This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Phononic-crystal cavity magnomechanics

D. Hatanaka daiki.hatanaka.hz@hco.ntt.co.jp    M. Asano    H. Okamoto    H. Yamaguchi NTT Basic Research Laboratories, NTT Corporation, Atsugi-shi, Kanagawa 243-0198, Japan
Abstract

Establishing a way to control magnetic dynamics and elementary excitations (magnons) is crucial to fundamental physics and the search for novel phenomena and functions in magnetic solid-state systems. Electromagnetic waves have been developed as means of driving and sensing in magnonic and spintronics devices used in magnetic spectroscopy, non-volatile memory, and information processors Jungwirth et al. (2012); Chumak et al. (2015); Tabuchi et al. (2015); Li et al. (2020); Awschalom et al. (2021). However, their millimeter-scale wavelengths and undesired cross-talk have limited operation efficiency and made individual control of densely integrated magnetic systems difficult. Here, we utilize acoustic waves (phonons) to control magnetic dynamics in a miniaturized phononic crystal micro-cavity and waveguide architecture. We demonstrate acoustic pumping of localized ferromagnetic magnons, where their back-action allows dynamic and mode-dependent modulation of phononic cavity resonances. The phononic crystal platform enables spatial driving, control and read-out of tiny magnetic states and provides a means of tuning acoustic vibrations with magnons. This alternative technology enhances the usefulness of magnons and phonons for advanced sensing, communications and computation architectures that perform transduction, processing, and storage of classical and quantum information Chumak et al. (2015); Awschalom et al. (2021).

Refer to caption
Figure 1: PnC cavity magnomechanics. (a) Schematic diagram of PnC magnomechanical system and measurement configuration. The device consists of a GaAs/Al0.7Ga0.3As heterostructure and the PnC is formed in a free-standing GaAs membrane. Rayleigh SAWs are excited by applying an alternating voltage via the piezoelectric effect to inter-digit transducers (IDT) and they transform into asymmetric Lamb waves in the PnC membrane. The acoustic waves excite a PnC cavity resonance through a line-defect waveguide, and the resonant vibrations are measured by an optical interferometer at room temperature and in a moderate vacuum. A ferromagnetic Ni film deposited on the cavity sustains magnon oscillations that are driven by cavity vibrations via magnetostriction. (b) Electron microscope images of the magnomechanical cavity-waveguide geometry (top) and cross-section (bottom). The scale bar is 4 μ\mum. The cavity is formed by removing two air holes from the phononic lattice. The Ni film on the defect is false-colored in yellow. The inset zooms in on the snowflake lattice, and the red dashed line denotes the unit structure. (c) Unit cell of PnC lattice designed with a=a= 4.0 μ\mum, b=b= 3.6 μ\mum, w=w= 1.0 μ\mum and t=t= 1.0 μ\mum. (d) Dispersion relation calculated by FEM (COMSOL Multiphysics). A complete phononic bandgap forms between 0.5 and 0.8 GHz, except around 0.61-0.62 GHz (highlighted in yellow). (e) Spectral response of PnC cavity measured with an optical interferometer. From left to right: well-defined acoustic resonance peaks appear at 0.593 GHz (L mode), 0.552 GHz (SL mode), and 0.581 GHz (S mode) with quality-factors Qa=Q_{a}= 550, 680 and 1,540. (f) Experimental (top) and simulated (bottom) modal shape of displacement amplitude of L, SL and S modes in the left, middle and right panels.
Refer to caption
Figure 2: Magnon-phonon interaction. (a) Left: xyxy-coordinate system in which the external magnetic field (μ0Hex\mu_{0}H_{\rm ex}), magnetization (m3m_{3}), and in-plane anisotropic filed (BuB_{\rm u}) respectively have an angle of ϕh\phi_{\rm h}, ϕm\phi_{\rm m} and ϕu\phi_{\rm u} away from the xx-axis. The inset shows the alternative 11-, 22- and 33-axis coordinate system. The 11- and 22-axis define the magnetization precession plane normal to the 33-axis. The spin oscillation with frequency ωm\omega_{\rm m} and damping κm\kappa_{\rm m} is sustained by the Ni thin film. Right: Acoustic resonant vibrations in the PnC cavity whose frequency and quality factor change due to magnetostrictive coupling. (b) Equilibrium magnetization component projected to the field orientation at ϕh\phi_{\rm h} = 0 plotted as a function of the bias field (μ0Hex\mu_{0}H_{\rm ex}); the green and pink solid lines are experimental and calculated results. The magnetization versus μ0Hex\mu_{0}H_{\rm ex} curve was obtained by magneto-optical Kerr microscopy. (c) (top): Field dependence of the magnetization angle (ϕm\phi_{\rm m}), (bottom): spin-wave resonance frequency (ωm/(2π)\omega_{\rm m}/(2\pi)). The black solid line in the bottom panel is the acoustic resonant frequency (ωa/(2π)\omega_{\rm a}/(2\pi)). (d)-(f) Spatial distribution of strain components |ϵxxϵyy|\left|\epsilon_{\rm xx}-\epsilon_{\rm yy}\right| (left) and |2ϵxy|\left|2\epsilon_{\rm xy}\right| (right) in L (0.593 GHz, (d)), SL (0.552 GHz, (e)), and S (0.581 GHz) modes simulated by FEM. The color scales of the SL and S modes are normalized by the maximum strain of |2ϵxy|\left|2\epsilon_{\rm xy}\right|, and that of the L mode by |ϵxxϵyy|\left|\epsilon_{\rm xx}-\epsilon_{\rm yy}\right|. (g)-(i) Left: Field response of magnetoelastic mode coupling (VcV_{\rm c}) at ϕh\phi_{\rm h} = 0. Right: ϕh\phi_{\rm h} dependence of VcV_{\rm c} versus μ0Hex\mu_{0}H_{\rm ex} response.
Refer to caption
Figure 3: Magnetoelastic modulation of acoustic cavity resonances. (a) and (b) Field (μ0Hex\mu_{0}H_{\rm ex}) response of acoustic resonance frequency shift Δωa/(2π)\Delta\omega^{\prime}_{\rm a}/(2\pi) (left) and quality factor QaQ_{a}^{\prime} (right) in L and S modes. Here, Δωa=ωaωa0\Delta\omega^{\prime}_{\rm a}=\omega^{\prime}_{\rm a}-\omega_{\rm a0} where ωa0/(2π)\omega_{\rm a0}/(2\pi) is the frequency at |μ0Hex|\left|\mu_{0}H_{\rm ex}\right| = 60 mT. The solid circles are experimental data taken from a Lorentzian fitting to the measured cavity spectra. The solid lines are theoretical results based on equation (3).
Refer to caption
Figure 4: Magnetoelastic modulation of acoustic cavity to the field orientation. (a)-(d) Field angle (ϕh\phi_{\rm h}) response of the resonant frequency shift (Δωa\Delta\omega^{\prime}_{\rm a}) and normalized acoustic damping (κa\kappa^{\prime}_{\rm a}) of L and S modes, where the normalized acoustic damping is defined as κa=κa/κa0\kappa_{\rm a}^{\prime}=\kappa_{\rm a}/\kappa_{\rm a0} where κa0\kappa_{\rm a0} is the value at |μ0Hex|\left|\mu_{0}H_{\rm ex}\right| = 60 mT. The left and right panels of each figure are experimental and calculated results. Here, we used κa\kappa_{\rm a}^{\prime} but not QaQ_{\rm a}^{\prime} in order to compare only the magnetoelastic back-action effects between different ϕh\phi_{\rm h}.

