noise spectral densities of the ampli- at the quantum limit, the matrix s satisfies: fier. Details and intermediate steps of these calculations s = −s (H3a) may be found in Appendix I.6. 11 22 1 As usual, to see if our amplifier can reach the quantum s12 = s21 (H3b) limit when used as a (weakly-coupled) op-amp , we need G to see if it optimizes the quantum noise constraint of The second equation also carries over to the op-amp pic- Eq. (5.88). We consider the optimal situation where both ture; at the quantum limit, one has: the auxiliary modes of the amplifier (ˆu andv ˆ† ) are in 1 in in λ0 = λ (H4) the vacuum state. The surprising upshot of our analysis I G V (see Appendix I.6) is the following: if we include a small One particularly simple limit is the case where there amount of reflection in the cold load line with the correct are no mirrors in the input and output line (θ = θ = 0), phase, then we can reach the quantum limit, irrespective x z only a mirror in the cold-load arm. When this mirror is of the mirrors in the input and output lines. In particular, √ √ tuned to reach the quantum limit (i.e. sin θ = −1/ G), if sin θ = −1/ G, our amplifier optimizes the quantum y y the scattering matrix takes the simple form: noise constraint of Eq. (5.88) in the large gain (i.e. large √ ! G) limit, independently of the values of θx and θy. Note 0 1/ G s = √ (H5) that tuning θy to reach the quantum limit does not have G 0 a catastrophic impact on other features of our amplifier. One can verify that this tuning only causes the voltage In this case, the principal effect of the weak mirror in gain λV and power gain GP to decrease by a factor of the cold-load line is to introduce a small amount of re- two compared to their θy = 0 values (cf. Eqs. (I60) and verse gain. The amount of this reverse gain is exactly (I64)). This choice for θy also leads to Zin  Za ∼ ZN what is needed to have the input impedance diverge (cf. (I62)), in keeping with our general expectations. (cf. Eq. (6.7d)). It is also what is needed to achieve an

Physically, what does this precise tuning of θy corre- optimal, noise-canceling feedback in the amplifier. To spond to? A strong hint is given by the behaviour of the see this last point explicitly, we can re-write the ampli- ˜ amplifier’s cross-correlation noise S¯VI [ω] (cf. Eq. (I65c)). fier’s back-action current noise (I) in terms of its original ¯ noises I˜ and V˜ (i.e. what the noise operators would have In general, we find that SVI [ω]√ is real and non-zero. 0 0 However, the tuning sin θy = −1/ G is exactly what is been in the absence of the mirror). Taking the relevant needed to have S¯VI vanish. Also note from Eq. (I65a) limit of small reflection (i.e.r ˜ ≡ sin θy goes to zero as that this special tuning of θy decreases the back-action |G| → ∞), we find that the modification of the current current noise precisely by a factor of two compared to noise operator is given by: √ its value at θy = 0. A clear physical explanation now 2 Gr˜ V˜0 emerges. Our original, reflection-free amplifier had cor- I˜ ' I˜0 + √ (H6) relations between its back-action current noise and out- 1 − Gr˜ Za put voltage noise (cf. Eq. (6.18c)). By introducing nega- As claimed, the presence of a small amount of reflection tive feedback of the output voltage to the input current r˜ ≡ sin θy in the cold load arm “feeds-back” the original (i.e. via a mirror in the cold-load arm), we are able to ˜ voltage noise of√ the amplifier V0 into the current. The use these correlations to decrease the overall magnitude choicer ˜ = −1/ G corresponds to a negative feedback, of the current noise (i.e. the voltage fluctuations V˜ par- and optimally makes use of the fact that I˜0 and V˜0 are tially cancel the original current fluctuations I˜). For an correlated to reduce the overall fluctuations in I˜. optimal feedback (i.e. optimal choice of θy), the current While it is interesting to note that one can reach the noise is reduced by a half, and the new current noise is quantum limit with no reflections in the input and output not correlated with the output voltage noise. Note that arms, this case is not really of practical interest. The this is indeed negative (as opposed to positive) feedback– 0 reverse√ current gain in this case may be small (i.e. λI ∝ it results in a reduction of both the gain and the power 1/ G), but it is not small enough: one finds that because gain. To make this explicit, in the next section we will 0 of the non-zero λI , the amplifier’s input impedance is map the amplifier described here onto a standard op-amp strongly reduced in the presence of a load (cf. Eq. (6.6)). with negative voltage feedback. There is a second simple limit we can consider which is more practical. This is the limit where reflections in the input-line mirror and output-line mirror are both strong. 1/8 Imagine we take θz√= −θx = π/2 − δ/G . If again 2. Explicit examples we set sin θy = −1/ G to reach the quantum limit, the scattering matrix now√ takes the form (neglecting terms To obtain a more complete insight, it is useful to go which are order 1/ G): back and consider what the reduced scattering matrix of ! +1 0 our system looks like when θy has been tuned to reach s = δ2G1/4 (H7) the quantum limit. From Eq. (I58), it is easy to see that 2 −1 86

(B = 0), an analysis of the circuit equations for our op- Va + amp system yields: Gf Vb - Gf R1 R2 λV = (H11a) 1 + B · Gf

0 B λI = (H11b) 1 + B · Gf FIG. 21 Schematic of a voltage op-amp with negative feed- Zb Zout = (H11c) back. 1 + B · Gf

Zin = (1 + B · Gf )Za (H11d) 2 Gf /2 In this case, we see that at the quantum limit, the reflec- GP = (H11e) B · Gf + 2Zb/Za tion coefficients s11 and s22 are exactly what is needed to have the input impedance diverge, while the reverse Again, Gf represents the gain of the amplifier in the ab- gain coefficient s12 plays no role. For this case of strong sence of any feedback, Za is the input impedance at zero reflections in input and output arms, the voltage gain is feedback, and Zb is the output impedance at zero feed- reduced compared to its zero-reflection value: back. Transforming this into the scattering picture yields a r  2 Zb δ 1/4 scattering matrix s satisfying: λV → G (H8) Za 2 BGf (Za − (2 + BGf )Zb) s11 = −s22 = − 2 (H12a) The power gain however is independent of θ , θ , and is BGf Za + (2 + BGf ) Zb x z √ still given by G/2 when θ is tuned to be at the quantum y 2 ZaZbGf (1 + BGf ) limit. s21 = − 2 (H12b) BGf Za + (2 + BGf ) Zb B s12 = s21 (H12c) 3. Op-amp with negative voltage feedback Gf

