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

Controllable Fano-type optical response and four-wave mixing via magnetoelastic coupling in a opto-magnomechanical system

Amjad Sohail [email protected] Department of Physics, Government College University, Allama Iqbal Road, Faisalabad 38000, Pakistan.    Rizwan Ahmed Physics division, Pakistan institute of nuclear science and technology (PINSTECH), Nilore, Islamabad 45650 Pakistan    Jia-Xin Peng [email protected] State Key Laboratory of Precision Spectroscopy, Quantum Institute for Light and Atoms, Department of Physics, East China Normal University, Shanghai 200062, People’s Republic of China    Aamir Shahzad Department of Physics, Government College University, Allama Iqbal Road, Faisalabad 38000, Pakistan.    Tariq Munir Department of Physics, Government College University, Allama Iqbal Road, Faisalabad 38000, Pakistan.    S. K. Singh [email protected] Graphene and Advanced 2D Materials Research Group (GAMRG), School of Engineering and Technology, Sunway University, No. 5, Jalan Universiti, Bandar Sunway, 47500 Petaling Jaya, Selangor, Malaysia    Marcos César de Oliveira [email protected] Instituto de Física Gleb Wataghin, Universidade Estadual de Campinas, Campinas, SP, Brazil
Abstract

We analytically investigate the Fano-type optical response and four-wave mixing (FWM) process by exploiting the magnetoelasticity of a ferromagnetic material. The deformation of the ferromagnetic material plays the role of mechanical displacement, which is simultaneously coupled to both optical and magnon modes. We report that the magnetostrictively induced displacement demonstrates Fano profiles, in the output field, which is well-tuned by adjusting the system parameters, like effective magnomechanical coupling, magnon detuning, and cavity detuning. It is found that the magnetoelastic interaction also gives rise to the FWM phenomenon. The number of the FWM signals mainly depends upon the effective magnomechanical coupling and the magnon detuning. Moreover, the FWM spectrum exhibits suppressive behavior upon increasing (decreasing) the magnon (cavity) decay rate. The present scheme will open new perspectives in highly sensitive detection and quantum information processing.

I Introduction

Cavity optomechanics is a rapidly growing area of theoretical and experimental research that explores the coherent coupling between mechanical and optical modes through the radiation pressure force of photons being trapped inside optomechanical cavities Mey . It has been responsible for significant advances in quantum technology, particularly when non-linearities are available. We mention quantum information processing Fio , quantum walk implementation Jalil , ultrahigh-precision measurementTeu ; AliM , higher-order sidebands generation Qian , optical nonlinearity Lud , single photon blockade SS1 ; SS2 , quantum entanglement AM0 ; MYP ; SKS ; BTT , gravitation-wave detectionArva ; AliM2 , optomechanically induced transparency (OMIT) MYPOM1 ; OM1 ; OM2 , and optomechanically induced absorption (OMIA) MYPOM2 including optomechanically induced stochastic resonance Faraz . In OMIT, analogously to electromagnetic induced transparency (EIT) Imam , the radiation pressure-induced mechanical oscillations lead to destructive interference and hence significantly alter the optical response as well as its group delay Saif ; Saiff .

Another important and fascinating non-linear phenomenon is the interference and quantum coherence, called Fano resonance Yuri , where the prominent feature is a steep and sharp asymmetric line profile. Fano resonance were theoretically discovered by Ugo Fano Fano , who found that the interaction of a discrete excited state of an atom with a continuum of scattering states, is entirely different from the resonance phenomena delineated by the Lorentzian formula Fano . Fano interference has important applications such as in lasing without population inversion Gav , as a probe to study decoherence Kuh , and as a way to improve the efficiency of heat engines Chap . Moreover, the radiating two coupled-resonators model, in which the two resonators have quite different lifetime can be used to explain the Fano line shapes Tass ; Alim2 . So the lifetimes of these modes are very significant in observing the Fano line shapes Gal and can be easily obtained in plasmonic systems Gall and metamaterial structures Ali ; AliM5 . Fano resonance has been recently studied in cavity magnomechanics Kamran and magnon-qubit system Sabur .

Four-wave mixing (FWM), as the most engrossing nonlinear optical effect, is widely studied in optomechanical systems FWM1 . FWM is based on interference and quantum coherence and is extensively used to enlarge the frequency span of coherent light sources to ultraviolet and infrared FWM2 ; AliM3 . Multi-wave or FWM-mixing processes allow many vital physical processes, such as realizing optical bistability FWM3 , normal mode splitting FWM4A or generating two coexisting pairs of narrowband biphotons FWM5 . Consequently, the FWM effect has gained much attention from researchers. Jiang et al. examined the controllability and tunability of the FWM process by driving the two cavities in a two-mode cavity optomechanical system at their respective red sidebands FWM6 . Wang and Chen studied the enhancement of the FWM response FWM7 . It must be pointed out that FWM is a weak phenomenon, which makes it hard to be realized in experiments. Therefore new directions are being proposed to improve the strength of those nonlinear phenomena.

The hybridization of different modes in cavity magnomechancis (CMM) opens up a promising platform to explore numerous non-linear quantum effects Tang . Magnomechanical systems consider a geometry such that an optical mode directly (indirectly) couples with a magnon mode (mechanical mode). They are based totally on magnons-quanta of spin waves in Ferrimagnetic materials Heyroth , such as yttrium iron garnet (YIG), which provide a versatile and new platform to investigate strong light-matter interactions Gros . Strong interactions are due to two main reasons: (1) The Spin density of the YIG material is very high; (2) The damping rate of the YIG material is very low. Therefore, in the last decade, CMM has been explored in theoretical as well as an experimental domain because of its potential applications for observing macroscopic quantum states Jiee , quantum sensing Usami and magnomechanically induced transparency (MMIT) Kamran ; AliM4 ; XLi . In CMM systems, the microwave (MW) cavity field is coupled to a magnon mode in a ferromagnetic YIG sphere MYP as well as the magnon mode also couples to a (vibrational) mechanical mode of the sphere through the magnetostrictive force Deng . Such magnomechanical interaction is dispersive Tang ; Rome in a similar fashion to radiation pressure interaction, leading to several theoretical and experimental investigations to study various quantum phenomena in cavity magnomechanical systems analogous to cavity optomechanics, such as magnomechanically induced transparency (MMIT) Tang ; Kamran , magnetically tunable slow light Cui ; XLi , exceptional points Den , tripartite magnon-photon-phonon entanglement and Einstein-Podolsky-Rosen steerable nonlocality Tann , cooling of mechanical motion Asjad , phonon lasing Cong and parity-time-related phenomena Annl .

In this paper, we propose an unconventional and advantageous opto-magnomechanical system (OMMS) setup which is a composite architecture based on the indirect coupling of optical photons and magnon modes via the phononic mode (vibrational mode of the magnon) as shown in Fig. 1(a). The magnetoelastic coupling is responsible for the geometric deformation of the magnon in a YIG bridge which forms the vibrational motion. We consider that the phonon frequency is much smaller than the frequency of the magnon and in this situation, the dispersive interaction between the phonon and magnon modes becomes dominant (See Appendix A for details). This present scheme can be well fabricated by placing a high reflectivity small size mirror on the surface (edge) of the YIG bridge and therefore, the vibrational motion of the magnon can be easily coupled to an optical cavity through the radiation pressure, assembling an optomechanical cavity. However, the tiny size of the mirror does not affect the vibrations of the YIG bridge. In addition, such kind of indirect coupling between optical photons and magnons can easily be realized by assuming the low mechanical damping rate γb\gamma_{b} both in the optomechanical and magnomechanical cooperativities Com=Gc2κcγb>1C_{om}=\frac{G_{c}^{2}}{\kappa_{c}\gamma_{b}}>1 (κc\kappa_{c} being the cavity decay rate), and Cmm=Gm2κmγb>1C_{mm}=\frac{G_{m}^{2}}{\kappa_{m}\gamma_{b}}>1 (κm\kappa_{m} being the magnon decay rate) respectively, under present-day technology Mey ; Tang ; CAPE . Moreover, experimentally feasible parameters have been taken for the current scheme. In this paper, we have presented several interesting results for generating/controlling Fano profiles and the FWM by varying many system parameters.

The current article is arranged as follows. In Sect. 2, we present the theoretical model and the corresponding Hamiltonian of the OMMS. Section 3 designates the dynamics of the OMMS by exploiting the standard quantum Langevin approach, to get the expression for the optical response and the FWM process. In Sect. 4, we analytically discuss the optical response and the FWM process in detail. Finally, the conclusion is given in the last Sect. 5.

