<<

Cavity-assisted measurement and coherent control of collective atomic spin oscillators

Jonathan Kohler,1, ∗ Nicolas Spethmann,1, 2 Sydney Schreppler,1 and Dan M. Stamper-Kurn1, 3, † 1Department of , University of California, Berkeley, CA 94720, USA 2Department of Physics and Research Center OPTIMAS, Technische Universit¨atKaiserslautern, 67663 Kaiserslautern, Germany 3Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA (Dated: January 12, 2017) We demonstrate continuous measurement and coherent control of the collective spin of an atomic ensemble undergoing Larmor in a high-finesse optical cavity. The coupling of the pre- cessing spin to the cavity field yields phenomena similar to those observed in cavity optomechanics, including cavity amplification, damping, and optical spring shifts. These effects arise from au- tonomous optical feedback onto the atomic spin dynamics, conditioned by the cavity spectrum. We use this feedback to stabilize the spin in either its high- or low-energy state, where, in equilib- rium with measurement back-action heating, it achieves a steady-state temperature, indicated by an asymmetry between the Stokes and anti-Stokes scattering rates. For sufficiently large Larmor frequency, such feedback stabilizes the spin ensemble in a nearly pure quantum state, in spite of continuous measurement by the cavity field.

PACS numbers: 42.50.Pq, 42.50.Lc, 76.70.Hb, 42.65.Dr

Quantum systems are invariably perturbed by mea- tions a feedback loop that allows the accumulated noise surement. For a repeated or continuous measurement, from measurement back-action to be suppressed and the the measurement back-action perturbing the system gen- collective spin to be stabilized near either its lowest- or erally adds noise to the subsequent measurement record, highest-energy state, chosen by the detuning of the cav- reducing the precision of state estimation and measure- ity probe light from the cavity resonance. We find that, ment sensitivity of external forces [1]. However, this for sufficiently large Larmor frequency, such a feedback- additional noise carries useful information about the stabilized spin oscillator remains in a nearly pure quan- measurement-induced perturbation, allowing for feed- tum state, in spite of continuous interaction with the back that suppresses the effects of back-action and con- probe field. trols the evolution of the quantum system [2]. Such Consider an ensemble of N identical in their feedback may be either measurement-based, where the electronic ground state, each with spin f. The dimen- quantum system is driven externally based on classical sionless collective spin, Fˆ, is sensed optically via the cir- information gleaned from quantum measurements [3–8], cular birefringence of the atoms. Specifically, a beam of or autonomous, where feedback is implemented through circularly polarized light propagating along the z axis ac- coherent [9] or dissipative [10–12] dynamics inherent to quires a phase shift proportional to Fˆz, the projection of the system itself. the collective spin onto the axis of optical angular mo- Atomic spin ensembles in cavity quantum electrody- mentum. namic systems have been demonstrated as prototypical An applied magnetic field B induces the Hamiltonian examples of quantum measurement and control. Many ˆ B = ~γB F, where γ is the atomic gyromagnetic ra- recent experiments with spin-cavity systems have focused Htio. When− B·is oriented away from z – say, along x (see on avoiding back-action to overcome standard quantum Fig. 1a) – the light measures oscillations of one trans- limits [13] and to prepare squeezed states for improved verse spin component, allowing real-time observation of metrology through quantum non-demolition [14–18] or Larmor precession of the collective spin. The light acts back-action evading [19] measurements. Here, we focus back on the precessing spin as an effective magnetic field instead on continuous weak detection of non-stationary Bopt z, proportional to the probe light intensity. Mea- k arXiv:1607.07941v2 [physics.-ph] 11 Jan 2017 observables in order to study quantum-limited back- surement back-action arises from quantum fluctuations action and autonomous stabilization effects using the of the light intensity, and, hence, of Bopt. The resulting cavity’s finite lifetime. fluctuations in the produced by Bopt affect both In this Letter, we report the observation of back-action the phase of Larmor precession and the energy of the spin effects from weak continuous measurement of the Larmor ensemble. precession of an atomic spin ensemble within a driven Inside a high-finesse cavity, optical fluctuations persist high-finesse optical cavity. By recirculating the light for a finite life-time, allowing them to coherently alter probing the spin ensemble, the cavity allows the optical the subsequent spin dynamics. In particular, when the modulation induced by the precessing spin to act back on cavity is driven off-resonance, the precessing spin induces the subsequent spin dynamics. The cavity thereby condi- amplitude modulation of the probe field. This amplitude 2

Fˆ B nˆ(t) modulation acts back on the precessing spin through the a) b) z opt | ∝ B effective magnetic field Bopt, which varies synchronously opt | (ar b. ) with the Larmor precession. The finite cavity lifetime + (ar b. )