Acoustic phonons are possibly means for controlling magnons on-chip because of their micro-/nanometer wavelengths (similar to those of magnons), low-loss property, and negligible cross-talk Li et al. (2021); Zhang et al. (2016); Kikkawa et al. (2016); Berk et al. (2019); An et al. (2020); Potts et al. (2021). The pioneering studies on magnomechanical technology used surface acoustic wave (SAW) devices Weiler et al. (2011); Dreher et al. (2012); Kobayashi et al. (2017); Sasaki et al. (2017); Herna´\acute{\rm a}ndez-Mi´\acute{\rm i}nguez et al. (2020); Xu et al. (2020); Kawada et al. (2021); Hatanaka et al. (2022). They succeeded in generating various magnetoelastic phenomena like acoustic spin pumping Kobayashi et al. (2017); Matsuo et al. (2013) and nonreciprocal transport Sasaki et al. (2017); Xu et al. (2020). However, their large cavity structures are unsuitable for system integration and have difficulty taking full advantage of phonons Hatanaka et al. (2022).
    A phononic crystal (PnC) is a promising platform that enables acoustic phonons to be guided and trapped in a tiny wavelength-scale acoustic cavity and waveguide Maldovan (2013); Benchabane et al. (2006); Mohammadi et al. (2009); Otsuka et al. (2013); Pourabolghasem et al. (2018); Baboly et al. (2018); Hatanaka and Yamaguchi (2020). Individual cavities are efficiently and finely driven and thereby can be used to control magnetic elements embedded in a PnC circuit. Moreover, the cavity sustains various spatial distributions of vibrational strains among multiple acoustic resonances, so it can be used to adjust the magnetoelastic effect Hatanaka and Yamaguchi (2021). We consider that PnCs will enable us to make full use of phonons in magnomechanical technology.
      Here, we demonstrate a PnC-based magnomechanical system sustained by cavity-waveguide coupled systems, as shown in Fig. 1(a). The nanomechanical vibrations confined in the cavity, excited through the waveguide, generate spin waves (magnons) in a nickel (Ni) film placed on its surface via magnetostriction. The acoustic spin pumping reversibly induces a frequency shift and damping modulation of the cavity resonances. Moreover, the magnon-phonon interaction can be tailored by selectively driving an appropriate cavity mode with specific strain distributions. PnC cavity magnomechanics is useful for on-chip control of magnons and phonons as well as their hybridized states (magnon polaron), and shows promise for extending the capabilities of classical and quantum information technologies.
    The PnC is fabricated in free-standing GaAs, as shown in Fig.1(b) (the details of the fabrication and structure are presented in the Methods.). It consists of a snowflake triangular lattice with full bandgaps between 0.5 and 0.8 GHz (Fig. 1(c) and 1(d)) Hatanaka and Yamaguchi (2020); Safavi-Naeini et al. (2014). Acoustic waves at frequencies within the bandgap propagate in a line-defect waveguide and drive a line-defect cavity. Measuring the cavity’s spectral response reveals three acoustic resonances (Fig. 1(e)). Their modal shapes exhibit complete confinement of the vibrations in the defect (top panels of Fig. 1(f)). Numerical calculations with the finite-element method (FEM) reproduce these modal shapes and verify the origin of the observed peaks. In this way, the PnC cavity can strongly confine vibrations that are remotely driven through the waveguide.
    The Ni film on the cavity surface has a magnetization whose precession is acoustically excited via magnetostriction. The magnetostrictive force that induces the precession is divided into two components μ0h1am\mu_{0}h^{\rm am}_{1} and μ0h2am\mu_{0}h^{\rm am}_{2} on the 11- and 22-axis, whose definitions are given in Fig. 2(a). The equilibrium magnetization axis (m3m_{3}) can be decomposed into in-plane xx and yy components and is at an angle (ϕm\phi_{\rm m}) from the waveguide direction xx. The out-of-plane magnetostrictive force (μ0h1am\mu_{0}h^{\rm am}_{1}) is negligibly small because out-of-plane shear strains such as ϵxz\epsilon_{\rm xz} and ϵyz\epsilon_{\rm yz} vanish in the Ni at that location, whereas in-plane force (μ0h2am\mu_{0}h^{\rm am}_{2}) can be expressed as Dreher et al. (2012)

μ0h2am=bam{(ϵxxϵyy)sin2ϕm2ϵxycos2ϕm)},\mu_{0}h^{\rm am}_{2}=b_{\rm am}\{(\epsilon_{\rm xx}-\epsilon_{\rm yy})\sin{2\phi_{\rm m}}-2\epsilon_{\rm xy}\cos{2\phi_{\rm m}})\}, (1)

where bamb_{\rm am} is the magnetostrictive coefficient and ϵij\epsilon_{\rm ij} is the vibrational strain component. Thus, the magnetostriction is governed by two major factors, vibrational strain (ϵxx\epsilon_{\rm xx}, ϵyy\epsilon_{\rm yy} and ϵxy\epsilon_{\rm xy}) and the magnetization direction (ϕm\phi_{\rm m}), whose contributions to the system are theoretically investigated below.
    The magnetization angle (ϕm\phi_{\rm m}) can be predicted from the magnetic free-energy density normalized by the saturation magnetization (MsM_{s}), given by Dreher et al. (2012)

G=μ0𝑯𝐞𝐱𝒎Bu(𝒎𝒖)2+Bdmz2,G=-\mu_{0}\bm{H_{\rm ex}}\cdot\bm{m}-B_{\rm u}(\bm{m}\cdot\bm{u})^{2}+B_{\rm d}m_{\rm z}^{2}, (2)

where μ0𝑯𝐞𝐱\mu_{0}\bm{H_{\rm ex}} and 𝒎\bm{m} are the external magnetic field and unit vector of magnetization, respectively. The thin-film structure of the Ni results in a perpendicular magnetic anisotropy (BdB_{\rm d}). The angle ϕm\phi_{\rm m} is determined by estimating the minimum of GG. In this calculation, the in-plane anisotropic field (BuB_{\rm u}) and its unit vector (𝒖\bm{u}) are introduced so as to reproduce the experimental magnetization curve (Fig. 2(b)). For instance, the response of ϕm\phi_{\rm m} as a function of μ0Hex\mu_{0}H_{\rm ex} at ϕh\phi_{\rm h} = 0 is shown in the top panel of Fig. 2(c). The magnetization is parallel to μ0Hex\mu_{0}H_{\rm ex}, i.e. ϕm=0\phi_{\rm m}=0^{\circ}, when the field strength stays in the high field region |μ0Hex||\mu_{0}H_{\rm ex}| >> 20 mT. However, it undergoes a rotation to ϕm=90\phi_{\rm m}=90^{\circ} in the low field region |μ0Hex|<|\mu_{0}H_{\rm ex}|< 20 mT before reversing. In this way, the magnetization experiences a rotation and reversal while sweeping μ0Hex\mu_{0}H_{\rm ex}. This change in magnetization determines the magnon resonance frequency (ωm/(2π)\omega_{\rm m}/(2\pi)). The field response is shown in the bottom panel of Fig. 2(c). The frequency monotonically decreases with decreasing μ0Hex\mu_{0}H_{\rm ex} and approaches the acoustic resonant frequency (ωa/(2π)\omega_{\rm a}/(2\pi)) at |μ0Hex||\mu_{0}H_{\rm ex}| = 7 mT, where the magnon-phonon frequency mismatch is minimized.
    Another aspect determining the magnetostriction is the spatial distribution and direction of vibration strains. The observed acoustic resonances can be decomposed into three strain components (Fig. 2(d)-2(f) shows the spatial profiles of the longitudinal (|ϵxxϵyy|\left|\epsilon_{\rm xx}-\epsilon_{\rm yy}\right|) and shear strain components (|2ϵxy|\left|2\epsilon_{\rm xy}\right|) of the resonances). Shear (longitudinal) strain is dominant in the resonances at 0.581 GHz (0.593 GHz), whereas both strains are comparable at 0.552 GHz. The desired strain distributions can thus be generated by selectively actuating an appropriate resonance. Hereafter, the shear- and longitudinal-strain modes will be labeled S and L, while the mode with comparable strains will be labeled SL.
    The magnetization dynamics and spatial strain profiles allow us to estimate the magnetostrictive coupling mode volume (VcV_{\rm c}), which characterizes the interaction efficiency. determined by ϕm\phi_{\rm m} and the magnon-phonon spatial mode overlap (expression and derivation in the Methods). Figures 2(g)-2(i) show the simulated field dependence of VcV_{\rm c} of the L, SL, and S modes at ϕh\phi_{\rm h} = 0 and for various ϕh\phi_{\rm h} between 00^{\circ} and 9090^{\circ}. For the L mode at ϕh=0\phi_{\rm h}=0^{\circ}, VcV_{\rm c} mostly vanishes in the high field region because ϕh=ϕm=0\phi_{\rm h}=\phi_{\rm m}=0^{\circ} and |2ϵxy|\left|2\epsilon_{\rm xy}\right| is negligibly small. In contrast, it increases dramatically as μ0Hex\mu_{0}H_{\rm ex} decreases below 20 mT. This enhancement at μ0Hex=6\mu_{0}H_{\rm ex}=6 mT is caused by the magnetization rotation from ϕm\phi_{\rm m} = 0 to 45; thereby, the term (ϵxxϵyy)sin2ϕm(\epsilon_{\rm xx}-\epsilon_{\rm yy})\sin{2\phi_{\rm m}} becomes non-zero. VcV_{\rm c} temporarily returns to almost zero as ϕm\phi_{\rm m} reaches 90 just before the magnetization reverses and then approaches the original value after the increase at ϕm=225\phi_{\rm m}=225^{\circ} and μ0Hex=6\mu_{0}H_{\rm ex}=-6 mT. In contrast, the S mode exhibits the opposite field dependency, in which a finite VcV_{\rm c} in the high field region is reduced in the low field region because the dominant magnetostrictive term is 2ϵxycos2ϕm2\epsilon_{\rm xy}\cos{2\phi_{\rm m}}, not (ϵxxϵyy)sin2ϕm(\epsilon_{\rm xx}-\epsilon_{\rm yy})\sin{2\phi_{\rm m}}. Since the SL mode has almost equal contributions from both strain components, the variation with respect to μ0Hex\mu_{0}H_{\rm ex} is moderate compared with the other two modes (Methods). As ϕh\phi_{\rm h} increases, the susceptibility to μ0Hex\mu_{0}H_{\rm ex} is cyclically modulated (right panels of Fig. 2(g)-2(i)). The change in direction of VcV_{\rm c} while sweeping ϕh\phi_{\rm h} is opposite between S and L modes, resulting from different dominant strain components. These results indicate that the cavity mode structures as well as the external field can be used to tune the magnetostrictive interaction.
    To experimentally show the magnetostrictive interaction, the field response of the cavity resonance was measured at ϕh=0\phi_{\rm h}=0^{\circ}. The observed resonant frequency shift (Δωa/(2π)\Delta\omega^{\prime}_{\rm a}/(2\pi)) and quality factor (QaQ_{\rm a}^{\prime}) are plotted as a function of μ0Hex\mu_{0}H_{\rm ex} in Fig. 3(a) and 3(b), where L and S modes are chosen for their opposite and distinct field susceptibilities. The response on the L mode exhibits dual dips at |μ0Hex||\mu_{0}H_{\rm ex}| = 5 mT with a reduction in Δωa/(2π)\Delta\omega^{\prime}_{\rm a}/(2\pi) and QaQ_{\rm a}^{\prime} in the low field region. This behavior can be understood from the theoretical formula,