II The model and the dynamics

The opto-magnomechanical system under consideration consists of an optical mode cc and a slab of YIG nanoresonator in which collective excitation of a number of spins are coupled to mechanical vibrations Heyroth . It is vital to mention here that a micrometer size YIG bridge can well support a gigahertz (megahertz) magnon mode (vibrational mode) Heyroth . Therefore, such a magnomechanical system has a strong magnon-phonon dispersive coupling disp2 ; mag ; Kittel . The present scheme can be accomplished by placing a small mirror on the edge (surface) of the YIG bridge as shown in Fig. 1 (a), and in this way, an optical field can be coupled to the magnetostriction-induced mechanical displacement via radiation pressure force. The fabricated mirror should be highly refractive, light, and small such that it will not affect the mechanical properties of the ferrimagnet. Therefore, in this way the magnomechanical displacement can be probed by light exploiting the optomechanical interaction, which is further used to determine the magnon population ZRCS . Moreover, the deformation mode (displacement of the mechanical resonator) is along the direction of the sticky surface area of the slab, such that the tightly coupled YIG bridge and the mirror resonate with the same frequency. In addition, the dispersive magnetostrictive interaction between magnon and phonon modes depends on the resonance frequencies of these coupled modes Zha .

It is important to mention here that Fig.1 is only a diagrammatic sketch of one of the possible configurations to couple the magnomechanically induced displacement to an optical cavity. In addition, one doesn’t have to attach a mirror (on the surface of the microbridge) to form the optical cavity. Alternatively, one can also adopt the analogy of taking a ”membrane-in-the-middle” configuration by positioning the YIG sample inside the optical cavity. In this setup, the magnomechanically induced displacement can also be dispersively coupled to the optical cavity. In this case, no additional damping of the mechanical mode is needed. In the current study, we consider a large size crystal such that the magnon frequency is considered to be much greater than the vibrational frequency of the phononic mode such that the dispersive interaction between magnon and phonon modes, becomes dominant Tang ; Ballestero . The Hamiltonian of the present OMMS reads

H/=H+Hin+Hdr,H/\hbar=H_{\circ}+H_{in}+H_{dr}, (1)

where

H\displaystyle H_{\circ} =\displaystyle= ωccc+ωmmm+ωb2(q2+p2),\displaystyle\omega_{c}c^{\dagger}c+\omega_{m}m^{\dagger}m+\frac{\omega_{b}}{2}(q^{2}+p^{2}), (2)
Hin\displaystyle H_{in} =\displaystyle= gmbmmqgcbccq,\displaystyle g_{mb}m^{\dagger}mq-g_{cb}c^{\dagger}cq, (3)
Hdr\displaystyle H_{dr} =\displaystyle= iΩ(meiωtmeiωt)\displaystyle i\Omega(m^{\dagger}e^{-i\omega_{\circ}t}-me^{i\omega_{\circ}t}) (4)
+if=L,pεf(ceiωftceiωft).\displaystyle+i\sum_{f=L,p}\varepsilon_{f}(c^{\dagger}e^{-i\omega_{f}t}-ce^{i\omega_{f}t}).
Refer to caption
Figure 1: (Color Online) (a) Schematic diagram of the opto-magnomechanical system where mechanical displacement is simultaneously coupled to the magnonic mode via the dispersive magnetostrictive interaction and to an optical cavity through radiation pressure force. (b) Schematic diagram of the Kittel and deformation modes.

where c(m)c(m) and c(m)c^{{\dagger}}\left(m^{{\dagger}}\right) are the respective annihilation and creation operators for the cavity (magnon) mode. pp and qq are the dimensionless momentum and position of the phononic mode (deformation vibrational mode), considered as a mechanical resonator, which simultaneously couples to the cavity (magnon) mode via radiation pressure interaction gcbg_{cb} (dispersive magnetostrictive interaction gmbg_{mb}). The details of the optomechanical coupling gcbg_{cb} can be seen in CGEN while the detailed discussion about the magnomechanical coupling gmbg_{mb} is present in Appendix A. In addition, ωc\omega_{c}(ωm\omega_{m})[ωb\omega_{b}] is the resonance frequencies of the cavity mode (magnon mode) [phononic mode]. ωm\omega_{m} can be flexibly tuned via ωm=γ0Hm\omega_{m}=\gamma_{0}H_{m}, where HmH_{m} is the bias magnetic field and γ0\gamma_{0} is the gyromagnetic ratio. The Rabi frequency Ω=(5/4)γ0NsB0\Omega=\left(\sqrt{5}/4\right)\gamma_{0}\sqrt{N_{s}}B_{0} Jiee represents the coupling strength of the input laser drive field with frequency ω0\omega_{0}, where γ0=28\gamma_{0}=28GHz/T, B0=3.9×109B_{0}=3.9\times 10^{-9}T and the total number of spins Ns=ρVN_{s}=\rho V, where ρ=4.22×1027m3\rho=4.22\times 10^{27}m^{-3} denotes the spin density and VV being the volume of the YIG bridge. In addition, we apply the Holstein-Primakoff transformation to the bosonic operators for the magnon (mm and mm^{\dagger}), which basically form from the collective motion of the spins. In fact, Ω\Omega can be derived by the basic low-lying excitations assumption i.e., mm2Ns\langle m^{\dagger}m\rangle\ll 2Ns, where NN (s=52s=\frac{5}{2}) is the total number of spin states (the spin number of the ground state Fe3+ ion) in YIG Holstein . Furthermore, we also drive the optical cavity through external laser field with amplitude εL=κcL/ωL\varepsilon_{L}=\sqrt{\kappa_{c}\wp_{L}/\hbar\omega_{L}}, where κc\kappa_{c} is the decay rate of the optical cavity mode, ωL\omega_{L} (L\wp_{L}) is the frequency (power) of the external input laser field. At the same time, the cavity mode is also driven by a weak probe field with amplitude εL\varepsilon_{L} and frequency p\wp_{p}. The corresponding Hamiltonian of the OMMS, after the rotating wave approximation at the coupling frequency ωL\omega_{L}, is given by

H/\displaystyle H/\hbar =\displaystyle= Δc0cc+Δm0mm+ωb2(q2+p2)+gmbmmq\displaystyle\Delta_{c}^{0}c^{{\dagger}}c+\Delta_{m}^{0}m^{{\dagger}}m+\frac{\omega_{b}}{2}(q^{2}+p^{2})+g_{mb}m^{{\dagger}}mq (5)
gcbccq+iΩ(mm)+iεL(cc)\displaystyle-g_{cb}c^{\dagger}cq+i\Omega(m^{{\dagger}}-m)+i\varepsilon_{L}(c^{\dagger}-c)
+iεp(ceiδtceiδt),\displaystyle+i\varepsilon_{p}(c^{\dagger}e^{-i\delta t}-ce^{i\delta t}),

where Δm0=ωmωL\Delta_{m}^{0}=\omega_{m}-\omega_{L}, Δc0=ωcωL\Delta_{c}^{0}=\omega_{c}-\omega_{L} and δ=ωpωL\delta=\omega_{p}-\omega_{L}.

Now, we discuss the dynamics of this OMMS by writing the corresponding QLEs, which are given by

q˙\displaystyle\dot{q} =\displaystyle= ωbp,\displaystyle\omega_{b}p,
p˙\displaystyle\dot{p} =\displaystyle= ωbqγbpgmbmm+gcbcc+ξ,\displaystyle\omega_{b}q-\gamma_{b}p-g_{mb}m^{\dagger}m+g_{cb}c^{\dagger}c+\xi,
c˙\displaystyle\dot{c} =\displaystyle= (iΔc0+κc)c+igcbcq+εL+εpeiδt+κccin,\displaystyle-(i\Delta_{c}^{0}+\kappa_{c})c+ig_{cb}cq+\varepsilon_{L}+\varepsilon_{p}e^{-i\delta t}+\sqrt{\kappa_{c}}c^{in},
m˙\displaystyle\dot{m} =\displaystyle= (iΔm0+κm)migmbmq+Ω+κmmin,\displaystyle-(i\Delta_{m}^{0}+\kappa_{m})m-ig_{mb}mq+\Omega+\sqrt{\kappa_{m}}m^{in}, (6)

where κc\kappa_{c}(κm\kappa_{m}) is the decay rate of the cavity mode (magnon mode), and γb\gamma_{b} is the damping rate of the vibrational mode. Furthermore, ξ\xi, cinc^{in} and minm^{in} are, respectively, the input noise operators for the vibrational mode, cavity modes and magnon mode. The above QLEs can be linearized by writing z=zs+δzz=z_{s}+\delta z, (z=p,q,c,m)(z=p,q,c,m), where zsz_{s} (δz\delta z) is the steady state value (fluctuation) of any operator. For a sufficiently large amplitude of magnon and cavity modes, average value should be much greater than the fluctuation. Substituting the above ansatz into Eq.(6) and ignoring the higher order perturbations, the average values of the dynamical operators are given by

ps\displaystyle p_{s} =\displaystyle= 0,\displaystyle 0,
qs\displaystyle q_{s} =\displaystyle= gcb|cs|2ωbgcb|ms|2ωb,\displaystyle g_{cb}\frac{\left|c_{s}\right|^{2}}{\omega_{b}}-g_{cb}\frac{\left|m_{s}\right|^{2}}{\omega_{b}},
cs\displaystyle c_{s} =\displaystyle= εL(iΔc+κc),\displaystyle\frac{\varepsilon_{L}}{(i\Delta_{c}+\kappa_{c})},
ms\displaystyle m_{s} =\displaystyle= Ω(iΔm+κm),\displaystyle\frac{\Omega}{(i\Delta_{m}+\kappa_{m})}, (7)

where Δc=Δc0gcbqs\Delta_{c}=\Delta_{c}^{0}-g_{cb}q_{s} and Δm=Δm0+gmbqs\Delta_{m}=\Delta_{m}^{0}+g_{mb}q_{s} are the effective cavity and magnon mode detunings, respectively, which include the slight shift of frequency. The corresponding linearized QLEs, after ignoring quantum and thermal noise terms, are:

Refer to caption
Figure 2: (Color Online) (A) The real part Re(ϵT\epsilon_{T}) as a function of δ/ωb\delta/\omega_{b} and Δm/ωb\Delta_{m}/\omega_{b} when Δc/ωb=1\Delta_{c}/\omega_{b}=1. Contour plot of real part Re(ϵT\epsilon_{T}) as a function of δ/ωb\delta/\omega_{b} and Δm/ωb\Delta_{m}/\omega_{b} when (B) Δc=1.05ωb\Delta_{c}=1.05\omega_{b} and (C) Δc=0.95ωb\Delta_{c}=0.95\omega_{b}. The rest of the parameters are same as in Table .1.
δq˙\displaystyle\delta\dot{q} =\displaystyle= ωbδp,\displaystyle\omega_{b}\delta p,
δp˙\displaystyle\delta\dot{p} =\displaystyle= ωbδqγbδpGmδmGmδm+Gcδc+Gcδc,\displaystyle\omega_{b}\delta q-\gamma_{b}\delta p-G_{m}\delta m^{\dagger}-G_{m}^{\ast}\delta m+G_{c}\delta c^{\dagger}+G_{c}^{\ast}\delta c,
δc˙\displaystyle\delta\dot{c} =\displaystyle= (iΔc+κc)δc+iGcδq+εpeiδt,\displaystyle-(i\Delta_{c}+\kappa_{c})\delta c+iG_{c}\delta q+\varepsilon_{p}e^{-i\delta t},
δm˙\displaystyle\delta\dot{m} =\displaystyle= (iΔm+κm)δmiGmδq.\displaystyle-(i\Delta_{m}+\kappa_{m})\delta m-iG_{m}\delta q. (8)

Here, Gmb=2gmbmsG_{mb}=\sqrt{2}g_{mb}m_{s} and Gcb=2gcbcsG_{cb}=\sqrt{2}g_{cb}c_{s} are the effective magnomechanical and optomechanical coupling, respectively. The obtained fluctuation operators can then be solved by setting: δz=zeiδt+z+eiδt\delta z=z_{-}e^{-i\delta t}+z_{+}e^{i\delta t} where zz_{-} and z+z_{+} (with z=q,p,c,mz=q,p,c,m) are much smaller than zsz_{s}. It is noteworthy to mention that the fluctuation operators contain many Fourier components after expansion. Moreover, the high-order terms can safely be neglected in the limit of weak signal field. Substituting the above ansatz into Eq. (8) and taking the lowest order in εp\varepsilon_{p}, we establish the final solution for the coefficients of optical response and FWM process as

c=[ωb2δ2iΘn]εp[ωb2δ2iΘn]Ω2ciωbGcb2,c_{-}=\frac{[\omega_{b}^{2}-\delta^{2}-i\Theta_{n}]\varepsilon_{p}}{[\omega_{b}^{2}-\delta^{2}-i\Theta_{n}]\Omega_{2}^{c}-i\omega_{b}G_{cb}^{2}}, (9)

and

c+=iGcb2ωbΘpεp[ωb2δ2+iΘp](Ω1c)iωbGcb2,c_{+}=\frac{iG_{cb}^{2}\omega_{b}\Theta_{p}\varepsilon_{p}}{[\omega_{b}^{2}-\delta^{2}+i\Theta_{p}]\left(\Omega_{1}^{c}\right)^{\ast}-i\omega_{b}G_{cb}^{2}}, (10)

where

Θn\displaystyle\Theta_{n} =\displaystyle= δγbωbχ12,Θp=δγb+ωbα12,\displaystyle\delta\gamma_{b}-\omega_{b}\chi_{12},\ \ \ \ \ \ \ \ \ \Theta_{p}=\delta\gamma_{b}+\omega_{b}\alpha_{12},
χ12\displaystyle\chi_{12} =\displaystyle= Gmb2χ1+Gcb2χ2,χ1=[Ω2mΩ1m][Ω1mΩ2m]1,\displaystyle G_{mb}^{2}\chi_{1}+G_{cb}^{2}\chi_{2},\ \ \ \chi_{1}=[\Omega_{2}^{m}-\Omega_{1}^{m}][\Omega_{1}^{m}\Omega_{2}^{m}]^{-1},
χ2\displaystyle\chi_{2} =\displaystyle= [Ω1c]1,α2=[(Ω2c)]1,\displaystyle[\Omega_{1}^{c}]^{-1},\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \alpha_{2}=[\left(\Omega_{2}^{c}\right)^{\ast}]^{-1},
α12\displaystyle\alpha_{12} =\displaystyle= Gmb2α1+Gab2α2,\displaystyle G_{mb}^{2}\alpha_{1}+G_{ab}^{2}\alpha_{2},
α1\displaystyle\alpha_{1} =\displaystyle= [(Ω1m)(Ω2m)][(Ω1m)(Ω2m)]1,\displaystyle[\left(\Omega_{1}^{m}\right)^{\ast}-\left(\Omega_{2}^{m}\right)^{\ast}][\left(\Omega_{1}^{m}\right)^{\ast}\left(\Omega_{2}^{m}\right)^{\ast}]^{-1},
Ω1m\displaystyle\Omega_{1}^{m} =\displaystyle= κmi(Δm+δ),Ω2m=κm+i(Δmδ),\displaystyle\kappa_{m}-i(\Delta_{m}+\delta),\ \ \ \ \ \ \ \ \ \Omega_{2}^{m}=\kappa_{m}+i(\Delta_{m}-\delta),
Ω1c\displaystyle\Omega_{1}^{c} =\displaystyle= κci(Δc+δ),Ω2c=κc+i(Δcδ).\displaystyle\kappa_{c}-i(\Delta_{c}+\delta),\ \ \ \ \ \ \ \ \ \ \ \ \ \Omega_{2}^{c}=\kappa_{c}+i(\Delta_{c}-\delta).

III Output field of OMMS

To investigate the optical properties of the output field and the FWM process in OMMS, we start with the following standard input-output relation DFW

cout(t)=cin(t)2κcc(t).c_{out}(t)={c}_{in}(t)-\sqrt{2\kappa_{c}}c(t). (11)

In the above formula, cin{c}_{in} and coutc_{out} are, respectively the input field and output field operators. Furthermore, the expectation value of the output field is given by

cout(t)\displaystyle c_{out}(t) =\displaystyle= (2κcc0εL2κc)eiωLt\displaystyle\left(\sqrt{2\kappa_{c}}c_{0}-\frac{\varepsilon_{L}}{\sqrt{2\kappa_{c}}}\right)e^{-i\omega_{L}t} (12)
+(2κccεp2κc)ei(δ+ωL)t\displaystyle+\left(\sqrt{2\kappa_{c}}c_{-}-\frac{\varepsilon_{p}}{\sqrt{2\kappa_{c}}}\right)e^{-i(\delta+\omega_{L})t}
+2κcc+ei(δωL)t,\displaystyle+\sqrt{2\kappa_{c}}c_{+}e^{i(\delta-\omega_{L})t},
=\displaystyle= (2κcc0εL2κc)eiωLt\displaystyle\left(\sqrt{2\kappa_{c}}c_{0}-\frac{\varepsilon_{L}}{\sqrt{2\kappa_{c}}}\right)e^{-i\omega_{L}t}
+(2κccεp2κc)eiωpt\displaystyle+\left(\sqrt{2\kappa_{c}}c_{-}-\frac{\varepsilon_{p}}{\sqrt{2\kappa_{c}}}\right)e^{-i\omega_{p}t}
+2κcc+ei(ωp2ωL)t,\displaystyle+\sqrt{2\kappa_{c}}c_{+}e^{i(\omega_{p}-2\omega_{L})t},

where the second (third) term corresponds to the components at the frequencies ωp\omega_{p} (2ωLωp2\omega_{L}-\omega_{p}) representing the optical (FWM) response of the system induced by the anti-Stokes (Stokes) scattering process. In addition, cout(t)c_{out}(t) can also be expanded as

