Dynamical Backaction Magnomechanics

C.A. Potts,1 E. Varga,1 V.A.S.V. Bittencourt,2 S. Viola Kusminskiy,2, 3 and J.P. Davis1 1Department of Physics, University of Alberta, Edmonton, Alberta T6G 2E9, Canada∗ 2Max Planck Institute for the Science of Light, Staudtstr. 2, PLZ 91058 Erlangen, Germany 3Institute for Theoretical Physics, University Erlangen-Nuremberg, Staudtstr. 7, PLZ 91058 Erlangen, Germany Dynamical backaction resulting from radiation pressure forces in optomechanical systems has proven to be a versatile tool for manipulating mechanical vibrations. Notably, dynamical back- action has resulted in the cooling of a mechanical to its ground-state, driving lasing, the generation of entangled states, and observation of the optical-spring effect. In certain magnetic materials, mechanical vibrations can interact with magnetic excitations (magnons) via the magnetostrictive interaction, resulting in an analogous magnon-induced dynamical backaction. In this article, we directly observe the impact of magnon-induced dynamical backaction on a spherical magnetic sample’s mechanical vibrations. Moreover, dynamical backaction effects play a crucial role in many recent theoretical proposals; thus, our work provides the foundation for future experimental work pursuing many of these theoretical proposals.

I. INTRODUCTION ing (damping) — leading to phonon lasing [42, 43] and noise squashing [44] — and the first observation of the Hybrid cavity systems hold great promise for explor- magnonic spring effect [37]. These results are enabled, ing a wide variety of physical phenomena. One broad ex- in part, by homodyne detection of the mechanics, with ample of this is the rapid maturation of cavity optome- reduced clamping and viscous damping. Our observa- chanics, i.e. coupling electromagnetic cavities with me- tions of dynamical backaction in a cavity magnomechan- chanical degrees of freedom [1]. The coupling of electro- ical system open the door for many of the above theoret- magnetic cavities with magnonic systems has also gener- ical proposals, in particular those that involve entangle- ated significant interest [2–4], with theoretical proposals ment [23–28], squeezed states [29–31], and thermometry for magnetometry [5] and axion detection [6–8], and ex- [37]. With reasonable improvements, future experiments periments demonstrating strong-coupling [9–13], magnon could realize, for example, magnon-mediated cooling of Fock state detection [14, 15], coupling to superconduct- the mechanics into the ground-state, which would achieve ing qubits [16, 17], bidirectional microwave to optical conversion [18, 19], Floquet electromagnonics [20], and non-reciprocity [21]. Combining these concepts yields the field of cav- (a) (b) 푎ො 푚ෝ ity magnomechanics: coupling of magnetic excitations (magnons) with both an electromagnetic cavity as well as a mechanical resonator [22], as illustrated in Fig. 1(b). Furthermore, an important triple-resonance condition is 푏෠ possible, where the phonon frequency matches the differ- ence in frequencies between the hybrid cavity-magnon modes [22], allowing selective cavity enhancement of scattering processes. This triple-resonance system has sparked significant interest, resulting in theoretical pro- posals for the generation of non-classical entangled states FIG. 1. Experimental set-up. (a) Photograph of half of the [23–28], squeezed states [29–31], classical and quan- microwave cavity, made from oxygen-free copper. The YIG tum information processing [32–36], quantum correlation sphere is placed – free to move – within a glass capillary, inner thermometry [37], and exploring PT -symmetry [38–41]. dimension of approximately 300 microns. A set of permanent Many of these proposals rely on the ability of the ex- neodymium magnets attached to a pure iron yoke provide the bias magnetic field, a solenoid (not shown) allows the bias field ternal drive to act on the mechanical motion, so called arXiv:2104.11218v2 [cond-mat.mes-hall] 9 Jul 2021 to be varied dynamically. Inset: Optical micrograph of the dynamical backaction. Yet, despite the large number of YIG sphere inside of the glass capillary. Scale bar is 500 mi- theoretical papers, there remains to date a single exper- crons. (b) Schematic of the coupled magnomechanical system. imental observation of cavity magnomechanics [22]. aˆ – Numerical simulation of the TE101 microwave magnetic Here, we report cavity magnomechanical detection field distribution.m ˆ – Schematic representation of the Kittel of all of the fundamental dynamical backaction ef- magnon mode within the spherical YIG sample. ˆb – Numeri- fects, i.e. dynamical heating (amplification) and cool- cally simulated displacement of the mechanical mode that has the strongest magnomechanical coupling to the Kittel mode. The solid line indicates linear magnon- coupling, while the dashed line represents parametric magnon-phonon cou- ∗ [email protected], [email protected] pling. 2 the largest mechanical system to date to be taken into (a) (b) its quantum ground state, and possibly allow tests of gravitational decoherence currently being pursued with levitated spheres [45–48]. 휔+