We now show that a conventional op-amp with feed- Note the connection between these equations and the nec- back can be mapped onto the amplifier described in essary form of a quantum limited s-matrix found in the the previous subsection. We will show that tuning the previous subsection. strength of the feedback in the op-amp corresponds to Now, given a scattering matrix, one can always find a tuning the strength of the mirrors, and that an optimally minimal representation of the noise operators Fa and Fb tuned feedback circuit lets one reach the quantum limit. which have the necessary commutation relations. These This is in complete correspondence to the previous sub- are given in general by: section, where an optimal tuning of the mirrors also lets Fˆ = p1 − |s |2 − |s |2 + |l|2 · uˆ + l · vˆ† (H13) one reach the quantum limit. a 11 12 in in ˆ p 2 2 † More precisely, we consider a scattering description of Fb = |s21| + |s22| − 1 · vˆin (H14) a non-inverting op-amp amplifier having negative voltage s s∗ + s s∗ l = 11 21 12 22 (H15) feedback. The circuit for this system is shown in Fig. 21. p|s |2 + |s |2 − 1 A fraction B of the output voltage of the amplifier is fed 21 22 back to the negative input terminal of the op-amp. In Applying this to the s matrix for our op-amp, and then practice, B is determined by the two resistors R and R † 1 2 taking the auxiliary modesu ˆin andv ˆ to be in the vac- used to form a voltage divider at the op-amp output. The in uum state, we can calculate the minimum allowed S¯VV op-amp with zero feedback is described by the “ideal” and S¯II for our non-inverting op-amp amplifier. One can amplifier of Sec. VI.B: at zero feedback, it is described then calculate the product S¯VV S¯II and compare against by Eqs. (6.11a)- (6.11d). For simplicity, we consider the the quantum-limited value (S¯IV is again real). In the relevant case where: case of zero feedback (i.e. B = 0), one of course finds that this product is twice as big as the quantum limited Zb  R1,R2  Za (H9) value. However, if one takes the large Gf limit while keeping B non-zero but finite, one obtains: In this limit, R1 and R2 only play a role through the feedback fraction B, which is given by:  2! ¯ ¯ 2 2B 1 SVV SII → (~ω) 1 − + O (H16) R Gf Gf B = 2 (H10) R1 + R2 Thus, for a fixed, non-zero feedback ratio B, it is possible Letting Gf denote the voltage gain at zero feedback to reach the quantum limit. Note that if B does not tend 87 to zero as Gf tends to infinity, the voltage gain of this quantum limit. amplifier will be finite. The power gain however will be proportional to Gf and will be large. If one wants a large voltage gain, one could set B to go to zero with Appendix I: Additional Technical Details 1 Gf i.e. B ∝ √ . In this case, one will still reach Gf the quantum limit in the large Gf limit, and the voltage This appendix provides further details of calculations p presented in the main text. gain will also be large (i.e. ∝ Gf ). Note that in all these limits, the reflection coefficients s11 and s22 tend to −1 and 1 respectively, while the reverse gain tends to 0. This is in complete analogy to the amplifier with 1. Proof of quantum noise constraint mirrors considered in the previous subsection, in the case where we took the reflections to be strong at the input Note first that we may write the symmetrized Iˆ and Fˆ and at the output (cf. Eq. (H7)). We thus see yet again noise correlators defined in Eqs. (4.4a) and (4.4b) as sums how the use of feedback allows the system to reach the over transitions between detector energy eigenstates:

¯ X ˆ 2 SFF [ω] = π~ hi|ρˆ0|ii · |hf|F |ii| [δ(Ef − Ei + ~ω) + δ(Ef − Ei − ~ω)] (I1) i,f ¯ X ˆ 2 SII [ω] = π~ hi|ρˆ0|ii · |hf|I|ii| [δ(Ef − Ei + ~ω) + δ(Ef − Ei − ~ω)] (I2) i,f

Here,ρ ˆ0 is the stationary density matrix describing the we see that the noise correlators S¯II and S¯FF may be state of the detector, and |ii (|fi) is a detector energy written as: eigenstate with energy E (E ). Eq. (I1) expresses the i f ¯ noise at frequency ω as a sum over transitions. Each SII [ω]dω = h~v,~viω (I6) transition starts with an an initial detector eigenstate S¯FF [ω]dω = h~w, ~wiω (I7) |ii, occupied with a probability hi|ρ0|ii, and ends with a final detector eigenstate |fi, where the energy difference We may now employ the Cauchy-Schwartz inequality: between the two states is either + ω or − ω . Further, ~ ~ h~v,~vi h~w, ~wi ≥ |h~v, ~wi |2 (I8) each transition is weighted by an appropriate matrix el- ω ω ω ement. A straightforward manipulation shows that the real To proceed, we fix the frequency ω > 0, and let the part of h~v, ~wiω is determined by the symmetrized cross- ¯ index ν label each transition |ii → |fi contributing to the correlator SIF [ω] defined in Eq. (4.4c): noise. More specifically, ν indexes each ordered pair of Re h~v, ~wi = Re eiδS¯ [ω] dω (I9) detector energy eigenstates states {|ii, |fi} which satisfy ω IF

Ef − Ei ∈ ±~[ω, ω + dω] and hi|ρ0|ii= 6 0. We can now In contrast, the imaginary part of h~v, ~wiω is independent ˆ ˆ consider the matrix elements of I and F which contribute of S¯IF ; instead, it is directly related to the gain χIF and ¯ ¯ to SII [ω] and SFF [ω] to be complex vectors ~v and ~w. reverse gain χFI of the detector: Letting δ be any real number, let us define: Im h~v, ~wi = ~Re eiδ χ [ω] − [χ [ω]]∗ dω (I10) ω 2 IF FI ˆ [~w]ν = hf(ν)|F |i(ν)i (I3) Substituting Eqs. (I10) and (I9) into Eq. (I8), one im- ( −iδ ˆ mediately finds the quantum noise constraint given in e hf(ν)|I|i(ν)i if Ef(ν) − Ei(ν) = +~ω, [~v] = Eq. (4.16). As in the main text, we letχ ˜ = χ −χ∗ . ν eiδhf(ν)|Iˆ|i(ν)i if E − E = − ω. IF IF FI f(ν) i(ν) ~ Maximizing the RHS of this inequality with respect to (I4) the phase δ, one finds that the maximum is achieved for ˜ δ = δ0 = − arg(˜χIF ) + δ0 with ¯ Introducing an inner product h·, ·iω via: ˜ |SIF | sin 2φ tan 2δ0 = − ¯ (I11) (~/2)|χ˜IF | + |SIF | cos 2φ ~ X ∗ where φ = arg(S¯ χ˜ ∗). At δ = δ , Eq. (4.16) becomes h~a, biω = π hi(ν)|ρˆ0|i(ν)i · (aν ) bν , (I5) IF IF 0 ν the final noise constraint of Eq. (4.11). 88

The proof given here also allows one to see what must where the plus sign corresponds to Eb > Ea, the mi- be done in order to achieve the “ideal” noise condition nus to Ea > Eb. If now the final state |bi was also oc- of Eq. (4.17): one must achieve equality in the Cauchy- cupied (i.e. hb|ρˆ0|bi= 6 0), then the reverse transition Schwartz inequality of Eq. (I8). This requires that the |i = bi → |f = ai) would also contribute to the noise. vectors ~v and ~w be proportional to one another; there The proportionality condition of Eq. (I13) would now re- must exist a complex factor α (having dimensions [I]/[F ]) quire: such that: ha|Iˆ|bi = e∓iδαha|Fˆ|bi (I20) ~v = α · ~w (I12) As Iˆ and Fˆ are both Hermitian operators, and as α must Equivalently, we have that have an imaginary part in order for there to be gain, we ( have a contradiction: Eq. (I19) and (I20) cannot both be eiδαhf|F |ii if E − E = + ω f i ~ true. It thus follows that the final state of a transition hf|I|ii = −iδ (I13) e αhf|F |ii if Ef − Ei = −~ω. contributing to the noise must be unoccupied in order for Eq. (I13) to be satisfied and for the detector to have for each pair of initial and final states |ii, |fi contributing ¯ ¯ ideal noise properties. Note that this necessary asym- to SFF [ω] and SII [ω] (cf. Eq. (I1)). Note that this not metry in the occupation of detector energy eigenstates the same as requiring Eq. (I13) to hold for all possible immediately tells us that a detector or amplifier cannot states |ii and |fi. This proportionality condition in turn reach the quantum limit if it is in equilibrium. implies a proportionality between the input and output (unsymmetrized) quantum noise spectral densities:

2 2. Proof that a noiseless detector does not amplify SII [ω] = |α| SFF [ω] (I14) It thus also follows that the imaginary parts of the input With the above results in hand, we can now prove as- and output susceptibilities are proportional: sertions made in Sec. IV.A.4 that detectors which evade the quantum noise constraint of Eq. (4.11) and simply 2 Im χII [ω] = |α| Im χFF [ω], (I15) satisfy as well as the symmetrized input and output noise (i.e. 2 S¯FF S¯II = |S¯IF | (I21) Eq. (4.18)). Finally, one can also use Eq. (I13) to relate the unsymmetrized I-F quantum noise correlator SIF [ω] are at best transducers, as their power gain is limited to to SFF [ω]: (cf. Eq. (4.7)): being at most one. The first way to make the RHS of Eq. (4.11) vanish is ( −iδ ∗ e α SFF [ω] if ω > 0, ∗ to have χIF = χFI . We have already seen that when- SIF [ω] = iδ ∗ (I16) e α SFF [ω] if ω < 0 ever this relation holds, the detector power gain cannot be any larger than one (c.f. Eq. (5.53)). Now, imagine Note that SFF [ω] is necessarily real and positive. that the detector also has a minimal amount of noise, Finally, for a detector with quantum-ideal noise prop- i.e. Eq. (I21) also holds. This latter fact implies that the erties, the magnitude of the constant α can be found from proportionality condition of Eq. (I13) also must hold. In Eq. (4.18). The phase of α can also be determined from: this situation, the detector must have a power gain of unity, and is thus a transducer. There are two possibil- −Im α ~|χ˜IF |/2 ˜ ¯ ¯ = cos δ0 (I17) ities to consider here. First, SFF and SII could both |α| p ¯ ¯ 2 SII SFF be non-zero, but perfectly correlated: |S¯IF | = S¯FF S¯II . In this case, the proportionality constant α must be real For zero frequency or for a large detector effective tem- (c.f. Eq. (I17)). Using this fact along with Eqs. (I16) and perature, this simplifies to: (4.8b), one immediately finds that SFF [ω] = SFF [−ω]. This implies the back-action damping γ associated with −Im α ~χ˜IF /2 = p (I18) the detector input vanishes (c.f. Eq. (2.12)). It thus fol- |α| S¯ S¯ II FF lows immediately from Eq. (5.52) and Eq. (5.53) that the Note importantly that to have a non-vanishing gain power gain GP,rev (defined in a way that accounts for the and power gain, one needs Im α 6= 0. This in turn places reverse gain) is exactly one. The detector is thus simply a ∗ a very powerful constraint on a quantum-ideal detectors: transducer. The other possibility here is that χIF = χFI ¯ ¯ all transitions contributing to the noise must be to final and one or both of SII , SFF are equal zero. Note that if states |fi which are completely unoccupied. To see this, the symmetrized noise vanishes, then so must the asym- imagine a transition taking an initial state |ii = |ai to metric part of the noise. Thus, it follows that either the a final state |fi = |bi makes a contribution to the noise. damping induced by the detector input, γ, or that in- For a quantum-ideal detector, Eq. (I13) will be satisfied: duced by the output, γout (c.f. Eq. (5.48)) (or both) must be zero. Eqs. (5.52) and (5.53) then again yield a power ±iδ hb|Iˆ|ai = e αhb|Fˆ|ai (I19) gain GP,rev = 1. We thus have shown that any detector 89

∗ ¯ ¯ ¯ 2 which has χFI = χIF and satisfies SFF SII = |SIF | must We now consider the relevant limit of a large detector necessarily be a transducer, with a power gain precisely power gain GP . GP is determined by Eq. (5.56); the only equal to one. way this can become large is if kBTeff /(~ω) → ∞ while A second way to make the RHS of Eq. (4.11) vanish is Im α does not tend to zero. We will thus take the large to have S¯IF /χ˜IF be purely imaginary and larger in mag- Teff limit in the above equations while keeping both α and nitude than ~/2. Suppose this is the case, and that the the phase of λ fixed. Note that this means the parameter ˜ detector also satisfies the minimal noise requirement of δ must evolve; it tends to zero in the large Teff limit. In Eq. (I21). Without loss of generality, we takeχ ˜IF to be this limit, we thus find for λ and S¯IF : real, implying that S¯IF is purely imaginary. Eqs. (I10) iδ " " 2## and (I11) then imply that the phase factor e appear- ~λ[ω] −iδ ~ω = −2e kBTeff γ[ω] (Im α) 1 + O ing in the proportionality relation of Eq. (I16) is purely 2 kBTeff imaginary, while the constant α is purely real. Using this (I28) proportionality relation in Eq. (4.8b) forχ ˜ yields: IF    −iδ ~ω α S¯IF [ω] = 2e kBTeff γ[ω] (Re α) 1 + O χ˜IF = (SFF [ω] − SFF [−ω]) kBTeff ~ (I29) = 2α [−Im χFF [ω]] (I22)

Using this result and the relation between χ and χ Thus, in the large power-gain limit (i.e. large Teff FF II ¯ in Eq. (I15), we can write the power gain in the absence limit), the gain λ and the noise cross-correlator SIF have ¯ of reverse gain, GP (c.f. Eq. (5.52)), as the same phase: SIF /λ is purely real. 2 GP = 1/ |1 − χFI /χIF | (I23) 4. Derivation of non-equilibrium Langevin equation If the reverse gain vanishes (i.e. χFI = 0), we immedi- ately find that GP = 1: the detector has a power gain of In this appendix, we prove that an oscillator weakly one, and is thus simply a transducer. If the reverse gain coupled to an arbitrary out-of-equilibrium detector is de- is non-zero, we must take the expression for GP above scribed by the Langevin equation given in Eq. (5.41), an and plug it into Eq. (5.53) for the power gain with re- equation which associates an effective temperature and verse gain, GP,rev. Some algebra again yields that the damping kernel to the detector. The approach taken here full power gain is at most unity. We again have the con- is directly related to the pioneering work of Schwinger clusion that the detector does not amplify. (Schwinger, 1961). We start by defining the oscillator matrix Keldysh green function: 3. Simplifications for a quantum-limited detector ! GK (t) GR(t) In this appendix, we derive the additional constraints Gˇ(t) = (I30) GA(t) 0 on the property of a detector that arise when it satisfies the quantum noise constraint of Eq. (4.17). We focus on where GR(t − t0) = −iθ(t − t0)h[ˆx(t), xˆ(t0)]i, GA(t − t0) = the ideal case where the reverse gain χ vanishes. FI iθ(t0 −t)h[ˆx(t), xˆ(t0)]i, and GK (t−t0) = −ih{xˆ(t), xˆ(t0)}i. To start, we substitute Eq. (I16) into Eqs. (4.8b)- At zero coupling to the detector (A = 0), the oscillator (4.8a); writing S [ω] in terms of the detector effective FF is only coupled to the equilibrium bath, and thus Gˇ has temperature T (cf. Eq. (2.8)) yields: 0 eff the standard equilibrium form:

~λ[ω] −iδ = −e [−Im χ [ω]] (I24)  ω  ! ~ FF −2Im g0[ω] coth ~ g0[ω] 2 ~ 2kBTbath Gˇ0[ω] =   ω   m g [ω]∗ 0 (Im α) coth ~ + i (Re α) 0 2kBTeff (I31) ¯ −iδ where: SIF [ω] = e ~ [−Im χFF [ω]] (I25)     1 ~ω g [ω] = (I32) (Re α) coth − i (Im α) 0 2 2 ω − Ω + iωγ0/m 2kBTeff To proceed, let us write: and where γ0 is the intrinsic damping coefficient, and Tbath is the bath temperature. λ ˜ e−iδ = e−iδ (I26) We next treat the effects of the coupling to the detector |λ| in perturbation theory. Letting Σˇ denote the correspond- ing self-energy, the Dyson equation for Gˇ has the form: The condition that |λ| is real yields the condition: !   A Re α ~ω  ˇ −1  ˇ −1 0 Σ [ω] tan δ˜ = tanh (I27) G[ω] = G0[ω] − R K (I33) Imα 2kBTeff Σ [ω]Σ [ω] 90

