Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Magnomechanically induced transparency and tunable slow-fast light via a levitated micromagnet

Open Access Open Access

Abstract

In this paper, we theoretically investigate the magnomechanically induced transparency (MIT) phenomenon and slow-fast light propagation in a microwave cavity-magnomechanical system which includes a levitated ferromagnetic sphere. Magnetic dipole interaction determines the interaction between the photon, magnon, and center of mass motion of the cavity-magnomechanical system. As a result, we find that apart from coupling strength, which has an important role in MIT, the levitated ferromagnetic sphere’s position provides us a parameter to manipulate the width of the transparency window. In addition, the control field’s frequency has crucial influences on the MIT. Also this hybrid magnonic system allows us to demonstrate MIT in both the strong coupling and intermediate coupling regimes. More interestingly, we demonstrate tunable slow and fast light in this hybrid magnonic system. In other words, we show that the group delay can be adjusted by varying the control field’s frequency, the sphere position, and the magnon-photon coupling strength. These parameters have an influence on the transformation from slow to fast light propagation and vice versa. Based on the recent experimental advancements, our results provide the possibility to engineer hybrid magnonic systems with levitated particles for the light propagation, and the quantum measurements and sensing of physical quantities.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Coherent interaction between laser radiation and multilevel atoms exhibits fascinating phenomena such as electromagnetically induced transparency (EIT) [16]. EIT is an approach for turning an opaque medium to a transparent one around resonance. EIT was first observed in cold atomic ensembles [1] and has attracted a great deal of interest due to its distinctive application in nonlinear optics and information processing. A remarkable characteristic of EIT is considerable reduction in the group velocity of light passing through the material, accomplished inside a practically lossless transparency window [2,3]. A noteworthy example of light beam control in optomechanical systems is optomechanically induced transparency (OMIT), which was initially theoretically predicted in April 2010 [6] and experimentally observed a few months later [4]. In analogy to the EIT and OMIT, magnomechanically induced transparency (MIT) is one of the promising phenomena in a Hybrid magnonic system [79], which allows us to manipulate light and explore the slow light in cavity magnomechanical systems by a nonlinear phonon-magnon interaction.

Recently, magnetic materials have offered a promising alternative to achieve strong light-matter interaction. It means a hybrid magnonic system is a novel class of hybrid quantum system that is based on magnons (collective spin excitations) in a YIG–Y$_3$ Fe$_5$ O$_{12}$ – crystal that has a low damping rate and high spin density. It should be noted that YIG is ferrimagnetic. However, YIG has a high Curie temperature as $559 K$, so it is ferromagnetic below this temperature [10]. Magnons have been extensively investigated because of their long coherence times [1114] and their potential for coupling with other quantum excitations, such as other magnons [15], deformation vibration phonons through magnetostrictive force [16,17], spin qubits [18,19], optical [20,21] and microwave photons [12,13,19,22] which may open up a new way to thoroughly study quantum phenomena and promising applications in quantum communication, computing, and information. In other words, various magnonic-based hybrid quantum systems have been led to some intriguing effects, e.g. spin current control [23,24], magnon-phonon entanglement, and magnon-squeezed states [2529], coherent optical-to-microwave conversion [19,30], ground-state cooling of the mechanical vibration mode [31,32], quantum state storage and retrieval [33], and magnon-induced nonreciprocity [34,35]. Additionally, interesting applications have been investigated in such hybrid systems, such as precision measurements [3638], long-time memory [39,40], ultra slow light engineering [8,41,42], magnon laser [43,44], quantum thermometry [45], and magnon and photon manipulating using exceptional points [4649].

Levitated objects provide an excellent platform for discovering the classical-quantum boundary with massive objects [50,51] and sensing weak forces [52,53], due to their weak coupling to the environment and the ability to cool their center of mass motion. Intriguing possibilities have been proposed for interacting systems comprising levitated particles, including quantum friction measurement [54], dark matter detection [55], and measurement of quantum gravity [51]. Additionally, levitated systems are a great platform for investigating non-equilibrium physics [56], as well as, employing these systems for ultrasensitive force detection opens up new possibilities for commercial sensing applications [57].

Motivated by the above-mentioned features of magnonic and levitated systems, we are interested in achieving controllable MIT and changing dispersive properties for manipulating light propagation in a hybrid magnonic system with a levitated particle. It should be noted that a magnet particle can be trapped in a number of ways, for example with an ion trap [58], magnetic trap [59,60], optical trap [61], Paul trap [62], or by being clamped to an ultrahigh-Q mechanical resonator [63,64]. Moreover, other ways of levitation such as magnetic levitation in a microwave cavity [65] or floating above a superconductor in free space [66], have the potential to levitate and cool millimeter-sized spherical magnets [50,67,68]. However, cooling and trapping a large particle, is still extremely challenging for the levitation of large objects [69,70]. Another advantage of a hybrid magnonic system is that the magnon excitations could couple to the center of mass motion (CM) [16,71]. Notably, a microwave cavity-magnomechanical system has been proposed [71], in which by effectively cooling the center of mass motion of a macroscopic YIG sphere, the phonon occupation 6 orders of magnitude smaller than standard cooling is obtained for large particles. Therefore, in the present contribution, we consider a microwave cavity magnomechanical system, composed of a levitated YIG sphere in which, the magnonic excitation in microwave regimes is coupled to the external CM quantum excitations of levitated particle. We illustrate how the coupling strength, the particle position, and the frequency of the control laser lead to the enhancement of the transparency window’s width. Besides, we find that MIT occurs for different coupling strength regimes i.e., the strong coupling and the intermediate coupling regimes. In addition, we demonstrate tunable and controllable slow and fast light emerges as a natural consequence of tunable MIT window. These properties demonstrate the capability of the cavity-magnomechanical systems with levitated ferromagnets as the manipulating and engineering tools for MIT.

The paper is organized as follows. In Sec. 2 the physical system is introduced, then in Sec. 3 the dynamics of the system is investigated through the quantum Langevin-Heisenberg equations of motion and the basic theory of MIT is presented. The results and parameters that play an important role in MIT are provided in Sec. 4. In Sec. 5 we discuss the group delay for slow and fast light propagation. Finally, the conclusion and outlook are mentioned in Sec. 6.

2. System and Hamiltonian

In this section, we consider a cavity-magnomechanical system with a levitated YIG sphere whose center of mass (CM) is trapped in a harmonic oscillator potential. In addition, a homogeneous magnetic field is applied on the sphere to excite the magnon modes [7274]. Hamiltonian of the studied hybrid system is given by

$$\begin{aligned} {\hat{H}}/{\hbar}=& \omega_a \hat{a}^{\dagger}\hat{a}+ \omega_b \hat{b}^{\dagger}\hat{b}+\frac{\omega_c}{2} \left(x^2+p_x^2 \right)\\ &+g_{ab} \left(\hat{a}^{\dagger}\hat{b}+\hat{a}\hat{b}^{\dagger}\right)\cos(k\, x)+i \left(s_{in}(t) \hat{a}^{\dagger} -s_{in}^*(t) \hat{a} \right). \end{aligned}$$

The free Hamiltonian of the cavity field and the magnonic modes are described by the first two terms, with the bosonic annihilation operator $\hat {a}$ ($\hat {b}$) corresponds to the cavity (magnon) modes with frequency $\omega _a$ ($\omega _b$). In third term, $x$ and $p_x$ are the position and momentum of the magnetic particle CM that oscillates at $\omega _c$, the frequency of the trap. The fourth term describes the interaction between the magnon and photon modes. This coupling is characterized by the magnetic dipole interaction [911,75] where $g_{ab}=\frac {\gamma }{2}\sqrt {\frac {\hbar \omega _a \mu _0}{ V_a}}\sqrt {2\rho _s V \, s }$ is the coupling strength in which $\gamma$ is the gyromagnetic ratio, $V_a$ and $V$ are the mode volume of the cavity mode and volume of the YIG sphere, $\mu _0$ is the vacuum permeability, $\rho _s=4.22\times 10^{27} m^{(-3)}$ is the spin density of the YIG sphere, and $s=\frac {5}{2}$ is the spin number of the YIG’s ground-state. As is known, the light-matter interaction is realized in different regimes depending on the comparison between coupling constant and damping rates. In this study, we demonstrate induced transparency in the strong coupling (SC) as well as intermediate coupling (IMC) regimes that the rotating wave approximation (RWA) is valid in both of these regimes. Moreover, we assume that the cavity mode along $\hat {y}$ direction has the form as $\vec {B}_{cav}=\hat {y} B \, \cos (kx)$ where $k$ is the microwave field wave number [71]. The magnetic dipole interaction $H_{int}=-\gamma \vec {S}\,.\,\vec {B}_{cav}$, in which $\vec {S}$ being the collective spin angular momentum, can be used to describe the coupling between the microwave cavity $\vec {B}_{cav}$ and Kittel magnon. It is worth mentioning that the magnon frequency is determined by the external magnetic field $B_0$ and the gyromagnetic ratio $\gamma /2\pi =28\, \rm {GHz}/{\rm T}$, that is, $\omega _m=\gamma B_0$. Here, we focus on the Kittle magnon mode which is the homogeneous ground state of the magnonic mode and can be simply tuned by $B_0$. The last term in Hamiltonian represents the driving of the cavity mode with a time-dependent classical field, $s_{in}(t)$. In the following, we shall investigate the magnomechaniclly induced transparency (MIT) in our hybrid system. Two situations are considered: first, as a general form, we assume that the cavity is driven by a control laser, and also, by a weak probe laser, as illustrated in Fig. 1. Second, as a limiting case, we suppose that the YIG sphere is fixed and the cavity mode is only coherently driven by a weak probe field.

 figure: Fig. 1.