σ z ˆ leads to a delay between the oscillating spin displace- F ˆ t ment Fz(t) and the modulation of the intra-cavity photon B numbern ˆ(t) (Fig. 1b). c) F The modulation of Bopt has two effects, which become clear when considered in a frame co-rotating with the τ x z precessing spin around B [20]. Under the rotating-wave z F approximation, the modulation component in-phase with y y z Bopt x y the oscillation of Fˆz generates a net torque tangential to the instantaneous trajectory of the precessing spin (Fig. FIG. 1. (color online) (a) Schematic of the experimental 1c), shifting the effective Larmor frequency. The out-of- system. Atoms trapped at an anti-node of the cavity field phase component generates an average torque acting per- experience an effective magnetic field Bopt due to interaction pendicular to this trajectory, causing the atomic spin to with circularly polarized light. A large external magnetic field nutate toward either the low-energy or the high-energy B along x defines the Larmor frequency ωL and high- and pole, depending on the sign of the optical modulation. low-energy stable poles of the dynamics. (b) The oscillat- ˆ The relative phase between the spin precession and the ing transverse spin component Fz couples to the cavity field, cavity field modulation is controlled by the probe detun- causing amplitude modulation for a cavity driven either above ing from cavity resonance ∆ , and by the ratio of ω to (blue) or below (red) resonance. The optical response is de- pc L layed due to the finite cavity linewidth κ. (c) The average the cavity half-linewidth κ = 2π 1.82 MHz. × torque τ acting on the spin due to this optical feedback, in This Letter extends concepts from cavity optomechan- a rotating frame. The in-phase modulation generates torque ics [21] to the cavity-based detection and control of col- tangential to the spin trajectory, shifting the Larmor preces- lective spin modes, which was discussed theoretically in sion frequency. The out-of-phase component generates torque Ref. [20], and also recently in terms of solid-state cavity perpendicular to this trajectory, causing spin nutation toward opto-magnonics [22–24]. The autonomous feedback driv- or away from the poles. ing the collective spin toward or away from either the low- or high-energy states is analogous to cavity-induced ferent TEM mode of the cavity, whose resonance fre- damping [25–28] or amplification [29] of motion. We also 00 quency ω is detuned by ∆ /2π = 42 GHz from the observe feedback-induced shifts in the Larmor frequency c ca 87Rb D2 transition (with wavelength− 780 nm). At this that correspond to the optical spring effect [30, 31]. Fur- large detuning, the atom-cavity interaction is predom- thermore, similar to experiments in cavity optomechanics inately dispersive. Because the detuning is also much [32, 33], we observe asymmetry between sidebands gener- larger than the excited-state hyperfine splitting, the in- ated by optical Stokes and anti-Stokes processes, allowing teraction is dominated by scalar and vector terms, and us to characterize the effective spin temperature reached higher-order tensor interactions are negligible. by the balance of coherent and incoherent back-action The cavity supports modes of two independent polar- effects. izations at frequencies near ω . Driving the cavity with a Our experiments are performed on gases of about c weak coherent beam of one circular polarization of light, N = 3500 87Rb atoms placed within a Fabry-P´erotcav- with helicity along either z, and neglecting the small ity [34]. The atoms are evaporatively cooled to around 5 linear birefringence of our± cavity [35], the resulting evo- µK and trapped within a couple adjacent sites of a 64 µK lution is described by the effective system Hamiltonian deep, one-dimensional, spin-independent optical lattice, created by driving a TEM00 mode of the cavity with light 2 † ˆ ~g0 † ˆ at a wavelength of 860 nm. A uniform magnetic field B = ~ωcaˆ aˆ + ~ωLFx + aˆ aˆ(α0N α1Fz), (1) H ∆ ± is applied along x, transverse to the cavity axis (Fig. 1a), ca and its magnitude sets the Larmor frequency ωL = γB , obtained by adiabatically eliminating the atomic excited | | which we vary from ωL/2π = 100 kHz to several MHz. states. Here,a ˆ is the creation operator for cavity pho- The atoms are prepared initially in the f = 2, mF = +2 tons, and the coefficients α0 = 2/3 and α1 = 1/6 describe hyperfine state (with the quantization axis| along B). Thei the relative strength of the scalar and vector parts of the collective atomic spin, with total spin F = 2N 7000, is ac Stark shift, determined by summing over all excited- thus prepared in its highest energy state, since∼γ < 0 for state hyperfine levels. The position-averaged vacuum 87 the f = 2 hyperfine level of Rb. Alternately, the en- Rabi coupling g0/2π = 13 MHz is evaluated by consider- semble can be rotated to its lowest-energy spin state by ing the geometry of the cavity mode [34]. In this work, applying a π pulse to the atoms using a radio-frequency atoms are trapped near antinodes of the cavity probe magnetic field. field to minimize linear optomechanical effects. The cou- We probe the ensemble through its influence on a dif- pling of the cavity field to the environment is further de- 3

2000 a) iniscent of optical pumping [36]. However, unlike optical

1000 pumping, the dynamics in our experiment cannot spon- i

z taneously generate spin polarization. In addition, while

^ 0 F h optical pumping uses circularly polarized light to pump -1000 into an atomic gas, the asymmetric -2000 fluctuation spectrum of the cavity optical field is used to 2000 b) pump energy into the atomic system. Indeed, we confirm

1000 that the dynamics are quantitatively the same for both i

z circular probe polarizations (Fig. 3a-b).