휔− 2푔푎푚 II. THEORETICAL BACKGROUND Ω푏 Hybrid magnomechanical systems are composed of three parts: the electromagnetic field (confined in the (c) (d) microwave cavity), the magnetic excitations of a dielec- tric, and its mechanical vibrations (both hosted in the dielectric YIG sphere). To study the dynamical response Ω푏 Ω푏 of the coupled cavity magnomechanical system, we con- sider three interacting bosonic modes, as shown in Fig. 1, with annihilation operatorsa ˆ (microwave cavity mode), mˆ (magnon mode), and ˆb (phonon mode), with frequen- cies ωa, ωm, and Ωb respectively. It should be noted while we have used a quantum description of the theory, all ex- perimental results presented in the article are classical. Before tackling the full interacting Hamiltonian, we FIG. 2. Normal mode spectrum. (a) Measured normal mode first study the microwave-magnon interaction. The linear spectrum as a function of static magnetic field. Dashed lines interaction between magnons and microwaves is the hall- correspond to the spectrum in (b-d). (b) Cavity reflection mark of cavity magnonic systems [9–11], and the Hamil- spectrum when the magnon is resonant with the bare-cavity tonian describing a magnon mode interacting with a sin- mode. The normal mode splitting, 2gam is smaller than the gle driven microwave cavity mode reads (in the frame phonon frequency, Ωb = 2π × 12.6278 MHz. (c) Cavity re- rotating with the driving frequency) [10, 11] flection spectrum when the magnon frequency is smaller than the bare-cavity resonance frequency. The system was set-up such that the normal mode spacing matches the mechani- H = −~∆aaˆ†aˆ − ~∆mmˆ †mˆ √ (1) cal frequency. Here the lower mode is ‘magnon-like’ and the + ~gam(ˆamˆ † +a ˆ†mˆ ) + i~d κext(ˆa − aˆ†). upper mode is ‘photon-like.’ (d) Cavity reflection spectrum when the magnon frequency is larger than the bare-cavity The magnon-photon coupling rate is given by g , the resonance frequency, similar to the detuning in (c). Here the p am external drive is d = P/~ωd (where P is the external upper mode is ‘magnon-like’ and the lower mode is ‘photon- drive power at the device), κext is the external coupling like.’ rate, and ~ is the reduced Planck’s constant. The cav- ity and magnon detunings are defined as ∆a = ωd − ωa, and ∆m = ωd − ωm respectively, where ωd is the ex- is resonant with the magnon mode, ωa = ωm, the nor- ternal drive frequency. The magnon frequency is given mal modes are a maximal hybridization of magnons and by ωm = γ|B0|, where γ/2π = 28 GHz/T is the gyro- . Otherwise, the normal modes describe partial magnetic ratio, and B0 is the applied static magnetic hybridization; one of the modes is ‘magnon-like’, and the field [49]. The first two terms in the Hamiltonian rep- other ‘photon-like’. The normal mode splitting can be di- resent the occupancy of the photon and magnon modes, rectly measured via the reflected microwave signal. This respectively. The third term describes the linear magnon- can be seen in Fig. 2, which depicts the normal mode photon coupling, and the final term describes the external splitting in our experiment. drive. The value of the coupling rate can be extracted by per- Due to the linear coupling, microwaves and magnons forming a fit to the normal modes presented in Fig. 2(a), hybridize: the normal modes of the interacting Hamilto- using the theory from Ref. [10]. For our experimental nian Eq. (1) are superpositions of magnons and photons. configuration we extract a coupling rate gam/2π = 5.43 We label these modes as + and −, and the difference MHz. Furthermore, both the magnon and microwave between their frequencies is given by modes are subject to decay. The microwave cavity de- cay rate is composed of both an intrinsic cavity de- p 2 2 ω+ − ω ≡ ∆ω = 4gam + ∆am, (2) cay and the coupling to external coaxial cables (used to − drive/measure the cavity). The main source of magnon where ∆am = ωa − ωm is the magnon-photon detun- damping is intrinsic Gilbert damping [50], which includes ing. Since an externally applied bias field can tune the dissipation processes associated with -lattice cou- magnon frequency, the hybridization of the modes is con- pling. Other possible sources of damping include two- trollable: e.g., by varying the current through a solenoid magnon scattering processes between the magnetostatic providing the bias field. Furthermore, when the cavity mode and the spin-wave continuum [51]. Those pro- 3 cesses yield an inhomogeneous broadening of the magnon (for ωa = ωm) and the phonon frequency are not perfectly linewidths for different magnon modes [52], which are less matched; indeed, the phonon frequency is slightly larger prominent in well-polished spheres. From the measured than the normal mode splitting, Fig. 2(b). However, by data we extract the magnon decay rate γm/2π = 1.01 detuning the magnon frequency slightly from the bare MHz, and the total cavity decay rate κ/2π = 3.87 MHz, microwave cavity frequency (i.e. ∆am =6 0), the normal placing our experiment well within the strong-coupling mode splitting can be tuned to exactly match the phonon regime, gam > {κ, γm}. frequency, see Fig. 2(c,d). This results in the system Besides the coupling to microwave photons, magnons becoming triply resonant, which significantly enhances couple to the mechanical vibrations of the material. The the magnon-phonon coupling. Indeed, it has been shown magnon-phonon coupling is mediated by the magne- in Ref. [22] when compared with off-resonance driving the tostrictive interaction [22, 53, 54], which couples magne- cooperativity of a fully-hybridized triple-resonant system tization to mechanical strain. The magnetoelastic cou- will be enhanced by a factor of pling originates from the effects of lattice strain combined with exchange interactions, dipole-dipole interactions,  Ω 2 F = 16 b . (4) and spin-orbit coupling [55]. Under strain, the distance κ + γm between magnetic /ions changes and, as a conse- quence of the aforementioned interactions, the orbital Finally, when the magnon-phonon interaction Hamil- and spin configurations are modified. The overall effect tonian Eq. (3) is combined with the microwave-magnon is a coupling between magnetic order and strain/stress, Hamiltonian, Eq. (1), a more complex scenario occurs. which is obtained from the magnetic anisotropy energy Since the magnon-phonon coupling is weak compared density [53, 55]. For small vibrations and magnetization to the microwave-magnon coupling, we can describe the oscillations, this energy density term yields two types of magnon mode as a superposition of the normal modes magnon-phonon couplings: a linear coupling, relevant for discussed above. The magnon-phonon Hamiltonian can resonant excitations, and a parametric coupling [22]. Our then be understood as describing a scattering process be- experiments investigate the later, which has a form anal- tween the normal modes, mediated by a phonon. For ex- ogous to the photon-phonon coupling in optomechanical ample, a drive photon resonant with the higher frequency cavities, where the optical cavity frequency is modulated normal mode is scattered into the lower frequency normal by mechanical vibrations, for example, of a movable mir- mode and a phonon is generated. Such a process is reso- ror [1]. In our experiment, the magnetic element is a YIG nant if the phonon frequency matches the normal mode sphere, which acts as both the magnon and phonon res- splitting, fulfilling the triple-resonance condition. The onator. YIG is the material of choice for magnomechan- sphere diameter was specifically chosen for our system to ical experiments since it possesses excellent magnetic [9– operate near the triple-resonance. 11] and mechanical properties [56]. In fact, YIG spheres have mechanical decay rates similar to spheres of silicon and quartz [57]. III. EXPERIMENTAL CONFIGURATION The parametric magnetostrictive coupling is described by a radiation-pressure-like parametric Hamiltonian [22]: The system used in our experiment consists of a three- 0 ˆ ˆ dimensional microwave cavity machined from oxygen-free Hint = ~g mˆ †mˆ (b + b†). (3) mb high-conductivity copper, seen in Fig. 1(a). The mi- Here, the single magnon-phonon coupling strength is crowave cavity has inner dimensions 30 × 30 × 6 mm3, 0 given by gmb. The exact value of the coupling rate de- resulting in the TE101 mode having a frequency of pends on the mode overlap between the specific magnon ωa/2π = 7.074 GHz. The intrinsic decay rate of the and phonon modes. The lowest order spherical modes are microwave cavity is κint/2π = 1.56 MHz. Coupling to described by S1,l,m, where l and m are the angular and the cavity is achieved through a pair of coaxial cables azimuthal mode numbers, respectively. For this work, we with external coupling rates of κ1/2π = 1.11 MHz, and specifically focus on the spherical S1,2,0 mode, shown in κ2/2π = 1.20 MHz, resulting in a total cavity linewidth Fig. 1(b), since it possesses the largest magnon-phonon of κ/2π = (κint + κ1 + κ2)/2π = 3.87 MHz. coupling rate. We used a commercial YIG sphere, nom- The single crystal YIG sphere is placed free to move inally 250 microns in diameter [58]. The phonon fre- – to avoid mechanical clamping losses – within a 300- quency and magnon-phonon coupling rate can be esti- micron inner diameter capillary [57]. The sphere is lo- mated numerically using COMSOL multiphysics. For the cated near the magnetic field maximum of the microwave mode of interest, for a 269-micron diameter sphere, the cavity and is held in place, and oriented along its easy 0 coupling rate is predicted to be gmb/2π ≈ 4.73 mHz, and axis, by the applied static magnetic field. A pair of the phonon frequency is estimated to be Ωb/2π = 12.66 neodymium magnets provide the static magnetic field, MHz. Note, the coupling rate could be increased by de- seen in Fig. 1(a). Tunability of the static magnetic field creasing the YIG sphere’s diameter; however, this has the is provided via a ∼ 104 turn solenoid, wrapped around a effect of decreasing the magnon-photon coupling rate. pure iron core and connected to the permanent magnets In our experiment, the resonant normal mode splitting using an iron yoke [16]. 4

To determine the mechanical frequency, and ensure sitic thermal effects. We further see a phenomena known 0 constancy of the magnomechanical coupling rate gmb as noise squashing resulting from interference between between various measurement techniques, we first per- the thermal magnon bath and excitations scattered via formed a series of magnomechanically induced trans- the magnomechanical interaction, which will be discussed parency (MMIT) measurements similar to that presented further in Section IV C. in Ref. [22]. From this set of measurements (described in Appendix A) we are able to extract a phonon frequency Ωb/2π = 12.6270 MHz, a single magnon-phonon coupling Furthermore, it is possible to observe mechanical mo- 0 tion without relying on the triple-resonance condition, rate of gmb/2π = 4.38 mHz, and an intrinsic mechanical linewidth of Γb/2π = 286 Hz. one can apply the drive tone on the blue sideband of the In the rest of the experiment, we measure the YIG upper normal mode (or the red sideband of the lower sphere’s mechanical vibrations without resorting to the normal mode). Although this sideband driving has some MMIT window, which more easily enables the observa- similarities with procedures commonly adopted in cavity tion of dynamical backaction effects. In this scheme, optomechanics [1], the composition of the normal modes a microwave signal was sent to the hybrid magnome- can be changed by varying the magnon-photon detun- chanical system, as shown in Fig. 3(a). The transmit- ing. This in turn results in each normal mode experienc- ted signal was demodulated using an IQ-mixer, and the ing a detuning-dependent coupling rate, decay rate, and low-frequency mechanical signal was digitized using an effective phonon coupling rate. Therefore, there exists analog-to-digital (ADC) converter. The reflected signal an optimal detuning for mechanical measurements. The was passed through a directional coupler and measured magnon and photon components of the normal mode pro- using a VNA to characterize the normal mode spectrum. vide two distinct operations, the magnon-like component During the mechanical measurements, the VNA was not couples directly to the mechanical motion, whereas the exciting the cavity to avoid potential beat frequencies applied microwave tone can drive the photon component. from obfuscating the mechanics. To balance the homo- Thus, the competition between these two effects needs to dyne circuit, the DC component of the demodulated sig- be balanced for optimal mechanical detection. The me- nal was continually measured and locked dynamically – chanical power spectrum while driving above (below) the at a rate of 1 kHz – by adjusting the local oscillator phase. upper (lower) normal mode, with optimal detunings, are Data was taken in three atmospheric conditions: ambi- shown in Fig. 3(d) and (e), respectively. Due to the small ent pressure, a partial pressure of pure helium gas (∼ 15 magnomechanical coupling (compared to the microwave- Torr), and vacuum (< 1 mTorr). magnon coupling), a drive power of ∼ 5 mW was required to resolve the mechanical spectrum. All powers in this article are quoted as power at the device, which were IV. HOMODYNE MECHANICAL DETECTION carefully calibrated for each experimental configuration.