Fig. 1. Schematic of the cavity which consists of the levitated YIG sphere and is driven by a control field and a weak input probe field.

Download Full Size | PDF

3. Dynamics of the system

3.1 MIT via a levitated particle under the control laser

In this section, we shall proceed to investigate the MIT in our hybrid system. First we note that for a YIG sphere that is trapped with low-frequency trap $(\sim 50 kHz)$ and with a YIG sphere with the size of $R>1 nm$, the thermal variance in it’s CM position $\Delta x=\sqrt {(k_b T/(m \omega _c^2 )}$ is much smaller than the microwave wavelength [71]. Therefore, by expanding $\cos (kx)$ up to first order in $x$ around the minimum position of the trap, $x_0$, the Hamiltonian (1) can be obtained as follows:

$$\begin{aligned} { \hat{H}}/{\hbar}=& \omega_a \hat{a}^\dagger \hat{a}+\omega_b \hat{b}^\dagger \hat{b} +\omega_c \hat{c}^\dagger \hat{c}+g_{ab} \cos (kx_0) (\hat{a} \hat{b}^\dagger{+}\hat{a}^\dagger \hat{b})\\ &-g_{abc} \sin(k x_0)(\hat{a} \hat{b}^\dagger{+}\hat{a}^\dagger \hat{b})(\hat{c}^\dagger{+}\hat{c})+i \left(s_{in}(t) \hat{a}^{\dagger} -s_{in}^*(t) \hat{a} \right), \end{aligned}$$
where, $\hat {c}$ is the annihilation operator of the CM oscillation, the coupling strength is determined by $g_{abc}=g_{ab}k \sqrt {\frac {\hbar }{2 \rho _mV\omega _c}}$ with the YIG sphere’s mass density $\rho _m=5170 kgm^{-3}$, and $g_{abc}$ is independent of the sphere’s size. In addition, we assume the cavity is driven by an external electromagnetic field and an input weak probe field as follows:
$$s_{in}(t)=\left(\varepsilon_d+\varepsilon_p e^{{-}i(\omega_p-\omega_d) t} \right) e^{{-}i\omega_d t}=s_{in}^\prime e^{{-}i\omega_d t},$$
where $\omega _{d (p)}$ is the input drive (probe) field frequency and $\omega _p$ is close to the cavity frequency of the magnomechanical system with amplitude $\varepsilon _{d (p)}=\sqrt {\frac {2 \gamma _a P_{d (p)}}{\hbar \omega _{d(p)}}}$, $\gamma _a$ and $P_{d (p)}$ are the cavity decay rate and power of input drive (probe) field, respectively. Thus, the Hamiltonian in the frame rotating at the frequency $\omega _d$, is given by:
$$\begin{aligned} \hat{H}/{\hbar}=&\Delta_a \hat{a}^\dagger \hat{a}+\Delta_b \hat{b}^\dagger \hat{b}+\omega_c \hat{c}^\dagger \hat{c}+g_{ab} \cos (kx_0) \left(\hat{a} \hat{b}^\dagger{+}\hat{a}^\dagger \hat{b} \right)\\ &-g_{abc} \sin(k x_0) \left(\hat{a} \hat{b}^\dagger{+}\hat{a}^\dagger \hat{b} \right) \left(\hat{c}^\dagger{+}\hat{c} \right)+i \left(s_{in}^\prime \hat{a}^\dagger{-}s_{in}^{\prime *} \hat{a} \right), \end{aligned}$$
where $\Delta _j=\omega _j-\omega _d$ $(j=a, b)$ is the detuning between the photon and magnon excitations with respect to the driving field. It should be noted that the Kittel magnon frequency can be adjusted to rather close to on-resonance with the cavity modes. In the following, we assume the same frequency detunings $\Delta _a=\Delta _b=\Delta$ in the case of the magnon-photon on-resonance: $\omega _a=\omega _b$.

3.1.1 Dynamics of the system

The dynamics of cavity, magnonic and levitated particle CM subsystems is described by quantum Langevin equations through adding damping and noise terms to the Heisenberg equations as follows:

$$\begin{aligned} \hat{\dot{a}}=& -(\gamma_a +i \Delta_a) \hat{a}+i\,g_{abc}\, \sin(kx_0)\, \hat{b}\,(\hat{c}+\hat{c}^\dagger)-i\, g_{ab} \cos (kx_0) \,\hat{b}+s_{in}^\prime+ \sqrt{2\gamma_a}\hat{a}^{in},\\ \hat{\dot{b}}=& -(\gamma_b +i \Delta_b) \hat{b}+i\,g_{abc}\, \sin(kx_0) \,\hat{a}\,(\hat{c}+\hat{c}^\dagger)-i\,g_{ab} \cos (kx_0) \,\hat{a} + \sqrt{2\gamma_b}\hat{b}^{in},\\ \hat{\dot{c}}=& -(\gamma_c +i \omega_c) \hat{c}+i\,g_{abc}\, \sin(kx_0)\, (\hat{a}\hat{b}^\dagger{+}\hat{a}^\dagger\hat{b}) + \sqrt{2\gamma_c}\hat{c}^{in}. \end{aligned}$$
where $\gamma _o (o=a,b,c)$ is the related damping rate (we suppose the external damping rate and ignore intrinsic dissipation) and $o^{in}$ represents the noise operator which is assumed to satisfy the following correlation functions
$$\begin{aligned} <o^{in}(t) o^{in\dagger}(t^\prime)>{=}&(\bar{n}_o+1)\delta(t-t^\prime),\\ <o^{in\dagger}(t) o^{in}(t^\prime)>{=}&\bar{n}_o \delta(t-t^\prime), \end{aligned}$$
where $\bar {n}_o=(e^{(\frac {\hbar \omega _o}{k_B T})}-1)^{-1}$ is the bosonic mode's mean thermal occupation, $T$ is the temperature of the thermal environment and $k_B$ is the Boltzmann constant. Equations (5) are the non-linear quantum Langevin equations (QLEs), so, if we assume a weak nonlinearity in the system, we can decompose each operator as the sum of its steady-state value and a small fluctuation as: $\hat {o}=o_0+\delta \hat {o}$ and ignore small second-order fluctuation terms. The equations of the steady state mean values of the system are obtained as:
$$\begin{aligned} \dot{a}_0=&-({\gamma_a}+i \Delta_a) a_0-i(G_{ab} +2G_{abc}\,Re[c_0])b_0+\varepsilon_d=0,\\ \dot{b}_0=&-({\gamma_b}+i \Delta_b) b_0-i(G_{ab}+2G_{abc}\,Re[c_0])a_0=0,\\ \dot{c}_0=&-({\gamma_c}+i \omega_c) c_0-iG_{abc}(a_0 b^*+a_0^*b_0)=0, \end{aligned}$$
where $G_{ab}=g_{ab}\cos (kx_0)$ and $G_{abc}=-g_{abc}\sin (kx_0)$. So the steady-state mean values of $\hat {a}$, $\hat {b}$ and $\hat {c}$ are given by:
$$\begin{aligned} a_0=&\frac{\varepsilon_d (\gamma_b+i\Delta_b)}{(\gamma_a+i\Delta_a)(\gamma_b+i\Delta_b)+(G_{ab}+2G_{abc}Re[c_0])^2},\\ b_0=&\frac{-i\, \varepsilon_d (G_{ab}+2G_{abc}Re[c_0])}{(\gamma_a+i\Delta_a)(\gamma_b+i\Delta_b)+(G_{ab}+2G_{abc}Re[c_0])^2},\\ c_0=&\frac{2i G_{abc}\Delta_b(G_{ab}+2G_{abc}Re[c_0])}{(\gamma_c+i\omega_c)(\gamma_b^2+\Delta_b^2)}\vert a_0\vert^2. \end{aligned}$$