Refer to caption
Figure 3: (Color Online) The real part Re(εT\varepsilon_{T}) as a function of normalized detuning for different value of Δm\Delta_{m} and GmbG_{mb}. In figure(a-f), solid orange represents Δm=0\Delta_{m}=0, dashed red represents Δm=0.3ωb\Delta_{m}=0.3\omega_{b}, dot-dashed green represents Δm=0.7ωb\Delta_{m}=0.7\omega_{b} and solid gray represents Δm=0.9ωb\Delta_{m}=0.9\omega_{b}. We take Gmb=0.2ωbG_{mb}=0.2\omega_{b} for the left panel and Gmb=0.5ωbG_{mb}=0.5\omega_{b} for the right panel. Moreover, we take Δc=0.9ωb\Delta_{c}=0.9\omega_{b} in (a-b), Δc=ωb\Delta_{c}=\omega_{b} in (c-d), Δc=1.1ωb\Delta_{c}=1.1\omega_{b} in (e-f). The rest of the parameters are same as in Table .1.
Refer to caption
Figure 4: (Color Online) The Real part Re(εT\varepsilon_{T}) as a function of normalized detuning for different value of Δc\Delta_{c}. In figure(a-c), solid orange represents Δc=ωb\Delta_{c}=\omega_{b}, dashed red represents Δc=0.9ωb\Delta_{c}=0.9\omega_{b} and dot-dashed black represents Δc=0.8ωb\Delta_{c}=0.8\omega_{b}.In figure(d-f),solid orange represents Δc=ωb\Delta_{c}=\omega_{b}, dashed red represents Δc=1.1ωb\Delta_{c}=1.1\omega_{b} and dot-dashed black represents Δc=1.2ωb\Delta_{c}=1.2\omega_{b}. We take, Δm=0.1ωb\Delta_{m}=0.1\omega_{b} in (a)(d), Δm=0.3ωb\Delta_{m}=0.3\omega_{b} in (b)(e) and Δm=0.5ωb\Delta_{m}=0.5\omega_{b} in (c)(f). The rest of the parameters are same as in Table .1.
cout(t)\displaystyle\left\langle c_{out}(t)\right\rangle =\displaystyle= cout,0eiωLt+cout,ei(δ+ωL)t\displaystyle c_{out,0}e^{-i\omega_{L}t}+c_{out,-}e^{-i(\delta+\omega_{L})t} (13)
+cout,+ei(δωL)t,\displaystyle+c_{out,+}e^{i(\delta-\omega_{L})t},

Comparing Eq. (12) and Eq. (13), we easily arrive at

cout,\displaystyle c_{out,-} =\displaystyle= 2κccεp2κc,\displaystyle\sqrt{2\kappa_{c}}c_{-}-\frac{\varepsilon_{p}}{\sqrt{2\kappa_{c}}},
cout,+\displaystyle c_{out,+} =\displaystyle= 2κcc+.\displaystyle\sqrt{2\kappa_{c}}c_{+}. (14)

Finally, we obtain the following total output field relation at the probing frequency ωp\omega_{p}

εT\displaystyle\varepsilon_{T} =\displaystyle= 2κccout,+εpεp=2κccεp=Λ+iΛ~,\displaystyle\frac{\sqrt{2\kappa_{c}}c_{out,-}+\varepsilon_{p}}{\varepsilon_{p}}=\frac{2\kappa_{c}c_{-}}{\varepsilon_{p}}=\Lambda+i\tilde{\Lambda}, (15)

where Λ\Lambda=Re[εT\varepsilon_{T}] and Λ~\tilde{\Lambda}=Im[εT\varepsilon_{T}] are, respectively, represent the in-phase and out-of-phase quadratures of the output probe field. In addition, one can obtained the absorptive (dispersive) behavior of the output probe field from Λ\Lambda(Λ~\tilde{\Lambda}) AMEPJD .

Similarly, at Stokes scattering, the FWM intensity can be obtained from output stokes field as

εFWM=|2κccout,+εp|2=|2κcc+εp|2.\varepsilon_{FWM}=\left|\frac{\sqrt{2\kappa_{c}}c_{out,+}}{\varepsilon_{p}}\right|^{2}=\left|\frac{2\kappa_{c}c_{+}}{\varepsilon_{p}}\right|^{2}. (16)
Parameters Symbol Value
Decay rate of the cavity mode κc\kappa_{c} 0.1ωb0.1\omega_{b}
Decay rate of the magnon mode κm\kappa_{m} 0.1κc\simeq 0.1\kappa_{c}
Mechanical damping rate γb\gamma_{b} 105ωb10^{-5}\omega_{b}
Effective Cavity coupling GcbG_{cb} 0.05ωb0.05\omega_{b}
Effective magnon coupling GmbG_{mb} 0.5ωb0.5\omega_{b}
Table 1: The parameters used in our calculations, taken from recent the experiments in Ref Tang ; disp2 ; AMPS .

IV Numerical results and discussions

IV.1 Fano-type optical response

In this section, we present the frequency response and the FWM process in OMMS by employing the system parameters. For the present OMMS, we have selected parameters which are well experimentally realizable and are given in Table 1 Tang ; disp2 ; AMPS . The practical feasibility of these parameters is given in AMPS ; ASIJTP ; AMPS2 . It is noteworthy that the cavity mode is weakly driven as compared to the driven magnon, leading to |cs|2|ms|2\left|c_{s}\right|^{2}\ll\left|m_{s}\right|^{2} and therefore, GcbGmbG_{cb}\ll G_{mb}. Hence, one can safely claim that the megnetostrictive interaction is paramount over optomechanical interaction, i.e., qsgcb|ms|2ωbq_{s}\simeq-g_{cb}\frac{\left|m_{s}\right|^{2}}{\omega_{b}} (See Eq. (7)). In addition, it has already been proved that effective magnomechanical coupling, GmbG_{mb} is almost 1010 times greater than effective magnomechanical coupling, GcbG_{cb} ASIJTP ; ZRCS . Therefore for simplicity, we have assumed Gmb=(0.20.5)ωbG_{mb}=(0.2-0.5)\omega_{b} and Gcb=0.05ωbG_{cb}=0.05\omega_{b}.

Fano resonances emerge due to the interaction between the cavity/magnon mode and the vibrational mode of the resonator. We have realized the resonances in the output spectra viz. εT=Λ+iΛ~\varepsilon_{T}=\Lambda+i\tilde{\Lambda}. Since OMIT originates by the constructive and destructive interference of different frequency contributions, it is well known that we can expect Fano profiles under some conditions KQGS . We first note that when the magnon (or cavity) detuning is out of resonance with the vibrational frequency, for instance Δm<ωb\Delta_{m}<\omega_{b} and Δcωb\Delta_{c}\neq\omega_{b} (or even Δc=ωb\Delta_{c}=\omega_{b}), consequently Fano profiles emerge in OMIT spectra, in agreement with previous reports KQGS ; MYPOM1 . To discuss the emergence and tunability of Fano profiles in OMMS, we first show the absorption spectra as a function of normalized detuning δ/ωb\delta/\omega_{b} for a fixed value of magnomechanical coupling rate GmbG_{mb}, when Δc=ωb\Delta_{c}=\omega_{b} in Fig. 2. Starting from Δm/ωb=0\Delta_{m}/\omega_{b}=0, we tuned the frequency Δm\Delta_{m} in such a way that it approached the vicinity of the mechanical mode frequency ωb\omega_{b} (see the green curve in Fig. 2 (a)). In the absence of magnon interaction, two dips in absorption can be found, which appear around δ=±ωb\delta=\pm\omega_{b} as shown by the magenta line in Fig. Fig. 2 (a) (Also in inset). This is because in the present system, we have taken extremely low value of the damping rate and therefore the lowest value of the output field amplitude exist when ωb2δ20\omega_{b}^{2}-\delta^{2}\simeq 0 (or δ=±ωb\delta=\pm\omega_{b}). Consequently, we observe a typical OMIT windows around these points. However, the magnomechanical interaction involve to exhibit asymmetric Fano profiles as shown by the yellow, red and green lines in Fig. 2(a). In addition, we obtain a curve almost similar to Lorentzian when Δm=ωb\Delta_{m}=\omega_{b}. Hence, the interaction of magnomechanical interaction leads to enhance the absorption around δ=ωb\delta=\omega_{b} around Δm=ωb\Delta_{m}=\omega_{b}. However the peak of the Fano profiles gradually decrease with the increase of Δm\Delta_{m}. The inset 3D figure in Fig. (a) exhibits the same absorption spectra in the blue-detuned region, which shows the small amplification of each curve occurs where we observe the dip in the red-detuned region. For example, the absorption dips (amplification dips) of the green line in red-detuned region occur at δ=0.55ωb\delta=0.55\omega_{b} (δ=0.55ωb\delta=-0.55\omega_{b}) and δ=1.3ωb\delta=1.3\omega_{b} (δ=1.3ωb\delta=-1.3\omega_{b}) (-ve values can be seen in inset of Fig. 2(a)). To further shed light on the Fano profiles, we show the Fano profile spectra for a specific range of magnon detuning as density plots in Fig. 2 (b) and Fig. 2 (c) at Δc=1.05ωb\Delta_{c}=1.05\omega_{b} and Δc=0.95ωb\Delta_{c}=0.95\omega_{b} respectively. The Fano profiles become more prominent when we slightly shift the cavity detuning to higher frequency (see Fig. 2 (b)) as compared to the Fano profile spectra for the lower cavity detuned region (see Fig. 2 (c)).