ua(ω)=fdω2+ωa2iκaω+Vcbam2χm(ω)/(ρVa),u_{\rm a}(\omega)=\frac{f_{\rm d}}{-\omega^{2}+\omega_{\rm a}^{2}-i\kappa_{\rm a}\omega+V_{\rm c}b^{2}_{\rm am}\chi_{\rm m}(\omega)/(\rho V_{\rm a})}, (3)

where ρ\rho, VaV_{\rm a}, fdf_{\rm d} and ω\omega are mass density, acoustic mode volume, driving force density and angular frequency, and χm\chi_{\rm m} is magnetic susceptibility (Methods). The theoretical predictions are in agreement with the experimental results, indicating that the acoustic modulation is due to the magnon-phonon interaction, as described by our model. This interaction enables acoustic excitation of spin-wave oscillations in Ni, which exerts back-action force on the cavity resonance and tunes ωa\omega_{\rm a}^{\prime} and QaQ_{\rm a}^{\prime}.
    Remarkably, ferromagnetic magnons were able to be driven by phonons in the tiny PnC cavity. The effective mode volume of the L mode is estimated to be Va6.6V_{\rm a}\approx 6.6 μ\mum3 = 0.54 λ2t\lambda^{2}t with an acoustic wavelength λ\lambda of 3.5 μ\mum, which is 10510610^{5}-10^{6} times smaller than that of a SAW-based magnomechanical cavity system Hatanaka et al. (2022). The tiny-energy vibrations are confined by the high QQ resonance, so they hardly affect surrounding systems, unlike conventional electromagnetic-wave-based magnetic devices. We believe that the PnC cavity-waveguide system would be a building block for a magnomechanical system and could be used as a local magnon driver and phonon modulator.
    Strong back-action effects were also observed in the S mode, where shear strain dominate magnetostriction. A comparison with the effect in the L mode reveals the impact of the mode strain profiles on the interaction. Figure 3(b) shows that the dual dips in this mode at |μ0Hex|\left|\mu_{0}H_{\rm ex}\right| = 10 mT are wider than those in the L mode because of the finite VcV_{\rm c} in the high field region. In addition, a distinct center dip occurs around μ0Hex=1\mu_{0}H_{\rm ex}=-1 mT, an effect of the magnetization rotation. The field response is distinctly different from that in the L mode and can be simulated with our model. Thus, our cavity geometry is also used to selectively drive the magnetization dynamics utilizing the difference in strain distributions.
      To examine how the strain distribution affects the magnetoelastic modulation, the field responses of the acoustic resonant frequency (Δωa/(2π)\Delta\omega^{\prime}_{\rm a}/(2\pi)) and damping rate (κa\kappa_{\rm a}^{\prime}) on the L and S modes were investigated at ϕm\phi_{\rm m} ranging from 00^{\circ} to 9090^{\circ} (Fig. 4(a)-4(d)). The magnetoelastic modulation regions in Δωa\Delta\omega^{\prime}_{\rm a} and κa\kappa_{\rm a}^{\prime} broaden as ϕh\phi_{\rm h} increases from 0 to 60 in the L mode and shrink toward ϕh\phi_{\rm h} = 90. In contrast, the S mode shows the opposite dependency, in which the modulation regions become narrow around ϕh\phi_{\rm h} = 45. The theoretical calculations reproduce the experimental variations in both modes; here, the cyclic modulation while changing ϕh\phi_{\rm h} is governed by VcV_{\rm c}, so the L and S modes show the opposite behaviors. Note that the experiment and theoretical model showed a moderate response to the magnetoelastic effect in the SL mode (Methods), supporting the validity of our model. We also found similar mode-sensitive magnetoelastic modulation in a PnC cavity with a different defect geometry (Methods). These results verify that the mode-tunable magnetostriction allows us to control magnomechanical states acoustically.
    In conclusion, the ease of designing the PnC cavity and the small spatial leakage of its vibrations are useful for constructing integrated magnomechanical systems, in which microwave signal operations such as sensing, memory and processing are performed using magnons and phonons. Moreover, it is possible to build a tiny magnomechanical system of magnon wavelength size to efficiently control and read-out the magnetic state of a micro-/nano-ferromagnetic system such as a magnetic tunnel junction Yuasa et al. (2004). We believe that PnC cavity magnomechanics will expand the use of magnons and phonons and related technologies.

Acknowledgments

This work was partially supported by JSPS KAKENHI(S) Grant Number JP21H05020.

Author contributions

D.H. fabricated the device and performed the measurements and the data analysis. M.A. made theoretical model, and D.H. and M.A conducted the simulations with support from H. Y. and H.O.. D.H. and M.A. wrote the manuscript. All authors discussed the results during preparation of the paper.