Clearly, the fifth-order equation of the $I_a=\vert a_0\vert ^2$ is satisfied by the coupled equations stated above, and the magnonic system may exhibit multistability for some of the parameters. In this paper, we have chosen the parameters such that a stable solution exists. To achieve the results, we have used the following steady-state values (which are valid because of the small value of $G_{abc}$):

$$\begin{aligned} a_0 \simeq & \frac {\varepsilon_d (\gamma_b + i\Delta_b)}{(\gamma_a + i \Delta_a)(\gamma_b + i\Delta_b)+G_{ab} ^2},\\ b_0 \simeq & \frac {-i \varepsilon_d G_{ab}}{(\gamma_a + i \Delta_a)(\gamma_b + i \Delta_b)+G_{ab} ^2},\\ c_0 \simeq & \frac {2iG_{ab}G_{abc}\Delta_b}{(\gamma_c + i\omega_c)(\gamma_b^2 + \Delta_b^2)}\vert a_0\vert^2. \end{aligned}$$

Moreover, the time evolution of the quantum fluctuations are obtained as:

$$\begin{aligned} \delta\dot{\hat{a}}(t)=&-(\gamma_a+i\Delta_a)\delta\hat{a}-i\, \Gamma \delta \hat{b}-iG_{abc}b_0(\delta\hat{c}+\delta\hat{c}^{\dagger})+\varepsilon_p e^{{-}i\delta t}+\sqrt{2\gamma_a}\hat{a}^{in}\\ \delta \dot{\hat{b}}(t)=&-(\gamma_b+i\Delta_b)\delta\hat{b}-i\, \Gamma \delta \hat{a}-i G_{abc}a_0(\delta\hat{c}+\delta\hat{c}^{\dagger})+\sqrt{2\gamma_b}\hat{b}^{in},\\ \delta\dot{\hat{c}}(t)=&-(\gamma_c+i\omega_c)\delta\hat{c}-i\,G_{abc}(a_0\delta\hat{b}^{\dagger}+a_0^*\delta\hat{b}+b_0\delta\hat{a}^{\dagger}+b_0^*\delta\hat{a})+\sqrt{2\gamma_c}\hat{c}^{in}. \end{aligned}$$

In the above set of equations we have defined: $\Gamma =G_{ab}+2G_{abc}Re[c_0]$ and $\delta =\omega _p-\omega _d$. In the following, we are interested in the mean response of the hybrid system, so the mean values of noise operators are zero. In order to solve this set of coupled equations, we use the following ansatz for the solutions [5,76]:

$$\begin{aligned} \delta \hat{a}(t)=&A^- e^{{-}i\delta t}+A^+ e^{i\delta t},\\ \delta \hat{b}(t)=&B^- e^{{-}i\delta t}+B^+ e^{i\delta t},\\ \delta \hat{c}(t)=&C^- e^{{-}i\delta t}+C^+ e^{i\delta t}. \end{aligned}$$

By inserting these relations into Eqs. (10) and equating the coefficients with the same frequency, we arrive at the following relations:

$$\begin{aligned} [{\gamma}_{a}+i(\Delta_{a}-{\delta})]A^-{=}&-i\,\Gamma B^-{-}i\,G_{abc}b_0\,Q^-{+}\varepsilon_p,\\ [\gamma_a-i(\Delta_a+\delta)]A^{-,*}=&i\,\Gamma B^{-,*}+i\,G_{abc}b_0^*\,Q^- ,\\ [\gamma_b+i(\Delta_b-\delta)]B^{-}=&-i\,\Gamma A^{-}-i\,G_{abc}a_0\,Q^-,\\ [\gamma_b-i(\Delta_b+\delta)]B^{-,*}=&i\,\Gamma A^{-,*}+i\,G_{abc}a_0^*\,Q^-, \end{aligned}$$
and
$$Q^-{=}-2 G_{abc}\omega_c\frac{ a_0^* B^{-}+a_0 B^{-,*}+b_0^* A^-{+}b_0 A^{-,*} }{(\gamma_c-i\delta)^2+\omega_c^2}.$$

Here, $(C^-+C^{-,*})=Q^-$. Since we are going to calculate the amplitude of the cavity output field, by using the input-output relation $\hat {a}^{out}+\hat {a}^{in}=\sqrt {2 \gamma _a} \hat {a}$ [77], we have

$$\varepsilon_{out}+\varepsilon_d e^{{-}i\omega_d t}+\varepsilon_p e^{{-}i\omega_p t}=2 \gamma_a (a_0+\delta \hat{a}) e^{{-}i\omega_d t}.$$

By substituting Eq. (11) in Eq. (14), it is clear that $A^-$ and $A^+$ oscillate at the frequencies $\omega _p$ and $2\omega _d-\omega _p$, respectively. The total output field $\varepsilon _t$, at the probe frequency, is given by

$$\varepsilon_t=\frac{2 \gamma_a \vert A^-\vert }{\varepsilon_p}=\frac{2 \gamma_a \left[\gamma_b+i(\Delta_b-\delta)\right]}{f(\delta)+\Gamma^2},$$
where we have
$$f(\delta)=\left[\gamma_a+i(\Delta_a-\delta)\right] \left[\gamma_b+i(\Delta_b-\delta)\right]+ \frac{\zeta(\delta)}{\xi(\delta)} \left[\Gamma a_0+ib_0(\gamma_b+i(\Delta_b-\delta)) \right],$$
with
$$\begin{aligned} \zeta(\delta) =&-2 G_{abc}^2\omega_c \left[b_0^*-\frac{i\Gamma a_0^*}{\gamma_b+i(\Delta_b-\delta)} \right],\\ \xi(\delta) =& 2 G_{abc}^2 \omega_c \left[a_0\beta(\delta)+b_0\alpha(\delta) -\frac{i \vert a_0\vert^2}{\gamma_b+i(\Delta_b-\delta)} \right]+(\gamma_c-i\delta)^2+\omega_c^2, \end{aligned}$$
and
$$\alpha(\delta) =\frac{-\Gamma a_0^*+i\,b_0^*(\gamma_b-i(\Delta_b+\delta))}{(\gamma_a-i(\Delta_a+\delta)) (\gamma_b-i(\Delta_b+\delta))+\Gamma^2}, \qquad \beta(\delta) =\frac{i\, (\Gamma\alpha(\delta)+a_0^*)}{\gamma_b - i(\Delta_b+\delta)}.$$

Now, as a limiting case, if we suppose our system is driven only with an input probe field and assume that we are dealing with a mounted YIG sphere (rather than a levitated sphere, i.e. $g_{abc}\ll g_{ab}$), the mentioned hybrid system reduces to an electromagnonic system [12,21]. As it is clear, in this simple case, the output entire fields $\varepsilon _t$, at the probe frequency is simplified to the following form as

$$\varepsilon_t = \frac{2 \gamma_a \vert A^-\vert }{\varepsilon_p}=\frac{2 \gamma_a }{\gamma_a+i(\omega_a-\omega_p)+ \frac{g_{ab}^2 \cos ^2 (kx_0)}{\gamma_b+i(\omega_b-\omega_p)}}.$$

The output field’s absorptive and dispersive behaviors at the probe frequency are given by the real part $\varepsilon _r$ and imaginary part $\varepsilon _i$ of the field amplitude $\varepsilon _t$, respectively. In the following, we investigate the MIT in both cases.

4. Magnomechanically induced transparency

In this section, we investigate the induced transparency in the hybrid magnomechanical system consisting of a levitated YIG sphere in the microwave cavity in interaction with magnon, photon and CM. It should be noted that the hybrid magnomechanical system stability is controlled by the Routh-Hurwitz criteria [78], i.e. real part of the eigenvalues of the drift matrix should be negative. According to this criteria, we have selected the experimentally realizable parameters so that the cavity-magnomechanical system with a levitated YIG sphere is stable in our studies.