As already discussed, the existence of Fano profiles in OMMS is due to the nonresonant interactions of magnon (or cavity) mode, the symmetry of the typical OMIT window is transfigured into asymmetric Fano profiles. Therefore, in the present OMMS, we investigate the Fano profile which can be tuned by the effective magnomechanical coupling strength GmbG_{mb} and magnon detuning Δm\Delta_{m} for both resonance (Δc=ωb\Delta_{c}=\omega_{b}) as well as a non-resonance case (Δcωb\Delta_{c}\neq\omega_{b}). In Fig. 3, we show a typical behavior of absorption spectra as a function of the normalized probe-coupling detuning for different values of Δm\Delta_{m} and GmbG_{mb}. We obtained a typical OMIT window in the absence of magnon interaction as show by the orange solid line in Fig. 3 (c-d). However, the asymmetry of the absorption spectra incenses with the increase of Δm\Delta_{m} as shown by the dashed red, dot-dashed green and solid black lines. Similarly, Fig. 3 (a-b) shows the absorption spectra when Δc=0.9ωb\Delta_{c}=0.9\omega_{b} and Fig. 3 (e-f) shows the absorption spectra when Δc=1.1ωb\Delta_{c}=1.1\omega_{b}. The height of the Fano profiles decrease with the increase of Δm\Delta_{m}. In addition, one can see that the width between the central absorption peak and sharp Fano profile for left panel of Fig. 3, is not prominently increase while for the right panel, it sufficiently increases. Therefore, it is clear from the above discussion, that Δm\Delta_{m} (GmbG_{mb}) controls the peak/dip (width) of the Fano profile.

In order to observe the combined effect of magnon and cavity detunings, we plot the absorption spectra by varying Δm/ωb\Delta_{m}/\omega_{b} and Δc/ωb\Delta_{c}/\omega_{b} in Fig. 4. We now present the absorption profile as a function of normalized detuning δ/ωb\delta/\omega_{b} at the fixed effective coupling strengths GmbG_{mb} and GcbG_{cb}. For a fixed value of Δm/ωb\Delta_{m}/\omega_{b}, we continuously tuned the cavity detuning from Δc=0.8ωb\Delta_{c}=0.8\omega_{b} to Δc=1.2ωb\Delta_{c}=1.2\omega_{b} as shown in Fig. 4. It is noted that the Lorentzian type peak move left (right) as cavity detuning Δc/ωb\Delta_{c}/\omega_{b} decreases (increases). However, the peak of the sharp Fano profile depends mainly on the magnon detuning Δm/ωb\Delta_{m}/\omega_{b} i.e., peak (dip) decreases (move slightly upward) with the increase of Δm/ωb\Delta_{m}/\omega_{b}.

IV.2 Four-wave mixing process

Refer to caption
Figure 5: (Color Online) FWM intensity IFWMI_{FWM} as a function of δ/ωb\delta/\omega_{b} and Δm/ωb\Delta_{m}/\omega_{b}. We take Δc=ωb\Delta_{c}=\omega_{b} in (a-b) and Δm=0.4ωb\Delta_{m}=0.4\omega_{b} in (c). The rest of the parameters are same as in Table .1.
Refer to caption
Figure 6: (Color Online) FWM intensity IFWMI_{FWM} as a function of (A) Δc/ωb\Delta_{c}/\omega_{b} and δ/ωb\delta/\omega_{b} and (B) Gmb/ωbG_{mb}/\omega_{b} and δ/ωb\delta/\omega_{b}. We take Δm=0.5ωb\Delta_{m}=0.5\omega_{b} in (A) and Δm=ωb\Delta_{m}=\omega_{b} in (B). The rest of the parameters are same as in Table .1.

Compared with the typical cavity optomechanical system comprising a mechanical resonator coupled with an optical cavity via radiation pressure force, the OMMS under consideration allows a greater controllability on the FWM process. Here, we concentrate on different parameters to tune the FWM signals, i.e., the cavity/magnon detuning, the decay rate and effective magnomechanical coupling strength. Fig. 5(A) plots the FWM spectrum as a function of the pump-probe detuning for different values of the magnon detuning. It is shown in Fig. 5(A) that the two peaks of FWM signals are obtained exactly at δ=±ωb\delta=\pm\omega_{b}, where OMMS exhibits transparency. The peaks at δ=±ωb\delta=\pm\omega_{b} represent the intensity of the stokes field produced by the coupling field with the vibrational mode frequency. In addition, for Δm<0.4ωb\Delta_{m}<0.4\omega_{b}, the FWM signal remains almost the same as shown in Fig. 5(A). However, one can see an extra peak located at δ=0\delta=0 along with the two peaks located at δ±0.1ωb\delta\simeq\pm 0.1\omega_{b} appear when Δm=0.5ωb\Delta_{m}=0.5\omega_{b}. Therefore, one can infer that the middle peak arises owing to the strong magnomechanical interaction. In addition, for Δm0.6ωb\Delta_{m}\geq 0.6\omega_{b}, the peak at δ=0\delta=0 further split to create two peaks at δ=±0.2ωb\delta=\pm 0.2\omega_{b} along with two peaks at δ=±1.2ωb\delta=\pm 1.2\omega_{b}. However, the strength of the two peaks at δ=±0.2ωb\delta=\pm 0.2\omega_{b} sufficiently smaller than the two peaks around δ=±ωb\delta=\pm\omega_{b}. It can be seen that as the magnomechanical interaction becomes strong (i.e., for Δm0.6ωb\Delta_{m}\geq 0.6\omega_{b}), the two peaks at δ=±0.2ωb\delta=\pm 0.2\omega_{b} become heighten as compared to the two peaks around δ=±ωb\delta=\pm\omega_{b} as shown in Fig. 5(B). This effect becomes more prominent in Fig. 6(a-b). Furthermore, the width between the peaks gets broader by increasing Δm\Delta_{m} as shown in Fig. 5(B). It is important to mention here that for increasing value of Δm\Delta_{m}, the FWM signals at δ=ωb\delta=\omega_{b} (δ=ωb\delta=-\omega_{b}) start moving to the right(left). Therefore, we infer that the signal at δ=±ωb\delta=\pm\omega_{b} is completely suppressed at due to the destructive interference (see the dark blue line). Fig. 5(C) plots the FWM spectrum as a function of the pump-probe detuning for different value of the cavity detuning which clearly shows the FWM signal becomes strong as we increase Δc\Delta_{c} and the width between the peaks remains constant.

Figure 6(A) shows the variation of the FWM spectrum as a function of the pump-probe detuning for different values of the effective magnon coupling strength GmbG_{mb} when Δm=0.5ωb\Delta_{m}=0.5\omega_{b}. It can be shown that the two less intensity middle peaks gets stronger and closer until Gmb0.4ωbG_{mb}\leq 0.4\omega_{b}. As the magnomechanical interaction becomes stronger, the two middle less intensified peaks then merged to form a single intensified peak for Gmb=0.5ωbG_{mb}=0.5\omega_{b}. Further enhancement of GmbG_{mb} not only brings the peak at δ=0\delta=0 to an end but also decrease the intensity of two side peaks. Figure 6(B) shows the variation of the FWM spectrum as a function of the pump-probe detuning for different value of the effective magnon coupling strength GmbG_{mb} when Δm=ωb\Delta_{m}=\omega_{b}. One can observe that the two distant peaks move further apart while the two middle peaks come closer by increasing the value of the GmbG_{mb}. In addition, the strength of the two distant peaks (middle peaks) decrease (increase) with the enhancement of GmbG_{mb}. From the above discussion, one can easily conclude that the FWM signals can be tuned (enhanced) by varying different system’s parameters.

The FWM process has been shown in cavity optomechanics, which strongly depends on the mean photon number. Varying the cavity decay rate κc\kappa_{c} (magnon decay rate κm\kappa_{m}) will directly modify the mean photon number (magnon number). Therefore, we plot the FWM signal by varying the cavity and magnon decay rates. As shown in Fig. 7(A), the FWM signal is strong and the two peaks can be seen at δ=±ωb\delta=\pm\omega_{b}. When the cavity decay rate is decreased, the two peaks of the FWM signal is sufficiently reduced. In addition, the two peaks located at δ=±ωb\delta=\pm\omega_{b} transfigured to four diminished peaks located at δ=±0.95ωb\delta=\pm 0.95\omega_{b} and δ=±1.05ωb\delta=\pm 1.05\omega_{b}. Similarly, in Fig. 7(B), we investigate the influence of the magnon decay rate κm\kappa_{m} on the FWM intensity. Figure 7(B) shows that the FWM intensity is enhanced sufficiently with the increase of the magnon decay rate κm\kappa_{m}. Hence, one can infer that cavity and magnon decay rates have a substantial impact on the FWM process FWM7 .