References

  • Jungwirth et al. (2012) T. Jungwirth, J. Wunderlich, and K. Olejník. Spin hall effect devices. Nature materials 11, 382 (2012).
  • Chumak et al. (2015) A. V. Chumak, V. I. Vasyuchka, A. A. Serga, and B. Hillebrands. Magnon spintronics. Nature Physics 11, 453 (2015).
  • Tabuchi et al. (2015) Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Nakamura. Coherent coupling between a ferromagnetic magnon and a superconducting qubit. Science 349, 405 (2015).
  • Li et al. (2020) Y. Li, W. Zhang, V. Tyberkevych, W.-K. Kwok, A. Hoffmann, and V. Novosad. Hybrid magnonics: Physics, circuits, and applications for coherent information processing. Journal of Applied Physics 128, 130902 (2020).
  • Awschalom et al. (2021) D. D. Awschalom, C. Du, R. He, J. Heremans, A. Hoffmann, J. Hou, H. Kurebayashi, Y. Li, L. Liu, V. Novosad, et al. Quantum engineering with hybrid magnonics systems and materials. IEEE Transactions on Quantum Engineering (2021).
  • Li et al. (2021) Y. Li, C. Zhao, W. Zhang, A. Hoffmann, and V. Novosad. Advances in coherent coupling between magnons and acoustic phonons. APL Mater. 9, 060902 (2021).
  • Zhang et al. (2016) X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang. Cavity magnomechanics. Sci. Adv. 2, e1501286 (2016).
  • Kikkawa et al. (2016) T. Kikkawa, K. Shen, B. Flebus, R. A. Duine, K.-i. Uchida, Z. Qiu, G. E. Bauer, and E. Saitoh. Magnon polarons in the spin seebeck effect. Phys. Rev. Lett. 117, 207203 (2016).
  • Berk et al. (2019) C. Berk, M. Jaris, W. Yang, S. Dhuey, S. Cabrini, and H. Schmidt. Strongly coupled magnon–phonon dynamics in a single nanomagnet. Nature communications 10, 1 (2019).
  • An et al. (2020) K. An, A. N. Litvinenko, R. Kohno, A. A. Fuad, V. V. Naletov, L. Vila, U. Ebels, G. de Loubens, H. Hurdequint, N. Beaulieu, et al. Coherent long-range transfer of angular momentum between magnon kittel modes by phonons. Phys. Rev. B 101, 060407(R) (2020).
  • Potts et al. (2021) C. A. Potts, E. Varga, V. A. Bittencourt, S. V. Kusminskiy, and J. P. Davis. Dynamical backaction magnomechanics. Physical Review X 11, 031053 (2021).
  • Weiler et al. (2011) M. Weiler, L. Dreher, C. Heeg, H. Huebl, R. Gross, M. S. Brandt, and S. T. B. Goennenwein. Elastically driven ferromagnetic resonance in nickel thin films. Phys. Rev. Lett. 106, 117601 (2011).
  • Dreher et al. (2012) L. Dreher, M. Weiler, M. Pernpeintner, H. Huebl, R. Gross, M. S. Brandt, and S. T. B. Goennenwein. Surface acoustic wave driven ferromagnetic resonance in nickel thin films: Theory and experiment. Phys. Rev. B 86, 134415 (2012).
  • Kobayashi et al. (2017) D. Kobayashi, T. Yoshikawa, M. Matsuo, R. Iguchi, S. Maekawa, E. Saitoh, and Y. Nozaki. Spin current generation using a surface acoustic wave generated via spin-rotation coupling. Phys. Rev. Lett. 119, 077202 (2017).
  • Sasaki et al. (2017) R. Sasaki, Y. Nii, Y. Iguchi, and Y. Onose. Nonreciprocal propagation of surface acoustic wave in ni/linbo 3. Physical Review B 95, 020407 (2017).
  • Herna´\acute{\rm a}ndez-Mi´\acute{\rm i}nguez et al. (2020) A. Herna´\acute{\rm a}ndez-Mi´\acute{\rm i}nguez, F. Macia`\grave{\rm a}, J. M. Herna`\grave{\rm a}ndez, J. Herfort, and P. V. Santos. Large nonreciprocal propagation of surface acoustic waves in epitaxial ferromagnetic/semiconductor hybrid structures. Phys. Rev. Appl. 13, 044018 (2020).
  • Xu et al. (2020) M. Xu, K. Yamamoto, J. Puebla, K. Baumgaertl, B. Rana, K. Miura, H. Takahashi, D. Grundler, S. Maekawa, and Y. Otani. Nonreciprocal surface acoustic wave propagation via magneto-rotation coupling. Science Adv. 6, eabb1724 (2020).
  • Kawada et al. (2021) T. Kawada, M. Kawaguchi, T. Funato, H. Kohno, and M. Hayashi. Acoustic spin hall effect in strong spin-orbit metals. Sci. Adv. 7, eabd9697 (2021).
  • Hatanaka et al. (2022) D. Hatanaka, M. Asano, H. Okamoto, T. Kunihashi, H. Sanada, and H. Yamaguchi. On-chip coherent transduction between magnons and acoustic phonons in cavity magnomechanics. Physical Review Applied 17, 034024 (2022).
  • Matsuo et al. (2013) M. Matsuo, J. Ieda, K. Harii, E. Saitoh, and S. Maekawa. Mechanical generation of spin current by spin-rotation coupling. Physical Review B 87, 180402 (2013).
  • Maldovan (2013) M. Maldovan. Sound and heat revolutions in phononics. Nature 503, 209 (2013).
  • Benchabane et al. (2006) S. Benchabane, A. Khelif, J.-Y. Rauch, L. Robert, and V. Laude. Evidence for complete surface wave band gap in a piezoelectric phononic crystal. Phys. Rev. E 73, 065601(R) (2006).
  • Mohammadi et al. (2009) S. Mohammadi, A. A. Eftekhar, W. D. Hunt, and A. Adibi. High-QQ micromechanical resonators in a two-dimensional phononic crystal slab. Appl. Phys. Lett. 94, 051906 (2009).
  • Otsuka et al. (2013) P. H. Otsuka, K. Nanri, O. Matsuda, M. Tomoda, D. M. Profunser, I. A. Veres, S. Danworaphong, A. Khelif, S. Benchabane, V. Laude, et al. Broadband evolution of phononic-crystal-waveguide eigenstates in real- and k-spaces. Sci. Rep. 3, 3351 (2013).
  • Pourabolghasem et al. (2018) R. Pourabolghasem, R. Dehghannasiri, A. A. Eftekhar, and A. Adibi. Waveguiding effect in the gigahertz frequency range in pillar-based phononic-crystal slabs. Phys. Rev. Appl. 9, 014013 (2018).
  • Baboly et al. (2018) M. G. Baboly, C. M. Reinke, B. A. Griffin, I. El-Kady, and Z. C. Leseman. Acoustic waveguiding in a silicon carbide phononic crystals at microwave frequencies. Appl. Phys. Lett. 112, 103504 (2018).
  • Hatanaka and Yamaguchi (2020) D. Hatanaka and H. Yamaguchi. Real-space characterization of cavity-coupled waveguide systems in hypersonic phononic crystals. Phys. Rev. Appl. 13, 024005 (2020).
  • Hatanaka and Yamaguchi (2021) D. Hatanaka and H. Yamaguchi. Mode-sensitive magnetoelastic coupling in phononic-crystal magnomechanics. APL Materials 9, 071110 (2021).
  • Safavi-Naeini et al. (2014) A. H. Safavi-Naeini, J. T. Hill, S. Meenehan, J. Chan, S. Gro¨\ddot{\rm o}blacher, and O. Painter. Two-dimensional phononic-photonic band gap optomechanical crystal cavity. Phys. Rev. Lett. 112, 153603 (2014).
  • Yuasa et al. (2004) S. Yuasa, T. Nagahama, A. Fukushima, Y. Suzuki, and K. Ando. Giant room-temperature magnetoresistance in single-crystal Fe/MgO/Fe magnetic tunnel junctions. Nature materials 3, 868 (2004).

