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

Nonlinear level attraction of cavity axion polariton in an antiferromagnetic topological insulator

Yang Xiao1,∗, Huaiqiang Wang2,3,∗, Dinghui Wang2, Ruifeng Lu4, Xiaohong Yan5, Hong Guo6, C. -M. Hu7, Ke Xia8, Haijun Zhang2,9,† and Dingyu Xing2,9 1College of Science, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China 2National Laboratory of Solid State Microstructures and Physics School, Nanjing University, Nanjing 210093, China 3Department of Physics, University of Zurich, Zurich 8057, Switzerland 4Department of Applied Physics, Nanjing University of Science and Technology, Nanjing 210094, China 5School of Material Science and Engineering, Jiangsu University, Zhenjiang, 212013, China 6Department of Physics, McGill University, Montreal, Quebec H3A 2T8, Canada 7Department of Physics and Astronomy, University of Manitoba, Winnipeg R3T 2N2, Canada 8Beijing Computational Science Research Center, Beijing 100193, China 9Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210093, China [email protected]
Abstract

Strong coupling between cavity photons and elementary excitations in condensed matters boosts the field of light-matter interaction and generates several exciting sub-fields, such as cavity optomechanics and cavity magnon polariton. Axion quasiparticles, emerging in topological insulators, were predicted to strongly couple with the light and generate the so-called axion polariton. Here, we demonstrate that there arises a gapless level attraction of cavity axion polariton in an antiferromagnetic topological insulator, which originates from the high-order interaction between axion and the odd-order resonance of a cavity. Such a novel coupling mechanism is essentially different from conventional level attractions with the linear-order interaction. Our results open up promising roads for exploring the axion polariton with cavity technologies.