^ 0 F h These dynamics may also be described in terms of cav- -1000 ity superradiance [37–41]. Consider an atomic spin en- -2000 semble initialized in the low-energy spin state. The op- 0 0.05 0.1 0.15 0.2 0.25 0.3 tically driven atoms lie in a virtually excited state from Time (ms) which they may decay by Raman scattering into the cav- ity mode. When the cavity is driven at a positive de- FIG. 2. (color online) Coherent damping and amplification tuning, the cavity Purcell effect induces Raman emission of Larmor precession of a spin ensemble with ωL/2π = 1.0 preferentially on the Stokes sideband, creating transverse MHz, observed in the phase modulation of transmitted light, averaged over 30-40 repetitions (blue). Cavity probe light (∆pc/2π = 1.0 MHz,n ¯ = 4 average intracavity photons)

drives the spin toward the high-energy pole. (a) Larmor pre- ) a) z 5

cession of a spin ensemble, displaced from the high-energy H k pole by a π/10 rf pulse, coherently damps back to the pole (

: 0

at a rate Γopt/2π = 4.9 ± 0.2 kHz. (b) A spin prepared near 2

= t

the low-energy pole, by application of a near π-pulse, is co- p o -5 herently amplified away at a rate Γopt/2π = −4.6 ± 0.4 kHz. !

Exponential rates are extracted by simultaneous fits (red) of )

both amplitude and phase quadratures. Insets show the har- z 5 b) H

monic nature of the Larmor precession signal and quality of fit k (

in the highlighted regions. The finite cavity linewidth causes : 0 2

the observed signal to saturate at around 2000. =

t

p o

! -5 / scribed using the standard input-output formalism, and -2 -1 0 1 2 the modulated probe field transmitted through the cav- "pc=2: (MHz) ity is measured with an optical heterodyne detector, with

) 8 c)

overall cavity photon detection efficiency  = 0.12. z H

The spin dynamics imprinted on the cavity output k 6 (

field are observed in the demodulated heterodyne sig- : 4

2 =

nal. For example, in Fig. 2, we compare the evolution of t p

o 2 spins prepared near either the high- or low-energy poles ! when the cavity is driven by a blue-detuned (∆pc > 0) 0 0 0.5 1 1.5 2 probe. In both cases, the probe drives the spin toward the ! =2: (MHz) high-energy pole. For a spin prepared initially near the L high-energy pole, cavity back-action coherently damps FIG. 3. (color online) (a) Optical damping rates and (b) the Larmor precession amplitude, analogous to cavity frequency shifts of Larmor precession as a function of probe optomechanical cooling. In comparison, the Larmor pre- detuning ∆pc, with fixed intracavity intensityn ¯ = 4 and cession of a spin prepared near the low-energy pole is ωL/2π = 1.0 MHz. Diamonds (blue) label results for an en- coherently amplified, analogous to regenerative optome- semble initially prepared near the high-energy pole, and cir- chanical amplification. At longer times (not shown in cles (red) for an ensemble initially near the low-energy pole. the Figure), in the latter case, the ensemble’s spin nu- Measurements repeated with either σ+ (solid symbols) or σ− tates past the equator of the Bloch sphere and also damps (open symbols) circularly polarized light demonstrate inde- pendence of optical helicity. (c) Peak damping rate as a back to the high-energy pole. If instead we drive the cav- function of Larmor frequency. Dotted vertical lines mark the ity with red-detuned probe light (∆pc < 0), we observe position of the cavity half-linewidth κ. All theory lines are similar behavior, with the collective spin instead driven plotted with no free parameters. Error bars reflect statistical toward and stabilized at the low-energy pole. uncertainties from the fits. Additional systematic errors in The light-induced driving of a spin ensemble to either the probe frequency stability and initial spin state prepara- tion predominately affect data at small probe detuning. the low- (∆pc < 0) or high-energy (∆pc > 0) pole is rem-

4 )

coherence in the ensemble. Such coherence stimulates . 1.4 N . a)

Raman scattering at an enhanced rate, driving the spins S f

exponentially away from the low-energy pole. o 1.2

s t A quantitative treatment for the coherent dynamics i

n 1 U

near both poles can be derived classically, as in the the- (

ory of linear cavity optomechanics [21], by neglecting the D S

quantum noise of the cavity field. We find that the in- P 1.4 b) e

verse susceptibility of the spin oscillator to torque τ(ω) n y

−1 2 2 d 1.2

at frequency ω is given by χs (ω) = ωL ω + Σ(ω), o r

− e

with t 1

e H 2 ωL ωL  Σ(ω) = 2βng¯ s , (2) -920 -900 -880 880 900 920 ω ∆pc + iκ − ω + ∆pc + iκ − Frequency (kHz) where β = sgn(Fˆx) labels the high-energy (β = +1) or -0.5 c)

low-energy (β = 1) state,n ¯ is the average photon num- -1 0.5 i − 2p L 1 ber in the cavity mode, and gs = α1g F/2/∆ca de- x

0 !