Methods

Appendix A Fabrication and measurement

The magnomechanical PnC was fabricated from GaAs (1.0 μ\mum)/ Al0.7Ga0.3As (3.0 μ\mum) heterostructure on a GaAs single-crystalline substrate. A periodic arrangement of snowflake-shaped air holes was formed by electron-beam lithography and dry etching. The GaAs layer, including the PnC lattice, was suspended by immersion in diluted hydrofluoric acid (5%\%). The PnC geometry gives rise to a complete bandgap between 0.45-0.60 GHz and 0.65-0.71 GHz. The acoustic waveguide was constructed by removing one line from the lattice, thereby enabling single-mode propagation at frequencies within the bandgap Hatanaka and Yamaguchi (2020). The resonator (cavity) formed by removing two holes was located at one side of the waveguide and sustains multiple resonant vibrations. A ferromagnetic thin film of nickel (Ni) with a thickness of 50 nm was deposited on the surface of the cavity and holds spin-wave (magnon) resonances. 5-nm-thick gold (Au) film was deposited on the Ni layer for preventing oxidization. The free-standing PnC slab is sandwiched by inter-digit transducers (IDT) made from Cr (5 nm) / Au (35 nm). The IDT consists of 100 transducers arrayed with a period of 4.9 μ\mum \sim 5.2 μ\mum.
      Acoustic waves were piezoelectrically excited by applying alternating voltages to one IDT and optically measured with an optical interferometer (Neoark, MLD-101). The data on spectral response of the PnC cavity were obtained with a time-gating technique with a network analyzer (Keysight E5080A) to remove undesired electrical cross-talk signals. The acoustic resonance frequencies (ωa/(2π)\omega^{\prime}_{\rm a}/(2\pi)) and quality factor (QaQ^{\prime}_{\rm a}) in Fig. 3(a) and 3(b) were obtained by making Lorentzian or exponential fittings to the spectral and temporal response results. The acoustic damping rates (κa\kappa_{\rm a}, κa0\kappa_{\rm a0}) in Fig. 4(b) and 4(d) were estimated from both ωa\omega^{\prime}_{\rm a} and QaQ^{\prime}_{\rm a}. The displacement amplitudes of the resonant mode profiles depicted in Fig. 1(f) were collected through frequency down-conversion followed by filtering with a lock-in amplifier (Stanford Research Systems, SR844). All experiments in this work were performed in a moderate vacuum (1010010-100 Pa) and at room temperature.

Appendix B Theory of magnetoelastic dyanmics in a phononic crystal cavity

The equation of motion of the acoustic mode is given by

ρ(u¨i+κau˙i+ωa2ui)=xkσik=ωa2ui+k={x,y,z}MskϵikGd\begin{split}\rho\left(\ddot{u}_{i}+\kappa_{a}\dot{u}_{i}+\omega_{\rm a}^{2}u_{i}\right)&=\partial_{x_{k}}\sigma_{ik}\\ &=\omega_{\rm a}^{2}u_{i}+\sum_{k=\{x,y,z\}}M_{\rm s}\partial_{k}\partial_{\epsilon_{ik}}G_{d}\end{split} (4)

where

Gd=l,n={x,y,z}bamϵlnmlmn.\displaystyle G_{\rm d}=\sum_{l,n=\{x,y,z\}}b_{\rm am}\epsilon_{ln}m_{l}m_{n}. (5)

is the magnetoelastic energy density. The magnetoelastic force density in the second term on the right-hand side is given by

fam,ikMskϵikGd=Msbamk,l,nδliδnkkmlmn=Msbam(mi𝒎).\displaystyle\begin{split}f_{{\rm am},i}\equiv\sum_{k}M_{\rm s}\partial_{k}\partial_{\epsilon_{ik}}G_{\rm d}&=M_{\rm s}b_{\rm am}\sum_{k,l,n}\delta_{li}\delta_{nk}\partial_{k}m_{l}m_{n}\\ &=M_{\rm s}b_{\rm am}(\nabla\cdot m_{i}{\bm{m}}).\end{split} (6)

In our setup, the magnetization is aligned in-plane of the Ni film and the field angle away from the xx-axis is defined as ϕm\phi_{\rm m}. Therefore, the conversion between xyzxyz-coordinate system and 123123-coordinate system is

(mxmymz)=(0sinϕmcosϕm0cosϕmsinϕm100)(m1m2m3)\displaystyle\left(\begin{matrix}m_{x}\\ m_{y}\\ m_{z}\end{matrix}\right)=\left(\begin{matrix}0&-\sin\phi_{m}&\cos\phi_{m}\\ 0&\cos\phi_{m}&\sin\phi_{m}\\ -1&0&0\end{matrix}\right)\left(\begin{matrix}m_{1}\\ m_{2}\\ m_{3}\end{matrix}\right) (7)
mx=\displaystyle m_{\rm x}= sinϕmm2+cosϕm\displaystyle-\sin\phi_{\rm m}m_{2}+\cos\phi_{\rm m} (8)
my=\displaystyle m_{\rm y}= cosϕmm2+sinϕm\displaystyle\cos\phi_{\rm m}m_{2}+\sin\phi_{\rm m} (9)
mz=\displaystyle m_{\rm z}= m1\displaystyle-m_{1} (10)

with approximations m1,m2m31m_{1},m_{2}\ll m_{3}\sim 1. Accordingly, under shear strains ϵizϵzi0\epsilon_{iz}\approx\epsilon_{zi}\approx 0, the magnetoelastic force density in equation (B3) becomes

fam,x/(Msbam)=kkmxmk=sin2ϕmxm2+cos2ϕmym2+𝒪(mi2)\begin{split}f_{\rm am,x}/(M_{\rm s}b_{\rm am})=&\sum_{k}\partial_{k}m_{\rm x}m_{k}\\ =&-\sin 2\phi_{\rm m}\partial_{x}m_{2}+\cos 2\phi_{\rm m}\partial_{y}m_{2}+\mathcal{O}(m_{i}^{2})\end{split} (11)

and

fam,y/(Msbam)=kkmymk=cos2ϕmxm2+sin2ϕmym2+𝒪(mi2)\begin{split}f_{\rm am,y}/(M_{\rm s}b_{\rm am})=&\sum_{k}\partial_{k}m_{\rm y}m_{k}\\ =&\cos 2\phi_{\rm m}\partial_{x}m_{2}+\sin 2\phi_{\rm m}\partial_{y}m_{2}+\mathcal{O}(m_{i}^{2})\end{split} (12)

where fam,xf_{\rm am,x} and fam,yf_{\rm am,y} are the xx and yy components of the magnetoelastic force. As a result, we can define a new magnetoelastic vector,

𝒗am𝒇amMsbam=(sin2ϕmcos2ϕmcos2ϕmsin2ϕm)Φm2(𝒓)\displaystyle{\bm{v}}_{\rm am}\equiv\frac{{\bm{f}}_{\mathrm{am}}}{M_{s}b_{\mathrm{am}}}=\left(\begin{matrix}-\sin 2\phi_{\rm m}&\cos 2\phi_{\rm m}\\ \cos 2\phi_{\rm m}&\sin 2\phi_{\rm m}\end{matrix}\right)\nabla\Phi_{m_{2}}({\bm{r}}) (13)

where Φm2(𝒓)\Phi_{m_{2}}({\bm{r}}) is the normalized amplitude of magnons at position rr. This magnetoelastic vector contributes to the coupling constant with the spatial integration of magnons. By redefining the acoustic mode as 𝒖(𝒓,t)=Ua(t)𝚿(𝒓){\bm{u}}({\bm{r}},t)=U_{a}(t){\bm{\Psi}}({\bm{r}}), we find that

U¨a(t)+κaU˙a(t)+ωa2Ua(t)=Msbamd𝒓𝚿(𝒓)𝒗amd𝒓ρ(𝒓)|𝚿(𝒓)|2m2(t)MsbamVc0meffm2(t)\begin{split}\ddot{U}_{\rm a}(t)+\kappa_{a}\dot{U}_{\rm a}(t)&+\omega_{\rm a}^{2}U_{\rm a}(t)\\ &=M_{\rm s}b_{\rm am}\frac{\int\mathrm{d}{\bm{r}}{\bm{\Psi}}({\bm{r}})\cdot{\bm{v}}_{\mathrm{am}}}{\int\mathrm{d}{\bm{r}}\rho({\bm{r}})|{\bm{\Psi}}({\bm{r}})|^{2}}m_{2}(t)\\ &\equiv M_{\rm s}b_{\rm am}\frac{V_{\rm c0}}{m_{\mathrm{eff}}}m_{2}(t)\end{split} (14)

where

meffρVa=ρVd3𝒓|𝚿(𝒓)|2\displaystyle m_{\mathrm{eff}}\equiv\rho V_{\rm a}=\rho\int_{V}\mathrm{d}^{3}{\bm{r}}|{\bm{\Psi}}({\bm{r}})|^{2} (15)

is the effective mass with max𝒓[|𝚿(𝒓)|2]=1\max_{\bm{r}}\left[|{\bm{\Psi}}({\bm{r}})|^{2}\right]=1 and

Vc0=Vd3𝒓𝚿(𝒓)𝒗am.\displaystyle V_{c0}=\int_{V}\mathrm{d}^{3}{\bm{r}}{\bm{\Psi}}({\bm{r}})\cdot{\bm{v}}_{\rm am}. (16)

is the magnetostrictive coupling mode volume. Note that Φm2(𝒓)\Phi_{m_{2}}({\bm{r}}) and m2(t)m_{2}(t) are non-dimensional variables.