To lowest order in A, Σ[ˇ ω] is given by: voltage amplifier, this means that we should have a short circuit at the amplifier input (i.e. no input voltage, ˇ 2 ˇ Σ[ω] = A D[ω] (I34) Va = 0), and we should have open circuit at the out- 2 Z put (i.e. I = 0, no load at the output drawing current). A iωt b ≡ dt e (I35) These two conditions define the uncoupled amplifier. Us- ~ ! ing the definitions of the voltage and current operators 0 iθ(−t)h[Fˆ(t), Fˆ(0)]i (cf. Eqs. (6.2a) and (6.2b)), they take the form: −iθ(t)h[Fˆ(t), Fˆ(0)]i −ih{Fˆ(t), Fˆ(0)}i aˆin[ω] = −aˆout[ω] (I39a) ˆ ˆ Using this lowest-order self energy, Eq. (I33) yields: bin[ω] = bout[ω] (I39b)

R ~ The scattering matrix equation Eq. (6.3) then allows us G [ω] = 2 2 2 R m(ω − Ω ) − A Re D [ω] + iω(γ0 + γ[ω]) to solve fora ˆin anda ˆout in terms of the added noise ˆ ˆ (I36) operators Fa and Fb. A  R ∗ G [ω] = G [ω] (I37) 1 − s22 s12 aˆin[ω] = − Fˆa[ω] − Fˆb[ω] (I40a) GK [ω] = −2iIm GR[ω] × D D s21 1 + s11  ω   ω  ˆb [ω] = − Fˆ [ω] + Fˆ [ω] (I40b) γ0 coth ~ + γ[ω] coth ~ in a b 2kBTbath 2kBTeff D D γ + γ[ω] 0 where D is given in Eq. (6.8), and we have omitted writ- (I38) ing the frequency dependence of the scattering matrix. Further, as we have already remarked, the commutators where γ[ω] is given by Eq. (2.12), and Teff [ω] is defined by of the added noise operators is completely determined by Eq. (2.8). The main effect of the real part of the retarded the scattering matrix and the constraint that output op- R Fˆ Green function D [ω] in Eq. (I36) is to renormalize the erators have canonical commutation relations. The non- oscillator frequency Ω and mass m; we simply incorporate vanishing commutators are thus given by: these shifts into the definition of Ω and m in what follows. h i If Teff [ω] is frequency independent, then Eqs. (I36) † 0 0 2 2 Fˆ [ω], Fˆ (ω ) = 2πδ(ω − ω ) 1 − |s11| − |s12| - (I38) for Gˇ corresponds exactly to an oscillator cou- a a pled to two equilibrium baths with damping kernels γ0 (I41a) and γ[ω]. The correspondence to the Langevin equa- h † i Fˆ [ω], Fˆ (ω0) = 2πδ(ω − ω0) 1 − |s |2 − |s |2 tion Eq. (5.41) is then immediate. In the more general b b 21 22 case where Teff [ω] has a frequency dependence, the cor- (I41b) relators GR[ω] and GK [ω] are in exact correspondence h i Fˆ [ω], Fˆ†(ω0) = −2πδ(ω − ω0)(s s∗ + s s∗ ) to what is found from the Langevin equation Eq. (5.41): a b 11 21 12 22 GK [ω] corresponds to symmetrized noise calculated from (I41c) Eq. (5.41), while GR[ω] corresponds to the response co- efficient of the oscillator calculated from Eq. (5.41). This The above equations, used in conjunction with again proves the validity of using the Langevin equation Eqs. (6.2a) and (6.2b), provide us with all all the infor- Eq. (5.41) to calculate the oscillator noise in the presence mation needed to calculate commutators between current of the detector to lowest order in A. and voltage operators. It is these commutators which en- ter into the linear-response Kubo formulas. As we will see, our calculation will crucially rely on the fact that 5. Linear-response formulas for a two-port bosonic amplifier the scattering description obeys causality: disturbances at the input of our system must take some time before In this appendix, we use the standard linear-response they propagate to the output. Causality manifests itself Kubo formulas of Sec. V.F to derive expressions for in the energy dependence of the scattering matrix: as a 0 the voltage gain λV , reverse current gain λI , input function of energy, it is an analytic function in the upper impedance Zin and output impedance Zout of a two-port half complex plane. bosonic voltage amplifier (cf. Sec. VI). We recover the same expressions for these quantities obtained in Sec. VI from the scattering approach. We stress throughout this a. Input and output impedances appendix the important role played by the causal struc- ture of the scattering matrix describing the amplifier. Eq. (5.84) is the linear response Kubo formula for In applying the general linear response formulas, we the input impedance of a voltage amplifier. Recall must bear in mind that these expressions should be ap- that the input operator Qˆ for a voltage amplifier is re- ˆ ˆ plied to the uncoupled detector, i.e. nothing attached to lated to the input current operator Ia via −dQ/dt = the detector input or output. In our two-port bosonic Iˆa (cf. Eq. (5.83)). The Kubo formula for the input 91 impedance may thus be re-written in the more familiar b. Voltage gain and reverse current gain form:  Z ∞  Within linear response theory, the voltage gain of the i i iωt Yin,Kubo[ω] ≡ − dth[Iˆa(t), Iˆa(0)]ie (I42) amplifier (λV ) is determined by the commutator between ω ~ 0 the “input operator” Qˆ and Vˆb (cf. Eq. (4.3); recall that ˆ ˆ ˆ where Yin[ω] = 1/Zin[ω], Q is defined by dQ/dt = −Ia. Similarly, the reverse 0 Using the defining equation for Iˆa (Eq. (6.1b)) and current gain (λI ) is determined by the commutator be- Eq. (I39a) (which describes an uncoupled amplifier), we tween Iˆa and Φ,ˆ where Φˆ is defined via dΦˆ/dt = −Vˆb obtain: (cf. Eq. (4.6)). Similar to the calculation of the input impedance, to properly evaluate the Kubo formulas for Yin,Kubo[ω] = (I43) the gains, we must make use of the causal structure of Z ∞ Z ∞ 0 0 2 iωt dω ω 0  −iω0t iω0t the scattering matrix describing our amplifier. dte Λaa(ω ) e − e Za 0 0 2π ω Using the defining equations of the current and volt- age operators (cf. Eqs. (6.1a) and (6.1b)), as well as where we have defined the real function Λaa[ω] for ω > 0 Eqs. (I39a) and (I39b) which describe the uncoupled am- via: plifier, the Kubo formulas for the voltage gain and reverse h † 0 i 0 current gain become: aˆin[ω], aˆin(ω ) = 2πδ(ω − ω )Λaa[ω] (I44) r Zb It will be convenient to also define Λ [ω] for ω < 0 λV,Kubo[ω] = 4 (I48) aa Z via Λ [ω] = Λ [−ω]. Eq. (I43) may then be written as: a aa aa Z ∞ Z ∞ 0  iωt dω 0 −iω0t ∞ ∞ 0 0 × dt e Re Λba(ω )e 2 Z Z dω ω 0 0 i(ω−ω )t 0 0 2π Yin,Kubo[ω] = dt Λaa(ω )e Za 0 −∞ 2π ω r 0 Zb Λ [ω] i Z ∞ ω0Λ (ω0)/Z λI,Kubo[ω] = −4 (I49) aa 0 aa a Za = + P dω 0 Za πω ω − ω Z ∞ Z ∞ 0  −∞ iωt dω 0 iω0t (I45) × dt e Re Λba(ω )e 0 0 2π

