quantum microwave radiometry with a superconducting qubit · 2019-09-30 · a classical microwave...

21
Quantum Microwave Radiometry with a Superconducting Qubit Zhixin Wang, 1, * Mingrui Xu, 2 Xu Han, 2 Wei Fu, 2 Shruti Puri, 1 S. M. Girvin, 1 Hong X. Tang, 2 S. Shankar, 1, and M. H. Devoret 1, 1 Department of Applied Physics and Physics, Yale University, New Haven, CT 06520, USA 2 Department of Electrical Engineering, Yale University, New Haven, CT 06520, USA (Dated: September 30, 2019) The interaction of photons and coherent quantum systems can be employed to detect electro- magnetic radiation with remarkable sensitivity. We introduce a quantum radiometer based on the photon-induced-dephasing process of a superconducting qubit for sensing microwave radiation at the sub-unit-photon level. Using this radiometer, we demonstrated the radiative cooling of a 1-K microwave resonator and measured its mode temperature with an uncertainty 0.01 K. We have thus developed a precise tool for studying the thermodynamics of quantum microwave circuits, which provides new solutions for calibrating hybrid quantum systems and detecting candidate particles for dark matter. Historically, radiometry experiments—measurements of the intensity of radiation—revealed the first invisible band of the electromagnetic spectrum [1], unveiled the law of radiative energy transfer [2, 3], provoked the gene- sis of quantum theory [4, 5], and provided the landmark evidence for the Big Bang model of the universe [68]. Contemporarily, a number of investigations in mesoscopic [9] and particle physics [10] are demanding microwave radiometry with single-photon sensitivity. In this article, we present a quantum-limited radiometry protocol based on the photon-induced dephasing of a superconducting qubit. By converting a small photon population to qubit coherence times, this mechanism is particularly suited for detecting microwave radiation beyond the Rayleigh– Jeans regime where single-photon energy exceeds the en- ergy of thermal fluctuations. A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector that converts the input power to an output electrical quan- tity. In reality, receiving circuits always add noise to the antenna radiation. This system noise, represented by a system temperature T sys , limits the sensitivity of a radiometer [12]. A common practice for reducing T sys is to equip the receiver with cryogenic low-noise ampli- fiers [13]. Following this design, space-based radiometers with T sys 10–100 K have been developed to map the anisotropies of the cosmic microwave background radi- ation [14, 15]. Alternatively, one can directly detect quantum radiation with novel microwave sensors. For in- stance, in circuit quantum electrodynamics (cQED) sys- tems [1619], one can deduce the coherent amplitude [20] or white-thermal-photon population [2127] of a micro- wave cavity from qubit coherence times—an effect known as photon-induced qubit dephasing [2830]. This method has been applied to extract the residual thermal popula- tions of qubit-readout cavities with the highest reported precision 10 -4 [27]. However, qubit radiometers have not yet been employed to probe radiative sources with pulsed emission or colored spectra. A frequently encountered type of thermal radiator is an electrical resonator at finite temperature, whose out- put field has a Lorentzian power spectral density cen- tered at its resonant frequency. Consider a resonator mode at frequency f coupled with rate κ c to a trans- mission line at temperature T c , and with rate κ i to an internal dissipative mechanism thermalized to a bath at T i >T c . The average thermal population of this res- onator mode is ¯ n =(κ c ¯ n c + κ i ¯ n i )/(κ c + κ i ), in which ¯ n c(i) =(e hf /kBT c(i) - 1) -1 . Radiative cooling n ¯ n c ) is manifested if κ c κ i , namely, when the resonator is over-coupled to a colder external bath. The same thermo- dynamic principle underlies nocturnal cooling on cloud- less nights [31, 32] when the earth’s surface, in contact with the 300-K atmosphere, is cooled by the 3-K cosmic microwave background. In this work, to test our qubit radiometer, we measured the thermal radiation from a 1-K microwave resonator, and unambiguously demon- strated its radiative cooling by a 0.1-K external coupling bath. Noise analysis yields T sys =0.31 K hf/2k B , which is dominated by dissipation and thermal leakage at low temperatures, and is a factor of two below the mini- mum output noise energy hf of an ideal quantum-limited phase-preserving linear amplifier, which includes both the amplified vacuum fluctuations accompanying the in- put signal and the minimum possible noise added by the amplifier. The well-understood performance and the ex- ceptional precision of this radiometer promise its applica- bility to a variety of problems, such as verifying ground- state cooling in hybrid quantum systems [33, 34] and up- grading microwave-cavity-based axion experiments [3537]. MEASUREMENT SETUP Although employing a different physical system, our experiment pays tribute to a classic measurement setup in radio astronomy invented by R. H. Dicke and R. Beringer [41], whose block diagram is sketched in Fig. 1(a). Enclosed in blue is a microwave radiometer arXiv:1909.12295v2 [quant-ph] 27 Sep 2019

Upload: others

Post on 10-Jul-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

Quantum Microwave Radiometry with a Superconducting Qubit

Zhixin Wang,1, ∗ Mingrui Xu,2 Xu Han,2 Wei Fu,2 Shruti Puri,1

S. M. Girvin,1 Hong X. Tang,2 S. Shankar,1, † and M. H. Devoret1, ‡

1Department of Applied Physics and Physics, Yale University, New Haven, CT 06520, USA2Department of Electrical Engineering, Yale University, New Haven, CT 06520, USA

(Dated: September 30, 2019)

The interaction of photons and coherent quantum systems can be employed to detect electro-magnetic radiation with remarkable sensitivity. We introduce a quantum radiometer based on thephoton-induced-dephasing process of a superconducting qubit for sensing microwave radiation atthe sub-unit-photon level. Using this radiometer, we demonstrated the radiative cooling of a 1-Kmicrowave resonator and measured its mode temperature with an uncertainty ∼ 0.01 K. We havethus developed a precise tool for studying the thermodynamics of quantum microwave circuits, whichprovides new solutions for calibrating hybrid quantum systems and detecting candidate particles fordark matter.

Historically, radiometry experiments—measurementsof the intensity of radiation—revealed the first invisibleband of the electromagnetic spectrum [1], unveiled thelaw of radiative energy transfer [2, 3], provoked the gene-sis of quantum theory [4, 5], and provided the landmarkevidence for the Big Bang model of the universe [6–8].Contemporarily, a number of investigations in mesoscopic[9] and particle physics [10] are demanding microwaveradiometry with single-photon sensitivity. In this article,we present a quantum-limited radiometry protocol basedon the photon-induced dephasing of a superconductingqubit. By converting a small photon population to qubitcoherence times, this mechanism is particularly suitedfor detecting microwave radiation beyond the Rayleigh–Jeans regime where single-photon energy exceeds the en-ergy of thermal fluctuations.

A classical microwave radiometer [11] typically consistsof a radio receiver followed by a square-law detector thatconverts the input power to an output electrical quan-tity. In reality, receiving circuits always add noise tothe antenna radiation. This system noise, representedby a system temperature Tsys, limits the sensitivity of aradiometer [12]. A common practice for reducing Tsys

is to equip the receiver with cryogenic low-noise ampli-fiers [13]. Following this design, space-based radiometerswith Tsys ∼ 10–100 K have been developed to map theanisotropies of the cosmic microwave background radi-ation [14, 15]. Alternatively, one can directly detectquantum radiation with novel microwave sensors. For in-stance, in circuit quantum electrodynamics (cQED) sys-tems [16–19], one can deduce the coherent amplitude [20]or white-thermal-photon population [21–27] of a micro-wave cavity from qubit coherence times—an effect knownas photon-induced qubit dephasing [28–30]. This methodhas been applied to extract the residual thermal popula-tions of qubit-readout cavities with the highest reportedprecision ∼ 10−4 [27]. However, qubit radiometers havenot yet been employed to probe radiative sources withpulsed emission or colored spectra.

A frequently encountered type of thermal radiator is

an electrical resonator at finite temperature, whose out-put field has a Lorentzian power spectral density cen-tered at its resonant frequency. Consider a resonatormode at frequency f coupled with rate κc to a trans-mission line at temperature Tc, and with rate κi to aninternal dissipative mechanism thermalized to a bath atTi > Tc. The average thermal population of this res-onator mode is n = (κcnc + κini)/(κc + κi), in whichnc(i) = (ehf/kBTc(i) − 1)−1. Radiative cooling (n → nc)is manifested if κc � κi, namely, when the resonator isover-coupled to a colder external bath. The same thermo-dynamic principle underlies nocturnal cooling on cloud-less nights [31, 32] when the earth’s surface, in contactwith the 300-K atmosphere, is cooled by the 3-K cosmicmicrowave background. In this work, to test our qubitradiometer, we measured the thermal radiation from a1-K microwave resonator, and unambiguously demon-strated its radiative cooling by a 0.1-K external couplingbath. Noise analysis yields Tsys = 0.31 K ∼ hf/2kB,which is dominated by dissipation and thermal leakage atlow temperatures, and is a factor of two below the mini-mum output noise energy hf of an ideal quantum-limitedphase-preserving linear amplifier, which includes boththe amplified vacuum fluctuations accompanying the in-put signal and the minimum possible noise added by theamplifier. The well-understood performance and the ex-ceptional precision of this radiometer promise its applica-bility to a variety of problems, such as verifying ground-state cooling in hybrid quantum systems [33, 34] and up-grading microwave-cavity-based axion experiments [35–37].

MEASUREMENT SETUP

Although employing a different physical system, ourexperiment pays tribute to a classic measurement setupin radio astronomy invented by R. H. Dicke andR. Beringer [41], whose block diagram is sketched inFig. 1(a). Enclosed in blue is a microwave radiometer

arX

iv:1

909.

1229

5v2

[qu

ant-

ph]

27

Sep

2019

Page 2: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

2

(a)

JPC

S I

SPA

cQED inProbe out

Probe in

Whitenoise in

JPCpump

Tref

Tadd: 0–2 K

Tatm

Superheterodynereceiver

: 1–2 KTVTS : 15 mKTMXC

Text

Text

cQED

ou

t

Tant

Ta

(b)

FIG. 1. (a) Diagram of an archetypal radio telescope. Bluebox: radio receiver. Orange box: earth’s atmosphere. Greentrident: receiving antenna. Blue arrow: extraterrestrial radi-ation. Red box: reference black-body radiator for tempera-ture calibration. (b) Setup of our quantum microwave radio-metric experiment. Blue area: mixing-chamber stage (MXC)of a dilution refrigerator. Orange box: variable-temperaturestage (VTS), which is attached to the still stage of the samedilution fridge through a weak thermal link. Green transmis-sion line: a niobium–titanium superconducting coaxial cableinductively coupled to a tunable antenna resonator (green).Blue arrow: residual thermal noise from the input line imping-ing on the resonator. Orange arrow: thermal leakage from theVTS to the MXC circuit. Red box: tunable broadband noisegenerator. JPC: Josephson parametric converter (depicted asan on–off switch), with three ports—signal (S), idler (I), andpump. cQED: a transmon qubit (purple) dispersively coupledto a readout cavity (blue). SPA: SNAIL parametric amplifier[38–40], which enables high-fidelity single-shot qubit readout.

of the Dicke type [42], characterized by a two-way switchat the receiver input selecting between a black-body ra-diator at a reference temperature Tref , and a receiv-ing antenna whose brightness is a mixture of extrater-restrial radiation and atmospheric background: Tant =γTatm + (1 − γ)Text, with γ denoting the atmosphericfractional absorption [43]. After calibrating Tant usingthe reference radiator and separately measuring Tatm andγ, one can extract Text—the microwave temperature ofthe astronomical object.

Our measurement setup is depicted in Fig. 1(b), withexperimental parameters listed in Table S1 of Ref. [44].At the system level, it can be viewed as a quantum-circuit realization of Dicke and Beringer’s radio telescope.The analogy is highlighted by the color code. Our “an-tenna,” with a tunable resonant frequency fa and a modetemperature Ta, is a superconducting LC resonator [45]anchored to a variable-temperature stage (VTS), whosetemperature TVTS can be adjusted between 1–2 K with-out affecting other parts of the dilution refrigerator. Theantenna mode is coupled with rate κa,i to the internaldissipative bath at TVTS, and with rate κa,c to a trans-mission line at Text ∼ 0.1 K [46]. Moreover, an external

broadband noise generator is able to raise the temper-ature of the transmission line by 0–2 K. In the quan-tum regime, i.e., kBTj . hfa with j denoting a mode orbath index, it is convenient to work with thermal photonpopulations in lieu of temperatures, which are linked bynj = (ehfa/kBTj − 1)−1. The thermal population of theantenna mode is then given by

na = γnVTS + (1− γ)(next + nadd) (1)

Here γ := κa,i/(κa,i+κa,c) = κa,i/κa ∼ 0.3 is analogous tothe “fractional absorption,” which is measured by a net-work analyzer through the ports Probe in and Probe out.During the experiment, TVTS is continuously monitoredby a calibrated thermometer. Therefore, the value of na

can be computed once Text is obtained.The detector of our quantum radiometer is a standard

cQED module in the strong dispersive regime, wherephoton-number fluctuations in a 3D readout cavity [47]induce extra dephasing in a superconducting transmonqubit [48–50]. The intensity of the cavity field can thenbe deduced given its photon statistics. For instance, withbroadband thermal noise at the cavity input, this photon-induced-dephasing rate is given by [30]

Γth(nthr ) =

κr

2

Re

√(1 +

κr

)2

+4iχnth

r

κr− 1

, (2)

where nthr is the cavity thermal population; κr = κr,c+κr,i

is the cavity linewidth; χ/2π is the dispersive shift of thequbit frequency per cavity photon. This dephasing ratecan be measured via the Ramsey-fringe technique. Dis-persive qubit readout [51] can be performed in reflectionthrough the ports cQED in and cQED out.