B.1 Equation of motion for magnon modes

The previous work by Dreher 𝑒𝑡\it{et} 𝑎𝑙.\it{al.} Dreher et al. (2012) derived the following equations of magnonic motion:

αγm˙1+(G11G3)m11γm˙2=\displaystyle\frac{\alpha}{\gamma}\dot{m}_{1}+(G_{11}-G_{3})m_{1}-\frac{1}{\gamma}\dot{m}_{2}= 0\displaystyle 0 (17)
αγm˙2+(G22G3)m2+1γm˙1=bam[sin2ϕm(xuxyuy)cos2ϕm(yux+xuy)].\begin{split}&\frac{\alpha}{\gamma}\dot{m}_{2}+(G_{22}-G_{3})m_{2}+\frac{1}{\gamma}\dot{m}_{1}\\ &=b_{\rm am}\left[\sin 2\phi_{\rm m}(\partial_{x}u_{\rm x}-\partial_{y}u_{\rm y})-\cos 2\phi_{\rm m}\left(\partial_{y}u_{\rm x}+\partial_{x}u_{\rm y}\right)\right].\end{split} (18)

To determine Ψmi\Psi_{m_{i}}, we have to diagonalize equations (B14) and (B15),

m˙1=\displaystyle\dot{m}_{1}= G1αm1+G2m2+γ1+α2Fam\displaystyle G_{1}\alpha m_{1}+G_{2}m_{2}+\frac{\gamma}{1+\alpha^{2}}F_{\rm am} (19)
m˙2=\displaystyle\dot{m}_{2}= G1m1+G2αm2+γα1+α2Fam\displaystyle-G_{1}m_{1}+G_{2}\alpha m_{2}+\frac{\gamma\alpha}{1+\alpha^{2}}F_{\rm am} (20)

where Giγ(G3Gii)/(1+α2)G_{i}\equiv\gamma(G_{3}-G_{ii})/(1+\alpha^{2}) and Fam=bam[sin2ϕm(xuxyuy)cos2ϕm(yux+xuy)]F_{\rm am}=b_{\rm am}\left[\sin 2\phi_{\rm m}(\partial_{x}u_{\rm x}-\partial_{y}u_{\rm y})-\cos 2\phi_{\rm m}\left(\partial_{y}u_{\rm x}+\partial_{x}u_{\rm y}\right)\right]. Here G11=2BdG_{11}=2B_{\rm d}, G22=2Busin2(ϕmϕu)G_{22}=-2B_{\rm u}\sin^{2}(\phi_{\rm m}-\phi_{\rm u}), and G3=μ0Hexcos(ϕmϕh)2Bucos2(ϕmϕu)G_{3}=-\mu_{0}H_{\mathrm{ex}}\cos(\phi_{\rm m}-\phi_{\rm h})-2B_{\rm u}\cos^{2}(\phi_{\rm m}-\phi_{\rm u}). Finally, we obtain

(m˙+m˙)=(λ+00λ)(m+m)+γ1+α2FamP1(1α)\displaystyle\left(\begin{matrix}\dot{m}_{+}\\ \dot{m}_{-}\end{matrix}\right)=\left(\begin{matrix}\lambda_{+}&0\\ 0&\lambda_{-}\end{matrix}\right)\left(\begin{matrix}m_{+}\\ m_{-}\end{matrix}\right)+\frac{\gamma}{1+\alpha^{2}}F_{\rm am}P^{-1}\left(\begin{matrix}1\\ \alpha\end{matrix}\right) (21)

where

λ±α(G1+G2)2G1G2(2+α2)+(G12+G22)α22\displaystyle\lambda_{\pm}\equiv\frac{\alpha(G_{1}+G_{2})\mp\sqrt{-2G_{1}G_{2}(2+\alpha^{2})+(G_{1}^{2}+G_{2}^{2})\alpha^{2}}}{2} (22)

and

γ1+α2FamP1(1,α)T(s+,s)Tγ1+α2Fam.\displaystyle\frac{\gamma}{1+\alpha^{2}}F_{\rm am}P^{-1}(1,\alpha)^{T}\equiv(s_{+},s_{-})^{T}\frac{\gamma}{1+\alpha^{2}}F_{\rm am}. (23)

s+s_{+} and ss_{-} are defined as

s+=\displaystyle s_{+}= iG1(2+α2)G2α2+αG12α2+G22α22G1G2(2+α2)2G12α2+G22α22G1G2(2+α2)\displaystyle-i\frac{G_{1}(2+\alpha^{2})-G_{2}\alpha^{2}+\alpha\sqrt{G_{1}^{2}\alpha^{2}+G_{2}^{2}\alpha^{2}-2G_{1}G_{2}(2+\alpha^{2})}}{2\sqrt{G_{1}^{2}\alpha^{2}+G_{2}^{2}\alpha^{2}-2G_{1}G_{2}(2+\alpha^{2})}} (24)
s=\displaystyle s_{-}= iG1(2+α2)G2α2αG12α2+G22α22G1G2(2+α2)2G12α2+G22α22G1G2(2+α2).\displaystyle i\frac{G_{1}(2+\alpha^{2})-G_{2}\alpha^{2}-\alpha\sqrt{G_{1}^{2}\alpha^{2}+G_{2}^{2}\alpha^{2}-2G_{1}G_{2}(2+\alpha^{2})}}{2\sqrt{G_{1}^{2}\alpha^{2}+G_{2}^{2}\alpha^{2}-2G_{1}G_{2}(2+\alpha^{2})}}. (25)

As a result, the equations of magnon motion becomes

m˙±=λ±m±+γs±1+α2Fam.\displaystyle\dot{m}_{\pm}=\lambda_{\pm}m_{\pm}+\frac{\gamma s_{\pm}}{1+\alpha^{2}}F_{\rm am}. (26)

By decomposing the temporal and spatial parts as m±m±(t)Φ±(𝒓)m_{\pm}\to m_{\pm}(t)\Phi_{\pm}({\bm{r}}), it can be expressed as

m˙±(t)Φ±(𝒓)=λ±m±(t)Φ±(𝒓)+bγs±1+α2U(t)[sin2ϕm(ΨxxΨyy)cos2ϕm(Ψxy+Ψyx)]\displaystyle\dot{m}_{\pm}(t)\Phi_{\pm}({\bm{r}})=\lambda_{\pm}m_{\pm}(t)\Phi_{\pm}({\bm{r}})+\frac{b\gamma s_{\pm}}{1+\alpha^{2}}U(t)\left[\sin 2\phi_{\rm m}\left(\frac{\partial\Psi_{\rm x}}{\partial x}-\frac{\partial\Psi_{\rm y}}{\partial y}\right)-\cos 2\phi_{\rm m}\left(\frac{\partial\Psi_{\rm x}}{\partial y}+\frac{\partial\Psi_{\rm y}}{\partial x}\right)\right] (27)

and

Φ+(𝒓)=Φ(𝒓)=1K[sin2ϕm(xΨxyΨy)cos2ϕm(yΨx+xΨy)]\displaystyle\Phi_{+}({\bm{r}})=\Phi_{-}({\bm{r}})=\frac{1}{K}\left[\sin 2\phi_{\rm m}(\partial_{x}\Psi_{\rm x}-\partial_{y}\Psi_{\rm y})-\cos 2\phi_{\rm m}\left(\partial_{y}\Psi_{\rm x}+\partial_{x}\Psi_{\rm y}\right)\right] (28)

where KK is a wavevector defined as Kmax𝒓[sin2ϕm(xΨxyΨy)cos2ϕm(yΨx+xΨy)]K\equiv\max_{\bm{r}}\left[\sin 2\phi_{\rm m}(\partial_{x}\Psi_{\rm x}-\partial_{y}\Psi_{\rm y})-\cos 2\phi_{\rm m}\left(\partial_{y}\Psi_{\rm x}+\partial_{x}\Psi_{\rm y}\right)\right]. Accordingly, the diagonalized equations can be simplified to

m˙±(t)=λ±m±(t)+bγs±K1+α2U(t).\displaystyle\dot{m}_{\pm}(t)=\lambda_{\pm}m_{\pm}(t)+\frac{b\gamma s_{\pm}K}{1+\alpha^{2}}U(t). (29)