Next, by making use of Eq. (I40a) and Eqs. (I41) for where we define the complex function Λba[ω] for ω > 0 the commutators of the added noise operators, we can via: explicitly evaluate the commutator in Eq. (I44) to cal- hˆ † 0 i 0 culate Λaa[ω]. Comparing the result against the result bin[ω], aˆin(ω ) ≡ (2πδ(ω − ω )) Λba[ω] (I50) Eq. (6.7d) of the scattering calculation, we find: We can explicitly evaluate Λba[ω] by using Eqs. (I40a)- Λaa[ω] (I41) to evaluate the commutator above. Comparing the = Re Yin,scatt[ω] (I46) Za result against the scattering approach expressions for the gain and reverse gain (cf. Eqs. (6.7a) and (6.7b)), one where Y [ω] is the input admittance of the ampli- in,scatt finds: fier obtained from the scattering approach. Returning to ∗ Eq. (I45), we may now use the fact that Yin,scatt[ω] is  0  Λba[ω] = λV,scatt[ω] − λI,scatt[ω] (I51) an analytic function in the upper half plane to simplify the second term on the RHS, as this term is simply a Note crucially that the two terms above have different Kramers-Kronig integral: analytic properties: the first is analytic in the upper half plane, while the second is analytic in the lower half plane. Z ∞ 0 0 1 0 ω Λaa(ω )/Za This follows directly from the fact that the scattering P dω 0 πω −∞ ω − ω matrix is causal. 1 Z ∞ ω0Re Y (ω0) At this stage, we can proceed much as we did in the = P dω0 in,scatt πω ω − ω0 calculation of the input impedance. Defining Λba[ω] for −∞ ∗ ω < 0 via Λba[−ω] = Λba[ω], we can re-write Eqs. (I48) = Im Yin,scatt[ω] (I47) and (I49) in terms of principle part integrals. It thus follows from Eq. (I45) that input impedance cal-  Z ∞ 0  Zb i 0 Λba(ω ) culated from the Kubo formula is equal to what we found λV,Kubo[ω] = Λba[ω] + P dω 0 previously using the scattering approach. Za π −∞ ω − ω  Z ∞ 0  The calculation for the output impedance proceeds in 0 Zb i 0 Λba(ω ) λI,Kubo[ω] = − Λba[ω] + P dω 0 the same fashion, starting from the Kubo formula given Za π −∞ ω + ω in Eq. (5.85). As the steps are completely analogous to (I52) the above calculation, we do not present it here. One again recovers Eq. (6.7c), as found previously within the Using the analytic properties of the two terms in Eq. (I51) scattering approach. for Λba[ω], we can evaluate the principal part integrals 92 above as Kramers-Kronig relations. One then finds that If we now do the time integrals and then take the limit the Kubo formula expressions for the voltage and cur- τ → 0+, we recover the results of the scattering approach rent gain coincide precisely with those obtained from the (cf. Eqs. (6.7a) and (6.7b)); in particular, λI = 0. Note scattering approach. that if we had set τ = 0 from the outset of the calculation, While the above is completely general, it is useful to we would have found that both λV and λI are non-zero! go through a simpler, more specific case where the role of causality is more transparent. Imagine that all the energy dependence in the scattering in our amplifier arises from 6. Details for the two-port bosonic voltage amplifier with the fact that there are small transmission line “stubs” of feedback length a attached to both the input and output of the amplifier (these stubs are matched to the input and out- In this appendix, we provide more details on the calcu- put lines). Because of these stubs, a wavepacket incident lations for the bosonic-amplifier-plus-mirrors system dis- on the amplifier will take a time τ = 2a/v to be either cussed in Sec. H. Given that the scattering matrix for reflected or transmitted, where v is the characteristic ve- each of the three mirrors is given by Eq. (H2), and that locity of the transmission line. This situation is described we know the reduced scattering matrix for the mirror-free by a scattering matrix which has the form: system (cf. Eq. (6.10)), we can find the reduced scattering matrix and noise operators for the system with mirrors. s[ω] = e2iωa/v · s¯ (I53) One finds that the reduced scattering matrix s is now given by: wheres ¯ is frequency-independent and real. To further simplify things, let us assume thats ¯ =s ¯ =s ¯ = 0. 1 11 22 12 s = × (I57) Eqs. (I51) then simplifies to M √ ! iωτ sin θz + G sin θx sin θy − cos θx cos θz sin θy Λba[ω] = = s21[ω] =s ¯21e (I54) √ √ G cos θx cos θz sin θx + G sin θy sin θz where the propagation time τ = 2a/v We then have: where the denominator M describes multiple reflection r Z Z ∞ b iω0t processes: λV [ω0] = 2 s¯21 dt e δ(t − τ) (I55) Za 0 √ r Z Z ∞ M = 1 + G sin θx sin θz sin θy (I58) b iω0t λI [ω0] = −2 s¯21 dt e δ(t + τ) (I56) Further, the noise operators are given by: Za 0

! √ ! ! Fa 1 cos θy cos θz G − 1 cos θz sin θx sin θy uin = √ √ † (I59) Fb M − G cos θx cos θy sin θz G − 1 cos θx vin