Refer to caption
Figure 1: Conventional LR (a), conventional LA (b) and high-order LA (c) depicted with solid red lines. The dashed blue lines generally present two uncoupled modes. (d) Schematic of cavity geometry with a dynamical axion insulator (DAI) film. The photon propagates from the left port to the right port through the DAI film. Depending on the mode number, the cavity resonance mode can be the even mode or the odd mode. As for the odd mode, the electric field at the DAI is negligibly small, indicating that the photon-axion coupling would be dominated by the high-order interaction. But for the even mode, a large electric field can lead to a linear photon-axion coupling. The static magnetic field is applied with an angle φ\varphi with respect to the zz axis and induces a coupling between the electric field 𝐄\mathbf{E} of photon and dynamical axion field (DAF) δθ\delta\theta. (e) The coupling strength bb between axion mode and the photon as a function of the mass parameter MM, distinguishing between topologically trivial (M>0M>0) and nontrivial (M<0M<0) DAI.
Refer to caption
Figure 2: (a) The axion angle, (b) coefficient 1/g1/g, (c) stiffness JJ, and (d) 1/(gJ)1/(g\sqrt{J}) as a function of the mass term MM, where M<0M<0 (M>0M>0) corresponds to the topologically nontrivial (trivial) phase with(without) band inversion.
Refer to caption
Figure 3: |S21||S_{21}| transmission spectra of a 0.1μm\mu m-thick DAI obtained from the numerical method. The regions in green box of left panels are zoomed-in and shown in the right panels. Top (a and b) and bottom (c and d) panels are for RR=0.99 and RR=0.999.
Refer to caption
Figure 4: |S21||S_{21}| transmission spectra of a 0.1μm\mu m-thick DAI for odd mode obtained from (left) the numerical method and (right) the formula Eq. (22). Top (a and b) and bottom (c and d) panels are for RR=0.99 and RR=0.999.
Refer to caption
Figure 5: (a) |S21||S_{21}| transmission spectra of the even mode as a function of the static magnetic field and the photon frequency, which is calculated based on Eq. (20) for the DAI film thickness ls=0.1μml_{s}=0.1\mu m. The strong coupling manifests the LR. (b) Spectrum at resonant magnetic field represented by the white dashed line in (a). (c) Absolute value of the term GnG_{n} in the expansion of G=n=0+GnG=\sum\limits_{n=0}^{+\infty}G_{n}. As for the even mode, one can see that the linear term (nn=1) dominates. (d-f) The same as (a-c) but for the odd mode and calculated using Eq. (21). In order to avoid divergence, a imaginary part is added into the frequency ω\omega. As for the odd mode, the first- and second-order terms are vanishing, indicating the high-order nature of the coupling.
Refer to caption
Figure 6: (a) |S21||S_{21}| transmission spectra of the odd mode calculated with the numerical method (without using Taylor expansion) for a DAI film thickness ls=0.1μml_{s}=0.1\mu m and the reflectivity R=0.99R=0.99 of the cavity. The reflectivity RR of electromagnetic wave at two ports determines the decay rate of intracavity field. (b) |S21||S_{21}| spectra for the reflectivity RR=0.95 (black), 0.97 (blue), 0.99 (red) at the resonant magnetic field. As the reflectivity RR increases, the dissipation is gradually suppressed, and the gap width is enhanced little by little. But we notice that the position of the gap remains unchanged. (c) The schematic mechanism of gaped LA. The approximated formula of S21S_{21} spectra is written as (ωm)/[(ωωc)(ωω1)](\omega-m^{\prime})/[(\omega-\omega^{\prime}_{c})(\omega-\omega^{\prime}_{1})] (S21,black), which can be considered as the combination of two functions 1/[(ωωc)(ωω1)]1/[(\omega-\omega^{\prime}_{c})(\omega-\omega^{\prime}_{1})] (F1, brown) and (ωm)(\omega-m^{\prime}) (F2, green). Here, ‘mm^{\prime}, ωc\omega^{\prime}_{c} and ω1\omega^{\prime}_{1}’ are renormalized by the dissipation.
Refer to caption
Figure 7: |S21||S_{21}| transmission spectra for a DAI film thickness lsl_{s}=0.1 μm\mu m with different coupling strength bb= 0(a), 1(b) and 2(c) meV. Controllable coupling strength bb can be obtained through applying the external magnetic field in topologically nontrivial DAI. We can see that an enhanced gap in the LA is induced by strong coupling. (d-f) The same as (a-c), but for a thicker DAI film lsl_{s}=0.5 μm\mu m. Note that high-order modes arise in the spectrum for which the high-order interaction dominates. For all data here, the reflectivity RR is fixed to be 0.99.
Refer to caption
Figure 8: |S21||S_{21}| transmission spectra of a 0.1μm\mu m-thick DAI for odd mode calculated with numerical method for RR=0.99 (left), RR=0.999 (middle) and RR=0.9999 (right). The regions in green box of top panels are zoomed-in and shown in the bottom panels. Notice that two figures of RR=0.99 (left top and left bottom) are the same for the sake of completeness.

I I. Introduction

Axion was first postulated as an elementary particle to solve the charge-parity puzzle in the strong interaction between quarks in the particle physics Peccei and Quinn (1977). But its existence in nature is another puzzle. Interestingly, axion emerges as a quasiparticle in three-dimensional (3D) topological insulators through the axion action 𝒮topo=(θ/2π)(e2/hc0)d3x𝑑t𝐄𝐁\mathcal{S}_{\text{topo}}=(\theta/2\pi)(e^{2}/hc_{0})\int d^{3}xdt\mathbf{E}\cdot{\mathbf{B}} from the topological field theory Qi and Zhang (2011); Qi et al. (2008), in which 𝐄\mathbf{E} and 𝐁\mathbf{B} are the electric field and magnetic induction inside the insulators, ee is the charge of an electron, hh is Plank’s constant, c0c_{0} is the speed of light in the vacuum, θ\theta (modulo 2π2\pi) is the dimensionless pseudoscalar parameter as the axion field. Once both time reversal symmetry and inversion symmetry are broken, the axion field θ\theta becomes dynamical due to the spin-wave excitation, denoted as a dynamical axion field (DAF) θ(𝒓,t)\theta(\bm{r},t) Li et al. (2010); Wang et al. (2011). Many exotic electromagnetic phenomena were predicted for DAF systems, such as, axion polariton Li et al. (2010); Marsh et al. (2019), the chiral magnetic effect Sekine and Nomura (2016); Sumiyoshi and Fujimoto (2016), and so on Ooguri and Oshikawa (2012); Taguchi et al. (2018); Imaeda et al. (2019); Gooth et al. (2019).