The cQED detector is only sensitive to incoming pho-tons in the vicinity of the qubit-readout frequency fr. Tomatch it to the antenna resonator, a Josephson paramet-ric converter (JPC) [52–54] is pumped at the frequencydifference of its S and I ports and operated as a loss-less microwave frequency converter [55] with a bandwidth∼ 100 MHz. Its conversion efficiency can be continuouslyadjusted from zero to unity. Without the pump, the JPCprovides over 40-dB isolation between its S and I ports,protecting the qubit from antenna radiation during itsinitialization, manipulation, and readout.

Our quantum radiometer shows clear advantages overits classical counterpart in sensing microwave radiationin the quantum regime. First, by mapping the cavitypopulation onto extra qubit dephasing, the cQED mod-ule directly measures the antenna radiation at the quan-tum level without passing it through any amplifier chainor demodulation circuit and without any sensitivity tovacuum noise. The system noise is thus significantly sup-pressed. Second, as qubit dephasing is caused by photon-number fluctuations in the readout cavity, our detectoressentially probes the excitations of an electromagnetic

Page 3: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

3

field instead of its quadrature amplitudes, and is thus im-mune to the one-photon vacuum fluctuation, which is theinherent noise limit of classical radiometers with phase-preserving linear amplifiers [56–59]. These two factorscontribute to the precision of the following experimentalresults.

RADIOMETRY PROTOCOL

Our radiometry measurements are based on theRamsey-interferometry experiment [60, 61]. As shown in

10.49 10.50

(b)

(c)

10.49 10.50

f

f rgf r

e

χ

fp

fa

/2πhot

cold

(a)Transmon

Readout

JPC pump

White noise

fp

= (frg + f r

ef r ) / 2τp

( )π2–Rφ( )π

2–R0

fqge

fqgeτ

FIG. 2. (a) Pulse sequences for noise-induced-dephasing mea-surements. For all data sets reported in this article, qubitrotations (purple) are completed in 200 ns; the second π/2pulse is applied after the JPC pump (gray) by τ − τp = 1 µs;the single-shot qubit readout (blue) lasts for 1.08 µs. (b)Frequency landscape of the experiment. The solid green peak(right) belongs to the antenna resonator. The light green ar-row indicates its frequency tunability. Along the frequencysweep, its down-converted image (dashed green peak) atfa−fp at crosses the two resonances of the qubit-readout cav-

ity (blue, fg(e)r = fr±χ/4π) separated by χ/2π. The red floor

refers to the added white noise. (c) Qubit-dephasing spectraat different added white-noise powers. The blue ribbon standsfor the range fe

r < fa − fp < fgr . The three zoomed-in plots,

from top to bottom, show the radiative heating, equilibrium,and cooling regimes, respectively. The accompanying picturesillustrate these regimes through the color (temperature) con-trasts between the amphorae and their backgrounds. For datain this subfigure, τp = 0.54 µs and TVTS = 1.03 K.

Fig. 2(a), two π/2 qubit pulses are separated by a fixeddelay τ . By varying their phase difference ϕ and pro-jectively measuring the qubit state afterwards, one canobserve 2π-periodic Ramsey fringes with an amplitudeequal to exp[−

∫ τ0

Γ(t) dt]/2, in which Γ(t) is the instan-taneous qubit decoherence rate [44]. When the JPC isin the isolation mode, Γ(t) equals the inherent Ramseydecoherence rate Γ2R = 1/T2R = (4±0.3)×104 s−1. Con-trarily, when the JPC is in the full-conversion mode, an-tenna radiation enters the cQED circuit while the resid-ual thermal photons at port I responsible for Γ2R areredirected to the port S. (Here we assume that Γ2R issolely caused by the residual photons coming from theinput line.) The ratio of the Ramsey amplitudes inthese two circumstances equals exp[

∫ τ0

Γa(t) dt− Γ2Rτp],in which Γa(t) is the net dephasing rate induced by theantenna radiation, and τp is the duration of the JPC-conversion pump, which also limits the frequency res-olution of the radiometry protocol to ∼ τ−1

p . We canthus experimentally obtain an average dephasing rateΓa = τ−1

p

∫ τ0

Γa(t) dt, in which the integral limit and the

prefactor are chosen such that Γa = Γth(nthr ) if the an-

tenna radiation is steady-state white noise with an aver-age thermal population nth

r . Finally, using Eq. (2), welink Γa to an effective cavity thermal population neff

r asif the dephasing is induced by steady-state white noise.Note that instead of measuring the frequency shifts ofthe zero-crossing points of Ramsey fringes, we essentiallyread out the fringe peaks where shot noise vanishes inthe absence of thermal photons or intrinsic qubit deco-herence.

In practice, we choose τp . min(T2R, T1)/10 to bothensure sufficient resolution on neff

r and mitigate qubit-relaxation events during the detection window. The sta-tistical error bars of Γa and neff

r are contributed by bothmeasurement imprecisions and the qubit T1 fluctuationsover time. The former scale as 1/

√Nrep, where Nrep is

the number of repetition of the Ramsey sequence. InRef. [44], we estimate the dynamic range of this radio-meter to be ∼ 50 dB.

RADIATIVE COOLING

Following this protocol, we measured neffr while sweep-

ing the frequency of the antenna resonator. Data areplotted in Fig. 2(c). When the down-converted antennaimage is detuned from fgr and fer , the qubit senses thewhite noise from the external bath with the thermal pop-ulation next + nadd reflected by the antenna. When theantenna image is aligned with fgr or fer , the qubit alsosenses the photons transmitted from the internal bathat TVTS = 1.03 K. As photons from a hotter bath leadto more qubit dephasing, three distinct regimes were ob-served: (i) If Tadd = 0, the dephasing increases when theresonances are aligned, indicating Text < TVTS. (ii) If

Page 4: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

4

FIG. 3. Radiative cooling and heating verified through clas-sical radiometry. The theoretical curves (light yellow) areplotted based on the experimental values: fa = 10.519 GHz,κa,c/2π = 0.30 MHz, κa,i/2π = 0.12 MHz, and TVTS =1.03 K.

Tadd ≈ 1 K, the dephasing is independent of fa, corre-sponding to next + nadd = nVTS. (iii) At any higher Tadd,the dephasing decreases when the resonances are aligned,suggesting next + nadd > nVTS. (i) and (iii) represent theradiative cooling and heating of the antenna mode by itsexternal bath, respectively.

Our method for comparing Text and TVTS by mea-suring thermal-photon-induced qubit dephasing is anal-ogous to the disappearing-filament technique in pyrom-etry [62, 63], which maps temperatures onto visible col-ors. Illustrated by the pictures on the right of Fig. 2(c),the peaked, flat, and dipped dephasing spectra (neff

r –fa curves) are associated with the amphorae [64] beingbrighter, indistinguishable, and darker compared to thebackground radiation, respectively. Temperature differ-ences are manifested by the data contrasts in both mea-surements, although one is performed in a kiln and theother in a dilution fridge.

To cross-check the above results, we performed a sep-arate classical radiometry experiment by probing thepower at Probe out [see Fig. 1(b)] with a commercialspectrum analyzer [65]. In this experiment, the JPC ispumped at the frequency sum of its S and I ports andoperated as a 23-dB quantum-limited phase-preservingamplifier. The receiver chain is further equipped with a33-dB HEMT amplifier on the 3.5-K stage of the dilu-tion fridge and two 28-dB RF amplifiers at room tem-perature. We first measured the background spectrawithin a 4-MHz detection window, and then tuned theantenna to the center of the window, took the spec-tra, subtracted the backgrounds, and referred the pow-ers to the output photon fluxes per unit bandwidthof the antenna resonator [44]. Results are plotted inFig. 3, together with the theoretical curves ∆nout

a [f −fa] = nVTSta[f − fa] + nadd(1 − ta[f − fa]), in which

ta[f −fa] = 4γ(1−γ)/[1+16π2(f−fa)2/κ2

a

]denotes the

Lorentzian transmission function of the resonator. Thetopmost, middle, and bottommost traces were measuredwith the same added white-noise powers as the threequbit-dephasing spectra in Fig. 2(c). The Lorentzianpeak evolves into a dip as nadd increases, which mani-fests the competition between the internal and externalbaths, and distinguishes the regimes of radiative coolingand heating. The agreement between the data and thetheory that assumes a constant TVTS = 1.03 K confirmsthe internal bath was not influenced while the externalbath was being heated by the white-noise source. There-fore, readouts of the VTS thermometer can be reliablyused in the calibration procedure of the radiometer.

NOISE ANALYSIS

We model our radiometer using the signal-flow dia-gram in Fig. 4(a) when the JPC is in the full-conversionmode, and write the photon flux per unit bandwidth at

(b) (c)

T loss

ta [f ] t loss

TVTS

Tadd Text

κa,c , κa,i

cQED

Antenna( )

VTSn( )

( )addn ( )extn ( )lossn

an outrn in

t leak

(a)

κr,c , κr,i( )

FIG. 4. (a) Noise model of the radiometer. Left beam split-ter (green): tunable antenna resonator. Right beam splitter(gray): ohmic loss along the coaxial cables and circulatorsbetween the resonator and the cQED module. Four coloredboxes represent the thermal baths. Yellow: internal bath inequilibrium with the VTS. Red: white-noise source. Blue:external bath in the transmission line. Gray: dissipation be-tween the antenna and the cQED cavity. (b) Normalizedslopes and (c) vertical intercepts of the neff

r –nadd lines at dif-ferent antenna frequencies with TVTS = 1.03 K. The solidred curve in (b) is the theoretical prediction based on theexperimental values of κa,c, κa,i, κr, and χ, with no fittingparameter. In (c), contributions of different baths are indi-cated using the same color code as that in (a). The orangeline separates the VTS leakage (below) and the transmittedradiation from the resonator (above). For all data in thisfigure, τp = 1.08 µs.

Page 5: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

5

the input of the qubit-readout cavity as

ninr [f−fa] = nout

a [f−fa]tloss + nloss(1− tloss), (3)

where nouta [f −fa] = nVTS(ta[f −fa] + tleak) + (next +

nadd)(1− ta[f−fa]) refers to the output photon flux perunit bandwidth of the antenna resonator. In Eq. (3),the term proportional to nloss represents the noise addedby the cables, connectors, and circulators between theresonator and the cQED module; the term containingnVTStleak represents the thermal leakage from the VTS.Together with the shot noise associated with the inher-ent qubit decoherence, they constitute the system noiseof the radiometer—the total noise at the output of theradiometer in the absence of input signals. Other terms,proportional to either ta or 1 − ta, are associated withphotons from the internal or external bath of the res-onator.

Unlike a standard square-law detector that measuresthe integral of impinging photon flux within its detectionbandwidth, our cQED module does not directly detect

ninr [f

g(e)r +fp−fa]. Instead, as we derive in Ref. [44], the

measured neffr in our experiment can be expressed as

neffr

[fr+fp−fa, τp

]={nVTS tloss ηa

[fr+fp−fa, τp

]

+ (next + nadd) tloss

(1− ηa

[fr+fp−fa, τp

])

+ nloss(1− tloss) + nVTS tleak tloss

}κ−2

r κr,cκa,

(4)

in which the dimensionless function ηa summarizes theresponse of the cQED detector to the pulsed noise witha Lorentzian spectrum. In Fig. 4(b), we plot the exper-imental data of ηa together with our theoretical predic-tions [44]. By measuring neff

r with different values ofnVTS and nadd, we calibrated the system parameters,obtained next, and computed na using Eq. (1). Theresults are listed in Table I, where the system noiseis given by nsys = npara + nshot, in which npara :=nVTStleak+nloss(1−tloss)/tloss denotes the parasitic whitenoise referred to the antenna input, and nshot, derivedin Ref. [44], is the shot noise due to the inherent qubitdecoherence. We obtain Tsys = 0.31 K ∼ hf/2kB,which is half of the minimum noise temperature of anideal quantum-limited phase-preserving linear amplifier.Radiative cooling is manifested again through next ∼nVTS/100 and na ∼ nVTS/3. As illustrated in Fig. 4(c),qubit dephasing is contributed by the white system noise(gray and lower orange), and the frequency-dependentradiation from the internal (upper orange) and external(blue) baths of the resonator. The shares of the lattertwo mechanisms in neff

r are proportional to ηa and 1−ηa,respectively.