IV.3 Experimental Feasibility

In the following, we discuss the experimental feasibility of our proposed scheme. There exist two subsystems: (i) Magnomechanical part (ii) Optomechanical part. The adopted YIG micro-bridge (the magnomechanical subsystem) was based on the experimental demonstration Heyroth . However, the experimental demonstration of the optomechanical part has been demonstrated in many research papers OM1 ; OM2 . As for as experimental parameters are concerned, we have chosen parameters which are can easily be achieved using present-day technology. In addition, we have considered the parametric regime which is valid only when the magnon occupation number is significantly smaller than the total number of spin i.e., mm<<2Ns=5Nm^{{\dagger}}m<<2N_{s}=5N, where 5N is the total spin number of the Fe3+Fe^{3+} ion in YIG s=5/2s=5/2. Our numerical results show that NN is of the order of 3.5×10163.5\times 10^{16} and Ω=712THz\Omega=712THz Tang ; AMPS . It has been observed in the present study that these conditions are fulfilled and our scheme is feasible for strong magnon-phonon coupling. This is due to the strong magnon pump field, which causes unwanted nonlinear effects. However, as shown in recent experiments, a larger value of magnon-phonon coupling would result in a working regime, where we can drop the higher order terms due to a strongly driven magnomechanical system AMPS . From the above discussion, we infer that we have chosen an experimentally feasible working regime. In near future, any experiment employing a planar cross-shaped cavity coupled to a magnon or can allow the implementation of the proposed scheme using a co-planar waveguide RJW ; LGQ .

Refer to caption
Figure 7: (Color Online) FWM intensity IFWMI_{FWM} as a function of δ/ωb\delta/\omega_{b} for different values of (A) κc\kappa_{c} and (B) κm\kappa_{m}. We take κm=0.01ωb\kappa_{m}=0.01\omega_{b} in (A) and κc=0.1ωb\kappa_{c}=0.1\omega_{b} in (B). The rest of the parameters are same as in Table .1.

V Conclusion

In conclusion, we have introduced a model consisting of an optomechanical system directly coupled with a ferromagnetic material. The deformation of the mechanical mode is taken simultaneously due to an optomechanical coupling and the magnetoelasticity of the ferromagnet. We report that the magnetostrictively induced displacement evolves to transfigure the standard OMIT window to Fano profiles in the output field at the probe frequency. The asymmetry of the Fano profiles can be well tuned/controlled by appropriate adjusting the magnon (cavity) detuning and the effective magnomechanical coupling. Remarkably, it is found that the magnetoelastic interaction also gives rise to the FWM phenomenon. FWM intensity can be resonantly tuned by the system’s parameters. For example, modulating the effective magnomechanical coupling can realize conversion among a double and triple peak signals. Similarly, in the case of magnon detuning, FWM exhibits the characteristics of symmetric double, triple and even quadruple peak signals. In addition, one can also control the strength of the FWM signals by employing the magnon (cavity) detuning and effective magnomechanical coupling. Moreover, the FWM spectrum exhibits suppressive behavior upon increasing (decreasing) the magnon (cavity) decay. The present OMMS opens new perspectives in highly sensitive detection and quantum information processing, which shall be addressed elsewhere.

Appendix A. Magnomechanical coupling

The magnetoelastic coupling is fully explained by the interaction between the elastic strain and magnetization of the magnetic material RBWD ; Kittel . The magnetoelastic energy density for a cubic crystal is given by KIT

Fme\displaystyle F_{me} =\displaystyle= 2B1Ms2(MxMyϵxy+MxMzϵxz+MyMzϵyz)\displaystyle\frac{2B_{1}}{M_{s}^{2}}\left(M_{x}M_{y}\epsilon_{xy}+M_{x}M_{z}\epsilon_{xz}+M_{y}M_{z}\epsilon_{yz}\right) (17)
+B2Ms2(Mx2ϵxx+My2ϵyy+Mz2ϵzz),\displaystyle+\frac{B_{2}}{M_{s}^{2}}\left(M_{x}^{2}\epsilon_{xx}+M_{y}^{2}\epsilon_{yy}+M_{z}^{2}\epsilon_{zz}\right),

where the magnetoelastic coupling coefficients are represented by B1B_{1} and B2B_{2}. MsM_{s} being the saturation magnetization. In addition, magnetization components are represented by MxM_{x}, MyM_{y} and MzM_{z}. Furthermore, ϵnm\epsilon_{nm} (n,mx,y,zn,m\in{x,y,z}) defines the strain tensor of the cubic magnetic crystal.

Now, the magnetization can be quantized using the Holstein-Primakoff transformation

m=V2γMs(MxiMy),m=\sqrt{\frac{V}{2\hbar\gamma M_{s}}}\left(M_{x}-iM_{y}\right), (18)

where VV is the volume of the cubic crystal and mm defines the magnon mode operator. After rearranging, we obtain

Mx\displaystyle M_{x} =\displaystyle= γMs2V(m+m),\displaystyle\sqrt{\frac{\hbar\gamma M_{s}}{2V}}\left(m+m^{\dagger}\right),
My\displaystyle M_{y} =\displaystyle= iγMs2V(mm),\displaystyle i\sqrt{\frac{\hbar\gamma M_{s}}{2V}}\left(m-m^{\dagger}\right), (19)

and

Mz=[Ms2Mx2My2]12MsγVmm.M_{z}=\left[M_{s}^{2}-M_{x}^{2}-M_{y}^{2}\right]^{\frac{1}{2}}\simeq M_{s}-\frac{\hbar\gamma}{V}m^{\dagger}m. (20)

Putting the value of Mx,y,zM_{x,y,z} in Eq. (17) and then integrating over the volume of the cubic magnetic crystal, the first term in Eq. (17) yields the semiclassical magnetoelastic Hamiltonian,

Ha\displaystyle H_{a} \displaystyle\simeq B1MsγVmm𝑑l3(ϵxx+ϵyy2ϵzz)\displaystyle\frac{B_{1}}{M_{s}}\frac{\hbar\gamma}{V}m^{\dagger}m\int dl^{3}\left(\epsilon_{xx}+\epsilon_{yy}-2\epsilon_{zz}\right) (21)
+B1Msγ2V(m2+m2)𝑑l3(ϵxxϵyy)\displaystyle+\frac{B_{1}}{M_{s}}\frac{\hbar\gamma}{2V}\left(m^{2}+m^{\dagger 2}\right)\int dl^{3}\left(\epsilon_{xx}-\epsilon_{yy}\right)
+B1Ms22γ2V2mmmm𝑑l3ϵzz,\displaystyle+\frac{B_{1}}{M_{s}^{2}}\frac{\hbar^{2}\gamma^{2}}{V^{2}}m^{\dagger}mm^{\dagger}m\int dl^{3}\epsilon_{zz},

and the second term of magnetoelastic energy density from Eq. (17) yields the Hamiltonian

Hb\displaystyle H_{b} \displaystyle\simeq B2MsγV(m2m2)𝑑l3ϵxy\displaystyle\frac{B_{2}}{M_{s}}\frac{\hbar\gamma}{V}\left(m^{2}-m^{\dagger 2}\right)\int dl^{3}\epsilon_{xy} (22)
+2B2MsγMs2V[MsγVmm]\displaystyle+\frac{2B_{2}}{M_{s}}\sqrt{\frac{\hbar\gamma M_{s}}{2V}}\left[M_{s}-\frac{\hbar\gamma}{V}m^{\dagger}m\right]
×[mdl3(ϵxz+iϵyz)+H.c].\displaystyle\times\left[m\int dl^{3}\left(\epsilon_{xz}+i\epsilon_{yz}\right)+H.c\right].

Hence, the total magnetoelastic Hamiltonian, which is sum of HaH_{a} and HbH_{b}, implies total magnon-phonon interactions

The magnetoelastic displacement can be defined as:

u(x,y,z)=l,m,nd(l,m,n)χ(l,m,n)(x,y,z),\vec{u}\left(x,y,z\right)=\sum\limits_{l,m,n}d^{(l,m,n)}\vec{\chi}^{(l,m,n)}\left(x,y,z\right), (23)

where χ(l,m,n)(x,y,z)\vec{\chi}^{(l,m,n)}\left(x,y,z\right) (d(l,m,n)d^{(l,m,n)}) denotes the normalized displacement eigenmode (the corresponding amplitude).

d(l,m,n)=dzpm(l,m,n)(bl,m,n+bl,m,n),d^{(l,m,n)}=d_{zpm}^{(l,m,n)}\left(b_{l,m,n}+b_{l,m,n}^{\dagger}\right), (24)

where dzpm(l,m,n)d_{zpm}^{(l,m,n)} represents the amplitude of the zero-point motion. In addition, bl,m,nb_{l,m,n} and bl,m,nb_{l,m,n}^{\dagger} are, respectively the corresponding phononic annihilation and creation operators. Substituting Eqs. (23) and (24) in Eq. (21), the dispersive kind of interaction can be obtained,

Ha=l,m,ngmb(l,m,n)mm(bl,m,n+bl,m,n),H_{a}=\hbar\sum\limits_{l,m,n}g_{mb}^{(l,m,n)}m^{\dagger}m\left(b_{l,m,n}+b_{l,m,n}^{\dagger}\right), (25)

where gmb(l,m,n)g_{mb}^{(l,m,n)} represents the dispersive coupling strength between magnon and phonon, is given by

gmb(l,m,n)\displaystyle g_{mb}^{(l,m,n)} =\displaystyle= B1MsγV𝑑l3dzmp(l,m,n)\displaystyle\frac{B_{1}}{M_{s}}\frac{\gamma}{V}\int dl^{3}d_{zmp}^{(l,m,n)} (26)
×[χx(l,m,n)x+χy(l,m,n)y2χz(l,m,n)z].\displaystyle\times\left[\frac{\partial\chi_{x}^{(l,m,n)}}{\partial x}+\frac{\partial\chi_{y}^{(l,m,n)}}{\partial y}-2\frac{\partial\chi_{z}^{(l,m,n)}}{\partial z}\right].