As mentioned, one remarkable aspect of this hybrid It should be noted that unlike many optomechanical magnomechanical system is that it can be brought into measurements, the observed mechanical motion is not triple-resonance. In this case, the red mechanical side- thermomechanical in nature. Due to the high drive pow- band of the upper normal mode has the same frequency ers, there is considerable backaction in the form of heat- as the lower normal mode and vice versa. Therefore, not ing (cooling) of the mechanical mode. However, the in- only is the drive tone cavity-enhanced, but one of the two trinsic mechanical properties can be extracted by consid- mechanical sidebands is simultaneously cavity-enhanced. ering the data presented in Fig. 3(b) and extrapolating to The Stokes (anti-Stokes) scattering process is strongly zero drive power. As a result, within a partial helium en- preferred, resulting in effective magnomechanical backac- vironment, we find the mechanical mode has a resonance tion heating (cooling). These two specific scenarios will frequency Ωb = 12.6278 MHz, and intrinsic mechanical be discussed in more detail in the next sections. decay rate Γb/2π = 98 Hz. The mechanical power spectrum under the triple- resonance condition is shown in Fig. 3(b,c). The normal mode spectrum for Fig. 3(b) is shown in Fig. 2(d), and Finally, to confirm the observed mechanical signal was the drive is tuned on resonance with the upper normal not a result of direct electromechanical coupling (i.e. cou- mode. With increasing drive power, two effects can be pling between photons and ), measurements were seen in Fig. 3(b): the frequency of the mode decreases, performed, shifting the magnon frequency far from the we attribute this to parasitic thermal effects that are dis- microwave resonance frequency. In this scenario, the nor- cussed in Appendix C, and a narrowing of the linewidth, mal mode spectrum disappears, and we are left with only which ultimately results in phonon lasing discussed in the microwave cavity resonance. Both OMIT and homo- Section IV B. Conversely, for Fig. 3(c), the normal mode dyne measurements were performed; no evidence of the spectrum is shown in Fig. 2(c), and the drive is tuned on mechanical motion was observed in either of these sce- resonance with the lower normal mode. Again we observe narios. Therefore we are confident the magnon indeed a power dependent frequency shift resulting from para- mediates the phonon interaction. 5

(a) (b) (d) RF LO Sync VNA

IQ ADC

(c) (e) 50Ω

FIG. 3. Homodyne mechanical detection. (a) Simplified schematic of the measurement setup: VNA, vector network analyzer; RF/LO, microwave generators; ADC, analog-to-digital converter; IQ, IQ-mixer, both the in-phase and quadrature ports were connected to the ADC. (b) Power spectral density of the mechanical motion. The normal mode spectrum was tuned to Fig. 2(d). The probe was tuned directly on resonance with the upper normal mode. With increasing drive power, a frequency shift, and linewidth narrowing can be observed. (c) Power spectral density of the mechanical motion; offset for clarity. The normal mode spectrum was tuned to Fig. 2(c). The probe was tuned directly on resonance with the lower normal mode. Due to interference between scattered excitations and the thermal magnon bath, the power spectrum lies below the noise floor, known as noise squashing. (d) Power spectral density of the mechanical motion, the drive tone was detuned one mechanical frequency above the upper normal mode. (e) Power spectral density of the mechanical motion, the drive tone was detuned one mechanical frequency below the lower normal mode. For both (d) and (e), the drive power was held constant at 12 mW. All data presented was obtained with the experimental setup in a partial pressure of helium gas.

A. Magnon Spring Effect article and would only break down for the highest on- resonance drive powers – the real and imaginary parts The magnomechanical interaction is given by the inter- of the self-energy describe a mechanical frequency shift action Hamiltonian Eq. (3). As seen above, this results in δΩb = −ReΣ[ω], the magnon-spring effect, and an addi- the formation of sidebands that carry information about tional magnomechanical damping rate Γmag = 2ImΣ[ω]. the mechanical vibrations. However, the interaction also In order to observe the small magnon-induced fre- results in a modification of the mechanical susceptibility quency shift, parasitic heating needed to be eliminated due to dynamical backaction from the interaction with (see Appendix C). To reduce heating of the sphere, a low magnons. We have previously described the full linear drive power was required; however, reducing the drive theory of the magnomechanical interaction in Ref. [37] power simultaneously reduces the frequency shift. There- and derived the following expression for the phonon self- fore, this experimental run was performed in a low pres- energy, sure (∼ 15 Torr) of pure helium gas to reduce the me- chanical linewidth, allowing the small frequency shift to Σ[ω] = i|g |2(Ξ[ω] − Ξ [−ω]). (5) mb ∗ be resolved. Helium was used because it possesses high 0 Here, gmb = gmbhmi is the cavity enhanced magnon- thermal conductivity; therefore, it provides good ther- phonon coupling rate, |hmi|2 is the coherent steady-state malization while limiting the mechanical damping of the 1 2 1 magnon population, and Ξ[ω] = [χm− [ω] + gamχa[ω]]− . sphere. Secondly, heating of the sphere was primarily due The magnon and cavity susceptibilities are given by, to magnon decay and not microwave photon absorption. χm[ω] = [−i(∆m + ω) + γm/2] and χa[ω] = [−i(∆a + Thus, unlike in Fig. 3(b), the drive tone was applied to ω) + κ/2], respectively. In the weak coupling limit, when the ‘photon-like’ normal mode and the interaction with gmb  κ, γm – which holds for all data presented in this phonons scattered excitations into the ‘magnon-like’ nor- 6

(a)

(b)