The next step is to convert the above into the op-amp Similarly, the output impedance only depends on the representation, and find the gains and impedances of the amount of reflection in the output line and the in the amplifier, along with the voltage and current noises. The cold-load line: voltage gain is given by: √ 1 + G sin θy 1 + sin θx r √ Zout = Zb √ · (I63) ZB 2 G 1 + sin θx 1 − sin θz 1 − G sin θy 1 − sin θx λV = √ · (I60) √ ZA 1 − G sin θy cos θx cos θz Note that as sin θy tends to −1/ G , both the input while the reverse gain is related to the voltage gain by admittance and output impedance tend to zero. the simple relation: Given that we now know the op-amp parameters of our amplifier, we can use Eq. (6.5) to calculate the amplifier’s 0 sin θy power gain GP . Amazingly, we find that the power gain λ = − √ λV (I61) I G is completely independent of the mirrors in the input and output lines: The input impedance is determined by the amount of reflection in the input line and in the line going to the G GP = 2 (I64) cold load: 1 + G sin θy √ √ 1 − G sin θy 1 + sin θz Note that at the special value sin θy = −1 G (which Zin = Za √ · (I62) 1 + G sin θy 1 − sin θz allows one to reach the quantum limit), the power gain 93 is reduced by a factor of two compared to the reflection Aleiner, I. L., N. S. Wingreen, and Y. Meir, 1997, Phys. Rev. free case (i.e. θy = 0). Lett. 79, 3740. Turning to the noise spectral densities, we assume the Allahverdyan, A. E., R. Balian, and T. M. Nieuwenhuizen, 2001, Phys. Rev. A 64, 032108. optimal situation where both the auxiliary modesu ˆin and † ˜ˆ Amin, M. H. S., and D. V. Averin, 2008, Phys. Rev. Lett. vˆin are in the vacuum state. We then find that both I 100, 197001. and V˜ˆ are independent of the amount of reflection in the Arcizet, O., P.-F. Cohadon, T. Briant, M. Pinard, and A. Hedimann, 2006, Nature 444, 71. output line (e.g. θx): Armen, M. A., J. K. Au, J. K. Stockton, A. C. Doherty, and   H. Mabuchi, 2002, Phys. Rev. Lett. 89, 133602. 2~ω 1 − sin θz S¯II = × Averin, D., 2003, in Quantum Noise in Mesoscopic Systems, Za 1 + sin θz edited by Y. Nazarov (Kluwer, Amsterdam), pp. 205–228.   Averin, D. V., 2000a, cond-mat/0010052v1 . G sin2 θ + cos(2θ ) Averin, D. V., 2000b, Fortschr. Phys. 48, 1055.  y y  (I65a)  √ 2  Bergeal, N., R. Vijay, V. E. Manucharyan, I. Siddiqi, R. J. G sin θy − 1 Schoelkopf, S. M. Girvin, and M. H. Devoret, 2008, eprint   0805.3452. ¯ 1 + sin θz SVV = ~ωZa × Bern´ad,J. Z., A. Bodor, T. Geszti, and L. Di´osi,2008, Phys. 1 − sin θz Rev. B 77, 073311. 3 + cos(2θ ) 1  Billangeon, P. M., F. Pierre, H. Bouchiat, and R. Deblock, y − (I65b) 4 2G 2006, Phys. Rev. Lett. 96, 136804. √ Blais, A., R.-S. Huang, A. Wallraff, S. M. Girvin, and R. J. 2 G(1 − 1/G) sin θy + cos θy Schoelkopf, 2004, Phys. Rev. A 69, 062320. S¯VI = √ (I65c) 1 − G sin θ Blencowe, M. P., and E. Buks, 2007, Phys. Rev. B 76, 014511. y Bocko, M. F., and R. Onofrio, 1996, Rev. Mod. Phys. 68, 755. As could be expected, introducing reflections in the input Bohm, D., 1989, Quantum Theory (Dover, New York). Boylestad, R., and L. Nashelsky, 2006, Electronic Devices and line (i.e. θz 6= 0) has the opposite effect on S¯II versus ¯ Circuit Theory (Prentice Hall, New Jersey), 9 edition. SVV : if one is enhanced, the other is suppressed. Braginsky, V. B., and F. Y. Khalili, 1992, Quantum Measure- It thus follows that the product of noise spectral ment (Cambridge University Press, Cambridge). densities appearing in the quantum noise constraint of Braginsky, V. B., and F. Y. Khalili, 1996, Rev. Mod. Phys. Eq. (5.88) is given by (taking the large-G limit): 68, 1. Braginsky, V. B., Y. I. Vorontsov, and K. P. Thorne, 1980, ¯ ¯ 2 Science 209, 547. SII SVV 2  1 + G sin θy = 2 − sin θy · (I66) van den Brink, A. M., 2002, Europhys. Lett. 58, 562. ( ω)2  √ 2 ~ 1 − G sin θy Brown, K. R., J. Britton, R. J. Epstein, J. Chiaverini, D. Leibfried, and D. J. Wineland, 2007, Phys. Rev. Lett. Note that somewhat amazingly, this product (and hence 99, 137205. the amplifier noise temperature) is completely indepen- Brun, T. A., 2002, Am. J. Phys. 70, 719. Buks, E., R. Schuster, M. Heiblum, D. Mahalu, and V. Uman- dent of the mirrors in the input and output arms (i.e. θz ¯ ¯ sky, 1998, Nature (London) 391, 871. and θx). This is a result of both SVV and SII having Burkhard, G., R. H. Koch, and D. P. DiVincenzo, 2004, Phys. no dependence on the output mirror (θx), and their hav- Rev. B 69, 064503. ing opposite dependencies on the input mirror (θz). Also Caldeira, A., and A. Leggett, 1983, Ann. Phys. (NY) 149, note that Eq. (I66) does indeed reduce to the result of 374. the last subsection: if θy = 0 (i.e. no reflections in the Callen, H. B., and T. A. Welton, 1951, Phys. Rev. 83, 34. line going to the cold load), the product S¯II S¯VV is equal Castellanos-Beltran, M. A., K. D. Irwin, G. C. Hilton, L. R. 2 Vale, and K. W. Lehnert, 2008, Nature Phys. 4, 929. to precisely twice√ the quantum limit value of (~ω) . For sin(θ ) = −1/ G, the RHS above reduces to one, imply- Caves, C. M., 1982, Phys. Rev. D 26, 1817. y Caves, C. M., and B. L. Schumaker, 1985, Phys. Rev. A 31, ing that we reach the quantum limit for this tuning of 3068. the mirror in the cold-load arm. Caves, C. M., K. S. Thorne, R. W. P. Drever, V. D. Sandberg, and M. Zimmermann, 1980, Rev. Mod. Phys. 52, 341. Cleland, A. N., J. S. Aldridge, D. C. Driscoll, and A. C. Gos- References sard, 2002, Appl. Phys. Lett. 81, 1699. Clerk, A. A., 2004, Phys. Rev. B 70, 245306. Aassime, A., G. Johansson, G. Wendin, R. J. Schoelkopf, and Clerk, A. A., 2006, Phys. Rev. Lett. 96, 056801. P. Delsing, 2001, Phys. Rev. Lett. 86, 3376. Clerk, A. A., M. H. Devoret, S. M. Girvin, and F. Mar- Abramovici, A., W. Althouse, R. Drever, Y. Gursel, S. Kawa- quardt, 2009, see supplementary material at http:// mura, F. Raab, D. Shoemaker, L. Sievers, R. Spero, link.aps.org/supplemental/10.1103/RevModPhys.82.1155. K. Thorne, R. Vogt, R. Weiss, et al., 1992, Science 256, Clerk, A. A., and S. M. Girvin, 2004, Phys. Rev. B 70, 121303. 325. Clerk, A. A., S. M. Girvin, A. K. Nguyen, and A. D. Stone, Aguado, R., and L. P. Kouwenhoven, 2000, Phys. Rev. Lett. 2002, Phys. Rev. Lett. 89, 176804. 84, 1986. Clerk, A. A., S. M. Girvin, and A. D. Stone, 2003, Phys. Rev. 94