Importantly, we have the relationship

(m+m)=1i((G1G2)α2G1G2(2+α2)+(G12+G22)α22G1(G1G2)α+2G1G2(2+α2)+(G12+G22)α22G111)1(m1m2)\begin{split}\left(\begin{matrix}m_{+}\\ m_{-}\end{matrix}\right)=&\frac{1}{i}\left(\begin{matrix}-\frac{(G_{1}-G_{2})\alpha-\sqrt{-2G_{1}G_{2}(2+\alpha^{2})+(G_{1}^{2}+G_{2}^{2})\alpha^{2}}}{2G_{1}}&-\frac{(G_{1}-G_{2})\alpha+\sqrt{-2G_{1}G_{2}(2+\alpha^{2})+(G_{1}^{2}+G_{2}^{2})\alpha^{2}}}{2G_{1}}\\ 1&1\end{matrix}\right)^{-1}\left(\begin{matrix}m_{1}\\ m_{2}\end{matrix}\right)\end{split} (30)

and thus, we use

m2i(m++m).\displaystyle m_{2}\approx i(m_{+}+m_{-}). (31)

B.2 Coupled mode equation

The above equations of motions of acoustic phonons and magnons lead to the following equation of motion of coupled modes,

U¨a(t)+κaU˙a(t)+ωa2Ua(t)=\displaystyle\ddot{U}_{\rm a}(t)+\kappa_{\rm a}\dot{U}_{\rm a}(t)+\omega_{\rm a}^{2}U_{\rm a}(t)= iMsbamVcKmeff(m+(t)+m(t))\displaystyle iM_{\rm s}b_{\rm am}\frac{V_{\rm c}}{Km_{\mathrm{eff}}}\left(m_{+}(t)+m_{-}(t)\right) (32)
m˙±(t)=\displaystyle\dot{m}_{\pm}(t)= λ±m±(t)+bamγs±K1+α2Ua(t)\displaystyle\lambda_{\pm}m_{\pm}(t)+\frac{b_{\rm am}\gamma s_{\pm}K}{1+\alpha^{2}}U_{\rm a}(t) (33)

where the coupling mode volume has been redefined as Vc=Vd3𝒓𝚿(𝒓)𝒗amV_{\mathrm{c}}=\int_{V}\mathrm{d}^{3}{\bm{r}}{\bm{\Psi}}({\bm{r}})\cdot{\bm{v}}_{\rm am} and

𝒗am=12(2Ψx+cos4ϕm(x2Ψxy2Ψx2xyΨy)+sin4ϕm(x2Ψyy2Ψy+2xyΨx)2Ψy+cos4ϕm(y2Ψyx2Ψy2xyΨx)+sin4ϕm(x2Ψxy2Ψx2xyΨy)).\displaystyle{\bm{v}}_{\rm am}=\frac{1}{2}\left(\begin{matrix}-\nabla^{2}\Psi_{\rm x}+\cos 4\phi_{\rm m}\left(\partial_{x}^{2}\Psi_{\rm x}-\partial_{y}^{2}\Psi_{\rm x}-2\partial_{x}\partial_{y}\Psi_{\rm y}\right)+\sin 4\phi_{\rm m}\left(\partial_{x}^{2}\Psi_{\rm y}-\partial_{y}^{2}\Psi_{\rm y}+2\partial_{x}\partial_{y}\Psi_{\rm x}\right)\\ -\nabla^{2}\Psi_{\rm y}+\cos 4\phi_{\rm m}\left(\partial_{y}^{2}\Psi_{\rm y}-\partial_{x}^{2}\Psi_{\rm y}-2\partial_{x}\partial_{y}\Psi_{x}\right)+\sin 4\phi_{\rm m}\left(\partial_{x}^{2}\Psi_{x}-\partial_{y}^{2}\Psi_{x}-2\partial_{x}\partial_{y}\Psi_{y}\right)\end{matrix}\right). (34)

By solving equations (B29) and (B30) with an additional driving force, fdf_{\rm d}, we obtain the acoustic displacement amplitude (Ua(t)=uaeiωtU_{\rm a}(t)=u_{\rm a}e^{i\omega t}) modulated by back-action from magnons, where

ua(ω)=fdω2+ωa2iκaω+Vcbam2χm(ω)/(ρVa),u_{\rm a}(\omega)=\frac{f_{\rm d}}{-\omega^{2}+\omega_{\rm a}^{2}-i\kappa_{\rm a}\omega+V_{\rm c}b^{2}_{\rm am}\chi_{\rm m}(\omega)/(\rho V_{\rm a})}, (35)

and the magnetic susceptibility is

χm(ω)=iMsγ1+α2(siωiωm+κm2+s+iω+iωm+κm2).\begin{split}\chi_{\rm m}(\omega)=-i\frac{M_{\rm s}\gamma}{1+\alpha^{2}}\left(\frac{s_{-}}{-i\omega-i\omega_{\rm m}+\frac{\kappa_{\rm m}}{2}}+\frac{s_{+}}{-i\omega+i\omega_{\rm m}+\frac{\kappa_{\rm m}}{2}}\right).\end{split} (36)

For the numerical calculation, we had to derive the value of Vc/meff\sqrt{V_{\rm c}/m_{\mathrm{eff}}} at which the spatial function appears as the same order |Ψ|2|\Psi|^{2}. This means that constant factors cancel out in their ratio. Magnetoelastic coupling coefficients (bamb_{\rm am}) of 5 T, 6T and 10 T were used for the simulations of the SL, S and L modes, respectively. The angle of the in-plane anisotropic field was set at Δϕu=ϕuϕh\Delta\phi_{\rm u}=\phi_{\rm u}-\phi_{\rm h} = 85, to reproduce the field response of the equilibrium magnetization as shown in Fig. 2(b). The table lists the other parameters of the calculations.
    By transforming into the rotating frame of the acoustic modes, i.e., Ua(t)=AeiωatU_{a}(t)=Ae^{-i\omega_{a}t}, the equation of motion of acoustic phonons becomes

A˙+κa2=MsbamVc2ωaKmeff(m+(t)+m(t)).\displaystyle\dot{A}+\frac{\kappa_{\rm a}}{2}=-\frac{M_{\rm s}b_{\rm am}V_{\rm c}}{2\omega_{\rm a}Km_{\mathrm{eff}}}\left(m_{+}(t)+m_{-}(t)\right). (37)

The coupled mode equation in the frequency domain is given by

(iω+κa2+MsbamVc2ωaKmeffMsbamVc2ωaKmeffbamγs+K1+α2iωλ+0bamγsK1+α20iωλ)(Ua(ω)m+(ω)m(ω))\displaystyle\left(\begin{matrix}-i\omega+\frac{\kappa_{\rm a}}{2}&+\frac{M_{\rm s}b_{\rm am}V_{\rm c}}{2\omega_{\rm a}Km_{\mathrm{eff}}}&\frac{M_{\rm s}b_{\rm am}V_{\rm c}}{2\omega_{\rm a}Km_{\mathrm{eff}}}\\ -\frac{b_{\rm am}\gamma s_{+}K}{1+\alpha^{2}}&-i\omega-\lambda_{+}&0\\ -\frac{b_{\rm am}\gamma s_{-}K}{1+\alpha^{2}}&0&-i\omega-\lambda_{-}\end{matrix}\right)\left(\begin{matrix}U_{\rm a}(\omega)\\ m_{+}(\omega)\\ m_{-}(\omega)\end{matrix}\right)\equiv (iω+κa2GA+GAG+Aiωλ+0GA0iωλ)(Ua(ω)m+(ω)m(ω))=(fd00).\displaystyle\left(\begin{matrix}-i\omega+\frac{\kappa_{\rm a}}{2}&G_{A+}&G_{A-}\\ -G_{+A}&-i\omega-\lambda_{+}&0\\ -G_{-A}&0&-i\omega-\lambda_{-}\end{matrix}\right)\left(\begin{matrix}U_{\rm a}(\omega)\\ m_{+}(\omega)\\ m_{-}(\omega)\end{matrix}\right)=\left(\begin{matrix}f_{\rm d}\\ 0\\ 0\end{matrix}\right).