FIG. 4. Magnon spring effect. Power spectral density of the mechanical motion within a partial pressure of helium gas and a constant probe power of 0.13 mW. The microwave drive was tuned to the normal mode shown in the inset. White curve is the theoretical prediction for the magnon-spring effect (c) [37], where the only fit parameter was the intrinsic mechanical frequency. mal mode. This indeed resulted in less heating of the YIG sphere; however, it has the unwanted secondary effect of reducing the detection efficiency due to the reduced ex- ternal coupling of the ‘magnon-like’ mode. Figure 4 shows the mechanical power spectrum, re- FIG. 5. Parametric instability. (a) Mechanical linewidth as vealing the magnon-induced frequency shift. To avoid a function of drive power in air (blue squares), helium par- complicating effects from heating or a slow-drift of the tial pressure ∼ 15 mTorr (purple triangles), and vacuum (red circles). The single magnon-phonon coupling rate extracted magnetic field, the magnet was re-adjusted between each 0 drive frequency, and the drive frequencies were applied from these curves was gmb/2π = 4.65, 4.66, 4.43 mHz, respec- tively. For this data the normal mode spectrum was tuned to in a randomized order. This data was taken with a drive (d) in Fig. 2, and the probe tone was tuned on resonance with power of 0.13 mW; at this power and drive detuning we the ‘magnon-like’ mode. Inset: Power spectral density of the observed negligible frequency shift due to heating. The mechanical motion for the green (drive power = 0.54 mW) and white curve is a theoretical prediction from Eq. (5), where orange (drive power = 0.68 mW) markers, respectively. (b) the only fit parameter used was the intrinsic mechanical Phase coherent oscillations are visible within the time-domain frequency, which increased slightly in the partial pressure signal; here, the drive power is set above the parametric in- of helium, Ωb/2π = 12.6278 MHz. All other parameters stability threshold. (c) With the drive power set below the were extracted using a fit to the reflected normal mode parametric instability threshold, the time-domain signal com- spectrum and the magnomechanical damping measure- prises primarily of phase-incoherent mechanical noise. ment discussed below. The theoretical prediction and the experimentally measured shift are in agreement, confirm- ing the direct observation of the magnon spring effect. mechanical damping rate. The backaction enhancement of damping will be discussed in the next section; here, we will focus on the case of anti-damping. B. Magnomechanical Anti-Damping We now consider the data presented in Fig. 3(b); the drive tone is on resonance with the upper normal mode, We can consider the effect magnon backaction has on and the splitting between the normal modes is tuned to the mechanical decay rate. As described above, the in- exactly one mechanical frequency. The additional mag- teraction with magnons causes additional damping of the nomechanical anti-damping is thus maximized due to the mechanical mode, Γmag = 2ImΣ[ω], which results in an triple-resonance enhancement and the total mechanical effective mechanical damping rate, damping should decrease linearly with drive power as predicted by Eq. (5). This behavior is confirmed by the ΓTot = Γb + Γmag. (6) experimental points shown in Fig. 5(a). Furthermore, Just as in cavity optomechanics, Γmag can be positive or fits to the total linewidth allow extraction of the intrinsic 0 negative [1], thus either increasing or decreasing the total linewidth, Γb, as well as gmb. Extrapolating to zero drive 7 power yields, Γb/2π = 247 Hz in air, Γb/2π = 98 Hz in a a¯th ≈ m¯ th ≈ 800. This thermal population produces a partial pressure of helium (∼ 15 Torr), and Γb/2π = 59 broad peak in the power spectrum, corresponding to the Hz in vacuum; suggesting the primary damping mech- hybrid system’s normal modes. In the triply resonant anism was viscous air damping. The magnomechanical situation, the mechanical mode lies directly in the cen- coupling rate can be determined using Eq. (5) and (6) ter of this broad peak in the power spectrum. Noise as well as the slope of the data presented in Fig. 5(a) squashing results from destructive interference between 0 resulting in a value gmb/2π = 4.58 mHz. As expected, upconverted drive excitations and the thermal excita- the magnon-phonon coupling rate is independent of the tions, causing the mechanical peak to appear below the intrinsic decay rate and is in good agreement with our detection noise floor. However, since the thermal peak numerical prediction and the result from the MMIT mea- is approximately constant over the mechanical mode’s surement (Appendix A). width, it is possible to extract the mechanical linewidth As the drive power is increased, a threshold will be by performing a fit to an inverted power spectral den- reached where Γb + Γmag becomes negative. In this situ- sity. The extracted linewidth from the data in Fig. 3(b,c) ation, the mechanical oscillations will grow exponentially are shown in Fig. 6(a). Extrapolating to zero drive in time and will ultimately be limited by higher-order power, the intrinsic linewidth in vacuum from the damp- nonlinear effects. This parametric instability is analo- ing and anti-damping data were Γb/2π = 58.5 Hz and gous to lasing and is often referred to as phonon las- Γb/2π = 59.7 Hz, respectively, which are in excellent ing [42, 43, 59, 60]. The onset of lasing can clearly be agreement. seen in Fig. 5(a) as the total decay rate approaches zero Finally, since our experiment lies well within the side- above a threshold drive power. Furthermore, the inset band resolved regime (i.e. {κ, γm}  Ωb), and the num- of Fig. 5(a) shows the mechanical power spectrum above ber of thermal excitations is small compared to the (orange) and below (green) the threshold power. The phonon population, we can extract the effective mode onset of mechanical lasing results in four orders of mag- temperature due to magnomechanical cooling. The effec- nitude increase of the mechanical power spectrum. The tive phonon mode temperature is given by an expression additional noise peaks in the lasing spectrum are a re- similar to that for driven cavity optomechanical systems sult of 60 Hz line noise captured by the solenoid being [1]: transduced via the mechanical mode. Finally, we ob-   Γb(1 + β(P)) serve the onset of mechanical lasing directly in the time- Tfinal = Tinit . (7) domain. The time-domain signal captured by the ADC ΓTot is plotted in Fig. 5(b,c). Below the lasing threshold, the Here, β(P) is a drive dependant variable; for derivation signal is mainly comprised of noise; however, above las- see Appendix D. The effective mode temperature is de- ing threshold, coherent oscillations at the mechanical fre- fined in a consistent way for a driven system via the quency are visible [56, 61]. The time-domain and power power spectrum of the mode [66]. The effective mode spectrum data provide unambiguous evidence of phonon temperature is shown in Fig. 6(b); at the highest drive lasing, which could be used for stable clock signals [62], power the mechanical mode was cooled to approximately or as the basis of sensitive mass and force sensors [63]. 65 kelvin from room temperature. With improvements to the experimental setup, such as smaller YIG spheres, and pre-cooling the experiment via cryogenics, which im- C. Magnomechanical Cooling

When driving on the red-sideband in our system, for (a) (b) example, Fig. 3(c,e), the mechanical spectrum dips be- low the measurement noise floor, a phenomenon known as noise squashing [1]. Noise squashing has been observed in optomechanics, primarily in the context of feedback cool- ing [64, 65]. In feedback cooling, noise squashing results from the detector noise and the noise-driven mechanical motion becoming correlated. In our experiment, the detector noise was not fed into the system and cannot correlate with the mechanical mo- tion. There was no feedback, and as a result, the noise squashing we have observed has a different origin. In- FIG. 6. Magnomechanical cooling. (a) Mechanical linewidth deed, it has a backaction-cooling origin, which has been as a function of drive power in vacuum. Red circles demon- observed in a microwave optomechanical nanobeam de- strate magnomechanical anti-damping and includes the data vice [44]; however, it has not been observed in mag- from Fig. 3(b), green crosses demonstrate magnomechanical nomechanics. Notably, because our experiment was per- damping and includes the data from Fig. 3(c). (b) Effective formed at room temperature, there exists a large num- mode temperature of the mechanical mode determined from ber of thermally excited gigahertz magnons and photons, Eq. (7). 8 proves κ,Γb, and reduces thermal noise, it may be pos- (a) (b) sible to achieve ground-state cooling of the mechanical Sync vibrations. Furthermore, with our current experimental RF values Eq. (4) predicts a triple-resonance cooperativity enhancement of F ≈ 100. Thus at cryogenic temper- VNA atures a cooperativity of C ≥ 1000 is expected, plac- ing our triply-resonant system among the state-of-the-art electromechanical experiments that recently have demon- strated macroscopic entanglement [67]. (c)

V. CONCLUSION

The magnomechanical interaction has in recent years been the focus of considerable theoretical work, yet ex- perimental progress has been surprisingly limited. Here, we demonstrated the direct detection of mechanical vi- bration within a sub-mm YIG sphere and explored the full suite of dynamical backaction effects. We have shown that the magnomechanical interaction can amplify and FIG. 7. Magnomechanically induced transparency. (a) cool the mechanical vibrations of the material effectively. Schematic illustration of the measurement setup: VNA, vec- As a consequence, we have observed phenomena such tor network analyzer; RF, microwave generator. The VNA as microwave-driven phonon lasing, and noise squashing and source are synced using a 10 MHz clock signal. Included due to correlations with thermal noise. Unlike previous is a rendering of the microwave cavity and magnetic yoke investigations, our experiment eases the observation of assembly. The frequency axis is negative due to the rela- such dynamical backaction effects due both to the detec- tive detuning between the drive tone and the normal mode. tion scheme and the specific configuration of the setup, (b) Normalized reflection spectrum centred around the lower- in which the magnetic element is free to move in vac- normal mode, the system is tuned near spectrum (c) from uum. These improvements have allowed the detection Fig. 2, but at a larger solenoid current to isolate the ‘magnon- like’ normal mode. A narrow transparency window opens as of magnon-induced mechanical frequency shifts, i.e. the ∆dp = −Ωb due to the magnomechanical coupling. (c) Zoom- magnon spring effect. The experimental results agree in of (b) showing a detailed spectrum of the magnomechanical well with our previous theoretical description of cavity induced transparency. magnomechanics and highlight the potential of cavity magnomechanical systems. With further improvement, microwave-driven magnomechanical cooling may be used to reach effective temperatures low enough to reach the Appendix A: Magnomechanically induced ground state of the mechanics. This would not only allow transparency the observation of quantum effects in a relatively massive system, but have also applications in quantum technolo- Since the mechanical mode probed in our experiment gies such as quantum memories. is different than in Ref. [22], to determine the mechan- ical frequency and extract the magnomechanical cou- pling rate, we calibrate the system using MMIT. MMIT is analogous to optomechanically induced transparency ACKNOWLEDGMENTS (OMIT) [68, 69], and is a consequence of the interfer- ence of sidebands generated by the parametric coupling Authors acknowledge helpful contributions from D. to phonons. Besides the natural response of the system Lachance-Quirion, D. Milling, and C. Doolin. This work at the normal mode frequencies, each mode has sidebands was supported by the University of Alberta; the Natu- shifted by the phonon frequency. MMIT is observed by ral Sciences and Engineering Research Council, Canada driving the cavity resonant with the red-sideband of one (Grants No. RGPIN-04523-16 and No. CREATE-495446- of the normal modes while sweeping the probe through 17); and the Alberta Quantum Major Innovation Fund. the normal mode resonance. The interference between V.A.S.V. Bittencourt and S. Viola Kusminskiy ac- the weak probe and the upconverted excitations via the knowledge financial support from the Max Planck So- annihilation of a phonon generates a transparency win- ciety and from the Deutsche Forschungsgemeinschaft dow. (DFG, German Research Foundation) through Project- Figure 7(a) illustrates a schematic of our MMIT mea- ID 429529648–TRR 306 QuCoLiMa (“Quantum Cooper- surement apparatus. A two-port microwave cavity was ativity of Light and Matter”). used, with the microwave drive at frequency ωd connected 9 to coupling port one and driven with a power between 1 the accurate determination of the power reaching the de- and 50 mW. The vector network analyzer (VNA) probe, vice from the microwave sources. All other variables have at frequency ωp, was connected to coupling port two, and been defined in the main text except θ ∈ [0, π/2], which the probe tone was held at a constant power of 0.03 mW. is defined as, To simplify our analysis, we decided to apply a red- 2g detuned drive tone on the lower normal mode. The re- tan(2θ) = am , (B3) flection spectrum as a function of the two-photon detun- ωm − ωa ing ∆dp = ωd − ωp is shown in Fig. 7(b,c). A sharp peak can be seen at ∆dp = −Ωb resulting from the coher- where θ describes the hybridization of the normal modes; ent magnomechanical interaction. A series of data was for maximally hybridized modes θ = π/4. taken at atmospheric pressure for various pump powers By varying the static magnetic field, and therefore the and magnon-photon detunings, and fit using the theory magnon-photon detuning, and measuring the cooperativ- presented in Ref. [22]; described in Appendix B. From ity at each detuning using Eq. (B1) it is possible to de- this we were able to extract the single magnon-phonon termine the magnon-phonon coupling rate. The cooper- 0 ativity as a function of the lower normal mode frequency coupling rate gmb/2π = 4.38 mHz, the mechanical fre- quency Ωb = 12.6270 MHz, and the intrinsic mechanical is shown in Fig. 8. decay rate Γb/2π = 286 Hz. Fig. 7(c) shows that there exists a second, higher fre- quency mechanical mode. This mode was not observed Appendix C: Magnon Heating in the homodyne mechanical detection scheme, described in the main text, except at the highest drive powers. In Fig. 9, we observe a drive-dependent phonon fre- However, at those drive powers magnon nonlinearities quency shift; however, this frequency is not entirely due resulted in the system becoming bistable and are there- to dynamical magnon backaction. The observed fre- fore not included in our analysis [70]. The second me- quency shift is mainly influenced by the heating of the chanical mode has a frequency Ωb = 12.637 MHz, and 0 YIG sample by the microwave drive. This is supported by a coupling rate gmb/2π = 2.41 mHz. The difference in the results plotted in Fig. 3(b,c); for zero drive detuning the transparency window height can be attributed to the in the triple-resonance scenario Eq. (5) predicts zero fre- increased damping rate, which is approximately an or- quency shift. However, both cases result in a softening of der of magnitude larger than the lower-frequency mode. the mechanical motion, likely due to a modification of the Numerical simulations reveal that clamping causes the Young’s modulus due to heating [71]. Magnetostrictive S1,2,0 mode to split into two nearly degenerate modes, resulting in the two modes observed here.