B 67, 165324. Goan, H. S., G. J. Milburn, H. M. Wiseman, and B. H. Sun, Clerk, A. A., F. Marquardt, and K. Jacobs, 2008, New J. 2001, Phys. Rev. B 63, 125326. Phys. 10, 095010. Gordon, J. P., W. H. Louisell, and L. R. Walker, 1963, Phys. Clerk, A. A., and A. D. Stone, 2004, Phys. Rev. B 69, 245303. Rev. 129, 481. Cohen-Tannoudji, C., J. Dupont-Roc, and G. Grynberg, 1989, Gottfried, K., 1966, Quantum Mechanics, Volume I: Funda- Photons and Atoms - Introduction to Quantum Electrody- mentals (W. A. Benjamin, Inc.). namics (Wiley, New York). Grassia, F., 1998, Fluctuations quantiques et thermiques dans Corbitt, T., Y. Chen, E. Innerhofer, H. Muller-Ebhardt, les transducteurs electromecaniques, Ph.D. thesis, Univer- D. Ottaway, H. Rehbein, D. Sigg, S. Whitcomb, C. Wipf, site Pierre et Marie Curie. and N. Mavalvala, 2007, Phys. Rev. Lett 98, 150802. Groblacher, S., J. B. Hertzberg, M. R. Vanner, G. D. Cole, Courty, J. M., F. Grassia, and S. Reynaud, 1999, Europhys. S. Gigan, K. C. Schwab, and M. Aspelmeyer, 2009, Nature Lett. 46, 31. Phys. 5, 485. Cover, T., and J. Thomas, 1991, Elements of Information Gurvitz, S. A., 1997, Phys. Rev. B 56, 15215. Theory (Wiley, New York). Haroche, S., and J.-M. Raimond, 2006, Exploring the Quan- Lupa¸scu,A., C. J. M. Verwijs, R. N. Schouten, C. J. P. M. tum Atoms, Cavities, and Photons (Oxford University Harmans, and J. E. Mooij, 2004, Phys. Rev. Lett. 93, Press, Oxford). 177006. Harris, J. G. E., B. M. Zwickl, and A. M. Jayich, 2007, Review Danilov, V. V., K. K. Likharev, and A. B. Zorin, 1983, IEEE of Scientific Instruments 78, 013107. Trans. Magn. 19, 572. Haus, H. A., 2000, Electromagnetic Noise and Quantum Op- Deblock, R., E. Onac, L. Gurevich, and L. P. Kouwenhoven, tical Measurements (Springer, New York). 2003, Science 301, 203. Haus, H. A., and J. A. Mullen, 1962, Phys. Rev. 128, 2407. Devoret, M., 1997, in Quantum Fluctuations (Les Houches Heffner, H., 1962, Proc. IRE 50, 1604. Session LXIII) (Elsevier, Amsterdam), pp. 351–86. Hertzberg, J. B., T. Rocheleau, T. Ndukum, M. Savva, A. A. Devoret, M. H., and R. J. Schoelkopf, 2000, Nature (London) Clerk, and K. C. Schwab, 2010, Nature Phys. 6, 213. 406, 1039. H¨ohberger-Metzger, C., and K. Karrai, 2004, Nature 432, Doherty, A. C., S. Habib, K. Jacobs, H. Mabuchi, and S. M. 1002. Tan, 2000, Phys. Rev. A 62, 012105. Il’ichev, E., N. Oukhanski, A. Izmalkov, T. Wagner, M. Gra- Doiron, C. B., B. Trauzettel, and C. Bruder, 2007, Phys. Rev. jcar, H.-G. Meyer, A. Y. Smirnov, A. Maassen van den B. 76, 195312. Brink, M. H. S. Amin, and A. M. Zagoskin, 2003, Phys. Duty, T., G. Johansson, K. Bladh, D. Gunnarsson, C. Wilson, Rev. Lett. 91, 097906. and P. Delsing, 2005, Phys. Rev. Lett. 95, 206807. Izmalkov, A., M. Grajcar, E. Il’ichev, T. Wagner, H.-G. Dykman, M. I., 1978, Sov. Phys. Solid State 20, 1306. Meyer, A. Y. Smirnov, M. H. S. Amin, A. M. van den Brink, Etaki, S., M. Poot, I. Mahboob, K. Onomitsu, H. Yamaguchi, and A. M. Zagoskin, 2004, Phys. Rev. Lett. 93, 037003. and H. S. J. V. der Zant, 2008, Nature Phys. 4, 785. Jacobs, K., and D. A. Steck, 2006, Contemporary Physics 47, Falferi, P., M. Bonaldi, M. Cerdonio, R. Mezzena, G. A. 279. Prodi, A. Vinante, and S. Vitale, 2008, Appl. Phys. Lett. Jayich, A. M., J. C. Sankey, B. M. Zwickl, C. Yang, J. D. 93, 172506. Thompson, S. M. Girvin, A. A. Clerk, F. Marquardt, and Flowers-Jacobs, N. E., D. R. Schmidt, and K. W. Lehnert, J. G. E. Harris, 2008, New Journal of Physics 10(9), 095008 2007, Phys. Rev. Lett 98, 096804. (28pp). Gambetta, J., A. Blais, D. I. Schuster, A. Wallraff, and R. J. Johansson, G., P. Delsing, K. Bladh, D. Gunnarsson, T. Duty, Schoelkopf, 2006, Phys. Rev. A 74, 042318. A. Kack, G. Wendin, and A. Aassime, 2003, in Quan- Gambetta, J., W. A. Braff, A. Wallraff, S. M. Girvin, and tum Noise in Mesoscopic Physics, edited by Y. V. Nazarov R. J. Schoelkopf, 2007, Phys. Rev. A 76, 012325. (Kluwer, Dordrecht), pp. 337–356. Gardiner, C. W., A. S. Parkins, and P. Zoller, 1992, Phys. Johansson, G., A. K¨ack, and G. Wendin, 2002, Phys. Rev. Rev. A 46, 4363. Lett. 88, 046802. Gardiner, C. W., and P. Zoller, 2000, Quantum Noise Johansson, G., L. Tornberg, and C. M. Wilson, 2006, Phys. (Springer, Berlin). Rev. B (R) 74, 100504. Gavish, U., Y. Levinson, and Y. Imry, 2000, Phys. Rev. B Jordan, A. N., and M. B¨uttiker, 2005a, Phys. Rev. Lett. 95, (R) 62, 10637. 220401. Gavish, U., B. Yurke, and Y. Imry, 2004, Phys. Rev. Lett. Jordan, A. N., and M. B¨uttiker, 2005b, Phys. Rev. B 71, 93, 250601. 125333. Gavish, U., B. Yurke, and Y. Imry, 2006, Phys. Rev. Lett. Jordan, A. N., and A. N. Korotkov, 2006, Phys. Rev. B 74, 96, 133602. 085307. Gea-Banacloche, J., N. Lu, L. M. Pedrotti, S. Prasad, M. O. Kleckner, D., and D. Bouwmeester, 2006, Nature 444, 75. Scully, and K. Wodkiewicz, 1990a, Phys. Rev. A 41, 369. Knight, P. L., and V. Buzek, 2004, in Quantum Squeezing, Gea-Banacloche, J., N. Lu, L. M. Pedrotti, S. Prasad, M. O. edited by P. D. Drummond and Z. Ficek (Springer, Berlin), Scully, and K. Wodkiewicz, 1990b, Phys. Rev. A 41, 381. pp. 3–32. Gerry, C. S., 1985, Phys. Rev. A 31, 2721. Knobel, R. G., and A. N. Cleland, 2003, Nature (London) Gigan, S., H. B¨ohm, M. Paternostro, F. Blaser, J. B. 424, 291. Hertzberg, K. C. Schwab, D. Bauerle, M. Aspelmeyer, and Koch, R. H., D. J. van Harlingen, and J. Clarke, 1980, Appl. A. Zeilinger, 2006, Nature 444, 67. Phys. Lett. 38, 380. Glauber, R. J., 2006, Rev. Mod. Phys. 78, 1267. Korotkov, A. N., 1999, Phys. Rev. B 60, 5737. Goan, H.-S., and G. J. Milburn, 2001, Phys. Rev. B 64, Korotkov, A. N., 2001a, Phys. Rev. B 63, 085312. 235307. Korotkov, A. N., 2001b, Phys. Rev. B 63, 115403. 95