Recently, several topological dynamical axion insulators (DAIs) were proposed to a host large DAF characterized by a nonzero spin Chern number Wang et al. (2020) and strongly couple with the light, for example, van der Waals layered material Mn2Bi2Te5 Zhang et al. (2020) and MnBi2Te4/Bi2Te3 superlattice Wang et al. (2020). Especially, some of us predicted tunable DAF in MnBi2Te4 films Zhu et al. (2021) which have been synthesized in experiments Gong et al. (2019); Otrokov et al. (2019); Deng et al. (2020); Liu et al. (2020); Chen et al. (2019a); Klimovskikh et al. (2020). The proposed DAIs provide a guarantee for the in-depth study of axion electrodynamics Nenno et al. (2020); Sekine and Nomura (2021). The axion polariton is one of the most interesting electromagnetic phenomena, where the axion quasiparticle in DAIs couples with free photons Li et al. (2010). However, the tunability of the free-photon setup turns out to be limited and no experimental evidence of axion polariton is reported. Compared to free photons, cavity photons provide a better choice due to large quality factor and enhanced light-matter interaction Cohen-Tannoudji et al. (2004); Haroche (2013); Savage (1989). Notably, the rapid progress in cavity fabrication has already made the THz cavity possible Scalari et al. (2012), which covers the typical energy scale (1\sim 1 meV) of axion quasiparticles in DAIs Li et al. (2010). This enables to study the axion polariton in a cavity, dubbed as cavity axion polariton. One expects that the cavity axion polariton will present unique properties compared to the axion polariton with free photons Li et al. (2010).

In this paper, we studied the electrodynamics of cavity axion polariton by embedding a DAI into a THz cavity. Intriguingly, in addition to the usual level repulsion (LR) Aspelmeyer et al. (2014); Agarwal (1984); Tabuchi et al. (2015), schematically shown in Fig. 1(a), a characteristic level attraction (LA), schematically shown in Fig. 1(c), was found. Based on numerical and analytical calculation, we found that such an LA originates from the high-order axion-photon interaction for which we call it high-order LA. This new mechanism, which is distinct from linear-order coupling of conventional LA Metelmann and Clerk (2015); Bernier et al. (2018); Harder et al. (2018); Yu et al. (2019); Tserkovnyak et al. (2005); Bhoi et al. (2019); Grigoryan et al. (2018); Boventer et al. (2020) schematically shown in Fig. 1(b), has so far not been revealed.

The remainder of this paper is organized as follows. In Sec. II, we provide a comprehensive description of theoretical and methodological details used in this work. For the convenience of readers, the detailed derivation of some formulae can be found in Appendix A\simC. Section III is devoted to the presentation and interpretation of our numerical and analytical results for the high-order LA.

II II. Theory and method

II.1 IIA. Axion electrodynamics

The axion-photon coupled system can be described by the effective action Li et al. (2010)

𝒮tot=𝒮 Maxwell +𝒮 topo +𝒮 axion =18πd3x𝑑t(𝐃𝐄𝐁𝐇)+α4π2(θ0+δθ)d3x𝑑t𝐄𝐁+g2Jd3x𝑑t[(tδθ)2(viiδθ)2m2δθ2],\begin{split}\mathcal{S}_{\mathrm{tot}}=&\mathcal{S}_{\text{ Maxwell }}+\mathcal{S}_{\text{ topo }}+\mathcal{S}_{\text{ axion }}\\ &=\frac{1}{8\pi}\int d^{3}xdt\left(\mathbf{D}\cdot\mathbf{E}-\mathbf{B}\cdot\mathbf{H}\right)\\ &+\frac{\alpha}{4\pi^{2}}\left(\theta_{0}+\delta\theta\right)\int d^{3}xdt\mathbf{E}\cdot\mathbf{B}\\ &+g^{2}J\int d^{3}xdt\left[\left(\partial_{t}\delta\theta\right)^{2}-\left(v_{i}\partial_{i}\delta\theta\right)^{2}-m^{2}\delta\theta^{2}\right],\end{split} (1)