Appendix B: MMIT Theory

Here, we will outline the equations used to fit the mag- nomechanically induced transparency data; however, for a full description of this theory see the supplementary material included with Ref. [22]. For the data presented in Section A, the drive was detuned from the lower nor- mal mode by the phonon frequency, ωd = ω − Ωb. In this scenario the transparency window will have− a peak reflectivity defined as

κ−,e 1 − 2 κ + C r = − . (B1) 1 + C The transparency window can be seen in Fig. 7(b,c). Here κ and κ ,e are the linewidth and external cou- pling rate− of the− lower normal mode, respectively, and C is the cooperativity. The cooperativity can be shown to have the form, FIG. 8. Magnomechanical cooperativity as a function of the lower normal mode frequency. For each measurement the 0 2 4 2 drive tone is detuned by the phonon frequency from the lower 4P(g ) κ1 sin (θ) cos (θ) C ≈ mb , (B2) normal mode, ω = ω − Ω . Blue circles are experimen- 2 2 2 d − b ~ωdΩbΓb κ cos (θ) + γm sin (θ) tally determined using Eq. (B1), and the dotted orange line 0 where P is the microwave power at the experimental de- is a numerical fit using Eq. (B2), where gmb is the only fit vice. All losses have been carefully calibrated to ensure parameter. 10 materials have been shown to exhibit a magnetic field de- tial pressure helium environment, and by applying the pendant Young’s modulus due to the ∆E-effect [72, 73]. drive tone to the ‘photon-like’ normal mode. Therefore, temperature-dependent modifications of the Additionally, in Fig. 9 one may expect that the signal- saturation magnetization and therefore internal static to-noise ratio should be symmetric about zero detun- magnetic field may cause the observed frequency shift ing, with variations resulting from the homodyne detec- [74]. Furthermore, for all detunings presented in Fig. 9, tion sensitivity. However, the primary determining fac- we observe a softening of the effective spring constant. tor of the signal-to-noise ratio is related to the triple- However, dynamical backaction does not predict soften- resonance condition. For example, in Fig. 9(a), and (c), ing in all cases. Specifically, in Fig. 9(c), we expect the the magnon-photon detuning was zero, such that the nor- magnon-spring effect to result in a hardening of the me- mal modes were fully hybridized, see Fig. 2(b). In this chanics. situation, the normal mode spacing is slightly smaller In all cases, the frequency shift follows the normal than the phonon frequency 2gam/2π = 10.86 MHz and mode shape; as the normal mode depth increases, the Ωb/2π = 12.627 MHz. The asymmetry in the signal-to- circulating power increases, and the frequency shift in- noise ratio in Fig. 9(a) can be understood by consider- creases. Moreover, the amplitude of the frequency shift ing the mechanical sideband created on the microwave is much larger than what is predicted by the calculated carrier. For negative drive detunings – relative to the phonon self-energy in Eq. (5) (see also [37]). The max- normal mode central frequency, i.e. ∆d < 0 – the me- imum frequency shift predicted is approximately an or- chanical sideband is at a lower frequency than the lower der of magnitude smaller than the one observed. Thus, normal mode and is therefore not resonantly enhanced. in this situation, any frequency shift resulting from the Conversely, a positive drive detuning (∆d > 0) results magnon-spring effect is overwhelmed by the additional in the lower mechanical sideband lying directly within frequency shift resulting from heating. the lower normal mode, resonantly enhancing this scat- The temperature dependence of the mechanical fre- ting process and improving the signal to noise ratio. A quency was measured by heating the experimental setup. similar argument can be made for the scattering process We observe a dependence of approximately −715 Hz/K regarding Fig. 9(c), resulting in a resonant enhancement for the frequency shift, suggesting that at the highest of the scattering process for negative detunings. drive powers, the temperature of the sphere was increased by approximately 1 kelvin due to the microwave drive. As described in the main text, this heating can be mitigated Appendix D: Phonon temperature by using lower drive powers, placing the sphere in a par- We now derive the effective phonon temperature and the approximations that lead to Eq. (7). The first step (a) (b) is to consider the linearized theory by writing the to- tal Hamiltonian in terms of the fluctuations defined as aˆ = hai + δaˆ,m ˆ = hmi + δmˆ and ˆb = hbi + δˆb, where the terms h·i denote the steady-state average values. We then discard terms in the Hamiltonian containing more than two fluctuation terms and obtain the time domain Langevin equations of motion, which in the frequency domain read [37]:

1 X √ ˆ (c) (d) χa− [ω]δaˆ[ω] = −igamδmˆ [ω] + κjξj[ω] j=int, 1, 2 1 χ− [ω]δmˆ [ω] = −ig δaˆ[ω] − ig (δˆb[ω] + δˆb†[ω]) m √am mb + γmηˆ[ω] 1 ˆ χb− [ω]δb[ω] = −i(gmbδmˆ †[ω] + gmb∗ δmˆ [ω]) p + Γbζˆ[ω], (D1) where the constants are defined within the main text. The total cavity decay rate κ = κint + κ1 + κ2 is a FIG. 9. Mechanical spectrum. (a-d) Mechanical power spec- sum of an intrinsic cavity decay κint and the decay rates into each of the ports κ , . Accordingly, the total tral density. Detuning is relative to the specific mode shown 1 2 √ within the inset. In plots (a) and (c) the probe power is held noise acting in the cavity is given by the terms κjξˆj, constant at 0.54 mW. In plots (b) and (d) the probe power is which includes the noises from each of the possible de- held constant at 0.27 mW. All data presented was obtained cay channels ξˆint,1,2. Within our convention, the relation with the experimental setup in atmospheric conditions. between time domain and frequency domain operators 11