4.1 MIT via a levitated particle

In this section, we explore MIT phenomena in the presence of a control field and demonstrate how to enhance the controllability of the system by choosing appropriate values for the magnomechanical system parameters. Similar to the optomechanical systems [5], we assume that the cavity field is driven by control and probe fields with different frequencies and amplitudes. Based on the input-output relation Eq. (14) and total output field Eq. (15), the real and imaginary parts of the field’s amplitude $\varepsilon _t$ versus the normalized frequency $\delta /\omega _c$ have been displayed in Fig. 2 for $\omega _a/2\pi =\omega _b/2\pi =30 \, GHz$, $\omega _c/2\pi =50\,kHz$, $\gamma _a/2\pi =1\, MHz$, $\gamma _b/2\pi =1\, MHz$, $\gamma _c/2\pi =10^{-6}\,Hz$, $kx_0=2\pi /3$, and $P=1\, \mu W$ [71]. In addition, for a bias field $B_0\sim 1 T$, the Kittel magnon can resonantly interact with microwave magnetic field of wavelength $\sim 1 cm$. Figure 2(a), corresponds to the absence of the magnon-photon coupling, i.e., $g_{ab}=0$, where the hybrid magnomechanical system releases to a standard bare cavity system. In Fig. 2(b), we see that the two symmetric Lorentzian peaks are also appeared. In other words, by the magnon-photon-CM interaction, a transparent window (i.e., $\varepsilon _r\simeq 0$) is appeared.

 figure: Fig. 2.

Fig. 2. The real (red line), $\varepsilon _r$, and the imaginary (blue line), $\varepsilon _i$, parts of the field amplitude, $\varepsilon _t$, versus the normalized frequency $\delta /\omega _c$ (a) without magnon-photon coupling $g_{ab}=0$ (b) with the magnon-photon coupling $g_{ab}/2\pi = 15.6\,MHz$. The parameters are selected as follows: $\omega _a/2\pi =\omega _b/2\pi =30 \, GHz$, $\omega _c/2\pi =50\,kHz$, $\gamma _a/2\pi =1\, MHz$, $\gamma _c/2\pi =10^{-6}\,Hz$, $k x_{0}=2\pi /3$, and $P=1\, \mu W$. To obtain this result, we have selected $\Delta _a=\Delta _b\equiv \Delta =\omega _c$ [71].

Download Full Size | PDF

Here, to acquire further insight, we demonstrate the role of magnon-photon-CM interaction in occurrence of the MIT in the introduced system. For this purpose, we assume that magnon-photon coupling $g_{ab}$ is small ($g_{ab}/2\pi \sim 1.5 MHz$) in such a manner that is not strong enough to lead to a perfect transparency window. In Fig. 3 the absorption profile, $\varepsilon _r$, has been plotted versus the normalized frequency $\delta /\gamma _b$ for different values of magnon-photon-CM coupling $g_{abc}/2\pi =10^{-4}Hz$ (green line), $12\times 10^{-2}Hz$ (blue line), $13\times 10^{-2}Hz$ (red line), $14\times 10^{-2}Hz$ (purple line). It should be noted that we assumed the experimentally realizable parameters in such a way that keep the system stable. As is shown in Fig. 3 , increasing the magnon-photon-CM coupling $g_{abc}$ enables us to achieve a wider transparency window. Thus, the additional coupling to the CM can play an important role to enhance the width of the transparency window. In order to obtain more insight, in the following we study three parameres that have important effects on the transparency window’s plots:

  • i) Coupling strength: In Fig. 4(a) the absorption profile has been plotted versus the normalized frequency $\delta /\omega _c$ and the magnon-photon coupling strength $g_{ab}$. we can see a Lorentzian peak at $\delta /\omega _c=0$ splits into two symmetric Lorentzian peaks by modifying the coupling strength due to the interaction in the hybrid magnon-photon-CM motion. As is evident, increasing coupling strength causes a wider transparency window. Besides, we can increase the coupling strength by enlarging the size of the YIG sphere or by employing other methods like modulation of coupling strength [26,37]. Physically, when we consider the photon-magnon interaction of the cavity-magnomechanical system, in analogy with the photon-phonon interaction in the linearized optomechanical system, we find out that the magnon modes can play the same role as the mechanical modes.
  • ii) Particle position: Since the YIG sphere is levitated, let us consider and investigate the impact of particle position on the induced transparency. In Fig. 4(b), we have plotted $\varepsilon _r$ versus the normalized frequency $\delta /\omega _c$ and particle position. As it is seen, a peak appears in the node position of the cavity magnetic field (CMF), that is: $k\, x_0=(2n+1)\pi /2$, and by moving along the antinode of the cavity magnetic field this peak splits into two peaks, and a transparency window appears. It is worth noting that the maximum transparency exactly happens in antinode ($k\, x_0=n\pi$) position. In order to gain more physical insight, we consider two special position for the YIG sphere: the node and antinode of the CMF in Hamiltonian (4). It is obvious that the fourth term in this Hamiltonian becomes zero in the node of the CMF, since $G_{ab}=0$ and $G_{abc}$ is very small (which has small experimental value), so that the transparency window disappears. Also, the MIT is maximum in the antinode of the CMF. In other words, by changing the particle position, it is possible to control the width of the transparency window.

    So far, we have looked into two significant parameters that contribute to MIT. It should be noted that in the MIT condition, the magnomechanical coupling $G_{ab}=g_{ab}\cos (kx_0)$ has an equivalent role to the Rabi frequency in the atomic EIT [4]. As a comparison, in Fig. 5, $\varepsilon _r$ has been plotted versus the normalized frequency $\delta /\omega _c$ and particle position ${kx_0}/{\pi }$ for $R=100\, \mu m$ and $R=50\, \mu m$. Although in both cases, we encounter with a transparency window, Figs. 5(a) and (b) clearly show that the width of the transparency window decreases by decreasing $R$, which means the decreasing of particle radius is accompanied by coupling strength decrement. Therefore, Fig. 5 illustrates greater coupling strength and antinodes allow us to achieve the maximum transparency window. Also, we can see considerable MIT for the small size of the YIG sphere, in the stability situation.

  • iii) Detuning frequency: It is worth emphasizing that the driving field provides us with more controllable parameters. In Fig. 6 the absorption profile is plotted versus the normalized frequency $\delta /\omega _c$ and normalized detuning frequency $\Delta /\omega _c$. By adjusting detuning frequency, an induced transparency window appears in various values of $\Delta _a=\Delta _b$. It is possible to shift the two Lorentzian peaks Fig. 6 around $\delta /\omega _c$ to the left or right, by selecting on negative or positive values for detuning frequency. In the other word, the driving field causes to exist MIT in the wide range of values as $\Delta _a=\Delta _b=\Delta$; e.g., $-100 \omega _c\,<\,\Delta <\,100 \omega _c$ and $\omega _a=\omega _b$.

    Until now we have studied MIT in the SC regime, i.e., $\gamma _{a(b)}<g_{ab}<\omega _{a(b)}$. Now we consider the IMC regime by selecting $\gamma _a/2\pi =14 MHz$, which means $\gamma _b<g_{ab}<\gamma _a$ and $g_{ab}<\omega _{a(b)}$. In Fig. 7(a) the absorption profile has been plotted versus the normalized frequency $\delta /\omega _c$ and particle position ${kx_0}/{\pi }$. As it is seen, although we have MIT in the antinode of the CMF, the transparency window decreases by increasing the photon and the magnon decay rates. In addition, in Fig. 7(b), the absorption profile has been plotted versus the normalized frequency $\delta /\omega _c$ and magnon-photon coupling $g_{ab}$. This figure clearly shows that a transparency window is appeared and enhanced by increasing the coupling strength. Therefore, MIT can be appear in our magnomechanical system, in both IMC and SC regimes.

 figure: Fig. 3.

Fig. 3. The real part of the field amplitude, $\varepsilon _r$, versus the normalized frequency $\delta /\gamma _b$ with $\gamma _c/2\pi =10 Hz$, and $P=100\, mW$, and the magnon-photon coupling $g_{ab}/2\pi = 1.5\,MHz$ for different magnon-photon-CM couplings: $g_{abc}/2\pi =10^{-4}Hz$ (green line), $12\times 10^{-2}Hz$ (blue line), $13\times 10^{-2}Hz$ (red line), $14\times 10^{-2}Hz$ (purple line). Other parameters are selected the same as Fig. (2).

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. The real $\varepsilon _r$ part of the output field versus the normalized frequency $\delta /\omega _c$ and (a) magnon-photon coupling $g_{ab}$ and (b) particle position ${kx_0}/{\pi }$. The parameters are selected the same as Fig. (2)

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. (a), (b) The real part, $\varepsilon _r$, of the output field versus the normalized frequency $\delta /\omega _c$ and particle position ${kx_0}/\pi$, (a) $R=100 \, \mu m$ and (b) $R=50 \, \mu m$. Other parameters are selected the same as Fig. (2)

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. The real $\varepsilon _r$ part of the output field versus the normalized frequency $\delta /\omega _c$ and the normalized detuning $\Delta /\omega _c$. Other parameters are selected the same as Fig. (2)

Download Full Size | PDF

 figure: Fig. 7.