In case of a one dimensional mechanical mode oscillation, we obtain the simplified hamiltonian as:

Ha=gmbmm(b+b),H_{a}=\hbar g_{mb}m^{\dagger}m\left(b+b^{\dagger}\right), (27)

which is the same as the first part of HiH_{i} (see Eq. (3)), since q=12(b+b)q=\frac{1}{\sqrt{2}}\left(b+b^{\dagger}\right). Note that we use the condition of low-lying magnon excitations to derive the above and therefore, the magnon excitation mmm^{\dagger}m second order term (third line of Eq. (21) ) can be safely eliminated.

Data availability

All data used during this study are available within the article.

References

  • (1) M. Aspelmeyer, P. Meystre, and K. Schwab, “Quantum optomechanics”. Phys. Today 65(7) 29 (2012).
  • (2) C. H. Dong, V. Fiore, M.C. Kuzyk, and H. L. Wang, “Optomechanical dark mode”. Science.; 338 1609 (2012).
  • (3) J. K. Moqadam, R. Portugal, M. C. de Oliveira, “Quantum walks on a circle with optomechanical systems”. Quantum Inf Process 14, 3595 (2015).
  • (4) J. D. Teufel, T. Donner, M. A. Castellanos-Beltran, J. W. Harlow and K. W. Lehnert, “Nanomechanical motion measured with an imprecision below that at the standard quantum limit”. Nature nanotechnology ; 4(12): 820 (2009).
  • (5) A. Motazedifard, A. Dalafi and M. H. Naderi, “Ultraprecision quantum sensing and measurement based on nonlinear hybrid optomechanical systems containing ultracold atoms or atomic Bose¨CEinstein condensate” AVS Quantum Sci. 3, 024701 (2021)
  • (6) Y. Jiao, H. Lu, J. Qian, Y. Li and H. Jing, “Nonlinear optomechanics with gain and loss: amplifying higher-order sideband and group delay”. New J. Phys. ; 18(8):083034 (2016).
  • (7) L. M. Ludwig, A. H. Safavi-Naeini, O. Painter and F. Marquardt, “Enhanced quantum nonlinearities in a two-mode optomechanical system”. Phys. Rev. Lett. ; 109(6):063601 (2012).
  • (8) S. K. Singh and C. H. Raymond Ooi, ¡Quantum correlations of quadratic optomechanical oscillator, J. Opt. Soc. Am. B 31, 2390 (2014).
  • (9) S. K. Singh and S. V. Muniandy, ¡Temporal Dynamics and Nonclassical Photon Statistics of Quadratically Coupled Optomechanical Systems,¡±\pm Int. J. Theor. Phys. 55, 287 (2016).
  • (10) A. Sohail, M. Rana, S. Ikram, T. Munir, T. Hussain, R. Ahmed, C. S. Yu. “Enhancement of mechanical entanglement in hybrid optomechanical system”. Quantum Information Processing ; 19(10):372 (2020).
  • (11) A. Sohail, et al., “Enhanced entanglement induced by Coulomb interaction in coupled optomechanical systems”. Phys. Scr. 95:035108 (2020).
  • (12) S. K Singh , J. X. Peng, M. Asjad and M. Mazaheri, “Entanglement and coherence in a hybrid Laguerre-Gaussian rotating cavity optomechanical system with two-level atoms”. J. Phys. B: At. Mol. Opt. Phys. 54:215502 (2021).
  • (13) Berihu Teklu, Tim Byrnes, Faisal Shah Khan. “Cavity-induced mirror-mirror entanglement in a single-atom Raman laser”. Phys. Rev. A 97:023829 (2018).
  • (14) A. Arvanitaki and A. A. Geraci, “Detecting high-frequency gravitational waves with optically levitated sensors”. Phys. Rev. Lett. 110(7):071105 (2013).
  • (15) M. S. Ebrahimi, A. Motazedifard, and M. Bagheri Harouni, “Single-quadrature quantum magnetometry in cavity electromagnonics” Physical Review A 103 (6), 062605 (2021)
  • (16) S. Weis, et al. Optomechanically induced transparency. Science 330, 1520 (2010).
  • (17) A. H. Safavi-Naeini, et al. Electromagnetically induced transparency and slow light with optomechanics. Nature (London) 472, 69 (2011).
  • (18) A. Sohail, Y. Zhang, G, Bary, C. S. Yu. “Tunable Optomechanically Induced Transparency and Fano Resonance in Optomechanical System with Levitated Nanosphere”. IJTP 57(9) 2814 (2018).
  • (19) A. Sohail, R. Ahmed, C. S. Yu. “Switchable and Enhanced Absorption via Qubit-Mechanical Nonlinear Interaction in a Hybrid Optomechanical System”. IJTP 60 739 (2021).
  • (20) F. Monifi1, J. Zhang, B. Peng1, Yu. Liu, F. Bo, Franco Nori and L. Yang, “Optomechanically induced stochastic resonance and chaos transfer between optical fields”. Nature Photon. 10(6):399 (2016).
  • (21) M. Fleischhauer, A. Imamoglu, and J. P. Marangos, “Electromagnetically induced transparency: Optics in coherent media”. Reviews of modern physics 77(2):633 (2005).
  • (22) S. K. Singh, M. Parvez, T. Abbas, J. X. Peng, M. Mazaheri, M. Asjad, “Tunable optical response and fast (slow) light in optomechanical system with phonon pump”. Phys. Lett. A 442 128181 (2022).
  • (23) S. K. Singh, M. Asjad and C. H. Raymond Ooi, “Tunable optical response in a hybrid quadratic optomechanical system coupled with single semiconductor quantum well”. Quantum Inf. Process. 21 47 (2022).
  • (24) A. E. Miroshnichenko, S. Flach, and Y. S. Kivshar, “Fano resonances in nanoscale structures”. Rev. Mod. Phys. 82(3):2257 (2010).
  • (25) U.Fano, “Effects of configuration interaction on intensities and phase shifts”. Phys. Rev. 124(6):1866 (1961).
  • (26) M. O. Scully, S. Y. Zhu, and A. Gavrielides, “Degenerate quantum-beat laser: Lasing without inversion and inversion without lasing”. Phys. Rev. Lett. 62(24):2813 (1989).
  • (27) A. Bärnthaler, S. Rotter, F. Libisch, J. Burgdorfer, S. Gehler, U. Kuhl, and H. J. Stockmann, Stöckmann H J. “Probing decoherence through Fano resonances”. Phys. Rev. Lett. 105(5):056801 (2010).
  • (28) M. O. Scully, K. R. Chapin, K. E. Dorfman, M. B. Kim, and A. A. Svidzinsky, “Quantum heat engine power can be increased by noise-induced coherence”. Proc. Natl. Acad. Sci. 108(37): 15097 (2011).
  • (29) P. Tassin, L. Zhang, R. Zhao, A. Jain, T. Koschny, and C. M. Soukoulis, “Electromagnetically induced transparency and absorption in metamaterials: the radiating two-oscillator model and its experimental confirmation”. Phys. Rev. Lett. 109(18):187401 (2012).
  • (30) Ali Motazedifard, A. Dalafi, M. H. Naderi, “Negative cavity photon spectral function in an optomechanical system with two parametrically-driven mechanical oscillators” arXiv:2205.15314 (2022).
  • (31) B. Gallinet and O. J. F. Martin, “Ab initio theory of Fano resonances in plasmonic nanostructures and metamaterials”. Phys. Rev. B 83(23):235427 (2011).
  • (32) B. Gallinet and O. J. F. Martin, “Influence of electromagnetic interactions on the line shape of plasmonic Fano resonances”. ACS nano 5(11):8999 (2011).
  • (33) A. Artar, A. Ali Yanik and H. Altug, “Directional double Fano resonances in plasmonic hetero-oligomers”. Nano Lett. 11(9):3694 (2011).
  • (34) S. M. Mozvashi, M. R. Givi, M. B. Tagani, “The effects of substrate and stacking in bilayer borophene” Scientific Reports 12 (1), 1-10 (2022).
  • (35) K. Ullah, M. T. Naseem , Ö. E.Müstecaplioglu, “Tunable multiwindow magnomechanically induced transparency, Fano resonances, and slow-to-fast light conversion”. Phys. Rev. A 102(3): 033721 (2020).
  • (36) A. B. Sabur , A. B. Bhattacherjee. “Magnomechanically induced absorption and switching properties in a dispersively coupled magnon-qubit system”. J. Appl. Phys. 132, 123104 (2022).