^

F

h 7

scribes the relevant single-photon, single-excitation cou- = !

T -5 B

pling [20]. The real and imaginary parts of Eq. 2 de- 5 F

h k scribe the cavity-induced Larmor frequency shift and -10 10 the exponential damping or amplification rate, given by   δωopt = Re[Σ(ωL)]/2ωL and Γopt = Im Σ(ωL) /ωL, 0.05 0.1 0.5 1 2 − respectively. !L=5 Averaging the coherent heterodyne signal from up to 40 repetitions of the experiment to reduce statistical FIG. 4. (color online) Sideband-asymmetry thermometry of noise, we extract both Γopt and δωopt from early times a spin oscillator. (a) Stokes and anti-Stokes sidebands ob- in the measured transient response (Fig. 3a-b). For this, served in the averaged heterodyne PSD, normalized by the we simultaneously fit both quadratures of the heterodyne shot-noise level, showing the sideband asymmetry of a low- energy spin, with ω /2π = 900 kHz, in equilibrium with signal, restricting the fit to spin dynamics near the poles L an optical damping tone detuned by ∆pc/2π = −1.5 MHz (within 0.8 rad, see Supplemental Material [42]). with average photon numbern ¯damp = 5.0, and measurement The close agreement between our measurements and back-action from an on-resonance probe with average pho- the theoretical predictions (Fig. 3) demonstrates the close ton numbern ¯probe = 1.0. (b) Inverted sideband asymme- analogy between cavity optomechanics and spin optody- try for a negative-temperature spin ensemble near its high- energy state, with ∆pc/2π = 1.5 MHz,n ¯damp = 4.3, and namics. In addition, for a spin oscillator, ωL can be tuned readily over a broad range. Therefore, we can observe op- n¯probe = 2.0. (c) Equilibrium temperature (left scale) and thermal occupation (right scale) of a high-energy spin oscil- todynamical effects in the transition from the unresolved- lator, measured by sideband asymmetry, as a function of the to the resolved-sideband regime. For example, we demon- Larmor frequency ωL. Lines indicate expected temperature at strate that the peak damping rate, measured for a range equilibrium between coherent damping and incoherent back- of Larmor frequencies, is maximized for ωL/κ > 1 (Fig. action. The solid line indicates the ideal limit in the absence 3c). of the on-resonance probe, and the dashed line includes the In addition to coherent cavity back-action, the collec- additional back-action for the average probe cooperativity. tive spin is also subject to quantum fluctuations of Bopt, which cause diffusion of the collective spin away from its stable state. The balance between coherent and incoher- the equilibrium temperature can be determined from the ent back-action effects leads the spin ensemble to achieve asymmetry in its Raman scattering spectrum, recorded a steady-state temperature. This equilibrium between in the power spectral density (PSD) of the heterodyne cavity-assisted damping and measurement back-action is signal. The amplitude of the Stokes sideband corre- similar to that achieved in cavity cooling of mechani- sponds to the energy absorption rate Γ+ and the anti- cal oscillators [44, 45]. Because the atomic spins are ex- Stokes sideband corresponds to the emission rate Γ−. tremely isolated from their environment, high spin opto- Assuming the collective spin state to be in thermal equi- dynamical cooperativity, where the optical coupling ex- librium at temperature T , the ratio of these rates satisfies ceeds the intrinsic damping, is achieved already at mini- the detailed balance relation Γ+/Γ− = exp(~ωL/kBT ). mal optical power. Therefore, the equilibrium spin tem- We characterize the steady state of the atomic spin perature is determined solely by the quantum fluctua- ensemble by driving the optical cavity with two coher- tions of the cavity field, which are shaped by the cavity ent, primarily shot-noise limited tones: one strong tone p 2 2 linewidth. at a detuning ∆pc = β κ + ωL that provides damp- Due to the quantization of the collective spin oscillator, ing toward either the high- or low-energy pole, and a 5 second weak tone on cavity resonance. The asymmetry Rev. Lett. 110, 133602 (2013). of sidebands generated on the resonant tone is observed [7] N. Behbood, G. Colangelo, F. Martin Ciurana, in our heterodyne receiver (Fig. 4a) and used to deter- M. Napolitano, R. J. Sewell, and M. W. Mitchell, Phys. mine T . Whereas a spin oscillator stabilized in its low- Rev. Lett. 111, 103601 (2013). [8] D. J. Wilson, V. Sudhir, N. Piro, R. Schilling, energy state yields the conventional Stokes/anti-Stokes A. Ghadimi, and T. J. Kippenberg, Nature 524, 325 sideband asymmetry, consistent with positive tempera- (2015). ture, for a spin oscillator in the high-energy state, the [9] H. Mabuchi, Phys. Rev. A 78, 032323 (2008). sideband asymmetry is reversed, indicating a negative [10] B. Kraus, H. P. B¨uchler, S. Diehl, A. Kantian, A. Micheli, temperature