where 𝐇\mathbf{H} and 𝐃\mathbf{D} are the vectors of magnetic field and electric displacement. The constitutive relations are 𝐁=μr𝐇\mathbf{B}=\mu_{r}\mathbf{H} and 𝐃=ϵ𝐄\mathbf{D}=\epsilon\mathbf{E} with ϵ\epsilon and μr\mu_{r} the dielectric constant and magnetic permeability. α=e2/c0\alpha=e^{2}/\hbar c_{0} is the fine-structure constant, the axion field θ=θ0+δθ\theta=\theta_{0}+\delta\theta with θ0\theta_{0} the static part of the axion field. δθ\delta\theta represents the DAF originating from the longitudinal spin-wave excitation with material-dependent stiffness JJ, velocity viv_{i}, mass mm, and coefficient gg Li et al. (2010). As given by the 𝒮 topo \mathcal{S}_{\text{ topo }} term, strong coupling between axion field (δθ\delta\theta) and electromagnetic field (𝐄\mathbf{E}) can be realized by applying a static magnetic field 𝐁0\mathbf{B}_{0} parallel to the electric field 𝐄\mathbf{E}.

In order to obtain the dynamical equations for the coupled axion-photon system, we employ the Euler-Lagrangian equation and obtain the modified Maxwell’s equations Eqs. (A1)\sim(A4) and the equation of motion for the axion field Eq. (A5). After some algebra and calculus, we further obtain

2𝐄t2c22𝐄+απϵ2δθt2𝐁0=0,\displaystyle\frac{\partial^{2}\mathbf{E}}{\partial t^{2}}-c^{2}\bigtriangledown^{2}\mathbf{E}+\frac{\alpha}{\pi\epsilon}\frac{\partial^{2}\delta\theta}{\partial t^{2}}\mathbf{B}_{0}=0, (2)
2δθt2+m2δθ+Γδθtα8π2g2J𝐁0𝐄=0,\displaystyle\frac{\partial^{2}\delta\theta}{\partial t^{2}}+m^{2}\delta\theta+\Gamma\frac{\partial\delta\theta}{\partial t}-\frac{\alpha}{8\pi^{2}g^{2}J}\mathbf{B}_{0}\cdot\mathbf{E}=0, (3)

where the speed of light in the DAI is c=c0/ϵμrc=c_{0}/\sqrt{\epsilon\mu_{r}}. Γ\Gamma is the intrinsic damping of axion mode. Equation (3) also implies that the mass mm of DAF provides the resonance frequency of axion mode. In this work, the axion field is uniform inside the DAI and the axion mode is of a very long wavelength. Therefore, the dispersion of axion mode is negligible and is henceforth not taken into account. Moreover, the magnetic induction 𝐁\mathbf{B} in Eq. (1) consists of both magnetic component of electromagnetic wave 𝐁ac\mathbf{B}_{ac} and static magnetic field 𝐁0\mathbf{B}_{0}. However, the former is in general four orders of magnitude smaller than the latter and thus only 𝐁0\mathbf{B}_{0} is present in the above two equations.

II.2 IIB. Axion mode and parallel pumping

In contrast to the dark axion in the universe, the axion mode studied in this work is a quasiparticle existing in an antiferromagnetic DAI. As predicted in Ref. Li et al. (2010), the axion quasiparticle arises from the longitudinal spin wave excitations. That is to say, δθδMz\delta\theta\propto\delta M_{z}^{-} with Mz=(MzAMzB)/2M_{z}^{-}=(M_{z}^{A}-M_{z}^{B})/2 the zz component difference of two sublattice magnetizations 𝐌A,B\mathbf{M}^{A,B} of an antiferromagnet. The saturation magnetizations of two sublattices are aligned along the zz direction. In general, there are a number of thermal spin wave inside the DAI but its tunability is uneasy. To better excite and manipulate the spin wave, we consider the light-driven spin wave which can be controlled by tuning the frequency and the power of light. There are two ways of exciting spin wave, i.e. perpendicular pumping Huebl et al. (2013); Tabuchi et al. (2014); Goryachev et al. (2014); Zhang et al. (2014); Bai et al. (2015); Wang et al. (2018) and parallel pumping Morgenthaler (1963). In the former case, the magnetic component 𝐁ac\mathbf{B}_{ac} of electromagnetic wave is perpendicular to the static magnetic field 𝐁0\mathbf{B}_{0}. Hence, only transverse magnetization Mx,yA,BM_{x,y}^{A,B} is excited. Both ferromagnetic resonance (FMR) and antiferromagnetic resonance (AFMR) belong to perpendicular pumping Huebl et al. (2013); Tabuchi et al. (2014); Goryachev et al. (2014); Zhang et al. (2014); Bai et al. (2015); Wang et al. (2018). The resonance frequency usually takes the form of ω=γB0z\omega^{\bot}=\gamma B_{0z} for FMR and ω=ω0±γB0z\omega^{\bot}=\omega_{0}\pm\gamma B_{0z} for AFMR. γ\gamma is the the gyromagnetic ratio, B0zB_{0z} is the zz component of 𝐁0\mathbf{B}_{0} and ω0\omega_{0} is the zero-field frequency of AFMR. Due to two sublattice magnetizations, two magnon modes and magnon frequencies occur in AFMR.