Fig. 7. The real part, $\varepsilon _r$, of the output field versus the normalized frequency $\delta /\omega _c$ and (a) particle position ${kx_0}/{\pi }$ and (b) magnon-photon coupling $g_{ab}$ for the intermediate coupling (IMC) regime $\gamma _b<g_{ab}<\gamma _a$ and $\gamma _a/2\pi =14\,MHz$. Other parameters are selected the same as Fig. (2)

Download Full Size | PDF

4.1.1 Limiting case: MIT without the control laser

Now, supposing a mounted YIG sphere in the absence of the control field, the output field $\varepsilon _t$ is given by Eq. (17). In Fig. 8, by considering the effective magnon-photon coupling, the absorption profile of the output field is plotted versus the normalized frequency $\Omega /\omega _b$ ($\Omega =\omega _a-\omega _p$) for $g_{ab}/2\pi =12.6,14.6,16.6,18.6 \,MHz$. Consequently, as a result of the effective magnon-photon interaction, a transparent window (i.e., $\varepsilon _r\simeq 0$) in the on-resonance condition $\omega _a=\omega _b$ is appeared. As it is seen, the transparency window for $g_{ab}/2\pi =18.6 \,MHz$ (green line) is larger than the transparency window for $g_{ab}/2\pi =12.6$ (pink line), which means that by increasing the magnon-photon coupling strength, the width of the transparency window is increased. In the following section, we present the numerical result of slow and fast light propagation based on the obtained MIT results.

 figure: Fig. 8.

Fig. 8. The real part, $\varepsilon _r$, of the output field versus the normalized frequency $\Omega /\omega _b$ and different magnon-photon couplings $g_{ab}/2\pi =12.6 MHz$ (pink line), $14.6 MHz$ (blue line), $16.6 MHz$ (red line), $18.6 MHz$ (green line). The parameters are selected the same as Fig. (2).

Download Full Size | PDF

5. Slow and fast light

The previous sections described MIT in various regimes (SC & IMC) as well as adjustable parameters that impact on changing the width of MIT window in magnomechanical system. As we know, MIT allows us to control and manipulate light propagation. Therefore, in this section we investigate the group delay of the total output field, $\varepsilon _t$, at the probe frequency. For this purpose, we use rapid phase dispersion $\Phi _t(\omega _p)=arg\left [\varepsilon _t(\omega _p)\right ]$, leading to define a group delay as [8]:

$$\tau_g=\frac{\partial \Phi _t(\omega_p)}{\partial \omega_p}=\text{Im} \left[\frac{1}{\varepsilon_t} \frac{\partial \varepsilon_t}{\partial \omega_p} \right].$$

The slow light is indicated by a positive group delay, $\tau _g>0$, and the fast light is implied by a negative group delay, $\tau _g<0$. Before investigating the group delay, in Fig. 9, the probe field phase $\Phi _t(\omega _p)$ has been plotted versus the normalized frequency $\delta /\omega _c$ and magnon-photon coupling $g_{ab}/\gamma _b$. It is clear, when the coupling strength is small, the rapid normal phase dispersion results in a positive group delay, corresponding to the generation of the slow light. Intriguingly, an anomalous dispersion of the probe field phase appears when we increase the coupling strength, which is corresponding to the fast light. Additionally, a critical point is illustrated in the vicinity of the magnon-photon coupling strength $g_{ab}/\gamma _b\sim 2$. The phase dispersion critical point induced by magnon-photon-CM interaction reminds us the possibility of realizing the transformation between the slow and the fast light in our levitated system.

 figure: Fig. 9.

Fig. 9. The output field phase versus the normalized frequency $\delta /\omega _c$ and magnon-photon coupling $g_{ab}/\gamma _b$. A critical point in the vicinity $g_{ab}/\gamma _b\sim 2$. The parameters are selected the same as Fig. (2).

Download Full Size | PDF

In Figs. 10(a)-(c), respectively, we have plotted the group delay, $\tau _g$, versus the magnon-photon coupling, the particle position and the detuning frequency, under the various values of the damping rate of microwave cavity field $\gamma _a$. Figure 10(a) illustrates that in the regime of $g_{ab}/\gamma _b \in (0,2)$, the group delay is positive which is corresponding to the generation of slow light. Furthermore, it shows that for suitable values of $g_{ab}$, it is possible to switch the group delay from superluminal to subluminal propagation and vice versa. In addition, the maximum value of the group delay $\tau _g$ is inversely proportional to the damping rate of the cavity field $\gamma _a$. It can be seen that from $\gamma _a/2\pi =1.1MHz$ to $0.27 MHz$, the maximum value of the group delay increases from $\tau _g=0.15\, \mu s$ (the blue solid line) to $\tau _g=0.6 \, \mu s$ (the purple solid line). Another important result of our study, as is illustrated in Fig. 10(b), is that the YIG sphere position in two critical points allows us to manipulate light propagation. Also, Fig. 10(c) clearly shows that, we can change the group delay and have slow light, in the wide range of the detuning frequency. According to these results, the magnomechanical system with a levitated particle is a promising candidate for manipulating the propagation of light.

 figure: Fig. 10.

Fig. 10. Group delay $\tau _g$ versus (a) the magnon-photon coupling strength $g_{ab}/\gamma _b$ (b) the particle position $kx_0/\pi$ (c) the detuning frequency $\Delta /\gamma _b$, under the various of the damping rate of microwave cavity field $\gamma _a$, respectively. The parameters are selected the same as Fig. (2).

Download Full Size | PDF

By comparing this result with other studies on slow-fast light in mounted particles [79], we find that apart from magnon-photon coupling and control field, the particle position enables us to manipulate the light propagation. On the other hand, the magnetic and electrostatic levitation development made it possible to control particle position. Hence, the ability to couple the motion of the levitated system to magnon and photon modes provides an opportunity for more control on light propagation, via controlling the particle position.

According to the obtained results, effective manipulation of slow-fast light can be accomplished by varying the control laser’s frequency, YIG sphere position, and coupling strength. As expected, if the damping rate can be further reduced, the maximum value of the group delay will be increased, which reminds us of the possibility of achieving a larger group delay in an ultrahigh-Q cavity photon-magnon coupling system.

6. Conclusion

We have studied theoretically the realization of MIT and light propagation in the cavity magnomechanical system with magnon, photon, and CM nonlinear interaction. It has been shown that both of them extremely depend on the coupling strength. Furthermore, the particle position and the frequency of the control laser provide us controllable parameters for modifying the transparency window. Another advantage of the introduced model is that MIT can occur in both SC and IMC coupling regimes. Interestingly, we obtained the slow and fast light effects and a transformation between subluminal and superluminal propagation of the output probe field. On the other hand, a comparison between the introduced system and an optomechanical system reveals an analogy between magnons and mechanical modes in the optomechanical system, and significantly influence the MIT and light propagation.

The obtained results of the cavity-magnomechanical system, with a levitated YIG sphere, open new perspectives for quantum interaction between the microwave, the magnetic, and the mechanical systems. Applications of these findings include optical data processing and communication. In addition, the introduced system in the present contribution has potential applications such as the sensing of weak forces or magnetic signals by assuming some nonlinearities in future investigations, such as the Kerr effect or magnon squeezing, and achieving more adjustable parameters [28].

Acknowledgments

The authors wish to thank the Office of Graduate Studies and Research Vice President of the University of Isfahan for their support.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. K.-J. Boller, A. Imamoğlu, and S. E. Harris, “Observation of electromagnetically induced transparency,” Phys. Rev. Lett. 66(20), 2593–2596 (1991). [CrossRef]  

2. M. Fleischhauer, A. Imamoglu, and J. P. Marangos, “Electromagnetically induced transparency: Optics in coherent media,” Rev. Mod. Phys. 77(2), 633–673 (2005). [CrossRef]  

3. M. Lukin and A. Imamoğlu, “Controlling photons using electromagnetically induced transparency,” Nature 413(6853), 273–276 (2001). [CrossRef]  

4. G. S. Agarwal and S. Huang, “Electromagnetically induced transparency in mechanical effects of light,” Phys. Rev. A 81(4), 041803 (2010). [CrossRef]  

5. S. Shahidani, M. Naderi, and M. Soltanolkotabi, “Control and manipulation of electromagnetically induced transparency in a nonlinear optomechanical system with two movable mirrors,” Phys. Rev. A 88(5), 053813 (2013). [CrossRef]  

6. S. Weis, R. Rivière, S. Deléglise, et al., “Optomechanically induced transparency,” Science 330(6010), 1520–1523 (2010). [CrossRef]  