for the spin ensemble (Fig. 4b). Accounting and P. Zoller, Phys. Rev. A 78, 042307 (2008). for shot-noise-driven heating of the spin oscillator from [11] H. Krauter, C. A. Muschik, K. Jensen, W. Wasilewski, both probe tones, the observed spin temperatures agree J. M. Petersen, J. I. Cirac, and E. S. Polzik, Phys. Rev. Lett. 107, 080503 (2011). well with the theoretical model (Fig. 4c). As predicted [12] S. Shankar, M. Hatridge, Z. Leghtas, K. M. Sliwa, [44, 45], autonomous feedback cooling yields increased A. Narla, U. Vool, S. M. Girvin, L. Frunzio, M. Mir- purity of the quantum state (kB T /~ωL < 1) as the sys- rahimi, and M. H. Devoret, Nature 504, 419 (2013). tem enters the resolved-sideband| regime.| [13] V. B. Braginsky, Y. I. Vorontsov, and K. S. Thorne, Science 209, 547 (1980). In summary, we have demonstrated cavity based mea- [14] I. D. Leroux, M. H. Schleier-Smith, and V. Vuleti´c,Phys. surement of Larmor precession of the collective spin of an Rev. Lett. 104, 073602 (2010). atomic ensemble and coherent control via autonomous [15] R. J. Sewell, M. Koschorreck, M. Napolitano, B. Dubost, feedback. These capabilities could be used to perform N. Behbood, and M. W. Mitchell, Phys. Rev. Lett. 109, quantum-limited measurements of the collective spin or 253605 (2012). to realize a phase-preserving, quantum-limited amplifier [16] J. G. Bohnet, K. C. Cox, M. A. Norcia, J. M. Weiner, for spin states. Furthermore, we demonstrate the ability Z. Chen, and J. K. Thompson, Nat. Photonics 8, 731 (2014). to stabilize the ensemble in a nearly-pure quantum state [17] O. Hosten, N. J. Engelsen, R. Krishnakumar, and M. A. with negative effective temperature. In this condition, Kasevich, Nature 529, 505 (2016). the system can be described as a near ground-state, neg- [18] K. C. Cox, G. P. Greve, J. M. Weiner, and J. K. Thomp- ative effective-mass oscillator. If measured jointly with son, Phys. Rev. Lett. 116, 093602 (2016). a mechanical oscillator, such a spin oscillator could al- [19] G. Vasilakis, H. Shen, K. Jensen, M. Balabas, D. Salart, low continuous QND position or force measurement via B. Chen, and E. S. Polzik, Nat. Phys. 11, 389 (2015). coherent quantum noise cancellation [46]. [20] N. Brahms and D. M. Stamper-Kurn, Phys. Rev. A 82, 041804 (2010). We thank L. Buchmann for helpful discussions and J. [21] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Gerber for assistance in the lab. This work was supported Rev. Mod. Phys. 86, 1391 (2014). by the Air Force Office of Scientific Research. N.S. was [22] A. Osada, R. Hisatomi, A. Noguchi, Y. Tabuchi, R. Ya- supported by a Marie Curie International Outgoing Fel- mazaki, K. Usami, M. Sadgrove, R. Yalla, M. Nomura, lowship, J.K. and S.S. by the U.S. Department of Defense and Y. Nakamura, Phys. Rev. Lett. 116, 223601 (2016). [23] T. Liu, X. Zhang, H. X. Tang, and M. E. Flatt´e,Phys. through the National Defense Science and Engineering Rev. B 94, 060405 (2016). Graduate Fellowship program. [24] S. Viola Kusminskiy, H. X. Tang, and F. Marquardt, Phys. Rev. A 94, 033821 (2016). [25] O. Arcizet, P.-F. Cohadon, T. Briant, M. Pinard, and A. Heidmann, Nature 444, 71 (2006). [26] S. Gigan, H. R. B¨ohm, M. Paternostro, F. Blaser, ∗ [email protected] G. Langer, J. B. Hertzberg, K. C. Schwab, D. B¨auerle, † [email protected] M. Aspelmeyer, and A. Zeilinger, Nature 444, 67 (2006). [1] C. M. Caves, K. S. Thorne, R. W. P. Drever, V. D. Sand- [27] A. Schliesser, P. Del’Haye, N. Nooshi, K. J. Vahala, and berg, and M. Zimmermann, Rev. Mod. Phys. 52, 341 T. J. Kippenberg, Phys. Rev. Lett. 97, 243905 (2006). (1980). [28] J. D. Thompson, B. M. Zwickl, A. M. Jayich, F. Mar- [2] H. M. Wiseman, Phys. Rev. A 49, 2133 (1994). quardt, S. M. Girvin, and J. G. E. Harris, Nature 452, [3] A. Kubanek, M. Koch, C. Sames, A. Ourjoumtsev, 72 (2008). P. W. H. Pinkse, K. Murr, and G. Rempe, Nature 462, [29] T. J. Kippenberg, H. Rokhsari, T. Carmon, A. Scherer, 898 (2009). and K. J. Vahala, Phys. Rev. Lett. 95, 033901 (2005). [4] C. Sayrin, I. Dotsenko, X. Zhou, B. Peaudecerf, [30] B. S. Sheard, M. B. Gray, C. M. Mow-Lowry, D. E. Mc- T. Rybarczyk, S. Gleyzes, P. Rouchon, M. Mirrahimi, Clelland, and S. E. Whitcomb, Phys. Rev. A 69, 051801 H. Amini, M. Brune, J.-M. Raimond, and S. Haroche, (2004). Nature 477, 73 (2011). [31] T. Corbitt, D. Ottaway, E. Innerhofer, J. Pelc, and [5] R. Vijay, C. Macklin, D. H. Slichter, S. J. Weber, K. W. N. Mavalvala, Phys. Rev. A 74, 021802 (2006). Murch, R. Naik, A. N. Korotkov, and I. Siddiqi, Nature [32] N. Brahms, T. Botter, S. Schreppler, D. W. C. Brooks, 490, 77 (2012). and D. M. Stamper-Kurn, Phys. Rev. Lett. 108, 133601 [6] P. Bushev, G. H´etet,L. Slodiˇcka, D. Rotter, M. A. Wil- (2012). son, F. Schmidt-Kaler, J. Eschner, and R. Blatt, Phys. [33] A. H. Safavi-Naeini, J. Chan, J. T. Hill, T. P. M. Alegre, 6