  • (37) L. Shang, B. Chen, L. L. Xing, J. B. Chen, H. B. Xue, and K. X. Guo, “Controllable four-wave mixing response in a dual-cavity hybrid optomechanical system”. Chin. Phys. B 30(5):295 (2021).
  • (38) M. Maeda, “General Remarks: Tunable Sources between Vacuum Ultraviolet and Far Infrared Region”. Rev. Laser Eng. 17(11):744 (1989).
  • (39) H. Allahverdi, Ali Motazedifard, A. Dalafi, D. Vitali, and M. H. Naderi, “Homodyne coherent quantum noise cancellation in a hybrid optomechanical force sensor” Physical Review A 106 (2), 023107 (2022)
  • (40) J. M. Yuan, W. K. Feng, P. Y. Li, Y. Q. Zhang, and Y. P. Zhang, “Controllable vacuum Rabi splitting and optical bistability of multi-wave-mixing signal inside a ring cavity”. Phys. Rev. A 86(6) 063820 (2012).
  • (41) H. J. Chen, H. W. Wu, J. Y. Yang, X. C. Li, Y. J. Sun, Y. Peng. “Controllable Optical Bistability and Four-Wave Mixing in a Photonic-Molecule Optomechanics”. Nanoscale Res. Lett. 14(1):73 (2019).
  • (42) K. Li, Y. Feng, X. Li, B. Gu, Y. Cai, and Y. Zhang, “Generation of competitive and coexisting two pairs of biphotons in single hot atomic vapor cell”. Ann. Phys. 532(8):2000283 (2020).
  • (43) C. Jiang, Y. Cui, H. Liu. “Controllable four-wave mixing based on mechanical vibration in two-mode optomechanical systems”. EPL (Europhysics Letters) 104(3):34004 (2013).
  • (44) X. F. Wang, B. Chen. “Four-wave mixing response in a hybrid atom-optomechanical system”. JOSA B 36(2):162 (2019).
  • (45) X. Zhang, C. L. Zhang, L. Zou, “Cavity magnomechanics”. Sci. Adv. 2 e1501286 (2016).
  • (46) F. Heyroth, et al., “Monocrystalline freestanding three-dimensional yttrium-iron-garnet magnon nanoresonators,” Phys. Rev. Applied. 12, 054031 (2019)
  • (47) H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, A. Marx, R. Gross, and S. T. B. Goennenwein, “High cooperativity in coupled microwave resonator ferrimagnetic insulator hybrids”. Phys. Rev. Lett. 111(12):127003 (2013).
  • (48) J. Li, S. Y. Zhu, and G.S. Agarwal, “Squeezed states of magnons and phonons in cavity magnomechanics”. Phys. Rev. A 99(2):021801 (2019).
  • (49) D. L. Quirion, Y. Tabuchi, A. Gloppe, K. Usami and Y. Nakamura, “Hybrid quantum systems based on magnonics”. Appl. Phys. Express 12(7):070101 (2019).
  • (50) Ali Motazedifard, A, Dalafi and M, H, Naderi, “A Green’s function approach to the linear response of a driven dissipative optomechanical system” Journal of Physics A: Mathematical and Theoretical 54 (21), 215301 (2021)
  • (51) Xiyun Li, et. al. “Phase control of the transmission in cavity magnomechanical system with magnon driving” J. Appl. Phys. 128, 233101 (2020).
  • (52) D. Zhang, X. M. Wang, T. F. Li, X. Q. Luo, W. Wu, F. Nori and J. You, “Cavity quantum electrodynamics with ferromagnetic magnons in a small yttrium-iron-garnet sphere”. npj Quantum Inf. 1(1):15014 (2015).
  • (53) C. Gonzalez-Ballestero, D. Hummer, J. Gieseler, and O. Romero-Isart, “Theory of quantum acoustomagnonics and acoustomechanics with a micromagnet”. Phys. Rev. B 101(12):125404 (2020).
  • (54) C. Kong, B. Wang, Z.-X. Liu, H. Xiong and Y. Wu, “Magnetically controllable slow light based on magnetostrictive forces”. Opt. express 27(4):5544 (2019).
  • (55) D. Zhang, X. Q. Luo, Y. P. Wang, T.-F. Li and J.Q. You, “Observation of the exceptional point in cavity magnon-polaritons”. Nat. commun. 8(1):1368 (2017).
  • (56) H. Tan, “Genuine photon-magnon-phonon Einstein-Podolsky-Rosen steerable nonlocality in a continuously-monitored cavity magnomechanical system”. Phys. Rev. Research 1(3):033161 (2019).
  • (57) M. Asjad, Jie Li, Shi-Yao Zhu and J. Q. You, “Magnon squeezing enhanced ground-state cooling in cavity magnomechanics”. arXiv 2022:2203.10767.
  • (58) M. S. Ding, L. Zheng and C. Li, “Phonon laser in a cavity magnomechanical system”. Sci. Rep. 9(1):15723 (2019).
  • (59) L. Wang, Z. X. Yang, Y. M. Liu, C. H. Bai, D. Y. Wang, S. Zhang and H. F. Wang, “Magnon blockade in a PT-symmetric-like cavity magnomechanical system”. Ann. Phys. 532(7):2000028 (2020).
  • (60) C. A. Potts, E. Varga, V. Bittencourt, S. V. Kusminskiy, and J. P. Davis, “Dynamical backaction magnomechanics” Phys. Rev. X 11, 031053 (2021).
  • (61) J. Li, S. Y. Zhu, G. S. Agarwal., “Squeezed states of magnons and phonons in cavity magnomechanics”. Phys. Rev. A 99(2):021801 (2019).
  • (62) J. Li, S. Y. Zhu, G. S. Agarwal., “Magnon-photon-phonon entanglement in cavity magnomechanics”. Phys. Rev. Lett. 121(20):203601 (2018).
  • (63) C. Kittel., “Interaction of spin waves and ultrasonic waves in ferromagnetic crystals”. Phys. Rev. 110(4):836 (1958).
  • (64) Z. Y. Fan , R. C. Shen , Y, P. Wang, J. Li, J. Q. You, “Optical sensing of magnons via the magnetoelastic displacement”. Phys. Rev. A 105 033507 (2022).
  • (65) X. Zhang, C. L. Zou, L. Jiang, H. X. Tang., “Strongly coupled magnons and cavity microwave photons”. Phys. Rev. Lett. 113(15), 156401 (2014).
  • (66) C. Gonzalez-Ballestero, D. Hümmer, J. Gieseler, Romero-Isart O. “Theory of quantum acoustomagnonics and acoustomechanics with a micromagnet”. Phys. Rev. B 101 (12):125404 (2020).
  • (67) C. Genesa, A. Marib, D. Vitali, P. Tombesi., “Quantum Effects in Optomechanical Systems”. 57:33-86 (2009).
  • (68) T. Holstein, H. Primakoff “Field dependence of the intrinsic domain magnetization of a ferromagnet”. Phys. Rev. 58:1098 (1940).
  • (69) D. F. Walls, G. J. Milburn, Quantum Optics (Springer), Berlin. Heidelberg (1994).
  • (70) A. Sohail, Y. Zhang, M. Usman, C. S. Yu, “Controllable optomechanically induced transparency in coupled optomechanical systems”. Eur. Phys. J. D 71(4):103 (2017).
  • (71) A. Sohail, R. Ahmed, R. Zainab, C. S. Yu,: “Enhanced entanglement and quantum steering of directly and indirectly coupled modes in a magnomechanical system”. Phys. Scr. 97075102 (2022)
  • (72) A. Sohail, et. al., “Magnon-Phonon-Photon Entanglement via the magnetoelastic Coupling in a Magnomechanical System”. IJTP 61:174 (2022).
  • (73) A. Sohail, A. Hassan, R. Ahmed, C. S. Yu,, “Generation of enhanced entanglement of directly and indirectly coupled modes in a two-cavity magnomechanical system”. Quantum Inf. Process. 21 207 (2022)
  • (74) K, Qu, G. S. Agarwal., “Fano resonances and their control in optomechanics”. Phys. Rev. A 87(6) 063813 (2013).
  • (75) R. Becker and W. Döring, “Ferromagnetismus” Springer, Berlin (1939).
  • (76) C. Kittel, “Physical Theory of Ferromagnetic Domains”. Rev. Mod. Phys. 21 541 (1949).
  • (77) J. W. Rao, et al, “Analogue of dynamic Hall effect in cavity magnon polariton system and coherently controlled logic device, Nat. Commun. 10 2934 (2019).
  • (78) G. Q. Luo , Z. F. Hu, Y. Liang, L. Y. Yu and L. L. Sun , “Development of low profile cavity backed crossed slot antennas for planar integration” IEEE Trans. Antennas Propag. 57 2972 (2009).