7. Y.-M. Huang, Z.-Q. Hua, Y.-Q. Yang, et al., “Magnon-photon induced transparency in a cavity-quantum-electrodynamics system coupling yttrium-iron-garnet sphere,” Results Phys. 15, 102516 (2019). [CrossRef]  

8. C. Kong, B. Wang, and Z.-X. Liu, “Magnetically controllable slow light based on magnetostrictive forces,” Opt. Express 27(4), 5544–5556 (2019). [CrossRef]  

9. T.-X. Lu, X. Xiao, and L.-S. Chen, “Magnon-squeezing-enhanced slow light and second-order sideband in cavity magnomechanics,” Phys. Rev. A 107(6), 063714 (2023). [CrossRef]  

10. D. Zhang, X.-M. Wang, and T.-F. Li, “Cavity quantum electrodynamics with ferromagnetic magnons in a small yttrium-iron-garnet sphere,” npj Quantum Inf. 1(1), 15014 (2015). [CrossRef]  

11. Y. Tabuchi, S. Ishino, and T. Ishikawa, “Hybridizing ferromagnetic magnons and microwave photons in the quantum limit,” Phys. Rev. Lett. 113(8), 083603 (2014). [CrossRef]  

12. H. Huebl, C. W. Zollitsch, and J. Lotze, “High cooperativity in coupled microwave resonator ferrimagnetic insulator hybrids,” Phys. Rev. Lett. 111(12), 127003 (2013). [CrossRef]  

13. X. Zhang, C.-L. Zou, L. Jiang, et al., “Strongly coupled magnons and cavity microwave photons,” Phys. Rev. Lett. 113(15), 156401 (2014). [CrossRef]  

14. L. Bai, M. Harder, and Y. Chen, “Spin pumping in electrodynamically coupled magnon-photon systems,” Phys. Rev. Lett. 114(22), 227201 (2015). [CrossRef]  

15. N. Lambert, J. Haigh, and S. Langenfeld, “Cavity-mediated coherent coupling of magnetic moments,” Phys. Rev. A 93(2), 021803 (2016). [CrossRef]  

16. C. Gonzalez-Ballestero, D. Hümmer, J. Gieseler, et al., “Theory of quantum acoustomagnonics and acoustomechanics with a micromagnet,” Phys. Rev. B 101(12), 125404 (2020). [CrossRef]  

17. C. Gonzalez-Ballestero, J. Gieseler, and O. Romero-Isart, “Quantum acoustomechanics with a micromagnet,” Phys. Rev. Lett. 124(9), 093602 (2020). [CrossRef]  

18. D. Lachance-Quirion, Y. Tabuchi, and S. Ishino, “Resolving quanta of collective spin excitations in a millimeter-sized ferromagnet,” Sci. Adv. 3(7), e1603150 (2017). [CrossRef]  

19. R. Hisatomi, A. Osada, and Y. Tabuchi, “Bidirectional conversion between microwave and light via ferromagnetic magnons,” Phys. Rev. B 93(17), 174427 (2016). [CrossRef]  

20. T. Liu, X. Zhang, H. X. Tang, et al., “Optomagnonics in magnetic solids,” Phys. Rev. B 94(6), 060405 (2016). [CrossRef]  

21. X. Zhang, N. Zhu, C.-L. Zou, et al., “Optomagnonic whispering gallery microresonators,” Phys. Rev. Lett. 117(12), 123605 (2016). [CrossRef]  

22. D. Lachance-Quirion, Y. Tabuchi, and A. Gloppe, “Hybrid quantum systems based on magnonics,” Appl. Phys. Express 12(7), 070101 (2019). [CrossRef]  

23. A. V. Chumak, V. I. Vasyuchka, A. A. Serga, et al., “Magnon spintronics,” Nat. Phys. 11(6), 453–461 (2015). [CrossRef]  

24. L. Bai, M. Harder, and P. Hyde, “Cavity mediated manipulation of distant spin currents using a cavity-magnon-polariton,” Phys. Rev. Lett. 118(21), 217201 (2017). [CrossRef]  

25. J. Li, S.-Y. Zhu, and G. Agarwal, “Magnon-photon-phonon entanglement in cavity magnomechanics,” Phys. Rev. Lett. 121(20), 203601 (2018). [CrossRef]  

26. M. Yu, H. Shen, and J. Li, “Magnetostrictively induced stationary entanglement between two microwave fields,” Phys. Rev. Lett. 124(21), 213604 (2020). [CrossRef]  

27. J. Li and S. Gröblacher, “Entangling the vibrational modes of two massive ferromagnetic spheres using cavity magnomechanics,” Quantum Sci. Technol. 6(2), 024005 (2021). [CrossRef]  

28. J. Li, S.-Y. Zhu, and G. Agarwal, “Squeezed states of magnons and phonons in cavity magnomechanics,” Phys. Rev. A 99(2), 021801 (2019). [CrossRef]  

29. W. Zhang, D.-Y. Wang, and C.-H. Bai, “Generation and transfer of squeezed states in a cavity magnomechanical system by two-tone microwave fields,” Opt. Express 29(8), 11773–11783 (2021). [CrossRef]  

30. C.-Z. Chai, Z. Shen, and Y.-L. Zhang, “Single-sideband microwave-to-optical conversion in high-q ferrimagnetic microspheres,” Photonics Res. 10(3), 820–827 (2022). [CrossRef]  

31. M.-S. Ding, L. Zheng, and C. Li, “Ground-state cooling of a magnomechanical resonator induced by magnetic damping,” J. Opt. Soc. Am. B 37(3), 627–634 (2020). [CrossRef]  

32. Z.-X. Yang, L. Wang, and Y.-M. Liu, “Ground state cooling of magnomechanical resonator in pt-symmetric cavity magnomechanical system at room temperature,” Front. Phys. 15(5), 52504 (2020). [CrossRef]  

33. B. Sarma, T. Busch, and J. Twamley, “Cavity magnomechanical storage and retrieval of quantum states,” New J. Phys. 23(4), 043041 (2021). [CrossRef]  

34. Y.-P. Wang, J. Rao, and Y. Yang, “Nonreciprocity and unidirectional invisibility in cavity magnonics,” Phys. Rev. Lett. 123(12), 127202 (2019). [CrossRef]  

35. Y. Zhao, J. Rao, and Y. Gui, “Broadband nonreciprocity realized by locally controlling the magnons radiation,” Phys. Rev. Appl. 14(1), 014035 (2020). [CrossRef]  

36. T. Trickle, Z. Zhang, and K. M. Zurek, “Detecting light dark matter with magnons,” Phys. Rev. Lett. 124(20), 201801 (2020). [CrossRef]  

37. M. S. Ebrahimi, A. Motazedifard, and M. B. Harouni, “Single-quadrature quantum magnetometry in cavity electromagnonics,” Phys. Rev. A 103(6), 062605 (2021). [CrossRef]  

38. D. Lachance-Quirion, S. P. Wolski, and Y. Tabuchi, “Entanglement-based single-shot detection of a single magnon with a superconducting qubit,” Science 367(6476), 425–428 (2020). [CrossRef]  

39. X. Zhang, C.-L. Zou, and N. Zhu, “Magnon dark modes and gradient memory,” Nat. Commun. 6(1), 8914 (2015). [CrossRef]  

40. R.-C. Shen, Y.-P. Wang, and J. Li, “Long-time memory and ternary logic gate using a multistable cavity magnonic system,” Phys. Rev. Lett. 127(18), 183202 (2021). [CrossRef]  

41. S.-N. Huai, Y.-L. Liu, and J. Zhang, “Enhanced sideband responses in a pt-symmetric-like cavity magnomechanical system,” Phys. Rev. A 99(4), 043803 (2019). [CrossRef]  

42. K. Ullah, M. T. Naseem, and Ö. E. Müstecaplıoğlu, “Tunable multiwindow magnomechanically induced transparency, fano resonances, and slow-to-fast light conversion,” Phys. Rev. A 102(3), 033721 (2020). [CrossRef]  

43. B. Wang, X. Jia, X.-H. Lu, et al., “Pt-symmetric magnon laser in cavity optomagnonics,” Phys. Rev. A 105(5), 053705 (2022). [CrossRef]  

44. Z.-X. Liu and H. Xiong, “Magnon laser based on brillouin light scattering,” Opt. Lett. 45(19), 5452–5455 (2020). [CrossRef]  

45. C. Potts, V. A. Bittencourt, S. V. Kusminskiy, et al., “Magnon-phonon quantum correlation thermometry,” Phys. Rev. Appl. 13(6), 064001 (2020). [CrossRef]  