A. Krause, and O. Painter, Phys. Rev. Lett. 108, 033602 [39] A. T. Black, H. W. Chan, and V. Vuleti´c,Phys. Rev. (2012). Lett. 91, 203001 (2003). [34] T. P. Purdy, D. W. C. Brooks, T. Botter, N. Brahms, [40] S. Slama, S. Bux, G. Krenz, C. Zimmermann, and P. W. Z.-Y. Ma, and D. M. Stamper-Kurn, Phys. Rev. Lett. Courteille, Phys. Rev. Lett. 98, 053603 (2007). 105, 133602 (2010). [41] M. A. Norcia, M. N. Winchester, J. R. K. Cline, and [35] The cavity’s linear birefringence splits the resonance J. K. Thompson, (2016), arXiv:1603.05671. frequencies of two orthogonal linear polarizations by [42] See Supplemental Material, which includes Ref. [43], for a ∆b/2π = 1.2 MHz. This leads to a small mixing between description of the fit procedure and derivations of plotted circular polarizations, with the lowest-order correction theoretical predictions. in photon number of the un-driven mode proportional to [43] T. Holstein and H. Primakoff, Phys. Rev. 58, 1098 2 (∆b/2κ) = 0.11, which was shown to have a negligible (1940). effect in numerical simulations. [44] F. Marquardt, J. P. Chen, A. A. Clerk, and S. M. Girvin, [36] W. Happer, Rev. Mod. Phys. 44, 169 (1972). Phys. Rev. Lett. 99, 093902 (2007). [37] R. H. Dicke, Phys. Rev. 93, 99 (1954). [45] I. Wilson-Rae, N. Nooshi, W. Zwerger, and T. J. Kip- [38] M. Gross, P. Goy, C. Fabre, S. Haroche, and J. M. Rai- penberg, Phys. Rev. Lett. 99, 093901 (2007). mond, Phys. Rev. Lett. 43, 343 (1979). [46] M. Tsang and C. M. Caves, Phys. Rev. X 2, 031016 (2012).