As a comparison, the classical radiometry experimentdescribed in the previous section using a quantum-limitedparametric amplifier yields ncl

ext = 0.02+0.06−0.02 and T cl

sys =(1.02 ± 0.01) K ∼ 2hf/kB [65]. The advantages of the

nloss Tloss (K) next Text (K)0.09 ± 0.01 0.20 ± 0.01 0.014 ± 0.005 0.11 ± 0.01

npara Tpara (K) na Ta (K)0.16 ± 0.02 0.25 ± 0.01 0.49 ± 0.01 0.45 ± 0.01

nsys Tsys (K) tleak tloss

0.25 ± 0.02 0.31 ± 0.01 0.046 ± 0.005 0.52 ± 0.06

TABLE I. Radiometric results. The values of npara (Tpara),nsys (Tsys), and na (Ta) are based on TVTS = 1.03 K (nVTS =1.59). Thermal populations and temperatures are linked here

by nj = (ehfa/kBTj − 1)−1.

qubit radiometer in precision and system noise are thusclearly demonstrated. Moreover, as shown in Ref. [44],viewed from the point of photon counting, the perfor-mance of our qubit radiometer is comparable to state-of-the-art microwave single-photon detectors in the limit ofsmall input-photon numbers.

PERSPECTIVES

Our radiometer comprises a tunable antenna resonatorand a fixed-frequency narrowband detector (cQED mod-ule) matched by a broadband frequency converter (JPC).Although this article reports the experiment of probingtwo white thermal baths at Text and TVTS, in whichthe JPC-conversion frequency was fixed while the an-tenna frequency was varied, one can also sweep the con-version frequency together with the antenna frequencyand thus use the qubit to sense a colored bath or radi-ator. The JPC-conversion curve then needs to be cal-ibrated at every pump frequency. In this “sweepingmode,” the frequency range of this radiometer is broad-ened from κr ∼ 1 MHz to the JPC-conversion bandwidth∼ 100 MHz. Gigahertz bandwidth would be achievableby if the JPC is replaced with a broadband three-wave-mixing parametric device [66].

Two applications are within sight: First, our setup canbe used to study the thermodynamics of hybrid quan-tum systems, such as verifying the ground-state coolingof an electro-optomechanical photon converter [33] or aspin ensemble [34]. Second, superconducting qubits havebeen envisaged to replace low-noise amplifiers as itinerantphoton detectors in microwave-cavity-based axion exper-iments [35–37]. Considering the resemblance of our mea-surement setup to an axion detector [67–69], this radio-metry protocol is useful for detecting this hypotheticalparticle and constraining theories of dark matter.

ACKNOWLEDGEMENTS

We acknowledge V. V. Sivak for fabricating the SNAILparametric amplifier, and P. Bertet, A. A. Clerk, L. Di-

Page 6: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

6

Carlo, Liang Jiang, S. K. Lamoreaux, and K. W. Lehnertfor insightful discussions. Z.W. thanks P. Reinhold andS. O. Mundhada for their assistance on FPGA electron-ics. Use of facilities was supported by the Yale Institutefor Nanoscience and Quantum Engineering and the YaleSchool of Engineering and Applied Science clean room.This research was supported by the US Army ResearchOffice (Grants No. W911NF-18-1-0020 and W911NF-18-1-0212), the US Air Force Office of Scientific Research(Grant No. FA9550-15-1-0029), and the National ScienceFoundation (Grant No. DMR-1609326).

[email protected]† Present address: Department of Electrical and Computer

Engineering, University of Texas, Austin, TX 78758,USA

[email protected][1] W. Herschel, “Experiments on the refrangibility of the

invisible rays of the sun,” Philos. Trans. Royal Soc. 90,284 (1800).

[2] P. L. Dulong and A. T. Petit, Recherches sur la mesuredes temperatures et sur les lois de la communication dela chaleur (Imprimerie royale, Paris, 1818).

[3] J. Tyndall, On Radiation: The “Rede” Lecture (Long-man, London, 1865).

[4] O. Lummer and E. Pringsheim, “Ueber die Strahlung desschwarzen Korpers fur lange Wellen,” Verh. Dtsch. Phys.Ges. 2, 163 (1900).

[5] H. Rubens and F. Kurlbaum, “Uber die Emission lang-welliger Warmestrahlen durch den schwarzen Korperbei verschiedenen Temperaturen,” Sitzungsber. Preuss.Akad. Wiss. 41, 929 (1900).

[6] A. A. Penzias and R. W. Wilson, “A measurement ofexcess antenna temperature at 4080 Mc/s,” Astrophys.J. 142, 419 (1965).

[7] P. G. Roll and D. T. Wilkinson, “Cosmic backgroundradiation at 3.2 cm—support for cosmic black-body ra-diation,” Phys. Rev. Lett. 16, 405 (1966).

[8] T. F. Howell and J. R. Shakeshaft, “Measurement of theminimum cosmic background radiation at 20.7-cm wave-length,” Nature 210, 1318 (1966).

[9] Ya. M. Blanter and M. Buttiker, “Shot noise in meso-scopic conductors,” Phys. Rep. 336, 1 (2000).

[10] P. W. Graham, I. G. Irastorza, S. K. Lamoreaux, A. Lind-ner, and K. A. van Bibber, “Experimental searches forthe axion and axion-like particles,” Annu. Rev. Nucl.Part. Sci. 65, 485 (2015).

[11] N. Skou and D. Le Vine, Microwave Radiometer Systems:Design and Analysis, 2nd ed. (Artech House, Norwood,Massachusetts, 2006).

[12] The sensitivity of a total-power radiometer is written as

δT = Tsys

√(Bτint)−1 + (δG/G)2, in which τint is the in-

tegration time; B is the detection bandwidth; G and δGare the receiver gain and its fluctuation.

[13] E. W. Bryerton, M. Morgan, and M. W. Pospieszal-ski, “Ultra low noise cryogenic amplifiers for radio as-tronomy,” in 2013 IEEE Radio and Wireless Symposium(2013) p. 358.

[14] N. Jarosik, C. L. Bennett, M. Halpern, G. Hinshaw,A. Kogut, M. Limon, S. S. Meyer, L. Page, M. Pospieszal-ski, D. N. Spergel, G. S. Tucker, D. T. Wilkinson, E. Wol-lack, E. L. Wright, and Z. Zhang, “Design, implemen-tation, and testing of the Microwave Anisotropy Proberadiometers,” Astrophys. J. Suppl. S. 145, 413 (2003).

[15] B. Aja, E. Artal, L. de la Fuente, J. P. Pascual, A. Medi-avilla, N. Roddis, D. Kettle, W. F. Winder, L. P. Cara,and P. de Paco, “Very low-noise differential radiometerat 30 GHz for the PLANCK LFI,” IEEE Trans. Micro.Theory Tech. 53, 2050 (2005).

[16] A. Blais, R.-S. Huang, A. Wallraff, S. M. Girvin, andR. J. Schoelkopf, “Cavity quantum electrodynamics forsuperconducting electrical circuits: An architecture forquantum computation,” Phys. Rev. A 69, 062320 (2004).

[17] A. Blais, J. Gambetta, A. Wallraff, D. I. Schuster,S. M. Girvin, M. H. Devoret, and R. J. Schoelkopf,“Quantum-information processing with circuit quantumelectrodynamics,” Phys. Rev. A 75, 032329 (2007).

[18] R. J. Schoelkopf and S. M. Girvin, “Wiring up quantumsystems,” Nature 451, 664 (2008).

[19] S. M. Girvin, “Circuit QED: superconducting qubitscoupled to microwave photons,” in Quantum Machines:Measurement and Control of Engineered Quantum Sys-tems, Lecture Notes of the Les Houches SummerSchool, Session XCVI, edited by M. Devoret, B. Huard,R. Schoelkopf, and L. F. Cugliandolo (Oxford UniversityPress, Oxford, 2014) pp. 113–255.

[20] D. I. Schuster, A. Wallraff, A. Blais, L. Frunzio, R.-S.Huang, J. Majer, S. M. Girvin, and R. J. Schoelkopf,“ac Stark shift and dephasing of a superconducting qubitstrongly coupled to a cavity field,” Phys. Rev. Lett. 94,123602 (2005).

[21] P. Bertet, I. Chiorescu, G. Burkard, K. Semba, C. J.P. M. Harmans, D. P. DiVincenzo, and J. E. Mooij,“Dephasing of a superconducting qubit induced by pho-ton noise,” Phys. Rev. Lett. 95, 257002 (2005).

[22] C. Rigetti, J. M. Gambetta, S. Poletto, B. L. T. Plourde,J. M. Chow, A. D. Corcoles, J. A. Smolin, S. T. Merkel,J. R. Rozen, G. A. Keefe, M. B. Rothwell, M. B. Ketchen,and M. Steffen, “Superconducting qubit in a waveguidecavity with a coherence time approaching 0.1 ms,” Phys.Rev. B 86, 100506 (2012).

[23] A. P. Sears, A. Petrenko, G. Catelani, L. Sun, HanheePaik, G. Kirchmair, L. Frunzio, L. I. Glazman, S. M.Girvin, and R. J. Schoelkopf, “Photon shot noise de-phasing in the strong-dispersive limit of circuit QED,”Phys. Rev. B 86, 180504 (2012).

[24] F. Yan, S. Gustavsson, A. Kamal, J. Birenbaum, A. P.Sears, D. Hover, T. J. Gudmundsen, D. Rosenberg,G. Samach, S. Weber, J. L. Yoder, T. P. Orlando,J. Clarke, A. J. Kerman, and W. D. Oliver, “The fluxqubit revisited to enhance coherence and reproducibil-ity,” Nat. Commun. 7, 12964 (2016).

[25] F. Yan, D. Campbell, P. Krantz, M. Kjaergaard, D. Kim,J. L. Yoder, D. Hover, A. Sears, A. J. Kerman, T. P. Or-lando, S. Gustavsson, and W. D. Oliver, “Distinguishingcoherent and thermal photon noise in a circuit quantumelectrodynamical system,” Phys. Rev. Lett. 120, 260504(2018).

[26] J.-H. Yeh, J. LeFebvre, S. Premaratne, F. C. Wellstood,and B. S. Palmer, “Microwave attenuators for use withquantum devices below 100 mK,” J. Appl. Phys. 121,224501 (2017).

Page 7: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

7

[27] Z. Wang, S. Shankar, Z. K. Minev, P. Campagne-Ibarcq,A. Narla, and M. H. Devoret, “Cavity attenuators forsuperconducting qubits,” Phys. Rev. Appl. 11, 014031(2019).

[28] P. Bertet, I. Chiorescu, C. J. P. M. Harmans, and J. E.Mooij, “Dephasing of a flux-qubit coupled to a harmonicoscillator,” arXiv:cond-mat/0507290.

[29] J. Gambetta, A. Blais, D. I. Schuster, A. Wallraff,L. Frunzio, J. Majer, M. H. Devoret, S. M. Girvin, andR. J. Schoelkopf, “Qubit–photon interactions in a cavity:Measurement-induced dephasing and number splitting,”Phys. Rev. A 74, 042318 (2006).

[30] A. A. Clerk and D. W. Utami, “Using a qubit to measurephoton-number statistics of a driven thermal oscillator,”Phys. Rev. A 75, 042302 (2007).

[31] D. Brunt, “Notes on radiation in the atmosphere. I,” Q.J. Royal Meteorol. Soc. 58, 389 (1932).

[32] P. Groen, “Note on the theory of nocturnal radiationalcooling of the earth’s surface,” J. Meteorol. 4, 63 (1947).

[33] C. Zhong, Z. Wang, C. Zou, M. Zhang, X. Han, W. Fu,M. Xu, S. Shankar, M. H. Devoret, H. X. Tang, andL. Jiang, “Heralded generation and detection of entan-gled microwave–optical photon pairs,” arXiv:1901.08228.

[34] B. Albanese, et al., in preparation.[35] S. K. Lamoreaux, K. A. van Bibber, K. W. Lehnert,

and G. Carosi, “Analysis of single-photon and linear am-plifier detectors for microwave cavity dark matter axionsearches,” Phys. Rev. D 88, 035020 (2013).

[36] H. Zheng, M. Silveri, R. T. Brierley, S. M. Girvin,and K. W. Lehnert, “Accelerating dark-matter ax-ion searches with quantum measurement technology,”arXiv:1607.02529.

[37] A. Dixit, A. Chou, and D. Schuster, “Detecting axiondark matter with superconducting qubits,” in MicrowaveCavities and Detectors for Axion Research, Springer Pro-ceedings in Physics 211, edited by G. Carosi, G. Rybka,and K. van Bibber (Springer, Cham, 2018) pp. 97–103.

[38] N. E. Frattini, U. Vool, S. Shankar, A. Narla, K. M.Sliwa, and M. H. Devoret, “3-wave mixing Josephsondipole element,” Appl. Phys. Lett. 110, 222603 (2017).

[39] N. E. Frattini, V. V. Sivak, A. Lingenfelter, S. Shankar,and M. H. Devoret, “Optimizing the nonlinearity anddissipation of a SNAIL parametric amplifier for dynamicrange,” Phys. Rev. Appl. 10, 054020 (2018).