46. D. Zhang, X.-Q. Luo, and Y.-P. Wang, “Observation of the exceptional point in cavity magnon-polaritons,” Nat. Commun. 8(1), 1368 (2017). [CrossRef]  

47. B. Wang, Z.-X. Liu, and C. Kong, “Magnon-induced transparency and amplification in Pt-symmetric cavity-magnon system,” Opt. Express 26(16), 20248–20257 (2018). [CrossRef]  

48. Y. Cao and P. Yan, “Exceptional magnetic sensitivity of p t-symmetric cavity magnon polaritons,” Phys. Rev. B 99(21), 214415 (2019). [CrossRef]  

49. H. Yuan, P. Yan, and S. Zheng, “Steady bell state generation via magnon-photon coupling,” Phys. Rev. Lett. 124(5), 053602 (2020). [CrossRef]  

50. U. Delić, M. Reisenbauer, K. Dare, et al., “Cooling of a levitated nanoparticle to the motional quantum ground state,” Am. Assoc. for Adv. Sci. 367(6480), 892–895 (2020). [CrossRef]  

51. C. Marletto and V. Vedral, “Gravitationally induced entanglement between two massive particles is sufficient evidence of quantum effects in gravity,” Phys. Rev. Lett. 119(24), 240402 (2017). [CrossRef]  

52. G. Ranjit, M. Cunningham, K. Casey, et al., “Zeptonewton force sensing with nanospheres in an optical lattice,” Phys. Rev. A 93(5), 053801 (2016). [CrossRef]  

53. T. Weiss, M. Roda-Llordes, E. Torrontegui, et al., “Large quantum delocalization of a levitated nanoparticle using optimal control: Applications for force sensing and entangling via weak forces,” Phys. Rev. Lett. 127(2), 023601 (2021). [CrossRef]  

54. R. Zhao, A. Manjavacas, F. J. G. de Abajo, et al., “Rotational quantum friction,” Phys. Rev. Lett. 109(12), 123604 (2012). [CrossRef]  

55. D. C. Moore and A. A. Geraci, “Searching for new physics using optically levitated sensors,” Quantum Sci. Technol. 6(1), 014008 (2021). [CrossRef]  

56. J. Gieseler, R. Quidant, C. Dellago, et al., “Dynamic relaxation of a levitated nanoparticle from a non-equilibrium steady state,” Nat. Nanotechnol. 9(5), 358–364 (2014). [CrossRef]  

57. C. Gonzalez-Ballestero, M. Aspelmeyer, and L. Novotny, “Levitodynamics: Levitation and control of microscopic objects in vacuum,” Science 374(6564), eabg3027 (2021). [CrossRef]  

58. A. A. Rahman, “Large spatial schrödinger cat state using a levitated ferrimagnetic nanoparticle,” New J. Phys. 21(11), 113011 (2019). [CrossRef]  

59. C. C. Rusconi, V. Pöchhacker, and K. Kustura, “Quantum spin stabilized magnetic levitation,” Phys. Rev. Lett. 119(16), 167202 (2017). [CrossRef]  

60. J. Gieseler, A. Kabcenell, and E. Rosenfeld, “Single-spin magnetomechanics with levitated micromagnets,” Phys. Rev. Lett. 124(16), 163604 (2020). [CrossRef]  

61. T. Seberson, P. Ju, and J. Ahn, “Simulation of sympathetic cooling an optically levitated magnetic nanoparticle via coupling to a cold atomic gas,” JOSA B 37(12), 3714–3720 (2020). [CrossRef]  

62. P. Huillery, T. Delord, and L. Nicolas, “Spin mechanics with levitating ferromagnetic particles,” Phys. Rev. B 101(13), 134415 (2020). [CrossRef]  

63. A. Vinante, G. Wijts, and O. Usenko, “Magnetic resonance force microscopy of paramagnetic electron spins at millikelvin temperatures,” Nat. Commun. 2(1), 572 (2011). [CrossRef]  

64. R. Fischer, D. P. McNally, and C. Reetz, “Spin detection with a micromechanical trampoline: towards magnetic resonance microscopy harnessing cavity optomechanics,” New J. Phys. 21(4), 043049 (2019). [CrossRef]  

65. N. K. Raut, J. Miller, J. Pate, et al., “Meissner levitation of a millimeter-size neodymium magnet within a superconducting radio frequency cavity,” IEEE Trans. Appl. Supercond. 31(5), 1–4 (2021). [CrossRef]  

66. C. Timberlake, G. Gasbarri, and A. Vinante, “Acceleration sensing with magnetically levitated oscillators above a superconductor,” Appl. Phys. Lett. 115(22), 1 (2019). [CrossRef]  

67. R. Diehl, E. Hebestreit, R. Reimann, et al., “Optical levitation and feedback cooling of a nanoparticle at subwavelength distances from a membrane,” Phys. Rev. A 98(1), 013851 (2018). [CrossRef]  

68. G. P. Conangla, F. Ricci, M. T. Cuairan, et al., “Optimal feedback cooling of a charged levitated nanoparticle with adaptive control,” Phys. Rev. Lett. 122(22), 223602 (2019). [CrossRef]  

69. A. Ashkin and J. Dziedzic, “Optical levitation by radiation pressure,” Appl. Phys. Lett. 19(8), 283–285 (1971). [CrossRef]  

70. F. Monteiro, W. Li, and G. Afek, “Force and acceleration sensing with optically levitated nanogram masses at microkelvin temperatures,” Phys. Rev. A 101(5), 053835 (2020). [CrossRef]  

71. A. Kani, B. Sarma, and J. Twamley, “Intensive cavity-magnomechanical cooling of a levitated macromagnet,” Phys. Rev. Lett. 128(1), 013602 (2022). [CrossRef]  

72. D. D. Stancil, A. Prabhakar, D. D. Stancil, et al., “Quantum theory of spin waves,” Spin Waves: Problems and Solutions pp. 21–35 (2021).

73. D. D. Stancil, Theory of magnetostatic waves (Springer Science & Business Media, 2012).

74. D. Mills, “Quantum theory of spin waves in finite samples,” J. Magn. Magn. Mater. 306(1), 16–23 (2006). [CrossRef]  

75. M. Goryachev, W. G. Farr, and D. L. Creedon, “High-cooperativity cavity qed with magnons at microwave frequencies,” Phys. Rev. Appl. 2(5), 054002 (2014). [CrossRef]  

76. P.-C. Ma, J.-Q. Zhang, and Y. Xiao, “Tunable double optomechanically induced transparency in an optomechanical system,” Phys. Rev. A 90(4), 043825 (2014). [CrossRef]  

77. C. W. Gardiner and P. Zoller, Quantum Noise, vol. 56 of Springer series in synergetics (Springer–Verlag Berlin, 2000), 97–98.