R iωt are Oˆ(t) = dωe− Oˆ[ω], and it should be noted that linked to the ability of the oscillator to emit/absorb en- R iωt Oˆ†[ω] ≡ dte Oˆ†(t) = (Oˆ[−ω])†. ergy [66, 75]. For our analysis it is sufficient to consider a white ther- Turning our attention to the full problem, we solve mal noise model described by the correlations [66]: the linear Langevin equations (D1) and obtain for the phonon mode ˆ ˆ hξi†[ω0]ξi[ω]i = 2πnth,aδ(ω + ω0), √ 1  ˆ X ˆ χb− [ω] − iΣ[ω] δb[ω] = A[ω] κiξi[ω] hξˆi[ω0]ξˆ†[ω]i = 2π(n , + 1)δ(ω + ω0), i th a i √ hηˆ†[ω0]ˆη[ω]i = 2πnth,mδ(ω + ω0), X + A˜[ω] κiξˆ†[ω] (D2) (D7) hηˆ[ω0]ˆη†[ω]i = 2π(n + 1)δ(ω + ω0), i th,m √ √ + B[ω] γ ηˆ[ω] + B˜[ω] γ ηˆ†[ω] hζˆ†[ω0]ζˆ[ω]i = 2πnth,bδ(ω + ω0), m m p p ˆ C˜ ˆ hζˆ[ω0]ζˆ†[ω]i = 2π(nth,b + 1)δ(ω + ω0). + Γbζ[ω] + [ω] Γbζ†[ω], The thermal occupancy of the baths is given by Bose- where Σ[ω] is the self energy given in the main text Einstein distributions Eq. (5) whose derivation is discussed in detail in Ref. [37]. The frequency-dependent coefficients have complicated 1 n = , and not elucidating forms that depend on the suscepti- th,(a,m) h ω i exp ~ (a,m) − 1 bilities and on the couplings. We then rewrite the above kB TBath (D3) expression as 1 nth,(b) = h i ,   exp ~Ωb − 1 ΓTot[ω] p kB TBath i(Ω˜ b[ω] − ω) + δˆb[ω] = Γ [ω]Υ[ˆ ω], (D8) 2 Tot where kB is the Boltzmann constant, TBath is the bath temperature (assumed to be the same for all the modes), where and we have assumed that the occupancy of the baths Ω˜ b[ω] = Ωb + δΩb[ω] = Ωb − ReΣ[ω], for all decay channels are the same and given in terms (D9) Γ [ω] = Γ + Γ [ω] = Γ + 2ImΣ[ω], of the cavity frequency ωa. This last assumption can be Tot b mag b readily generalized. are the phonon frequency and decay rates corrected by Before deriving the effective temperature of the phonon ˆ mode in the coupled and driven system described by the the self energy, and Υ[ω] is the combination of noises Eq. (D1), let us first consider why the noise spectral den- appearing in the right-hand side of Eq. (D7) divided by sity of the phonon mode provides information about that the total phonon decay rate. mode’s temperature. For that, we consider the simpler Recalling the relation of the spectral density Sδb†δb[ω] problem in which there is no magnon-phonon coupling. to the thermal number of phonons and using the noise In this case the phonon mode is driven only by thermal relations Eq. (D2), we obtain [76] noise, and the phonon component of Eq. (D1) reads ΓTot[ω]neff,b[ω] S † [ω] = , (D10) δb δb Γ2 [ω] p ˜ 2 Tot δˆb[ω] = χb[ω] Γbζˆ[ω], (ω + Ωb[ω]) + (D4) 4 ˆ p ˆ δb†[ω] = χb∗[−ω] Γbζ†[ω]. where the effective phonon number is given in terms of the frequency dependent coefficients of Eq. (D7) and of We then can consider the spectral density given by the thermal occupancy of the baths as Z ∞ iωt ˆ ˆ 2 Sδb†δb[ω] = dte hδb†(t)b(0)i, ΓTot[ω]neff,b[ω] = κ|A[−ω]| nth,a −∞ (D5) 2 Z + κ|A˜[−ω]| (nth,a + 1) ∞ dω0 = hδˆb†[ω]δˆb[ω0]i. 2 2π + γm|B[−ω]| nth,m (D11) −∞ |B˜ − |2 For the simple uncoupled case, the solutions of the fre- + γm [ ω] (nth,m + 1) 2 quency domain equations combined with the noise corre- + Γbnth,b + Γb|C˜[−ω]| (nth,b + 1). lations Eq. (D2) yields the simple relation, In deriving the above formula we have assumed that all Γbnth,b the decay channels of the cavity mode are related to ther- Sδb†δb[ω] = 2 , (D6) 2 Γb mal baths at the same temperature. The effective tem- (ω + Ωb) + 4 perature for the phonon mode is then given by which is given in terms of the bath occupancy nth,b and ˜    1 thus the temperature of the mode. The noise spectral ~Ωb[ω] neff,b[ω] + 1 − Teff,b[ω] = ln . (D12) density Sδb†δb[−Ωb] and its counterpart Sδbδb† [Ωb] are kB neff,b[ω] 12

We can make further approximations to the effective tem- our room temperature experiment, we can write perature formula. Since ω(a,m) are three orders of mag- nitude larger than the phonon frequency, nth,(a,m)  kBTeff,b[ω] neff,b[ω] ∼ , nth,b and we can discard the terms ∝ nth,(a,m) and Ω˜ [ω] ~ b (D16) ∝ n , a,m + 1. Furthermore, at room temperature th ( ) kBTBath nth,b  1, and Eq. (D11) simplifies to nth,(b) ∼ , ~Ωb 2 ΓTot[ω]neff,b[ω] = Γb(1 + |C˜[−ω]| )nth,b. (D13) and since the phonon frequency shift δΩb  Ωb, we can ˜ The remaining frequency-dependent coefficient is given take Ωb ∼ Ωb. Within those approximations, the phonon explicitly by mode effective temperature reads |g |2χ [−ω] (Ξ [−ω] − Ξ[ω]) Γ T (1 + |C˜[−ω]|2) C˜ mb b∗ ∗ b Bath [ω] = 2 , (D14) Teff,b[ω] ≈ . (D17) 1 + |gmb| χb∗[−ω] (Ξ∗[−ω] − Ξ[ω]) ΓTot[ω]

1 2 1 where Ξ[ω] = [χm− [ω] + gamχa[ω]]− , and Eq. (D14) A rough estimate of the effective temperature can be 2 2 depends on the driving power only through |gmb| = made by discarding the term ∝ |C˜[−ω]| , such that 0 2 |gmbhmi| where [37]: √ ΓbTBath id κext Teff,b[ω] ≈ . (D18) h i ΓTot[ω] m = 2 . (D15) (i∆a − κ/2)(i∆m − γm/2) + gam This has a familiar form of the effective phonon tempera- ˜ 2 At low powers, the contribution ∝ |C[−ω]| can be ture in driven optomechanical systems (c.f. [1, 76]), and safely discarded, but as the power increases, since χb[ω] in our case it is valid for low driving powers only. This is sharply peaked around Ωb, this contribution becomes can be seen in Fig. 10 which shows the effective tem- ˜ 2 prominent. In fact, |C[−ω]| goes from zero to its maxi- perature at the phonon frequency for the red detuning mum value of one. In the limit kBT  ~ωa,b,m, valid for scheme (driving at the lower normal mode) as given by Eq. (D12) and as given by Eq. (D18) for the parameters values corresponding to Fig. 6. The approximation is (a) 300 good for small driving powers but at powers larger than 250250 0.1 mW, the difference between the full formula and the Full Formula ∼ 200200 rough approximation can be 40 K. Thus, even though Approximation Eq. (D18) is a simple and practical approximation, it 150 Rough leads to an underestimate of the phonon effective tem- 100 Mode 100 perature. 5050 The improved approximation Eq. (D17) adds the con- Temperature (K) Temperature 00 tribution ∝ |C˜[−ω]|2 which, for a given detuning and at 5 4 3 2 1 1010-5 1010-4 0.00110 0.01010 0.10010 1 Ωb is a function of the driving power P given by (b) Driving Power (W) 2 2 2 P |C| |C˜[−Ωb]| = , (D19) 40 1 + 2PRe [C] + P2|C|2 30

20 00.08.08 10

erence (K) di ff erence 00.06.06 00 -55 -44 3 2 1 ective Temperature E ff ective Temperature 1010 10 0.00110 0.01010 0.10010 11 00.04.04 Driving Power (W) 00.02.02 erence (K) di ff erence FIG. 10. Comparison between the full formula for the ef- 0 0.00 ective Temperature E ff ective Temperature -5 5 -4 4 3 2 1 fective temperature Eq. (D12) and the approximate estimate 1010 1010 0.00110 0.01010 0.10010 1 Eq. (D18). (a) Effective temperature as given by equation Eq. (D12) (dashed line) and as given by Eq. (D18) (contin- Driving Power (W) uous line) as a function of the driving power. (b) Difference between the effective temperatures (Eq. (D12) - Eq. (D18)) as FIG. 11. Difference between the effective temperature given a function of the driving powers. For both plots we use the ex- by the full formula Eq. (D12) and the improved approxima- perimental parameters and driving scheme corresponding to tion Eq. (D17). For this plot we have used the experimental the results of Fig 6, and the bath temperature is TBath = 295 parameters and driving scheme corresponding to the results K. of Fig. 6, and the bath temperature is TBath = 295 K. 13 where C is a complex number that can be calculated mula Eq. (D12), and the approximation including the with Eq. (D14) and Eq. (D15) for a given set of pa- aforementioned power dependent factor Eq. (D17). In rameters. For conciseness, in the main text we define this case, the maximum difference is now ∼ 0.08 K, for 2 |C˜[−Ωb]| as β(P). Figure 11 shows the difference be- driving powers corresponding to temperatures of ∼ 150 tween the effective temperature obtained by the full for- K, and thus this approximation is more suitable for analysing the experimental data.