As for parallel pumping, 𝐁ac\mathbf{B}_{ac} is parallel to static magnetic field and two sublattice magnetizations. The time-varying nature of 𝐁ac\mathbf{B}_{ac} will induce the periodically-precessing magnetization and result in δMz\delta M_{z}^{-} and δθ\delta\theta. The frequency of δMz\delta M_{z}^{-} variation in parallel pumping is given by ω=2ω\omega^{\parallel}=2\omega^{\bot} Morgenthaler (1963). From a quantum point of view, the parallel pumping is a process that one photon splits into two magnons with the same frequency. Therefore, the resonance frequency of axion mode in Eq. (3) is written as m=2ω0±2γB0zm=2\omega_{0}\pm 2\gamma B_{0z}. In this work, we consider the case of large magnetic field and thus two frequencies of axion mode are separated. So, two axion modes can be regarded as independent. In the following, only the axion mode with the frequency m=2ω0+2γB0zm=2\omega_{0}+2\gamma B_{0z} is considered in this work.

II.3 IIC. Axion-photon coupling strength

As described in the previous section, the dynamic magnetic field 𝐁ac\mathbf{B}_{ac} is aligned along the zz axis in order to give rise to δMz\delta M_{z} and δθ\delta\theta. So, the dynamic electric field 𝐄\mathbf{E} is perpendicular to the zz axis. We assume that 𝐄\mathbf{E} is along the xx axis. One can see from Eq. (3) that the term of 𝐁0𝐄\mathbf{B}_{0}\cdot\mathbf{E} should be nonzero for the axion-photon coupling. This requires that 𝐁0\mathbf{B}_{0} should have an xx component. On the other hand, the frequency of axion mode is related to the zz component of 𝐁0\mathbf{B}_{0}. For these two reasons, we apply a static magnetic field 𝐁0\mathbf{B}_{0} that is oriented with an angle φ\varphi with respect to the zz axis in the xx-zz plane. In such case, Eqs. (2) and (3) can be rewritten in scalar forms as

2Et2c22E+αB0xπϵ2δθt2=0,\displaystyle\frac{\partial^{2}E}{\partial t^{2}}-c^{2}\bigtriangledown^{2}E+\frac{\alpha B_{0x}}{\pi\epsilon}\frac{\partial^{2}\delta\theta}{\partial t^{2}}=0, (4)
2δθt2+m2δθ+ΓδθtαB0x8π2g2JE=0.\displaystyle\frac{\partial^{2}\delta\theta}{\partial t^{2}}+m^{2}\delta\theta+\Gamma\frac{\partial\delta\theta}{\partial t}-\frac{\alpha B_{0x}}{8\pi^{2}g^{2}J}E=0. (5)

By assuming the time- and position-dependence of ejωt+j𝐤𝐫e^{-j\omega t+j\mathbf{k}\cdot\mathbf{r}} for EE and δθ\delta\theta, one can simultaneously solve Eqs. (4) and (5) and obtain the eigenfrequency