78. R. E. Bellman and R. E. Kalaba, “Selected papers on mathematical trends in control theory,” (No Title) (1964).

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1.
Fig. 1. Schematic of the cavity which consists of the levitated YIG sphere and is driven by a control field and a weak input probe field.
Fig. 2.
Fig. 2. The real (red line), $\varepsilon _r$, and the imaginary (blue line), $\varepsilon _i$, parts of the field amplitude, $\varepsilon _t$, versus the normalized frequency $\delta /\omega _c$ (a) without magnon-photon coupling $g_{ab}=0$ (b) with the magnon-photon coupling $g_{ab}/2\pi = 15.6\,MHz$. The parameters are selected as follows: $\omega _a/2\pi =\omega _b/2\pi =30 \, GHz$, $\omega _c/2\pi =50\,kHz$, $\gamma _a/2\pi =1\, MHz$, $\gamma _c/2\pi =10^{-6}\,Hz$, $k x_{0}=2\pi /3$, and $P=1\, \mu W$. To obtain this result, we have selected $\Delta _a=\Delta _b\equiv \Delta =\omega _c$ [71].
Fig. 3.
Fig. 3. The real part of the field amplitude, $\varepsilon _r$, versus the normalized frequency $\delta /\gamma _b$ with $\gamma _c/2\pi =10 Hz$, and $P=100\, mW$, and the magnon-photon coupling $g_{ab}/2\pi = 1.5\,MHz$ for different magnon-photon-CM couplings: $g_{abc}/2\pi =10^{-4}Hz$ (green line), $12\times 10^{-2}Hz$ (blue line), $13\times 10^{-2}Hz$ (red line), $14\times 10^{-2}Hz$ (purple line). Other parameters are selected the same as Fig. (2).
Fig. 4.
Fig. 4. The real $\varepsilon _r$ part of the output field versus the normalized frequency $\delta /\omega _c$ and (a) magnon-photon coupling $g_{ab}$ and (b) particle position ${kx_0}/{\pi }$. The parameters are selected the same as Fig. (2)
Fig. 5.
Fig. 5. (a), (b) The real part, $\varepsilon _r$, of the output field versus the normalized frequency $\delta /\omega _c$ and particle position ${kx_0}/\pi$, (a) $R=100 \, \mu m$ and (b) $R=50 \, \mu m$. Other parameters are selected the same as Fig. (2)
Fig. 6.
Fig. 6. The real $\varepsilon _r$ part of the output field versus the normalized frequency $\delta /\omega _c$ and the normalized detuning $\Delta /\omega _c$. Other parameters are selected the same as Fig. (2)
Fig. 7.
Fig. 7. The real part, $\varepsilon _r$, of the output field versus the normalized frequency $\delta /\omega _c$ and (a) particle position ${kx_0}/{\pi }$ and (b) magnon-photon coupling $g_{ab}$ for the intermediate coupling (IMC) regime $\gamma _b<g_{ab}<\gamma _a$ and $\gamma _a/2\pi =14\,MHz$. Other parameters are selected the same as Fig. (2)
Fig. 8.
Fig. 8. The real part, $\varepsilon _r$, of the output field versus the normalized frequency $\Omega /\omega _b$ and different magnon-photon couplings $g_{ab}/2\pi =12.6 MHz$ (pink line), $14.6 MHz$ (blue line), $16.6 MHz$ (red line), $18.6 MHz$ (green line). The parameters are selected the same as Fig. (2).
Fig. 9.
Fig. 9. The output field phase versus the normalized frequency $\delta /\omega _c$ and magnon-photon coupling $g_{ab}/\gamma _b$. A critical point in the vicinity $g_{ab}/\gamma _b\sim 2$. The parameters are selected the same as Fig. (2).
Fig. 10.
Fig. 10. Group delay $\tau _g$ versus (a) the magnon-photon coupling strength $g_{ab}/\gamma _b$ (b) the particle position $kx_0/\pi$ (c) the detuning frequency $\Delta /\gamma _b$, under the various of the damping rate of microwave cavity field $\gamma _a$, respectively. The parameters are selected the same as Fig. (2).

Equations (20)

Equations on this page are rendered with MathJax. Learn more.

H ^ / = ω a a ^ a ^ + ω b b ^ b ^ + ω c 2 ( x 2 + p x 2 ) + g a b ( a ^ b ^ + a ^ b ^ ) cos ( k x ) + i ( s i n ( t ) a ^ s i n ( t ) a ^ ) .
H ^ / = ω a a ^ a ^ + ω b b ^ b ^ + ω c c ^ c ^ + g a b cos ( k x 0 ) ( a ^ b ^ + a ^ b ^ ) g a b c sin ( k x 0 ) ( a ^ b ^ + a ^ b ^ ) ( c ^ + c ^ ) + i ( s i n ( t ) a ^ s i n ( t ) a ^ ) ,
s i n ( t ) = ( ε d + ε p e i ( ω p ω d ) t ) e i ω d t = s i n e i ω d t ,
H ^ / = Δ a a ^ a ^ + Δ b b ^ b ^ + ω c c ^ c ^ + g a b cos ( k x 0 ) ( a ^ b ^ + a ^ b ^ ) g a b c sin ( k x 0 ) ( a ^ b ^ + a ^ b ^ ) ( c ^ + c ^ ) + i ( s i n a ^ s i n a ^ ) ,
a ˙ ^ = ( γ a + i Δ a ) a ^ + i g a b c sin ( k x 0 ) b ^ ( c ^ + c ^ ) i g a b cos ( k x 0 ) b ^ + s i n + 2 γ a a ^ i n , b ˙ ^ = ( γ b + i Δ b ) b ^ + i g a b c sin ( k x 0 ) a ^ ( c ^ + c ^ ) i g a b cos ( k x 0 ) a ^ + 2 γ b b ^ i n , c ˙ ^ = ( γ c + i ω c ) c ^ + i g a b c sin ( k x 0 ) ( a ^ b ^ + a ^ b ^ ) + 2 γ c c ^ i n .
< o i n ( t ) o i n ( t ) > = ( n ¯ o + 1 ) δ ( t t ) , < o i n ( t ) o i n ( t ) > = n ¯ o δ ( t t ) ,
a ˙ 0 = ( γ a + i Δ a ) a 0 i ( G a b + 2 G a b c R e [ c 0 ] ) b 0 + ε d = 0 , b ˙ 0 = ( γ b + i Δ b ) b 0 i ( G a b + 2 G a b c R e [ c 0 ] ) a 0 = 0 , c ˙ 0 = ( γ c + i ω c ) c 0 i G a b c ( a 0 b + a 0 b 0 ) = 0 ,
a 0 = ε d ( γ b + i Δ b ) ( γ a + i Δ a ) ( γ b + i Δ b ) + ( G a b + 2 G a b c R e [ c 0 ] ) 2 , b 0 = i ε d ( G a b + 2 G a b c R e [ c 0 ] ) ( γ a + i Δ a ) ( γ b + i Δ b ) + ( G a b + 2 G a b c R e [ c 0 ] ) 2 , c 0 = 2 i G a b c Δ b ( G a b + 2 G a b c R e [ c 0 ] ) ( γ c + i ω c ) ( γ b 2 + Δ b 2 ) | a 0 | 2 .
a 0 ε d ( γ b + i Δ b ) ( γ a + i Δ a ) ( γ b + i Δ b ) + G a b 2 , b 0 i ε d G a b ( γ a + i Δ a ) ( γ b + i Δ b ) + G a b 2 , c 0 2 i G a b G a b c Δ b ( γ c + i ω c ) ( γ b 2 + Δ b 2 ) | a 0 | 2 .
δ a ^ ˙ ( t ) = ( γ a + i Δ a ) δ a ^ i Γ δ b ^ i G a b c b 0 ( δ c ^ + δ c ^ ) + ε p e i δ t + 2 γ a a ^ i n δ b ^ ˙ ( t ) = ( γ b + i Δ b ) δ b ^ i Γ δ a ^ i G a b c a 0 ( δ c ^ + δ c ^ ) + 2 γ b b ^ i n , δ c ^ ˙ ( t ) = ( γ c + i ω c ) δ c ^ i G a b c ( a 0 δ b ^ + a 0 δ b ^ + b 0 δ a ^ + b 0 δ a ^ ) + 2 γ c c ^ i n .
δ a ^ ( t ) = A e i δ t + A + e i δ t , δ b ^ ( t ) = B e i δ t + B + e i δ t , δ c ^ ( t ) = C e i δ t + C + e i δ t .
A = i Γ B i G a b c b 0 Q + ε p , [ γ a i ( Δ a + δ ) ] A , = i Γ B , + i G a b c b 0 Q , [ γ b + i ( Δ b δ ) ] B = i Γ A i G a b c a 0 Q , [ γ b i ( Δ b + δ ) ] B , = i Γ A , + i G a b c a 0 Q ,
Q = 2 G a b c ω c a 0 B + a 0 B , + b 0 A + b 0 A , ( γ c i δ ) 2 + ω c 2 .
ε o u t + ε d e i ω d t + ε p e i ω p t = 2 γ a ( a 0 + δ a ^ ) e i ω d t .
ε t = 2 γ a | A | ε p = 2 γ a [ γ b + i ( Δ b δ ) ] f ( δ ) + Γ 2 ,
f ( δ ) = [ γ a + i ( Δ a δ ) ] [ γ b + i ( Δ b δ ) ] + ζ ( δ ) ξ ( δ ) [ Γ a 0 + i b 0 ( γ b + i ( Δ b δ ) ) ] ,
ζ ( δ ) = 2 G a b c 2 ω c [ b 0 i Γ a 0 γ b + i ( Δ b δ ) ] , ξ ( δ ) = 2 G a b c 2 ω c [ a 0 β ( δ ) + b 0 α ( δ ) i | a 0 | 2 γ b + i ( Δ b δ ) ] + ( γ c i δ ) 2 + ω c 2 ,
α ( δ ) = Γ a 0 + i b 0 ( γ b i ( Δ b + δ ) ) ( γ a i ( Δ a + δ ) ) ( γ b i ( Δ b + δ ) ) + Γ 2 , β ( δ ) = i ( Γ α ( δ ) + a 0 ) γ b i ( Δ b + δ ) .
ε t = 2 γ a | A | ε p = 2 γ a γ a + i ( ω a ω p ) + g a b 2 cos 2 ( k x 0 ) γ b + i ( ω b ω p ) .
τ g = Φ t ( ω p ) ω p = Im [ 1 ε t ε t ω p ] .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.