Hence, the acoustic mode spectra is

Ua(ω)=fd[GA+G+A(λ++iω)GAGA(λ+iω)+(iω+κa2)]1.\displaystyle U_{\rm a}(\omega)=f_{\rm d}\left[-\frac{G_{A+}G_{+A}}{(\lambda_{+}+i\omega)}-\frac{G_{A-}G_{-A}}{(\lambda_{-}+i\omega)}+(-i\omega+\frac{\kappa_{\rm a}}{2})\right]^{-1}. (38)

Apparently, the symmetrized coupling strength is

G±GA±G±A=bamMsγs±2ωa(1+α2)Vcmeff.\displaystyle G_{\pm}\equiv\sqrt{G_{A\pm}G_{\pm A}}=b_{\rm am}\sqrt{\frac{M_{\rm s}\gamma s_{\pm}}{2\omega_{\rm a}\left(1+\alpha^{2}\right)}}\sqrt{\frac{V_{\rm c}}{m_{\mathrm{eff}}}}. (39)
ρ{\rho} mass density 8900 kg/m3
BdB_{\rm d} out-of-plane shape anisotropy 0.2 T
BuB_{\rm u} in-plane magnetic anisotropy 4 mT
α\alpha Gilbert damping factor 0.1
MsM_{\rm s} saturation magnetization 370 kA/m
γ\gamma gyromagnetic ratio 2.185μB/\mu_{\rm B}/\hbar
Table 1: Acoustic and magnetic parameters used in the calculations.

Appendix C Magnetoelastic modulation of SL mode in a PnC cavity

The left and right panels of Fig. 5(a) plot the resonant frequency shift (Δωa/(2π)\Delta\omega_{\rm a}^{\prime}/(2\pi)) and quality factor (QaQ_{\rm a}^{\prime}) as functions of the bias field (μ0Hex\mu_{0}H_{\rm ex}) for the SL mode in the ϕh=0\phi_{\rm h}=0^{\circ}. Their field responses reveal a dual dip structure due to the increased magnon-phonon interaction; the theoretical calculations (solid line) show the same dip. Compared with the other modes, the modulation magnitudes, i.e. dip depths, are small, because of the comparable contributions of the longitudinal (|ϵxxϵyy|\left|\epsilon_{\rm xx}-\epsilon_{\rm yy}\right|) and shear strain components (|2ϵxy|\left|2\epsilon_{\rm xy}\right|), and thus, the field dependency of the coupling (VcV_{\rm c}) is small. Figure 5(b) and 5(c) show the experimental (left) and simulated (right) field responses of Δωa\Delta\omega_{\rm a}^{\prime} and QaQ_{\rm a}^{\prime} while sweeping ϕh\phi_{\rm h} from 00^{\circ} to 9090^{\circ}. The field regions of the magnetoelastic modulation exhibit moderate variation with ϕh\phi_{\rm h} compared with the S and L modes. Clearly, the magnetoelastic dynamics are a consequence of the spatial strain distribution of this mode.

Refer to caption
Figure 5: Magnetoelastic modulation on an acoustic resonance of SL mode. (a) The field (μ0Hex\mu_{0}H_{\rm ex}) dependence of resonant frequency shift (Δωa/(2π)\Delta\omega_{\rm a}^{\prime}/(2\pi)) and quality-factor (QaQ_{\rm a}^{\prime}) in ϕh\phi_{\rm h} in the left and right panels respectively. (b) and (c) The ϕh\phi_{\rm h} dependence of the frequency Δωa/(2π)\Delta\omega_{\rm a}^{\prime}/(2\pi) and damping κa\kappa_{\rm a}^{\prime} versus μ0Hex\mu_{0}H_{\rm ex} respectively. The experimental and calculated results are shown in the left and right panels.

Appendix D Magnetoelastic modulation on an PnC cavity with a three-holes defect

A PnC cavity formed by removing three holes, shown in Fig. 6(a), holds two acoustic resonances at 0.579 GHz and 0.587 GHz with distinct modal shapes (Fig. 6(b) and 6(c)). These modes, labeled L and S, indicate that acoustic vibrations are confined in the defect (Fig. 6(d) and 6(e)). The numerically calculated spatial distributions of strains |ϵxxϵyy|\left|\epsilon_{\rm xx}-\epsilon_{\rm yy}\right| and |2ϵxy|\left|2\epsilon_{\rm xy}\right| are shown in the left and right panels of Fig. 7(a) and 7(b). Magnetoelastic coupling mode volume VcV_{\rm c} as function of the bias field strength μ0Hex\mu_{0}H_{\rm ex} at ϕh\phi_{\rm h} = 0 is plotted in the left panels of Fig. 7(c) and 7(d) for the L and S modes. These field responses indicate that the two modes show opposite coupling dynamics. Similarly, the field angle evolution of the VcV_{\rm c} response indicates the magnetoelastic modulation effect is sensitive to the field and magnetization orientation and the acoustic mode structure (right panels of Fig. 7(c) and 7(d)). Figure 8(a) and 8(c) indicate the frequency shifts of the L and S modes (left) and simulated responses (right). Figure 8(b) and 8(d) plot the normalized acoustic damping rates (left) and the simulated responses (right). The magnetoelastic coupling coefficient used in the calculations was bam=7b_{\rm am}=7 T for both modes.

Refer to caption
Figure 6: Acoustic dynamics of PnC cavity with three hole defect. (a) SEM image of the PnC cavity coupled to a waveguide. A Ni film (yellow) is placed on the surface of the cavity. (b) and (c) Acoustic resonant spectrum of L and S modes. (d) and (e) Experimental (left) and calculated (right) spatial profiles of the resonant vibration amplitude in L and S modes. The defect region is highlighted by yellow dotted lines in the left panel.
Refer to caption
Figure 7: Spatial strain profiles of cavity resonances and their magnetoelastic coupling mode volume. (a) and b Spatial distribution of longitudinal |ϵxxϵyy|\left|\epsilon_{\rm xx}-\epsilon_{\rm yy}\right| (left) and shear |2ϵxy|\left|2\epsilon_{\rm xy}\right| (right) strains on L and S modes, respectively. (c) and (d) Left: Simulated magnetoelastic coupling mode volume (VcV_{\rm c}) as a function of field strength (μ0Hex\mu_{0}H_{\rm ex}) at ϕh=0\phi_{\rm h}=0^{\circ}. Right: Corresponding field angle (ϕh\phi_{\rm h}) dependence of Vcμ0HexV_{\rm c}-\mu_{0}H_{\rm ex}.
Refer to caption
Figure 8: Mode-sensitive magnetoelastic cavity dynamics. (a)-(d) Experimental (left) and simulated (right) ϕh\phi_{\rm h} dependence of Δωa/(2π)\Delta\omega_{\rm a}^{\prime}/(2\pi) (a, c) and κa\kappa_{\rm a}^{\prime} (b, d) as a function of μ0Hex\mu_{0}H_{\rm ex} in L and S modes.