[40] V. V. Sivak, N. E. Frattini, V. R. Joshi, A. Lingenfelter,S. Shankar, and M. H. Devoret, “Kerr-free three-wavemixing in superconducting quantum circuits,” Phys. Rev.Appl. 11, 054060 (2019).

[41] R. H. Dicke and R. Beringer, “Microwave radiation fromthe sun and moon,” Astrophys. J. 103, 375 (1946).

[42] R. H. Dicke, “The measurement of thermal radiationat microwave frequencies,” Rev. Sci. Instrum. 17, 268(1946).

[43] R. H. Dicke, R. Beringer, R. L. Kyhl, and A. B. Vane,“Atmospheric absorption measurements with a micro-wave radiometer,” Phys. Rev. 70, 340 (1946).

[44] See Supplemental Material for further experimental de-tails and theoretical analysis of this work.

[45] M. Xu, X. Han, W. Fu, C.-L. Zou, M. H. Devoret, andH. X. Tang, “Frequency-tunable high-Q superconductingresonators via wireless control of nonlinear kinetic induc-tance,” Appl. Phys. Lett. 114, 192601 (2019).

[46] Text is in principle set by the temperature of the baseattenuator on the Probe in line. In practice, for the pur-

pose of a parallel experiment, this 30-dB cryogenic atten-uator is anchored to another variable-temperature stage(VTS) similar to the one holding the tunable antennaresonator, but connected to the 15-mK mixing-chamberstage through a weak thermal link. When the externalnoise generator is turned off, the ruthenium-oxide ther-mometer on the mixing-chamber VTS reads 70 mK. Con-sidering extra noise added by the following circulatorsand superconducting cables, we estimate Text ∼ 0.1 K,which is consistent with the radiometric data.

[47] H. Paik, D. I. Schuster, L. S. Bishop, G. Kirchmair,G. Catelani, A. P. Sears, B. R. Johnson, M. J. Reagor,L. Frunzio, L. I. Glazman, S. M. Girvin, M. H. De-voret, and R. J. Schoelkopf, “Observation of high coher-ence in Josephson junction qubits measured in a three-dimensional circuit QED architecture,” Phys. Rev. Lett.107, 240501 (2011).

[48] A. Cottet, Implementation of a quantum bit in a super-conducting circuit, PhD dissertation, Universite Paris VI(2002).

[49] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I.Schuster, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin,and R. J. Schoelkopf, “Charge-insensitive qubit designderived from the Cooper pair box,” Phys. Rev. A 76,042319 (2007).

[50] J. A. Schreier, A. A. Houck, Jens Koch, D. I. Schuster,B. R. Johnson, J. M. Chow, J. M. Gambetta, J. Ma-jer, L. Frunzio, M. H. Devoret, S. M. Girvin, and R. J.Schoelkopf, “Suppressing charge noise decoherence in su-perconducting charge qubits,” Phys. Rev. B 77, 180502(2008).

[51] A. Narla, Flying Qubit Operations in SuperconductingCircuits, PhD dissertation, Yale University (2017).

[52] N. Bergeal, R. Vijay, V. E. Manucharyan, I. Siddiqi, R. J.Schoelkopf, S. M. Girvin, and M. H. Devoret, “Ana-log information processing at the quantum limit with aJosephson ring modulator,” Nat. Phys. 6, 296 (2010).

[53] N. Bergeal, F. Schackert, M. Metcalfe, R. Vijay, V. E.Manucharyan, L. Frunzio, D. E. Prober, R. J. Schoelkopf,S. M. Girvin, and M. H. Devoret, “Phase-preserving am-plification near the quantum limit with a Josephson ringmodulator,” Nature 465, 64 (2010).

[54] N. Roch, E. Flurin, F. Nguyen, P. Morfin, P. Campagne-Ibarcq, M. H. Devoret, and B. Huard, “Widely tunable,nondegenerate three-wave mixing microwave device op-erating near the quantum limit,” Phys. Rev. Lett. 108,147701 (2012).

[55] B. Abdo, K. Sliwa, F. Schackert, N. Bergeal, M. Ha-tridge, L. Frunzio, A. D. Stone, and M. Devoret, “Fullcoherent frequency conversion between two propagatingmicrowave modes,” Phys. Rev. Lett. 110, 173902 (2013).

[56] J. R. Tucker and M. J. Feldman, “Quantum detectionat millimeter wavelengths,” Rev. Mod. Phys. 57, 1055(1985).

[57] L. P. Bolgiano, “Quantum fluctuations in microwaveradiometry,” IRE Trans. Microw. Theory Tech. 9, 315(1961).

[58] H. Heffner, “The fundamental noise limit of linear am-plifiers,” Proc. IRE 50, 1604 (1962).

[59] C. M. Caves, “Quantum limits on noise in linear ampli-fiers,” Phys. Rev. D 26, 1817 (1982).

[60] N. F. Ramsey, “A molecular beam resonance methodwith separated oscillating fields,” Phys. Rev. 78, 695(1950).

Page 8: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

8

[61] D. Vion, A. Aassime, A. Cottet, P. Joyez, H. Pothier,C. Urbina, D. Esteve, and M. H. Devoret, “Manipulatingthe quantum state of an electrical circuit,” Science 296,886 (2002).

[62] L. Holborn and F. Kurlbaum, “Uber ein optisches Py-rometer,” Ann. Phys. 315, 225 (1903).

[63] L. Michalski, K. Eckersdorf, J. Kucharski, andJ. McGhee, Temperature Measurement , 2nd ed. (Wiley,Chichester, West Sussex, 2001) p. 163.

[64] Photographed at the Museum of Fine Arts in Boston.Amphorae of this shape are representative of the Yang-shao culture (Banpo phase) in Neolithic China, c. 5000–4000 BCE.

[65] For more details on this classical radiometry method, seean accompanying publication by M. Xu, et al. (in prepa-ration).

[66] A. B. Zorin, “Josephson traveling-wave parametric am-plifier with three-wave mixing,” Phys. Rev. Appl. 6,034006 (2016).

[67] S. J. Asztalos, G. Carosi, C. Hagmann, D. Kinion,K. van Bibber, M. Hotz, L. J Rosenberg, G. Ry-bka, J. Hoskins, J. Hwang, P. Sikivie, D. B. Tanner,R. Bradley, and J. Clarke, “SQUID-based microwavecavity search for dark-matter axions,” Phys. Rev. Lett.104, 041301 (2010).

[68] N. Du et al. (ADMX Collaboration), “Search for invisi-ble axion dark matter with the axion dark matter exper-iment,” Phys. Rev. Lett. 120, 151301 (2018).

[69] B. M. Brubaker, L. Zhong, Y. V. Gurevich, S. B. Cahn,S. K. Lamoreaux, M. Simanovskaia, J. R. Root, S. M.Lewis, S. Al Kenany, K. M. Backes, et al., “First resultsfrom a microwave cavity axion search at 24 µeV,” Phys.Rev. Lett. 118, 061302 (2017).

Page 9: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

Supplemental Material for“Quantum Microwave Radiometry with a Superconducting Qubit”

I. EXPERIMENTAL DETAILS

A detailed drawing of our experimental setup is shownin Fig. S1. Low-temperature measurements were per-formed in a cryogen-free dilution refrigerator. Ourtransmon qubit consists of an Al/AlOx/Al Josephsonjunction connected to a pair of rectangular aluminumpads. The junction is defined on a sapphire chip us-ing the bridge-free electron-beam lithography technique[S1, S2] and fabricated via double-angle evaporation. Thereadout resonator is a 3D indium-plated copper cav-ity, which has a superconducting surface and a high-thermal-conductivity bulk at low temperatures. ThiscQED module is housed in a mu-metal (Amumetal 4K)magnetic shield and thermalized to the 15-mK mixing-chamber (MXC) stage. Homemade Eccosorb filters areinstalled on the input and output lines, and, in addi-tion, on the cavity-to-SMA coupler inside the mu-metalshield to block spurious high-frequency radiation and im-prove qubit coherence times [S3]. Using standard pulsesequences, we measured T1, Ramsey T2R, and Hahn echoT2e, and computed the inherent qubit dephasing rateΓφ = T−1

2R − (2T1)−1 = (3 ± 0.3) × 104 s−1. The factT2R ≈ T2e indicates that there is little low-frequencynoise and thus that qubit dephasing is primarily con-tributed by residual thermal photons in the readout cav-ity. Attributing Γφ to the residual thermal populationin the fundamental mode of the readout cavity, we ob-tain nth

r = (7± 0.7)× 10−3. Based on the results of ourprevious experiment [S4], we assume all these parasiticthermal photons solely come from the input line of thecQED module.

The Josephson parametric converter (JPC) and theSNAIL parametric amplifier (SPA) are enclosed in alu-minum cans inside their own mu-metal shields, and areboth thermalized to the MXC stage. Magnetic fluxes areapplied to the Josephson-junction loops through DC coilsmounted inside the aluminum cans. The bias fluxes arechosen such that the JPC idler resonator and the SPAresonator are both aligned to the qubit-readout cavity.Consequently, the JPC converts antenna radiation to thevicinity of the qubit-readout frequency fr = (fgr + fer )/2with a close-to-unity efficiency (see Sec. III). By pumpingthe SPA at 2fr + 10 MHz, we operate it as a quantum-limited degenerate phase-preserving amplifier with 14-dB gain for qubit-readout signals. The fidelity of thesingle-shot dispersive readout is optimized to be above0.95, which is mainly limited by the less-than-expectedSPA gain and can be improved by a higher-gain device[S5, S6], with which a readout fidelity of 0.975 has beenpreviously recorded in the same measurement setup.

Our tunable antenna resonator is anchored to a copperstage (variable-temperature stage, VTS), which is itself

Transmon qubit

fgeq (GHz) 4.6820

fefq (GHz) 4.4487

χ/2π (MHz) 3.1T1 (µs) 71± 2T2R (µs) 24± 2T2e (µs) 27± 2P inie 0.03± 0.002

cQED readout cavity

fr (GHz) 7.6011κr/2π (MHz) 0.83κr,i/2π (MHz) 0.06κr,c/2π (MHz) 0.77

Antenna resonator

fa (GHz) 10.48–10.52κa/2π (MHz) 0.37–0.42κa,i/2π (MHz) 0.10–0.14κa,c/2π (MHz) 0.26–0.30

JPC

fS (GHz) 10.52κS/2π (MHz) 56fI (GHz) 7.61

κI/2π (MHz) 83fp (GHz) 2.8935

TABLE S1. Summary of device parameters. fgeq and fef

q : ge-

and ef -transition frequencies of the transmon qubit. P inie :

excited-state population of the transmon in the absence ofdrives. Subscript “i”: internal dissipation; “c”: external cou-pling. Total linewidth of the readout cavity (antenna res-onator): κr(a) = κr(a),i + κr(a),c.

weakly thermalized to the still stage of the dilution fridgevia a section of copper wire. The temperature of the VTScan be raised by a heater and monitored in time by aruthenium-oxide thermometer. When the heater is off,we measured Tstill = 0.87 K and TVTS = 1.03 K. Whenthe heater current is changed, the VTS will reach itsnew stable temperature with a time constant ∼ 15 min.While varying TVTS between 1.03 K and 2.2 K, we did notobserve any change on the temperatures of other fridgestages beyond their normal fluctuations.

The last 30-dB attenuator on the antenna input line isanchored to a copper VTS that is constructed similarly tothe one holding the antenna resonator but is thermalizedto the MXC stage instead. In the absence of input sig-nals, the thermometer attached to the MXC-stage VTSreads 70 mK, which sets a lower bound of Text. Dur-ing the experiment, the heater on the MXC-stage VTSremained off. The temperature of the external bath ofthe antenna is thus only raised by the added noise with80-MHz bandwidth near the antenna frequency.

Key parameters of this experiment are listed in Ta-ble S1. Frequencies and linewidths of the JPC res-onators were characterized with weak probe tones with-out pumps.

arX

iv:1

909.

1229

5v2

[qu

ant-

ph]

27

Sep

2019

Page 10: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

2

0.1 K

0.9 K

3.5 K

300 K

50 K

HEMT

Legend

Mixer

=

=

Aluminumshield

Amumetalshield

Circulator Isolator

Low-pass�lter

Eccosorb�lter

Cryogenicattenuator

High-pass�lter

High-electron-mobility transistor

25 dBI

Q

I

Q

IF white noise (0–80 MHz)

25 dBI

Q

ge ef

signal

reference

readout

50 Ωload

SPA

Ecc

osor

b

Ecc

osor

b

-30

dB

-30

dB

Ecc

osor

b

Ecc

osor

b

-10

dB

-10

dBE

ccos

orb

Ecc

osor

b

Ecc

osor

b

-10

dB

-10

dB-1

0 dB

-10

dB

-10

dB

-10

dB

-10

dB-2

0 dB

-20

dB

-20

dB

-20

dB

-20

dB-1

0 dB

HEMT HEMT

P

P

JPC

S I

15 mK(MXC)

(Still)

JPCpump

SPApump

MXCVTS

Vectornetworkanalyzer

Whitenoise

StillVTS

64 dB

+50 MHzreadout

transmon

SA

SA

VNA

VNA

56 dB

IQ mixer

Superconductingcoaxial cable (Nb–Ti)

Microwavegenerator

Spectrumanalyzer

Microwaveswitch

cQEDin

cQEDout

Probeout

Probein

FIG. S1. Detailed cryogenic wiring diagram with simplified room-temperature electronics. Envelopes of qubit pulses aregenerated by the 500 Msample/s analog outputs of a field programmable gate array (FPGA) card and sent to the IQ mixers.Qubit-readout pulses are shaped by a high-frequency microwave switch. The 50-MHz signal and reference outputs of theheterodyne interferometer are digitized and processed by the same FPGA card.