SUPPLEMENTAL MATERIAL can be absorbed into the net probe detuning ∆pc. Neglecting these small static corrections, the equations S.1 DAMPING AND AMPLIFICATION RATES of motion for the collective spin, obtained from the lin- earized Hamiltonian, are given by

We reproduce here some results from Ref. [20], provid- ˙ † Fˆx = gc√n¯(ˆa +a ˆ)Fˆy, (S2a) ing a more detailed derivation of the connection to linear − ˙ † optomechanical theory and a foundation for the analysis Fˆy = ωLFˆz + gc√n¯(ˆa +a ˆ)Fˆx, (S2b) in the next section used to estimate the spin dynamics − ˆ˙ ˆ from the measured optical modulation. Fz = ωLFy. (S2c) To obtain an analytic solution for small-amplitude dy- For small excitations away from the stable poles, we ap- namics, we linearize the system Hamiltonian to describe proximate Fˆ βF , where β = 1 labels the high- and modulations of the spin around the magnetic poles and of x low-energy states,≈ and these simplify± to the cavity field about its mean value. Transforming into a cavity frame rotating at the probe laser frequency ωp, ˆ¨ 2 ˆ † Fz = ωLFz + βgc√nω¯ LF (ˆa +a ˆ). (S3) then expanding to first order about the average cavity − fielda ˆ a¯ +a ˆ(t) and spin projections Fˆz F¯z + Fˆz(t) The equation of motion for the cavity field, including the → → ˆ and Fˆx F¯x + Fˆx(t), dropping constant terms, the lin- vacuum noise at the cavity input ξ and dissipation of the earized→ Hamiltonian describing interactions with circu- cavity field, is larly polarized light with helicity along +z is given by ˙ ˆ aˆ = (i∆pc κ)ˆa igc√n¯Fˆz + √2κξ. (S4) − − † ˆ † ˆ = ~∆pcaˆ aˆ + ~ωLFx + ~gca¯(ˆa +a ˆ)Fz These equations closely resemble those describing lin- H − 2 ˆ ¯ † ear optomechanics, with the spin projection along the + ~gca¯ Fz + ~gcFzaˆ aˆ, (S1) cavity axis Fˆz representing the generalized position of 2 0 the oscillator. Near either pole, excitations of the collec- where gc = α1g0/∆ca = 2π 671 Hz and ∆pc = ωp ωc is the probe detuning from cavity× resonance, shifted− by the tive spin away from the high- or low-energy state can be 0 2 approximated as bosonic excitations of a harmonic oscil- scalar ac Stark shift ωc = ωc +α0Ng0/∆ca. We define the average field amplitudea ¯ = √n¯ to be real without loss of lator mode ˆb [43]. We define an effective unit-less spin dis- q generality. This linearized approximation is valid for spin ˆ† ˆ ˆ F placementz ˆs = b + b = Fz/∆SQL, where ∆SQL = 2 , amplitudes that modulate the cavity resonance by much the standard quantum limit for spin fluctuations, is the less than a cavity line-width, quantified by gcF¯z κ.  spin equivalent of the quantum zero-point motion. Un- The average cavity intensity interacts with the atoms der this approximation, the analogy to optomechanics is like an effective static magnetic field, represented in the exact, and the equations can be written as fourth term of the above Hamiltonian. This field ro- ¨ 2 † tates the static orientation of the spin slightly away from zˆs = ωLzˆs + 2β√ng¯ sωL(ˆa +a ˆ), (S5a) ¯ − the x axis, creating a small average projection Fz ˙ ˆ ' aˆ = (i∆ κ)ˆa i√ng¯ szˆs + √2κξ, (S5b) (gcn/ω¯ L)F¯x, which is negligible in the limit gcn/ω¯ L 1. − − This rotation also causes an additional static shift to the by defining the analogous single-photon, single-excitation cavity resonance, reflected in the fifth term above, which coupling gs = gc∆SQL. 7

Semi-classical solutions for the coherent dynamics of 2000 a)

the system can be found by dropping the vacuum in- 1000 i put noise and considering the spin displacement in re- z

^ 0 F sponse to a unit-less torque τ(ω) at frequency ω. The h -1000 complex susceptibility of the oscillator response χs(ω) = -2000 zs(ω)/τ(ω) is found by solving the Fourier transform of Equations S5, resulting in 2000 1000

−1 2 2 i χ (ω) = ω ω + Σ(ω), (S6) z

s L ^ 0 F

− h where Σ(ω) is given by Equation 2 in the main text. -1000 -2000

2000 S.2 SIMULTANEOUS QUADRATURE FIT b)

1000

i z

^ 0 F

To obtain an estimate of the coherent spin dynamics h Fˆz(t) from the recorded optical modulation, we solve -1000 h i the equation of motion for the cavity field, neglecting -2000 quantum fluctuations. Assuming the spin oscillator’s re- 2000 sponse takes the form Fz(t) = A(t) sin(ωLt), where A(t) 1000