[1] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, [15] D. Lachance-Quirion, Y. Tabuchi, S. Ishino, A. Noguchi, Cavity optomechanics, Rev. Mod. Phys. 86, 1391 (2014). T. Ishikawa, R. Yamazaki, and Y. Nakamura, Resolving [2] D. Lachance-Quirion, Y. Tabuchi, A. Gloppe, K. Us- quanta of collective spin excitations in a millimeter-sized ami, and Y. Nakamura, Hybrid quantum systems based ferromagnet, Sci. Adv. 3, e1603150 (2017). on magnonics, Appl. Phys. Express 12, 070101 (2019). [16] Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Ya- [3] D. D. Awschalom, C. Du, R. He, F. Heremans, A. Hoff- mazaki, K. Usami, and Y. Nakamura, Coherent coupling mann, J. Hou, H. Kurebayashi, Y. Li, L. Liu, V. Novosad, between a ferromagnetic magnon and a superconducting et al., Quantum Engineering With Hybrid Magnonics qubit, Science 349, 405 (2015). Systems and Materials, IEEE T. Quantum Eng. , 1 [17] Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Ya- (2021). mazaki, K. Usami, and Y. Nakamura, Quantum magnon- [4] Y. Li, W. Zhang, V. Tyberkevych, W.-K. Kwok, A. Hoff- ics: The magnon meets the superconducting qubit, C. R. mann, and V. Novosad, Hybrid magnonics: Physics, cir- Phys. 17, 729 (2016). cuits, and applications for coherent information process- [18] R. Hisatomi, A. Osada, Y. Tabuchi, T. Ishikawa, ing, J. Appl. Phys. 128, 130902 (2020). A. Noguchi, R. Yamazaki, K. Usami, and Y. Nakamura, [5] M. S. Ebrahimi, A. Motazedifard, and M. Bagheri, Ultra- Bidirectional conversion between microwave and light via precision single-quadrature quantum magnetometry in ferromagnetic magnons, Phys. Rev. B 93, 174427 (2016). cavity electromagnonics, arXiv:2011.06081 (2020). [19] N. Zhu, X. Zhang, X. Han, C.-L. Zou, C. Zhong, C.-H. [6] N. Crescini, D. Alesini, C. Braggio, G. Carugno, Wang, L. Jiang, and H. X. Tang, Waveguide cavity op- D. D’Agostino, D. Di Gioacchino, P. Falferi, U. Gam- tomagnonics for microwave-to-optics conversion, Optica bardella, C. Gatti, G. Iannone, et al., Axion search with 7, 1291 (2020). a quantum-limited ferromagnetic haloscope, Phys. Rev. [20] J. Xu, C. Zhong, X. Han, D. Jin, L. Jiang, and X. Zhang, Lett. 124, 171801 (2020). Floquet Cavity Electromagnonics, Phys. Rev. Lett. 125, [7] T. Ikeda, A. Ito, K. Miuchi, J. Soda, H. Kurashige, 237201 (2020). D. Lachance-Quirion, Y. Nakamura, and Y. Shikano, [21] Y.-P. Wang, J. W. Rao, Y. Yang, P.-C. Xu, Y. S. Gui, Axion search with quantum nondemolition detection of B. M. Yao, J. Q. You, and C.-M. Hu, Nonreciprocity and magnons, arXiv:2102.08764 (2021). unidirectional invisibility in cavity magnonics, Phys. Rev. [8] G. Flower, J. Bourhill, M. Goryachev, and M. E. To- Lett. 123, 127202 (2019). bar, Broadening frequency range of a ferromagnetic ax- [22] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Cavity ion haloscope with strongly coupled cavity–magnon polari- magnomechanics, Sci. Adv. 2, e1501286 (2016). tons, Phys. Dark Universe 25, 100306 (2019). [23] H.-J. Cheng, S.-J. Zhou, J.-X. Peng, A. Kundu, H.-X. [9] H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, M. Greifen- Li, L. Jin, and X.-L. Feng, Tripartite entanglement in a stein, A. Marx, R. Gross, and S. T. B. Goennenwein, Laguerre–Gaussian rotational-cavity system with an yt- High cooperativity in coupled microwave resonator ferri- trium iron garnet sphere, J. Opt. Soc. Am. B 38, 285 magnetic insulator hybrids, Phys. Rev. Lett. 111, 127003 (2021). (2013). [24] J. M. Nair and G. Agarwal, Deterministic quantum en- [10] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Strongly tanglement between macroscopic ferrite samples, Appl. coupled magnons and cavity microwave photons, Phys. Phys. Lett. 117, 084001 (2020). Rev. Lett. 113, 156401 (2014). [25] J. Li and S. Groeblacher, Entangling the vibrational [11] Y. Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, modes of two massive ferromagnetic spheres using cavity K. Usami, and Y. Nakamura, Hybridizing ferromagnetic magnomechanics, Quantum Sci. Tech. 6, 024005 (2021). magnons and microwave photons in the quantum limit, [26] J. Li and S.-Y. Zhu, Entangling two magnon modes via Phys. Rev. Lett. 113, 083603 (2014). magnetostrictive interaction, New J. Phys. 21, 085001 [12] M. Goryachev, W. G. Farr, D. L. Creedon, Y. Fan, (2019). M. Kostylev, and M. E. Tobar, High-cooperativity cav- [27] Z.-B. Yang, J.-S. Liu, A.-D. Zhu, H.-Y. Liu, and R.-C. ity QED with magnons at microwave frequencies, Phys. Yang, Nonreciprocal Transmission and Nonreciprocal En- Rev. Appl. 2, 054002 (2014). tanglement in a Spinning Microwave Magnomechanical [13] C. A. Potts and J. P. Davis, Strong magnon–photon cou- System, Ann. Phys. 532, 2000196 (2020). pling within a tunable cryogenic microwave cavity, Appl. [28] J. Li, S.-Y. Zhu, and G. S. Agarwal, Magnon-photon- Phys. Lett. 116, 263503 (2020). phonon entanglement in cavity magnomechanics, Phys. [14] D. Lachance-Quirion, S. P. Wolski, Y. Tabuchi, S. Kono, Rev. Lett. 121, 203601 (2018). K. Usami, and Y. Nakamura, Entanglement-based single- [29] J. Li, Y.-P. Wang, J. Q. You, and S.-Y. Zhu, Squeez- shot detection of a single magnon with a superconducting ing Microwave Fields via Magnetostrictive Interaction, qubit, Science 367, 425 (2020). arXiv:2101.02796 (2021). 14