II. FREQUENCY-TUNABLE ANTENNARESONATOR

The antenna of our radiometer is a 2D niobium-nitridesuperconducting LC resonator deposited on a sapphiresubstrate, which is inductively coupled to a niobium–titanium superconducting transmission line. As shown inFig. S2, square holes on the circular inductor arm allowthe kinetic inductance of the resonator to be increasedby an external magnetic field perpendicular to the deviceplane, down-shifting the resonator frequency [S7]. In thisexperiment, the antenna frequency can be lowered from

10.7 GHz to 10.4 GHz by a ∼ 1 mT magnetic field. Inthe upper panel of Fig. S3(a) is the magnetic-coil currentplotted against the antenna frequency within the rangeof our radiometry measurements. A frequency hysteresis∼ 30 kHz can be observed if one sweeps the magnetic fieldbi-directionally. To address this hysteresis, each antenna-frequency point in the qubit-dephasing spectra (neff

r –fa

curves) reported in the Main Text was measured by anetwork analyzer immediately before the Ramsey exper-iment. As shown in Fig. S3(b), κa,c and κa,i weakly de-pend on the antenna frequency, giving γ := κa,i/κa ∼ 0.3.

Page 11: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

3

0.1 mm

B

FIG. S2. Optical micrograph of the antenna resonator. Thetotal inductance of the device comprises a geometric part anda kinetic part. The zoomed-in view shows the square holeson the inductor arm responsible for the frequency tunabilityof the antenna resonator.

10.494 10.49810.49610.49 10.5010.495

(a) (b)

FIG. S3. (a) Antenna parameters measured during the fre-quency sweep of Figs. 4(b) and 4(c). Top: magnetic-coilcurrent versus antenna frequency, where I = 60 mA corre-sponds to B ∼ 1 mT perpendicular to the device plane. Bot-tom: κa,c and κa,i at different antenna frequencies. (b) An-tenna reflection at I = 48.765 mA, where fa = 10.4961 GHz,κa,c = 0.27 MHz, κa,i = 0.12 MHz, and γ = 0.31. Thesevalues are pointed out in (a) by the red arrows.

III. FREQUENCY CONVERSION THROUGH AJOSEPHSON PARAMETRIC CONVERTER

The JPC is a frequency-tunable non-degenerate three-wave-mixing device for converting or amplifying micro-wave signals close to the quantum limit [S8, S9]. It con-sists of a Josephson ring modulator (JRM) embedded intwo crossed λ/2 resonators denoted signal (S) and idler(I). To use the JPC as a quantum-limited frequency con-verter, we apply a pump at the frequency difference ofits S and I resonators: fp = fS − fI. Due to the nonlin-earity of the JRM, when the pump is applied, fS and fI

are down-shifted by ∼ 10 MHz from their linear-responsevalues recorded in Table S1. The pump frequency shouldthus be optimized within the range of a few megahertzto yield the maximum conversion efficiency between theantenna and the cQED module.

The JPC-conversion curves are calibrated with fp =

(a) (b)

FIG. S4. (a) JPC reflection and conversion coefficients asfunctions of the pump power Pp. Green (blue) curve: polyno-mial fit of the |SSS(SI)|2 data. (b) S-port reflection spectrum

at the full-conversion point (Pp = P fullp ).

2.8935 GHz. Using the method of Ref. [S10], we firstapplied to the port Probe in a continuous-wave signalat 10.4961 GHz and monitored the output signal at thesame frequency with a spectrum analyzer at the portProbe out. The green dots in Fig. S4(a) are the re-flection coefficients at different pump powers, which arecomputed by normalizing the output powers at Probeout, with the power measured with Pp = 0 mapped to|SSS| = 1 and the noise floor mapped to |SSS| = 0. Asshown in Fig. S4(b), we measured the reflection spec-trum with a network analyzer when the JPC is in thefull-conversion mode, and verified |SSS|2 < 0.001 withinthe antenna-frequency range of our experiment. To cal-ibrate the JPC-conversion curve (blue), we pumped theJPC at its half-conversion point (|SSS|2 = 0.5), appliedto the port cQED in a continuous-wave signal at 7.6026GHz, and adjusted its power such that the output powerat Probe out is the same given either input (Probe in orcQED in). Then by normalizing the converted power atProbe out in the same manner as the reflected power,we obtained the conversion coefficients plotted as theblue dots in Fig. S4(a). The reflection and conversioncurves are symmetric about |S|2 = 0.5. At the full-conversion point (Pp = P full

p ), the maximum conversion

efficiency is close to unity (|SSI|2 > 0.99). We thus modelthe JPC between its isolation and full-conversion modeas a near-ideal microwave switch with a response time∼ 1/κS(I) < 3 ns.

IV. NOISE-INDUCED-DEPHASINGEXPERIMENT

A. Ramsey experiment

The noise-induced-dephasing measurement in thiswork is based on the Ramsey-interferometry experiment.Following the pulse sequences in Fig. 2(a), we show twosinusoidal Ramsey-oscillation curves in Fig. S5, corre-sponding to the JPC switch being constantly off (blue)and turned on for duration τp (red), respectively. Their

Page 12: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

4

FIG. S5. Ramsey oscillations with and without antenna ra-diation. The JPC-conversion pump lasts for τp = 1.08 µs.

amplitudes

Aoff =1

2e−Γ2Rτ (S1)

and

Aon =1

2exp

{−∫ τ

0

[Γa(t) + Γ2R

]dt+ Γ2Rτp

}(S2)

are related to the additional qubit dephasing induced bythe antenna radiation via

Aoff

Aon= exp

[∫ τ

0

Γa(t) dt− Γ2Rτp

]= e(Γa−Γ2R)τp , (S3)

which is independent of the inherent qubit decoherencerate Γ2R = 1/T2R. In Eq. (S2), the term Γ2Rτp is addedto the exponent to account for the bi-directional conver-sion of the JPC. Namely, while the JPC is in the full con-version mode, the residual thermal photons causing Γ2R,which are assumed to come from the qubit input line, areconverted from port I to port S, and therefore have nocontribution to qubit dephasing during τp. We then ob-tain the effective thermal population of the qubit-readoutcavity from the average antenna-induced-dephasing rate

Γa =κr

2

Re

√(1+

κr

)2

+4iχneff

r

κr−1

neff

r �1−−−−→ χ2κrneffr

χ2 + κ2r

.

(S4)

If the antenna noise is white, this definition of neffr makes

it equal to the actual thermal population of the qubit-readout mode.

We note that the qubit excited-state population andthe infidelity of the single-shot qubit readout lower theRamsey-oscillation contrast from unity even in the ab-sence of dephasing channels. But we are able to cor-rect these offsets through dividing Aon by Aoff insteadof extracting neff

r directly from the Ramsey amplitudes.Moreover, as can be found in Fig. S5, the switch-on curveshows a phase shift compared to the switch-off curve.This demonstrates the AC Stark shift of the qubit fre-quency induced by the cavity photons. We can similarlyextract neff

r from Stark-shift data. The results match

those obtained using dephasing rates but have larger er-ror bars (data not shown). We therefore used the de-phasing data in the radiometry measurements of our ex-periment.

B. Photon-number calibration

We use Eq. (S91) to calibrate the temperature Tadd, orequivalently, the thermal population nadd of the externalwhite-noise source referred to the input of the antennaresonator. To do this, we adjust the antenna frequencysuch that |fa − fr − fp| � χ/2π, leaving nout

a and ninr

frequency-independent within the detection bandwidth

nouta = nVTStleak + next + nadd, (S5)

ninr = nout

a tloss + nloss(1− tloss) (S6)

= nparatloss + (next + nadd)tloss, (S7)

neffr =

κaκr,c

κ2r

ninr , (S8)

in which npara := nVTStleak + nloss(1− tloss)/tloss denotesthe parasitic white noise referred to the antenna input.As shown in Fig. S6(a), we operated JPC close to its full-conversion point, and measured neff

r as we ramped up thepower of added white noise at room temperature. UsingEqs. (S7) and (S8), we thus obtain the linear coefficientbetween Padd and neff

r , and convert neffr to nadd using

tloss = 0.57 (see Sec. VI). All the values of nadd reportedin the Main Text are calibrated using this method.

In Fig. S6(b), we fixed Padd = 2.5 mW and swept theJPC-pump power. The measured readout-mode popu-lations fall on a rescaled JPC-conversion curve [blue inFig. S4(a)], which shows that the JPC frequency conver-sion is noiseless in the sub-unit-photon regime.

(a) (b)

FIG. S6. White-noise-induced qubit dephasing. (a) neffr

against white-noise-generator power at room temperature.The JPC is close to its full-conversion mode. (b) neff

r ver-sus JPC pump power given Padd = 2.5 mW. The solid redcurve is a rescaled JPC-conversion curve.

Page 13: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

5

C. Dynamic range

Here we give an estimate of the dynamic range ofour radiometer. The detection limit of neff

r is set byδΓ2R/κr ∼ 10−3, in which δΓ2R is the fluctuation ofthe inherent qubit decoherence rate measured over 7hours. The upper bound of detectable neff

r is approx-imately 2π(fgeq −fefq )/χ ∼ 102. This restriction arisesfrom the phenomenological observation that the maxi-mum AC Stark shift of a transmon qubit (nrχ) appearsto be limited by the qubit anharmonicity. From these ap-proximations, we estimate a dynamic range of the radio-meter around 102/10−3 ∼ 50 dB.

V. THEORY OF COLORED-NOISE-INDUCEDQUBIT DEPHASING

Previously, theories of qubit dephasing induced by co-herent drives, white noise, and two-mode squeezed lightshave been studied in Refs. [S11], [S12], and [S13], re-spectively. However, in our experiment, the antennanoise incident on the radiometer is pulsed and frequency-dependent. Here we present two approaches for predict-ing the qubit dephasing induced by pulsed Lorentzianthermal noise. They quantitatively agree for the condi-tions of our experiment. For simplicity, only the resultsof the first approach, which is analytical, are comparedwith the experimental data in the Main Text.

A. Small-thermal-population approximation

In this subsection, we will derive an analytical expres-sion for the average dephasing rate Γa and the effectivethermal population neff

r as a function of the various bathpopulations if the antenna radiation has a Lorentzianspectrum—the condition of our radiometry experiment.

First, consider an ideal situation where npara = 0

(tleak = 0, tloss = 1). We use a and b to denote theantenna and the qubit-readout mode, respectively. TheHamiltonians of the two subsystems are written as

Ha

~= ωaa

†a+Hκa

~+Hda

~, (S9)

Hb

~= ωrb

†b+ωgeq

2σz −

χ

2b†bσz +

Hκb

~+Hdb

~. (S10)

In the above equations, ωj = 2πfj is the angular fre-quency of mode j; σz is the qubit-state operator, whoseeigenvalue ±1 refers to the qubit being in its e or g state;Hamiltonians with superscripts “κ” and “d” formally de-scribe dissipations and drives.

The Heisenberg–Langevin equations of the two cavity

modes are

˙a(t) = −(iωa+

κa

2

)a(t) +

√κa,iAin,i(t) +

√κa,cAin,c(t),

(S11)

˙b(t) = −

[i

(ωr −

χ

2σz

)+κr

2

]b(t)

+√κr,iBin,i(t) +

√κr,cBin,c(t). (S12)

Define the readout mode operator conditioned on thequbit state as

bσ(t) := 〈σ| b(t) |σ〉 , (S13)

in which σ = ±1 (e or g). Here we disregard the energyrelaxation of the qubit during the process, so that σ is aconstant of motion. We rewrite Eq. (S12) as

˙bσ(t) = −

(iωσr +

κr

2

)bσ(t) +

√κr,iBin,i(t) +

√κr,cBin,c(t),

(S14)

where ωσr = ωr − σχ/2 is the qubit-state-dependentreadout-cavity frequency.