is a slowly varying amplitude, the solution to Equation i z

^ 0 F

S4 is approximately h -1000 i√ng¯ -2000 a(t) = c A(t) 2 2 0 0.05 0.1 0.15 0.2 0.25 0.3 (κ i∆pc) + ωL − h i Time (ms) (κ i∆pc) cos ωLt + ωL sin ωLt . (S7) × − FIG. S1. Estimates of the spin dynamics for a (a) damped The cavity output field is determined by the bound- and (b) amplified spin oscillator, obtained from both the am- ary conditiona ˆin +a ˆout = √2κaˆ, where the input field plitude (green, lower) and phase (blue, upper) quadratures on the open port of our double-sided cavity is only vac- of the demodulated heterodyne signal. This shows the entire uum noise, which we also neglect in this semi-classical signal and simultaneous fit for the results shown in Fig. 2 of treatment. We perform a balanced heterodyne detection the main text. of this field, combining it with a local oscillator, with power PLO = 1.0 mW, derived from the same source and shifted ω0 relative to the probe. Demodulating the result- ing photocurrent, the detected optical power expected in the amplitude and phase quadratures is given by q The fits in the paper are performed by assuming A(t) = 2 2 IQ(t) = B κ + ω A(t) sin(ωLt + φPM), (S8a) −iΓt/2 h i L A0e , and simultaneously fitting Equations. S8a and II (t) = B∆pcA(t) sin(ωLt + φAM), (S8b) S8b directly to both demodulated quadratures, allowing h i the initial amplitude A0, exponential damping rate Γ, 0 with the relative phase between these responses shifted response frequency ωL and a common arbitrary tan(φPM φAM) = ωL/κ, and the common amplitude phase φ to vary. Since the relative signal to noise of scale − each quadrature varies with detuning, this naturally pro- vides the appropriate weighting of each quadrature in the s 2 2PLO~ωpnκg¯ c fit. Because the oscillator response and cavity coupling B = 2 2 2 2 2 2 . (S9) (κ + ∆pc + ωL) 4∆pcωL become inherently non-linear for large amplitude oscil- − lations, these solutions and the choice of an exponential This result shows that, for a detuned cavity probe, the model function are only valid for small excitations. To oscillator’s dynamics are encoded in both the amplitude consistently extract the damping (amplification) rates for and phase quadrature of the heterodyne signal, but with small amplitude oscillations across various conditions, we different phase. This phase difference prevents extraction iteratively scan the start (stop) time of the fit window, of all signal information by a simple linear combination choosing the longest time range within which the am- of the quadratures, so we instead perform a simultaneous plitude of the fit result remains below a fixed threshold, fit to both. Fz(t) < 1000. 8

S.3 CAVITY COOLING AND SIDEBAND Larmor frequency, showing that the spin equilibrates to ASYMMETRY THERMOMETRY the same temperature. For damping by a single coherent tone, one finds that The equilibrium spin temperature for a damped spin the equilibrium temperature is independent of the intra- oscillator is solely determined by the quantum fluctua- cavity intensity. However, the photon shot-noise from tions of the cavity photon number at the Larmor fre- the second, on-resonance tone causes additional heating, quency. For a spin precessing around a large magnetic with the time-averaged photon shot-noise spectrum given field along the x axis, the eigenstates have well-defined by angular momentum m along this axis. If the spin ex- 2κn¯ 2κn¯ periences a weak coupling to the cavity field of the form damp probe Snn(ω) = 2 2 + 2 2 , (S13) ˆ ˆ ˆ (ω + ∆pc) + κ ω + κ = ~gcnˆFz = ~gcnˆ(F+ +F−)/2, the photon fluctuations Hmediate transitions adding or removing units of angular where ∆ is the detuning of the damping tone which momentum along x. The transition rate for a collective pc has mean intra-cavity photon numbern ¯ , andn ¯ spin F from state m to m + 1 is given by damp probe | i | i is the corresponding intensity of the probe. Due to the 2 additional heating of this second tone, the resulting equi- gc 2 Γ↑ = Pm m + 1 Fˆ+ m Snn( ωL), (S10) librium depends on the balance between the intensity of 4 |h | | i| − the two tones. We choose this balance to achieve a fixed where Pm is the population in state m, and Snn(ω) is measurement cooperativity, defined by the power spectral density of photon number fluctuations 2 inside the cavity. The rate of the reverse process is 4gs n¯probe Cs = , (S14) κΓopt 2 gc 2 Γ↓ = Pm+1 m Fˆ− m + 1 Snn(ωL). (S11) 2  4 |h | | i| where Γopt = gs Snn(ωL) Snn( ωL) is the optical damping due to the off-resonance− − tone. The detuning Since this is the only process mediating energy exchange that minimizes the resulting temperature is given by with the spin ensemble, in equilibrium, detailed bal- p 2 2 ∆pc = β ωL + κ , which yields a minimum equilibrium ance requires these rates to be equal, and Pm+1 = temperature of Pm exp( ~ωL/kBT ), which implies − "s #−1 ω κ2 C κ2 Snn(ωL) ~ωL exp ~ L = 1 + 2 1 + + s 1 . (S15) = exp . (S12) k T ω2 κ2 + ω2 Snn( ωL) kBT B L L − − This result is also a statement of the effective tempera- The average measurement cooperativity for the data ture of the photon shot-noise in the cavity mode at the shown in Fig. 4c of the main text was Cs = 1.6.