ω±2=(c2k2+m2+b2jΓω)±(c2k2+m2+b2iΓω)24c2k2(m2jΓω)2\omega_{\pm}^{2}=\frac{(c^{2}k^{2}+m^{2}+b^{2}-j\Gamma\omega)\pm\sqrt{(c^{2}k^{2}+m^{2}+b^{2}-i\Gamma\omega)^{2}-4c^{2}k^{2}(m^{2}-j\Gamma\omega)}}{2} (6)

where b=α2B0x2/8π3ϵg2Jb=\sqrt{\alpha^{2}B_{0x}^{2}/8\pi^{3}\epsilon g^{2}J}. Equation (6) shows that the axion-photon coupling results in two polariton modes with frequency ω±\omega_{\pm}. At kk=0 and Γ\Gamma=0, ω+=m2+b2\omega_{+}=\sqrt{m^{2}+b^{2}} and ω=0\omega_{-}=0. As k+k\rightarrow+\infty, ω+ck\omega_{+}\rightarrow ck and ωm\omega_{-}\rightarrow m. Thus, there is an anticrossing gap between mm and m2+b2\sqrt{m^{2}+b^{2}} in which the wave can not propagate. This is in agreement with phonon, exciton and magnon polaritons Cohen-Tannoudji et al. (2004); Haroche (2013); Savage (1989).

Based on Eq. (6), the size of anticrossing gap in the spectrum is related to the parameter bb. The larger the parameter bb is, the larger the gap size is and the stronger the axion-photon coupling is. So, an important task is to determine the value of parameter bb. To do so, we consider the recently proposed MnBi2Te4 DAIs by some of the authors Wang et al. (2020); Zhang et al. (2020); Zhu et al. (2021), which are predicted to host a large DAF and may be potential candidates for observing the axion polariton.

The low-energy physics of the MnBi2Te4-based DAIs can be described by the following four-band Dirac Hamiltonian Wang et al. (2020); Zhang et al. (2020)

0(𝐤B)=E0(𝐤B)𝕀4×4+i=15di(𝐤B)Γi,d1,2,,5(𝐤B)=(A2kyB,A2kxB,A1kzB,MB(𝐤B),m5),\begin{split}\mathcal{H}_{0}(\mathbf{k}^{B})=&E_{0}(\mathbf{k}^{B})\mathbb{I}_{4\times 4}+\sum^{5}_{i=1}d_{i}(\mathbf{k}^{B})\Gamma^{i},\\ d_{1,2,...,5}(\mathbf{k}^{B})=&\Big{(}A_{2}k_{y}^{B},-A_{2}k_{x}^{B},A_{1}k_{z}^{B},M^{B}(\mathbf{k}^{B}),m_{5}\Big{)},\end{split} (7)

where 𝐤B\mathbf{k}^{B} is the wave vector in the Brillouin zone. E0(𝐤)=C+D1(kzB)2+D2((kxB)2+(kyB)2)E_{0}(\mathbf{k})=C+D_{1}(k_{z}^{B})^{2}+D_{2}((k_{x}^{B})^{2}+(k_{y}^{B})^{2}), MB(𝐤)=M+B1(kzB)2+B2((kxB)2+(kyB)2)M^{B}(\mathbf{k})=M+B_{1}(k_{z}^{B})^{2}+B_{2}((k_{x}^{B})^{2}+(k_{y}^{B})^{2}), and the five Dirac matrices are represented as Γ1,2,,5=(τxσx,τxσy,τyσ0,τzσ0,τxσz)\Gamma^{1,2,...,5}=(\tau_{x}\otimes\sigma_{x},\tau_{x}\otimes\sigma_{y},\tau_{y}\otimes\sigma_{0},\tau_{z}\otimes\sigma_{0},\tau_{x}\otimes\sigma_{z}), which satisfy the Clifford algebra {Γi,Γj}=2δi,j\{\Gamma^{i},\Gamma^{j}\}=2\delta_{i,j}. The parameters Bi=1,2B_{i=1,2} are assumed to be positive, and MM is the mass term related to spin-orbit coupling with M<0M<0 (M>0M>0) corresponding to the topological (trivial) antiferromagnetic insulator with(without) band inversion. The m5Γ5m_{5}\Gamma_{5} term results from the antiferromagnetic order, which breaks both time-reversal and inversion symmetries and is essential for the emergence of dynamical axion states. For convenience, in the numerical calculations, we choose the same values of the parameters Ai,Bi,C,DiA_{i},B_{i},C,D_{i} and m5m_{5} as those in Ref. Wang et al. (2020) for MnBi2Te4-Bi2Te3 heterostructure with C=0.0232C=0.0232 eV, D1=1.77D_{1}=1.77 eVÅ2\mathrm{eV\ \AA^{2}}, D2=10.82D_{2}=10.82 eVÅ2\mathrm{eV\ \AA^{2}}, A1=0.30A_{1}=0.30 eVÅ\mathrm{eV\ \AA}, A2=1.76A_{2}=1.76 eVÅ\mathrm{eV\ \AA}, B1=2.55B_{1}=2.55 eVÅ2\mathrm{eV\ \AA^{2}}, B2=14.20B_{2}=14.20 eVÅ2\mathrm{eV\ \AA^{2}}, m5=3.8m_{5}=3.8 meV. It is worth mentioning that the mass parameter MM can be tuned by changing the spin-orbit coupling strength through chemical doping, and is thus treated as a variable in our calculations.