The input–output relations [S14] conditioned on thequbit state are

Aout,c = Ain,c −√κa,c a = Bin,c, (S15)

Bσout,c = Bin,c −√κr,c bσ. (S16)

The input-mode operators Ain,i and Ain,c (Bin,i and

Bin,c) denote the thermal fluctuations of the internal andexternal baths of the antenna resonator (qubit-readout

cavity). Bin,i can be disregarded as we are calculatingonly the qubit dephasing induced by the antenna radia-tion incident on the qubit-readout cavity. We write thetime-correlation functions of the stationary white ther-mal baths

〈A†in,i(t)Ain,i(t′)〉 = nVTSδ(t− t′), (S17)

〈Ain,i(t)A†in,i(t

′)〉 = (nVTS + 1)δ(t− t′), (S18)

〈A†in,c(t)Ain,c(t′)〉 = (next + nadd)δ(t− t′), (S19)

〈Ain,c(t)A†in,c(t′)〉 = (next + nadd + 1)δ(t− t′), (S20)

and their spectral densities

〈A†in,iAin,i〉 [ω] = nVTS, (S21)

〈Ain,iA†in,i〉 [ω] = nVTS + 1, (S22)

〈A†in,cAin,c〉 [ω] = next + nadd, (S23)

〈Ain,cA†in,c〉 [ω] = next + nadd + 1, (S24)

which are related via

〈F1F2〉 [ω] =

∫ ∞

−∞〈F1(t+ τ)F2(t)〉 eiωτ dτ. (S25)

Page 14: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

6

Fourier-transform Eq. (S11) into the frequency domain:[κa

2− i(ω − ωa)

]a[f ] =

√κa,iAin,i[f ] +

√κa,cAin,c[f ].

(S26)

From Eqs. (S15) and (S26), we obtain

〈B†in,cBin,c〉 [ω] = 〈A†out,cAout,i〉 [ω]

= nVTSta[ω−ωa] + (next+nadd)(1−ta[ω−ωa]) (S27)

= (nVTS−next−nadd)ta[ω−ωa] + next + nadd, (S28)

in which ta is the Lorentzian transmission function

ta[ω−ωa] =4κa,iκa,c

κ2a + 4(ω−ωa)2

=4γ(1−γ)

1 + 4(ω−ωa)2/κ2a

, (S29)

where γ = κa,i/κa.We now set next = nadd = 0 in Eq. (S27) and fo-

cus on the effect of transmitted noise from the internalbath—the term proportional to ta. In the time domain,

the correlation functions of B†in,c and Bin,c have bilateralexponential forms. We define

L(t, t′) := κaγ(1−γ)e−κa2 |t−t′|eiωa(t−t′), (S30)

and can thus write

〈B†in,c(t)Bin,c(t′)〉 = nVTSL(t, t′), (S31)

〈Bin,c(t)B†in,c(t′)〉 = (nVTS + 1)L∗(t, t′), (S32)

〈Bin,c(t)Bin,c(t′)〉 = 〈B†in,c(t)B†in,c(t′)〉 = 0. (S33)

One can consider the input colored thermal noise as anincoherent mixture of coherent states. For a coherent-state input, the total cavity-induced qubit dephasing,namely, the off-diagonal element of the qubit density ma-trix after the cavity reaches a steady state is given bye−

∫Γcoh(t) dt, where [S11, S13]

Γcoh(t) =κr

2

∣∣βg(t)− βe(t)∣∣2 , (S34)

in which βg and βe are the (complex) coherent-state am-plitudes of the readout cavity given the qubit being in |g〉and |e〉. In the case of thermal noise input, the dephasingcan be obtained by averaging over all possible realizationsof the input coherent state 〈e−

∫Γcoh(t) dt〉. One can sim-

plify this expression through the cumulant expansion ofaverages, that is,⟨e−

∫Γcoh(t) dt

⟩= e−

∫〈Γcoh(t)〉dt+ 1

2

∫∫〈Γcoh(t)Γcoh(t′)〉 dtdt′+···

(S35)

Higher-order cumulants can be neglected in the limit ofsmall input thermal photon number, namely, γnVTS � 1.In this approximation, the total dephasing rate of thequbit is given by

Γa(t) = 〈Γcoh(t)〉 =κr

2

⟨∣∣βg(t)− βe(t)∣∣2⟩. (S36)

In the Heisenberg picture, the above equation can bewritten as

Γa(t) =κr

2〈D†(t)D(t)〉 , (S37)

in which

D(t) := bg(t)− be(t). (S38)

Consequently, the average dephasing rate defined in theMain Text is

Γa =κr

2τp

∫ τ

0

〈D†(t)D(t)〉dt (S39)

=κr

2τp

σ1,σ2

sgn[σ1σ2]

∫ τ

0

〈b†σ1(t)bσ2(t)〉dt, (S40)

in which τ is the separation between the two π/2 qubitpulses and τp is the duration of the pulsed JPC pump, asindicated in Fig. 2(a). To calculate the right-hand sideof Eq. (S39), we solve the dynamics of the readout modeoperators from Eq. (S14):

bσ(t) = bσ(0)e−(iωσr +κr2 )t

+√κr,c

∫ τ

0

Bin,ce−(iωσr +κr

2 )(t−t′) dt′.(S41)

In our experiment, before the JPC pump is applied, thereadout cavity is free from the antenna radiation,

〈b†σ1(0)bσ2

(0)〉 = 0. (S42)

Therefore,

〈b†σ1(t)bσ2(t)〉 = κr,ce

−κrtei(ωσ1r −ωσ2

r )t

×∫∫ τ

0

〈B†in,c(t′)Bin,c(t′′)〉 eκr2 (t′+t′′)ei(ω

σ2r t′′−ωσ1

r t′) dt′dt′′,

(S43)

in which

〈B†in,c(t′)Bin,c(t′′)〉 =

{nVTSL(t′, t′′), 0 < t′, t′′ < τp,

0, otherwise.

(S44)

This condition is due to the fact that the JPC is pulsed.Evaluating these integrals for t < τp yields

〈b†g(t)bg(t)〉 = N(t, κr,∆a−χ/2), (S45)

〈b†e(t)be(t)〉 = N(t, κr,∆a+χ/2), (S46)

〈b†g(t)be(t)〉 = N(t, κr−iχ,∆a), (S47)

〈b†e(t)bg(t)〉 = N(t, κr+iχ,∆a), (S48)

in which ∆a = ωa − ωr − ωp, and

N(t, κ,∆) = nVTSκaκr,cγ(1− γ)

×

(κa

κ −1)(

1−e−κt)

∆2 + (κa−κ2 )2

+

1− e(i∆−κa+κ

2 )t(∆+iκa

2

)2+ κ2

4

+ c.c.

.

(S49)

Page 15: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

7

When t > τp,

〈b†g(t)bg(t)〉 = 〈b†g(τp)bg(τp)〉 e−κr(t−τp), (S50)

〈b†e(t)be(t)〉 = 〈b†e(τp)be(τp)〉 e−κr(t−τp), (S51)

〈b†g(t)be(t)〉 = 〈b†g(τe)be(τp)〉 e−(κr−iχ)(t−τp), (S52)

〈b†e(t)bg(t)〉 = 〈b†e(τp)bg(τp)〉 e−(κr+iχ)(t−τp). (S53)

We plug these expressions into Eq. (S39) and evaluateΓa, which is linked to neff

r via Eq. (S91). Particularly,when neff

r � 1,

neffr =

χ2+κ2r

χ2κrΓa. (S54)

The dimensionless function ηa appearing in Eq. (4) andplotted in Fig. 4(b) is therefore defined as

ηa[∆a, τp] :=neff

r [∆a, τp]

nVTS

κ2r

κr,cκa, (S55)

which approaches zero when |∆a| � χ, κr. Similarly, ifwe set nVTS = 0 in Eq. (S27) and only study the qubitdephasing induced by the reflected thermal noise fromthe external bath, then we can get

neffr [∆a, τp] =

κaκr,c

κ2r

(next+nadd)(1−ηa[∆a, τp]

). (S56)

Taking the transmitted, reflected, and white parasiticnoise all into consideration, we thus obtain Eq. (4)—theresponse function of our quantum radiometer, which isthe theoretical basis for the noise analysis.

The results of this analytical approach are plotted inFig. 4(b) of the Main Text to be compared with theexperimental data. Furthermore, we show in Fig. S7for qubit-dephasing spectra (neff

r –fa relations) measuredwith different choices of JPC-pump duration τp, and plot-ted the theoretical predictions according to Eq. (4). Inthe theoretical model, the bath temperatures and tleak

are set to be their average values according to the radio-metry results shown in Table I, and we choose tloss = 0.57(within its 1-σ range: 0.52 ± 0.06) to yield the bestexperiment–theory agreement. As expected from time–frequency duality, the value of τp limits the frequencyresolution to ∼ 1/τp. With τp = 2.5 µs, the peak onthe qubit-dephasing spectrum at ∆a = −χ/2 is visiblylower than the one at ∆a = χ/2, which is due to thequbit-relaxation during the Ramsey experiment.

The main approximation made in deriving the expres-sions for the total qubit-dephasing rate is truncating thecumulant expansion after the first-order term. This ap-proximation breaks down as the number of input thermalphotons γnVTS increases. For large numbers of inputthermal photons, Eq. (S35) overestimates the dephasingrate, because the second-order correction in the cumulantexpansion is subtracted from the leading-order term. Inorder to verify our approximation, we compare the re-sults in this subsection with exact numerical solutions tothe master equation.

B. Beyond small-thermal-populationapproximation

In this subsection, we solve the master equation forcascaded systems using the P -function method.

To begin with, we write the density operator of theradiometer in the form

ρ(t) =∑

σ1,σ2=g,e

|σ1〉〈σ2| ⊗ ρσ1σ2(t), (S57)

in which |σ1,2〉 refers to the qubit state, and ρσ1σ2 is thejoint density operator of the antenna plus the readoutcavity conditioned on the qubit state.

First, we set next = nadd = 0 and study the qubit de-phasing induced by the transmitted noise from the inter-nal bath with a thermal population nVTS. The cascadedmaster equation of the system, written in the frame ro-tating at the antenna frequency ωa, is

˙ρ =1

i~

[ˆH, ρ

]+D

[√κr,cb

]ρ+D

[√κr,ib

]ρ+D

[√κa,ca

+D[√

(nVTS+1)κa,ia

]ρ+D

[√nVTSκa,ia

†]ρ

−√κa,cκr,c

([b†, aρ

]+[ρa†, b

]), (S58)

in which

ˆH

~= −

(∆a +

χ

2σz

)b†b+

ωge2σz (S59)

is the dispersive Hamiltonian in the rotating frame, and

D[A]ρ = AρA† − 1

2

{A†A, ρ

}(S60)

is the Lindblad superoperator. Moreover, we use a time-dependent coupling rate

κr,c(t) = κr,c

[Θ(0)−Θ(τp)

](S61)

to model the pulsed operation of the JPC switch, whereΘ(t) is the Heaviside step function. In our Ramsey ex-

periment, the qubit is initialized in (|g〉 + |e〉)/√

2 andthe readout cavity is initialized in vacuum. At τ = 0,we have Tr[ρge(0)] = 1/2. The information about qubitdephasing is therefore carried by ρge (or ρeg):

Aon

Aoff= 2

∣∣∣Tr[ρge(τ)

]∣∣∣ . (S62)

From Eqs. (S57) and (S58), we obtain the equation ofmotion for ρge:

˙ρge = i∆a

[b†b, ρge

]− iχ

2

{b†b, ρge

}

+D[√

κr,cb]ρge +D

[√κr,ib

]ρge +D

[√κa,ca

]ρge

+D[√

(nVTS+1)κa,ia

]ρge +D

[√nVTSκa,ia

†]ρge

−√κa,cκr,c

([b†, aρge

]+[ρgea

†, b])

, (S63)

Page 16: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

8

τp = 1.08 μs τp = 1.5 μs τp = 2.5 μsτp = 0.54 μs

FIG. S7. Qubit-dephasing spectra with different JPC-pump durations and nadd = 0. Parameters of the theoretical curves aretaken from Table I: nVTS = 1.59 (TVTS = 1.03 K), next = 0.014, nloss = 0.09, tleak = 0.046, and tloss = 0.57.

To solve Eq. (S63), we use the P -representation [S15] of

the two bosonic modes (α, α∗ ↔ a, a† and β, β∗ ↔ b, b†):

ρge(t) =

∫P (α, α∗, β, β∗, t) |α, β〉〈α, β| d2α d2β, (S64)

and write the Fokker–Planck equation [S16] for P :

∂P

∂t=

[κr + iχ

2− i∆a

]∂(βP )

∂β

+

[κr + iχ

2+ i∆a

]∂(β∗P )

∂β∗− iχ|β|2P

+κa

2

[∂(αP )

∂α+∂(α∗P )

∂α∗

]+ nVTSκa,i

∂2P

∂α∂α∗

+√κa,cκr,c

(α∂P

∂β+ α∗

∂P

∂β∗

), (S65)

whose initial condition is

P (α, α∗, β, β∗, 0) =1

πγnVTSe− |α|2γnVTS δ(β)δ(β∗). (S66)

FIG. S8. Predictions of the dimensionless detector responsefunction ηa by the approximate analytical theory comparedwith the numerical solutions of the master equations givendifferent added photon numbers. Here we set next = 0. Otherexperimental parameters are chosen to be the same as thoseused in Fig. 4 of the Main Text.