Korotkov, A. N., and D. V. Averin, 2001, Phys. Rev. B 64, Regal, C. A., J. D. Teufel, and K. W. Lehnert, 2008, Nature 165310. Phys. 4, 555. Korotkov, A. N., and M. A. Paalanen, 1999, Appl. Phys. Lett. Rugar, D., R. Budakian, H. J. Mamin, and B. W. Chui, 2004, 74, 4052. Nature (London) 430, 329. LaHaye, M. D., O. Buu, B. Camarota, and K. C. Schwab, Ruskov, R., and A. N. Korotkov, 2002, Phys. Rev. B 66, 2004, Science 304, 74. 041401. Lang, R., M. O. Scully, and J. Willis E. Lamb, 1973, Phys. Ruskov, R., K. Schwab, and A. N. Korotkov, 2005, Phys. Rev. Rev. A 7, 1788. B 71, 235407. Lesovik, G. B., and R. Loosen, 1997, JETP Lett. 65, 295. Santamore, D. H., A. C. Doherty, and M. C. Cross, 2004a, Levinson, Y., 1997, Europhys. Lett. 39, 299. Phys. Rev. B 70, 144301. Louisell, W. H., A. Yariv, and A. E. Siegman, 1961, Phys. Santamore, D. H., H.-S. Goan, G. J. Milburn, and M. L. Rev. 124, 1646. Roukes, 2004b, Phys. Rev. A 70, 052105. Lupa¸scu,A., E. F. C. Driessen, L. Roschier, C. J. P. M. Har- Sauter, T., W. Neuhauser, R. Blatt, and P. E. Toschek, 1986, mans, and J. E. Mooij, 2006, Phys. Rev. Lett. 96, 127003. Phys. Rev. Lett. 57, 1696. Lupa¸scu,A., C. J. P. M. Harmans, and J. E. Mooij, 2005, Schliesser, A., O. Arcizet, R. Riviere, G. Anetsberger, and Phys. Rev. B. 71, 184506. T. J. Kippenberg, 2009, Nature Phys. 5, 509. Lupa¸scu,A., S. Saito, T. Picot, P. C. de Groot, C. J. P. M. Schliesser, A., P. Del’Haye, N. Nooshi, K. J. Vahala, and T. J. Harmans, and J. E. Mooij, 2007, Nature Phys. 3, 119. Kippenberg, 2006, Phys. Rev. Lett. 97, 243905. Makhlin, Y., G. Sch¨on,and A. Shnirman, 2000, Phys. Rev. Schliesser, A., R. Rivi`ere,G. Anetsberger, O. Arcizet, and Lett. 85, 4578. T. J. Kippenberg, 2008, Nature Phys. 4, 415. Makhlin, Y., G. Schon, and A. Shnirman, 2001, Rev. Mod. Schoelkopf, R. J., P. J. Burke, A. A. Kozhevnikov, D. E. Phys. 73, 357. Prober, and M. J. Rooks, 1997, Phys. Rev. Lett. 78, 3370. Mallat, S., 1999, A Wavelet Tour of Signal Processing (Aca- Schoelkopf, R. J., A. A. Clerk, S. M. Girvin, K. W. Lehnert, demic, San Diego). and M. H. Devoret, 2003, in Quantum Noise in Mesoscopic Marquardt, F., J. P. Chen, A. A. Clerk, and S. M. Girvin, Physics, edited by Y. V. Nazarov (Kluwer, Dorderecht), 2007, Phys. Rev. Lett. 99, 093902. pp. 175–203. Marquardt, F., and S. M. Girvin, 2009, Physics 2, 40. Schoelkopf, R. J., P. Wahlgren, A. A. Kozhevnikov, P. Dels- Marquardt, F., J. G. E. Harris, and S. M. Girvin, 2006, Phys. ing, and D. E. Prober, 1998, Science 280, 1238. Rev. Lett. 96, 103901. Schuster, D. I., A. Wallraff, A. Blais, L. Frunzio, R.-S. Huang, Meystre, P., E. M. Wright, J. D. McCullen, and E. Vignes, J. Majer, S. M. Girvin, and R. J. Schoelkopf, 2005, Phys. 1985, J. Opt. Soc. Am. B 2, 1830. Rev. Lett. 94, 123602. Mollow, B. R., and R. J. Glauber, 1967a, Phys. Rev. 160, Schwinger, J., 1961, J. Math. Phys. 2, 407. 1076. Shnirman, A., and G. Sch¨on,1998, Phys. Rev. B 57, 15400. Mollow, B. R., and R. J. Glauber, 1967b, Phys. Rev. 160, Siddiqi, I., R. Vijay, F. Pierre, C. M. Wilson, M. Metcalfe, 1097. C. Rigetti, L. Frunzio, and M. H. Devoret, 2004, Phys. Mozyrsky, D., I. Martin, and M. B. Hastings, 2004, Phys. Rev. Lett. 93, 207002. Rev. Lett. 92, 018303. Sillanp¨a¨a, M. A., T. Lehtinen, A. Paila, Y. Makhlin, M¨uck, M., J. B. Kycia, and J. Clarke, 2001, Appl. Phys. Lett. L. Roschier, and P. J. Hakonen, 2005, Phys. Rev. Lett. 78, 967. 95, 206806. Nagourney, W., J. Sandberg, and H. Dehmelt, 1986, Phys. Sprinzak, D., E. Buks, M. Heiblum, and H. Shtrikman, 2000, Rev. Lett. 56, 2797. Phys. Rev. Lett. 84, 5820. Naik, A., O. Buu, M. D. LaHaye, A. D. Armour, A. A. Clerk, Stern, A., Y. Aharonov, and Y. Imry, 1990, Phys. Rev. A 41, M. P. Blencowe, and K. C. Schwab, 2006, Nature (London) 3436. 443, 193. Teufel, J. D., T. Donner, M. A. Castellanos-Beltran, J. W. von Neumann, J., 1932, Mathematische Grundlagen der Harlow, and K. W. Lehnert, 2009, eprint 0906.1212, URL Quantenmechanik (Springer, Berlin). http://arxiv.org/abs/0906.1212. Nogues, G., A. Rauschenbeutel, S. Osnaghi, M. Brune, J. M. Teufel, J. D., C. A. Regal, and K. W. Lehnert, 2008, New J. Raimond, and S. Haroche, 1999, Nature (London) 400, Phys. 10, 095002. 239. Thalakulam, M., Z. Ji, and A. J. Rimberg, 2004, Phys. Rev. Onac, E., F. Balestro, B. Trauzettel, C. F. J. Lodewijk, and Lett. 93, 066804. L. P. Kouwenhoven, 2006, Phys. Rev. Lett. 96, 026803. Thompson, J. D., B. M. Zwickl, A. M. Jayich, F. Marquardt, Oxtoby, N. P., J. Gambetta, and H. M. Wiseman, 2008, Phys. S. M. Girvin, and J. G. E. Harris, 2008, Nature (London) Rev. B 77, 125304. 452, 72. Oxtoby, N. P., H. M. Wiseman, and H.-B. Sun, 2006, Phys. Thorne, K. S., R. W. P. Drever, C. M. Caves, M. Zimmer- Rev. B 74, 045328. mann, and V. D. Sandberg, 1978, Phys. Rev. Lett. 40, 667. Pathria, R. K., 1996, Statistical Mechanics (Elsevier, Ams- Tittonen, I., G. Breitenbach, T. Kalkbrenner, T. Muller, terdam), second edition. R. Conradt, S. Schiller, E. Steinsland, N. Blanc, and N. F. Peres, A., 1993, Quantum Theory: Concepts and Methods de Rooij, 1999, Phys. Rev. A 59, 1038. (Kluwer, Dorderecht). Turek, B. A., K. W. Lehnert, A. Clerk, D. Gunnarsson, Pilgram, S., and M. B¨uttiker, 2002, Phys. Rev. Lett. 89, K. Bladh, P. Delsing, and R. J. Schoelkopf, 2005, Phys. 200401. Rev. B 71, 193304. Poggio, M., M. P. Jura, C. L. Degen, M. A. Topinka, H. J. Vinante, A., R. Mezzena, G. A. Prodi, S. Vitale, M. Cerdonio, Mamin, D. Goldhaber-Gordon, and D. Rugar, 2008, Nature P. Falferi, and M. Bonaldi, 2001, Appl. Phys. Lett. 79, Phys. 4, 635. 2597. 96

Wallraff, A., D. I. Schuster, A. Blais, L. Frunzio, R.-S. Huang, Wineland, D. J., C. Monroe, W. M. Itano, D. Leibfried, J. Majer, S. Kumar, S. M. Girvin, and R. J. Schoelkopf, B. King, and D. M. Meekhof, 1998, J. Res. NIST 103, 2004, Nature (London) 431, 162. 259. Walls, D. F., and G. J. Milburn, 1994, Quantum Optics Wiseman, H. M., and G. J. Milburn, 1993, Phys. Rev. Lett. (Springer, Berlin). 70, 548. Weiss, U., 1999, Quantum Dissipative Systems (Second Edi- Wiseman, H. M., and G. J. Milburn, 1994, Phys. Rev. A 49, tion) (World Scientific, Singapore). 1350. Wellstood, F. C., C. Urbina, and J. Clarke, 1994, Phys. Rev. Yurke, B., 1984, Phys. Rev. A 29, 408. B 49, 5942. Yurke, B., L. R. Corruccini, P. G. Kaminsky, L. W. Rupp, Wheeler, J. A., and W. H. Zurek, 1984, Quantum Theory and A. D. Smith, A. H. Silver, R. W. Simon, and E. A. Whit- Measurement (Princeton University Press, Princeton). taker, 1989, Phys. Rev. A 39, 2519. Wigner, E. P., 1955, Phys. Rev. 98, 145. Yurke, B., and J. S. Denker, 1984, Phys. Rev. A 29, 1419. Wilson-Rae, I., N. Nooshi, W. Zwerger, and T. J. Kippenberg, 2007, Phys. Rev. Lett. 99, 093901.