[30] J. Li, S.-Y. Zhu, and G. S. Agarwal, Squeezed states of [48] U. Deli´c,M. Reisenbauer, K. Dare, D. Grass, V. Vuleti´c, magnons and phonons in cavity magnomechanics, Phys. N. Kiesel, and M. Aspelmeyer, Cooling of a levitated Rev. A 99, 021801(R) (2019). nanoparticle to the motional quantum ground state, Sci- [31] W. Zhang, D.-Y. Wang, C.-H. Bai, T. Wang, S. Zhang, ence 367, 892 (2020). and H.-F. Wang, Generation and transfer of squeezed [49] P. C. Fletcher and R. O. Bell, Ferrimagnetic resonance states in a cavity magnomechanical system by two-tone modes in spheres, J. Appl. Phys 30, 687 (1959). microwave fields, Opt. Express 29, 11773 (2021). [50] T. L. Gilbert, A phenomenological theory of damping in [32] B. Sarma, T. Busch, and J. Twamley, Cavity magnome- ferromagnetic materials, IEEE T. Magn. 40, 3443 (2004). chanical storage and retrieval of quantum states, New J. [51] J. Nemarich, Contribution of the Two-Magnon Process to Phys. (2021). Magnetostatic-Mode Relaxation, Phys. Rev. 136, A1657 [33] X. Li, W.-X. Yang, T. Shui, L. Li, X. Wang, and Z. Wu, (1964). Phase control of the transmission in cavity magnome- [52] S. Klingler, H. Maier-Flaig, C. Dubs, O. Surzhenko, chanical system with magnon driving, J. Appl. Phys. 128, R. Gross, H. Huebl, S. T. B. Goennenwein, and 233101 (2020). M. Weiler, Gilbert damping of magnetostatic modes in a [34] Z.-B. Yang, J.-S. Liu, A.-D. Zhu, H.-Y. Liu, and R.- yttrium iron garnet sphere, Appl. Phys. Lett. 110, 092409 C. Yang, Nonreciprocal Transmission and Entanglement (2017). in a cavity-magnomechanical system, arXiv:2101.09931 [53] H. Keshtgar, M. Zareyan, and G. E. Bauer, Acoustic (2021). parametric pumping of spin waves, Solid State Commu. [35] J. Zhao, L. Wu, T. Li, Y.-x. Liu, F. Nori, Y. Liu, and 198, 30 (2014). J. Du, Phase-controlled pathway interferences and switch- [54] E. Callen, Magnetostriction, J. Appl. Phys. 39, 519 able fast-slow light in a cavity-magnon polariton system, (1968). Phys. Rev. Appl. 15, 024056 (2021). [55] A. C. Gurevich and G. A. Melkov, Magnetization Oscil- [36] C. Kong, B. Wang, Z.-X. Liu, H. Xiong, and Y. Wu, Mag- lations and Waves, Vol. 1 (CRC Press, 1996). netically controllable slow light based on magnetostrictive [56] E. G. Spencer and R. C. LeCraw, Magnetoacoustic res- forces, Opt. Express 27, 5544 (2019). onance in yttrium iron garnet, Phys. Rev. Lett. 1, 241 [37] C. A. Potts, V. A. S. V. Bittencourt, S. Viola Kusminskiy, (1958). and J. P. Davis, Magnon-Phonon Quantum Correlation [57] R. C. LeCraw, E. G. Spencer, and E. I. Gordon, Ex- Thermometry, Phys. Rev. Appl. 13, 064001 (2020). tremely low loss acoustic resonance in single-crystal gar- [38] M.-S. Ding, X.-X. Xin, S.-Y. Qin, and C. Li, Enhanced net spheres, Phys. Rev. Lett. 6, 620 (1961). entanglement and steering in PT-symmetric cavity mag- [58] See http://www.ferrisphere.com/ for information on nomechanics, Opt. Commun. , 126903 (2021). purchasing YIG spheres. (2020), accessed: 2020-05-05. [39] L. Wang, Z.-X. Yang, Y.-M. Liu, C.-H. Bai, D.-Y. Wang, [59] K. V. Kepesidis, S. D. Bennett, S. Portolan, M. D. Lukin, S. Zhang, and H.-F. Wang, Magnon Blockade in a PT- and P. Rabl, Phonon cooling and lasing with nitrogen- Symmetric-Like Cavity Magnomechanical System, Ann. vacancy centers in diamond, Phys. Rev. B 88, 064105 Phys. 532, 2000028 (2020). (2013). [40] Z.-X. Yang, L. Wang, Y.-M. Liu, D.-Y. Wang, C.-H. Bai, [60] U. Kemiktarak, M. Durand, M. Metcalfe, and J. Lawall, S. Zhang, and H.-F. Wang, Ground state cooling of mag- Mode competition and anomalous cooling in a multimode nomechanical resonator in PT-symmetric cavity mag- phonon , Phys. Rev. Lett. 113, 030802 (2014). nomechanical system at room temperature, Front. Phys. [61] S. Wang and T.-l. Hsu, Spin-wave experiments: paramet- 15, 1 (2020). ric excitation of acoustic waves and mode-locking of spin [41] M. Wang, D. Zhang, X.-H. Li, Y.-Y. Wu, and Z.-Y. Sun, waves, Appl. Phys. Lett. 16, 111 (1970). Magnon chaos in PT-symmetric cavity magnomechanics, [62] K. Vahala, M. Herrmann, S. Kn¨unz, V. Batteiger, IEEE Photonics J. 11, 1 (2019). G. Saathoff, T. H¨ansch, and T. Udem, A phonon laser, [42] I. Mahboob, K. Nishiguchi, A. Fujiwara, and H. Yam- Nat. Phys. 5, 682 (2009). aguchi, Phonon lasing in an electromechanical resonator, [63] V. B. Braginsky, Resolution in macroscopic measure- Phys. Rev. Lett. 110, 127202 (2013). ments: progress and prospects, Sov. Phys. Usp. 31, 836 [43] M.-S. Ding, L. Zheng, and C. Li, Phonon laser in a cavity (1988). magnomechanical system, Sci. Rep. 9, 15723 (2019). [64] P. Bushev, D. Rotter, A. Wilson, F. Dubin, C. Becher, [44] T. Rocheleau, T. Ndukum, C. Macklin, J. B. Hertzberg, J. Eschner, R. Blatt, V. Steixner, P. Rabl, and P. Zoller, A. A. Clerk, and K. Schwab, Preparation and detection of Feedback cooling of a single trapped ion, Phys. Rev. Lett. a mechanical resonator near the ground state of motion, 96, 043003 (2006). Nature 463, 72 (2010). [65] M. Poggio, C. L. Degen, H. J. Mamin, and D. Rugar, [45] O. Romero-Isart, Quantum superposition of massive ob- Feedback cooling of a cantilever’s fundamental mode be- jects and collapse models, Phys. Rev. A 84, 052121 low 5 mK, Phys. Rev. Lett 99, 017201 (2007). (2011). [66] A. A. Clerk, M. H. Devoret, S. M. Girvin, F. Marquardt, [46] O. Romero-Isart, A. C. Pflanzer, F. Blaser, and R. J. Schoelkopf, Introduction to quantum noise, R. Kaltenbaek, N. Kiesel, M. Aspelmeyer, and J. I. measurement, and amplification, Rev. Mod. Phys. 82, Cirac, Large quantum superpositions and interference of 1155 (2010). massive nanometer-sized objects, Phys. Rev. Lett. 107, [67] S. Kotler, G. A. Peterson, E. Shojaee, F. Lecocq, K. Ci- 020405 (2011). cak, A. Kwiatkowski, S. Geller, S. Glancy, E. Knill, [47] L. Childress, M. P. Schmidt, A. D. Kashkanova, C. D. R. W. Simmonds, et al., Direct observation of determin- Brown, G. I. Harris, A. Aiello, F. Marquardt, and J. G. E. istic macroscopic entanglement, Science 372, 622 (2021). Harris, Cavity optomechanics in a levitated helium drop, [68] S. Weis, R. Rivi`ere,S. Del´eglise,E. Gavartin, O. Arcizet, Phys. Rev. A 96, 063842 (2017). A. Schliesser, and T. J. Kippenberg, Optomechanically 15

induced transparency, Science 330, 1520 (2010). [73] J. J. Scheidler, V. M. Asnani, and M. J. Dapino, Dynam- [69] A. H. Safavi-Naeini, T. M. Alegre, J. Chan, M. Eichen- ically tuned magnetostrictive spring with electrically con- field, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, and trolled stiffness, Smart Mater. Struct. 25, 035007 (2016). O. Painter, Electromagnetically induced transparency and [74] P. Hansen, P. R¨oschmann, and W. Tolksdorf, Saturation slow light with optomechanics, Nature 472, 69 (2011). magnetization of gallium-substituted yttrium iron garnet, [70] Y.-P. Wang, G.-Q. Zhang, D. Zhang, T.-F. Li, C.-M. Hu, J. Appl. Phys. 45, 2728 (1974). and J. Q. You, Bistability of cavity magnon polaritons, [75] F. Marquardt, J. P. Chen, A. A. Clerk, and S. M. Girvin, Phys. Rev. Lett. 120, 057202 (2018). Quantum Theory of Cavity-Assisted Sideband Cooling of [71] D. F. Gibbons and V. G. Chirba, Acoustical Loss and Mechanical Motion, Phys. Rev. Lett. 99, 093902 (2007). Young’s Modulus of Yttrium Iron Garnet, Phys. Rev. [76] A. H. Safavi-Naeini, J. Chan, J. T. Hill, S. Gr¨oblacher, 110, 770 (1958). H. Miao, Y. Chen, M. Aspelmeyer, and O. Painter, Laser [72] K. Honda and T. Terada, II. On the change of elastic noise in cavity-optomechanical cooling and thermometry, constants of ferromagnetic substances by magnetization, New J. Phys. 15, 035007 (2013). London Edinburgh Philos. Mag. J. Sci. 13, 36 (1907).