Namely, the antenna resonator is in a thermal state withaverage photon population γnVTS, and the qubit readoutcavity is in vacuum.

We write the solution for P in the form of a Gaussianansatz:

P (α, α∗, β, β∗, t)=E(t)e−A(t)|α|2−B(t)|β|2−C(t)αβ∗−D(t)βα∗,(S67)

and obtain the following ordinary differential equationsfor its coefficients:

A = κaA− nVTSκa,iA2 +√κa,cκr,c(C +D), (S68)

B = (κr + iχ)B − nVTSκa,iCD + iχ, (S69)

C =

(κa+κr+iχ

2+ i∆a − nVTSκa,iA

)C +

√κa,cκr,cB,

(S70)

D =

(κa+κr+iχ

2− i∆a − nVTSκa,iA

)D +

√κa,cκr,cB,

(S71)

E =(κa + κr + iχ− nVTSκa,iA

)E. (S72)

Note that the set of differential equations above is closedand the Gaussian ansatz is therefore exact. These equa-tions can be numerically solved with the initial conditionwritten in Eq. (S66). Finally, the qubit dephasing can becalculated by integrating the P -function,

Aon

Aoff= 2

∣∣∣∣∫P (α, α∗, β, β∗, τ) d2α d2β

∣∣∣∣ , (S73)

The detector response function ηa can be subsequentlyobtained.

Similarly, we can set nVTS = 0 and study the qubit de-phasing induced by the reflected noise from the externalbath with the thermal population next + nadd. Accordingto the reasoning in the previous subsection, this calcula-tion should yield the same ηa function.

In Fig. S8, we set next = 0 and compare the predic-tions of ηa by numerically solving the master equation atdifferent values of nadd with the outcomes of the approx-imate analytical theory introduced in Sec. V A. We find

Page 17: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

9

that these two approaches are equivalent when nadd � 1.However, when nadd & 1 where photon bunching of ther-mal radiation cannot be neglected, the predictions ofthe approximate theory, in which qubit dephasing rateis always proportional to nadd, deviate from the master-equation results.

Finally, we note that the cascaded master equationEq. (S58) is equivalent to a set of optical Bloch equations

that describes the dynamics of 〈σ+〉, 〈a†aσ+〉, 〈b†bσ+〉,〈b†aσ+〉, and 〈a†bσ+〉. In the limit of small γnVTS, thesame approximation allowing us to neglect the higher-order cumulants in Sec. V A leaves these optical Bloch-equations to be a closed set of differential equations,which can be analytical solved using the Laplace trans-form in the limit of large χ. For the conditions of ourexperiment, the predictions of optical Bloch equationsquantitatively agree with those of the two approaches in-troduced in this section.

VI. RADIOMETER CALIBRATION

We calibrated our radiometer by measuring qubit-dephasing spectra while varying the thermal populationsof the external and internal antenna baths. To under-stand the procedure, we write Eq. (4) again,

neffr [∆a, τp] = κ−2

r κaκr,c

{nloss(1−tloss) + nVTStleaktloss

+ nVTStlossηa[∆a, τp] +(next+nadd)tloss

(1−ηa[∆a, τp]

)},

(S74)

in which ∆a, τp, nadd, and nVTS are tunable parametersin our experiment, while next, nloss, tloss, and tleak areto be extracted. We chose τp = 1.08 µs in the followingmeasurements. When |∆a| � χ, Eq. (S74) is reduced to

(a) (b)

FIG. S9. Experimental data and linear fits of neffr at different

(a) VTS temperatures and (b) added white-noise powers. In(a), Tadd = 0 (nadd = 0). In (b), TVTS = 1.03 K (nVTS =1.59). For all data in this figure, τp = 1.08 µs.

Eqs. (S5)–(S8) on white-noise-induced qubit dephasing,

neffr [∞] =

κaκr,c

κ2r

{nloss(1− tloss)

+ nVTStleaktloss + (next + nadd) tloss

}.

(S75)

Now we present our calibration protocol in three steps:(A) ηa[∆a]: We performed the Ramsey experiment and

measured neffr while varying nadd and fixing nVTS = 1.59

(TVTS = 1.03 K). Plotted in Fig. S9(a) are some rep-resentative data at three different values of ∆a, corre-sponding to the top, ridge, and floor of the right qubit-dephasing peak in Fig. S7. After normalizing the slopesof these neff

r –nadd lines with respect to the average slopeof those lines at |∆a| � χ, κr, we obtain the experimen-tal data of 1 − ηa[∆a] shown in Fig. 4(b), which is wellexplained by the theory in Sec. V.

(B) tloss and tleak: We measured neffr while varying

nVTS and fixing nadd = 0, and fitted the data using thelines neff

r = λnVTS + µ. As examples, data at the threeselected values of ∆a are plotted and fitted in Fig. S9(b).The slopes of the lines are equal to

λ[∆a] = λleak + λa[∆a]|∆a|�χ, κr−−−−−−−→ λleak, (S76)

in which

λleak := κ−2r κaκr,ctlosstleak, (S77)

λa[∆a] := κ−2r κaκr,ctlossηa[∆a]. (S78)

With κr, κr,c, and κa reported in Table S1, and ηa[∆a]measured in Step (A), we can deduce tloss from thefrequency-dependent λa[∆a], and subsequently obtaintleak from the frequency-independent λleak.

(C) next and nloss: The vertical intercepts of the neffr –

nVTS lines measured in Step (B) are equal to

µ[∆a] = µloss + µa[∆a]|∆a|�χ, κr−−−−−−−→ µloss, (S79)

in which

µloss := κ−2r κaκr,c

[nloss(1−tloss) + nexttloss

], (S80)

µa[∆a] := −κ−2r κaκr,cnexttlossηa[∆a]. (S81)

Therefore, with all the known parameters, we can extractnext from the data of µa[∆a], and then nloss from µloss.Results are listed in Table I.

VII. CLASSICAL RADIOMETRY DATA

Raw spectral density data underlying Fig. 3 are shownin Fig. S10. The bright traces were measured withfa = 10.519 GHz. The shadowed traces were takenwith the antenna resonator tuned away from the de-tection window. Their differences account for the dipsand peaks seen in Fig. 3, which manifest the radiativeheating and cooling effects, respectively. The floors of

Page 18: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

10

10.51910.51810.517 10.520 10.521

FIG. S10. Raw spectral density data underlying Fig. 3. Theresolution bandwidth of the spectrum analyzer was set to be1 kHz.

the spectra in Fig. 3 are acquired by averaging the dis-tances between the shadowed background traces here inFig. S10, with the trace at nadd = 0 being the reference.With the external noise generator calibrated through thewhite-noise-induced dephasing experiment (see Sec. IV),we can thus map Pout onto nout

a —the photon flux per unitbandwidth at the output of the antenna resonator. Thegain-saturation effect of the JPC amplifier is consideredwhen we convert the vertical axes.

This classical radiometry method is also capable ofmeasuring next, and therefore, na—the thermal popu-lation of the antenna mode. To achieve this, we shouldmeasure the power spectral density at Probe out whilevarying nVTS, similar to Step (B) in Sec. VI, and cali-brate the total gain of the amplifier chain. Results willbe reported in an accompanying article [S17].

VIII. COMPARISON WITHSINGLE-MICROWAVE-PHOTON DETECTORS

In the simplified linear model, a single-photon detectoris characterized by two important parameters—quantumefficiency η and dark-count rate Rdc. Assuming thePoisson counting statistics, the probability of getting a“click” event after a time interval τp is

Pclick(τp) = 1− e−(ηR0+Rdc)τp−P 0dc , (S82)

in which R0τp is the probability of the incoming fieldhaving nonzero photons during τp, and P 0

dc is the para-sitic dark-count probability. In our case, this parametermodels qubit initialization and readout errors as well asintrinsic qubit decoherence. Therefore, the total dark-count probability within the same detection window isgiven by Pdc = Rdcτp + P 0

dc.

Define the number of clicks during τp as

Nclick := ln

(1

1− Pclick

). (S83)

Then the quantum efficiency and the dark-count proba-bility can be formally defined as

η :=1

τp

dNclick

dR0

∣∣∣∣∣R0=0

, (S84)

Pdc :=Nclick|R0=0 . (S85)

The Ramsey experiment combined with single-shotqubit readout can be understood from the point of viewof photon detection. After each measurement sequence,the qubit readout yielding |g〉 or |e〉 corresponds to the“click” or “no click” event of a single-photon detector.Following this definition and the introduction to our mea-surement protocol in Sec. IV, we find for those sequencesduring which the JPC switch is on for time τp, the prob-ability of getting a “click” at the end of a sequence isequal to

Pclick(τp) =1

2−Aon(τp), (S86)

in which Aon is the amplitude of Ramsey oscillations. Ifthe antenna radiation is white noise, then

Pclick(τp) =1

2− 1

2e−Γthτp−Γ2R(τ−τp), (S87)

Γth is given by Eq. (2) in the Main Text.In reality, the contrast of Ramsey oscillations is smaller

than unity even when τ = τp = 0. Therefore, the expo-nential term in the above equation needs to be multipliedby an initial contrast factor A0:

Pclick(τp) =1

2− 1

2A0e

−Γthτp−Γ2Rτw . (S88)

in which τw = τ − τp is the waiting time between theend of the JPC pump and the second π/2 qubit pulse.In our experiment, A0 < 1 is mainly due to the readoutinfidelity and the imperfect qubit initialization, whichcause the probability of getting a |g〉 or |e〉 outcome whenτ = 0 to be

P 0g = P ini

g Pr(g|g) + P inie Pr(g|e), (S89)

P 0e = P ini

e Pr(e|e) + P inig Pr(e|g). (S90)

Here P inie = 1 − P ini

g = 0.03 denotes the qubit excited-state population after its initialization; Pr(e|g) = 1 −Pr(g|g) = 0.01 and Pr(g|e) = 1 − Pr(e|e) = 0.04 denotethe qubit readout infidelity. These numbers yield P 0

e =1− P 0

g = 0.0385, and A0 = P 0g − P 0

e = 0.923.In the limit of small incoming photon flux,

Γthnth

r �1−−−−→ χ2κrnthr

χ2 + κ2r

=χ2R0

χ2 + κ2r

. (S91)

Page 19: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

11

0.0 0.2 0.4 0.6 0.8 1.0R0τp

0.0

0.2

0.4

0.5N

clic

k

FIG. S11. “Number of clicks” of the Ramsey-sequence-basedqubit radiometer viewed as a photon detector. Experimentalparameters are taken from Fig. 4 in the Main Text.

Therefore, we obtain the quantum efficiency and thedark-count probability of our cQED detector based onthe Ramsey sequence

η =A0e

−Γ2Rτw

1 +A0e−Γ2Rτw

χ2

χ2 + κ2r

, (S92)

Pdc = ln

(2

1 +A0e−Γ2Rτw

). (S93)

Plugging in the experimental parameters (τp = 1.08 µsand τ = 2.08 µs), we get η = 0.44 and Pdc = 0.059, whichare comparable to the state-of-the-art single-microwave-photon detectors reported in recent years [S18–S23].

In Fig. S11, we plot Nclick as a function of R0τp usingthe same experimental parameters as Fig. 4 in the MainText. The quantum efficiency is shown as the slope atR0 → 0. Note that according to Eq. (S92), the efficiencyis limited to η < 0.5. This factor-of-two sacrifice arisesbecause a completely dephased qubit has equal probabil-ities of being in |g〉 and |e〉 after a Ramsey sequence. Thedark-count probability is represented by the vertical in-tercept. Here we remark that the Pdc = 0.059 reported inthe present experiment is limited by the qubit-state ini-tialization and the single-shot readout as well as the finitequbit T2, all of which can be improved with available ex-perimental techniques. For instance, measurement-basedqubit initialization plus a readout with fidelity 0.99 [S24]yields Pdc = 0.026 with the present value of T2R. Extend-ing T2R to 100 µs [S25], one can then expect a dark-countprobability Pdc = 0.010, which is close to the record valuereported in Ref. [S20]. In addition, the signal and pumpof our qubit-dephasing radiometer require no pulse shap-ing, which results in simpler operations and more flexi-bility in practical tasks.

The above comparison presents our qubit-dephasingradiometer as another candidate setup for microwave-cavity-based axion probes. As has been reasoned inRefs. [S26] and [S27], single-photon detectors show ad-vantages over quantum-limited linear amplifiers in searchof dark-matter axions at high frequencies and low tem-

peratures. These advantages are also shared by our qubitradiometer, which essentially has the similar photon-counting performance. Furthermore, the Peccei–Quinnmechanism [S28] predicts that the galactic halo dark-matter axions induce a steady-state microwave field,which resembles the antenna radiation in this work and isthus suited for the radiometry protocol based on photon-induced qubit dephasing. Practical challenges includeshielding the superconducting circuit from the strongmagnetic field, reducing the residual photon populationin the qubit-readout cavity [S4], and improving the sta-bilities of qubit coherence times.

IX. PRECISION ANALYSIS

In this section, we will compare the precisionsof radiometer models based on three different phys-ical principles—linear amplification, photon counting,and photon-induced qubit dephasing exploited in thisarticle—given a certain detection bandwidth and inte-gration time. The radiator is chosen to be a broadbandthermal source. The advantage of our quantum radio-meter over classical ones is quantitatively shown.

In the classical regime (hf � kBT ), the precision of atotal-power radiometer based on linear amplification plussquare-law detection is given by [S29]

δTlin

T linsys

=

√2π

Bτint+

(δG

G

)2

, (S94)

in which T linsys is the system noise temperature; τint is the

integration time; B is the detection bandwidth in theunit of rad/s; G and δG are the receiver gain and itsfluctuation. The precision of a Dicke radiometer has thesimilar expression but a different prefactor depending onthe specific type of modulation [S30]. In the quantumregime, if the gain fluctuations are neglected, the aboveequation can be rewritten using average photon numbers,

δnlin

nlinsys

=

√2π

Bτint. (S95)

Here the precision δnlin is defined as [S30]

δnlin :=σI

dI/dn, (S96)

in which I and σI are the average and the standard de-viation of the detector output signal, and n is the ther-mal population of the source under detection. For anideal quantum-limited phase-preserving linear amplifier,nlin∗

sys = 1, which comprises both the amplified vacuumfluctuations accompanying the input signal—1/2—andthe minimum possible noise added by the amplifier—another 1/2.

Now consider a single-photon detector with a quantumefficiency η and a dark-count probability Pdc within a de-tection time window τp. In the absence of input photons,

Page 20: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

12

the probability of getting a “click” within τp is Pdc, withthe standard deviation after N counts being

σI =

√Pdc(1− Pdc)

N(S97)

according to the binomial-distribution formula. Given asmall incoming photon flux that results in a probabilitydP0 of having nonzero photons at the detector input, theprobability of getting a “click” within τp is then Pdc +η dP0. The detector “signal” in this situation is thus

dI = (Pdc + η dP0)− Pdc = η dP0. (S98)

Following Eq. (S96), the precision δP0 is given by

δP0 =σI

dI/dP0

=1

η

√Pdc(1− Pdc)

N. (S99)

In the previous section, we have demonstrated that aqubit-dephasing radiometer can be viewed as a single-photon counter with an effective quantum efficiency anda dark-count rate [Eqs. (S92) and (S93)]. For the qubit-dephasing radiometer introduced in our experiment, wehave

dP0 = τp dR0 = κrτp dnthr . (S100)

Then the detection precision of nthr is

δnqu =σI

dI/dnthr

=1

ηκrτp

√Pdc(1− Pdc)

N. (S101)

Relating Eq. (S95) to Eq. (S101), we find the conditionfor the linear-amplifier-based and the qubit-dephasing-based radiometers to have the same precision, namely,δnlin = δnqu:

(nlinsys)

2

B/2π=Pdc(1− Pdc)

η2κ2r τp

, (S102)

in which we have assigned Nτp to be the effective integra-tion time τint for the dephasing radiometer. Therefore,for the two radiometers to have the same precision, thelinear-amplifier system should have a bandwidth

B =2πκrτp(nlin

sys)2η2

Pdc(1− Pdc)κr. (S103)

Equivalently, if the linear-amplifier system has the samebandwidth as the qubit-readout cavity, namely, B = κr,

then the qubit-dephasing radiometer outperforms by afactor of

δnlin

δnqu= nlin

sysη

√2πκrτp

Pdc(1− Pdc)(S104)

in terms of detection precision. Considering nlin∗sys = 1

for an ideal phase-preserving linear amplifier and takingin the values η = 0.44 and Pdc = 0.059 computed inSec. VIII, we obtain δn∗lin/δnqu = 11.

As is shown in Fig. 4(c) of the Main Text, when theJPC pump is on, the VTS thermal leakage and the dis-sipation between the antenna and the cQED cavity con-tribute to a parasitic white thermal background of npara

r

photon on average in the qubit-readout cavity. As a re-sult, the quantum efficiency and the dark-count rate ofthe whole setup are given by

η′ =A0e

−Γ2Rτw−Γth(nparar )τp

1 +A0e−Γ2Rτw−Γth(nparar )τp

χ2

χ2 + κ2r

, (S105)

P ′dc = ln

[2

1 +A0e−Γ2Rτw−Γth(nparar )τp

]. (S106)

In our experimental setup, we calibrated that nparar =

nparatlossκaκr,c/κ2r = 0.035, which yields η′ = 0.40, P ′dc =

0.14, and an outperforming factor of δn∗lin/δn′qu = 6.8

compared to an ideal quantum-limited amplifier. Usingthe calibrated value of nlin

sys = 1.54 of the classical ampli-fier chain (Probe out) [S17], we obtain δnlin/δn

′qu = 10.

X. SYSTEM NOISE

The single-photon-detector model of the cQED modulealso provides an intuitive way for us to understand thesystem noise of our qubit-dephasing radiometer. RewriteEq. (S106) as

P ′dc = ln

[2

1 + e−Γth(nsysr )τp

], (S107)

in which

χ2κrτpnsysr

χ2 + κ2r

=χ2κrτpn

parar

χ2 + κ2r

+ Γ2Rτw − lnA0. (S108)

Therefore, the system noise of our qubit radiometer,which includes the VTS thermal leakage, the dissipa-tion between the antenna and the cQED module, andthe qubit-decoherence shot noise, when referred to theantenna input, is given by

nsys = nsysr

κ2r

κr,cκatloss= npara + nshot, (S109)

in which

nshot =(χ2 + κ2

r )κr

χ2κr,cκatloss

Γ2Rτw − lnA0

τp. (S110)

In our experiment, nsys = 0.25±0.02—the value reportedin Table I of the Main Text.

Page 21: Quantum Microwave Radiometry with a Superconducting Qubit · 2019-09-30 · A classical microwave radiometer [11] typically consists of a radio receiver followed by a square-law detector

13

[S1] C. Rigetti, Quantum Gates for Superconducting Qubits,PhD dissertation, Yale University (2009).

[S2] F. Lecocq, I. M. Pop, Z. Peng, I. Matei, T. Crozes,T. Fournier, C. Naud, W. Guichard, and O. Buisson,“Junction fabrication by shadow evaporation without asuspended bridge,” Nanotechnology 22, 315302 (2011).

[S3] K. Serniak, S. Diamond, M. Hays, V. Fatemi,S. Shankar, L. Frunzio, R.J. Schoelkopf, and M.H. De-voret, “Direct dispersive monitoring of charge parity inoffset-charge-sensitive transmons,” Phys. Rev. Applied12, 014052 (2019).

[S4] Z. Wang, S. Shankar, Z. K. Minev, P. Campagne-Ibarcq,A. Narla, and M. H. Devoret, “Cavity attenuators forsuperconducting qubits,” Phys. Rev. Appl. 11, 014031(2019).

[S5] N. E. Frattini, V. V. Sivak, A. Lingenfelter, S. Shankar,and M. H. Devoret, “Optimizing the nonlinearity anddissipation of a snail parametric amplifier for dynamicrange,” Phys. Rev. Applied 10, 054020 (2018).

[S6] V. V. Sivak, N. E. Frattini, V. R. Joshi, A. Lingenfelter,S. Shankar, and M. H. Devoret, “Kerr-free three-wavemixing in superconducting quantum circuits,” Phys.Rev. Applied 11, 054060 (2019).

[S7] M. Xu, X. Han, W. Fu, C.-L. Zou, M. H. Devoret, andH. X. Tang, “Frequency-tunable high-q superconduct-ing resonators via wireless control of nonlinear kineticinductance,” Appl. Phys. Lett. 114, 192601 (2019).

[S8] E. Flurin, The Josephson mixer: a swiss army knife

for microwave quantum optics, PhD dissertation, Ecolenormale superieure (2014).

[S9] K. Sliwa, Improving the Quality of Heisenberg Back-Action of Qubit Measurements made with ParametricAmplifiers, PhD dissertation, Yale University (2016).

[S10] B. Abdo, K. Sliwa, F. Schackert, N. Bergeal, M. Ha-tridge, L. Frunzio, A. D. Stone, and M. Devoret, “Fullcoherent frequency conversion between two propagat-ing microwave modes,” Phys. Rev. Lett. 110, 173902(2013).

[S11] J. Gambetta, A. Blais, D. I. Schuster, A. Wallraff,L. Frunzio, J. Majer, M. H. Devoret, S. M. Girvin, andR. J. Schoelkopf, “Qubit–photon interactions in a cav-ity: Measurement-induced dephasing and number split-ting,” Phys. Rev. A 74, 042318 (2006).

[S12] A. A. Clerk and D. W. Utami, “Using a qubit to mea-sure photon-number statistics of a driven thermal oscil-lator,” Phys. Rev. A 75, 042302 (2007).

[S13] M. Silveri, E. Zalys-Geller, M. Hatridge, Z. Leghtas,M. H. Devoret, and S. M. Girvin, “Theory of remote en-tanglement via quantum-limited phase-preserving am-plification,” Phys. Rev. A 93, 062310 (2016).

[S14] A. A. Clerk, M. H. Devoret, S. M. Girvin, F. Marquardt,and R. J. Schoelkopf, “Introduction to quantum noise,measurement, and amplification,” Rev. Mod. Phys. 82,1155 (2010).

[S15] D. F. Walls and G. J. Milburn, Quantum Optics, 2nded. (Springer, Berlin and Heidelberg, 2008).

[S16] M. O. Scully and M. S. Zubairy, Quantum Optics (Cam-bridge University Press, Cambridge, 1997).

[S17] M. Xu, et al., in preparation.[S18] K. Inomata, Z. Lin, K. Koshino, W. D. Oliver, J.-

S. Tsai, T. Yamamoto, and Y. Nakamura, “Singlemicrowave-photon detector using an artificial L-typethree-level system,” Nat. Comm. 7, 12303 (2016).

[S19] S. Kono, K. Koshino, Y. Tabuchi, A. Noguchi, andY. Nakamura, “Quantum non-demolition detection ofan itinerant microwave photon,” Nat. Phys. 14, 546(2018).

[S20] R. Lescanne, S. Deleglise, E. Albertinale, U. Reglade,T. Capelle, E. Ivanov, T. Jacqmin, Z. Leghtas, andE. Flurin, “Detecting itinerant microwave photons withengineered non-linear dissipation,” arXiv:1902.05102.

[S21] A. Narla, S. Shankar, M. Hatridge, Z. Leghtas, K. M.Sliwa, E. Zalys-Geller, S. O. Mundhada, W. Pfaff,L. Frunzio, R. J. Schoelkopf, and M. H. Devoret, “Ro-bust concurrent remote entanglement between two su-perconducting qubits,” Phys. Rev. X 6, 031036 (2016).

[S22] P. Campagne-Ibarcq, E. Zalys-Geller, A. Narla,S. Shankar, P. Reinhold, L. Burkhart, C. Axline,W. Pfaff, L. Frunzio, R. J. Schoelkopf, and M. H. De-voret, “Deterministic remote entanglement of supercon-ducting circuits through microwave two-photon transi-tions,” Phys. Rev. Lett. 120, 200501 (2018).

[S23] J.-C. Besse, S. Gasparinetti, M. C. Collodo, T. Walter,P. Kurpiers, M. Pechal, C. Eichler, and A. Wallraff,“Single-shot quantum nondemolition detection of indi-vidual itinerant microwave photons,” Phys. Rev. X 8,021003 (2018).

[S24] A. Narla, Flying Qubit Operations in SuperconductingCircuits, PhD dissertation, Yale University (2017).

[S25] C. Rigetti, J. M. Gambetta, S. Poletto, B. L. T.Plourde, J. M. Chow, A. D. Corcoles, J. A. Smolin, S. T.Merkel, J. R. Rozen, G. A. Keefe, M. B. Rothwell, M. B.Ketchen, and M. Steffen, “Superconducting qubit in awaveguide cavity with a coherence time approaching 0.1ms,” Phys. Rev. B 86, 100506 (2012).

[S26] S. K. Lamoreaux, K. A. van Bibber, K. W. Lehnert,and G. Carosi, “Analysis of single-photon and linearamplifier detectors for microwave cavity dark matteraxion searches,” Phys. Rev. D 88, 035020 (2013).

[S27] L. Zhong, S. Al Kenany, K. M. Backes, B. M. Brubaker,S. B. Cahn, G. Carosi, Y. V. Gurevich, W. F. Kindel,S. K. Lamoreaux, K. W. Lehnert, et al., “Results fromphase 1 of the HAYSTAC microwave cavity axion ex-periment,” Phys. Rev. D 97, 092001 (2018).

[S28] R. D. Peccei and Helen R. Quinn, “CP conservation inthe presence of pseudoparticles,” Phys. Rev. Lett. 38,1440 (1977).

[S29] M. S. Hersman and G. A. Poe, “Sensitivity of the totalpower radiometer with periodic absolute calibration,”IEEE Trans. Microw. Theory Tech. 29, 32 (1981).

[S30] D. F. Wait, “The sensitivity of the Dicke radiometer,”J. Res. Natl. Bur. Stand. 71C, 127 (1967).