The value of θ\theta in this model can be simply calculated via the formula Li et al. (2010)

θ=14πd3k2|d|+d4(|d|+d4)2|d|3ϵijkldixdjydkzdl,\theta=\frac{1}{4\pi}\int d^{3}k\frac{2|d|+d_{4}}{(|d|+d_{4})^{2}|d|^{3}}\epsilon^{ijkl}d_{i}\partial_{x}d_{j}\partial_{y}d_{k}\partial_{z}d_{l}, (8)

where i,j,k,li,j,k,l take values from 1, 2, 3, 5 and |d|=n=15dn2|d|=\sqrt{\sum_{n=1}^{5}d_{n}^{2}} with the lattice-regularized components of d1,2,,5d_{1,2,...,5} in the continuum model. Figure 2(a) shows θ\theta as a function of MM, which approaches π\pi (0) for topologically nontrivial (trivial) phase with M<0M<0 (M>0M>0). In the presence of longitudinal magnetization fluctuations, the axion field becomes dynamical, and to the linear order, δθ(𝐱,t)=δm5/g\delta\theta(\mathbf{x},t)=\delta m_{5}/g, where δm5\delta m_{5} is proportional to the amplitude fluctuation of antiferromagnetic order along the zz-direction δMz\delta M_{z}, and the coefficient 1/g1/g is determined from Eq. (8). In Fig. 2(b), we plot 1/g1/g as a function of MM, and found that 1/g1/g is significantly enhanced near the topological transition point M0M\rightarrow 0^{-}.

In addition, the stiffness parameter JJ is obtained from the above effective model as Li et al. (2010)

J=d3k(2π)3di(𝐤)di(𝐤)16|d|5d3k(2π)3116|d|3=1Ω𝐤116|d|3,J=\int\frac{\mathrm{d}^{3}k}{(2\pi)^{3}}\frac{d_{i}(\mathbf{k})d^{i}(\mathbf{k})}{16|d|^{5}}\sim\int\frac{\mathrm{d}^{3}k}{(2\pi)^{3}}\frac{1}{16|d|^{3}}=\frac{1}{\Omega}\sum_{\mathbf{k}}\frac{1}{16|d|^{3}}, (9)

where (i=1,2,3,4)\quad(i=1,2,3,4) and Ω\Omega is the size of the unit cell. Figure 2(c) shows the value of JJ as a function of MM. Taking the values of JJ and 1/g1/g together, we plot the band-structure-dependent coefficient 1/(gJ)1/(g\sqrt{J}) of bb as a function of MM in Fig. 2(d), where the largest magnitude is achieved near the topological transition.

Finally, after choosing typical values of the permittivity ϵ20\epsilon\sim 20 and the static magnetic field B0z0.2B_{0z}\sim 0.2 T, the realistic axion-photon coupling strength bb as a function of MM can be evaluated. As shown by Fig. 1(e), it is distinguished between topologically nontrivial (M<0M<0) and trivial (M>0M>0) DAIs. The magnitude of the coupling strength bb in the nontrivial regime is significantly larger than that in the trivial regime. The strongest axion-photon coupling can be obtained near the topological transition (M0M\sim 0). Therefore, an antiferromagnetic topological insulator with a large topological DAF is an ideal candidate for studying the strong axion-photon coupling in a cavity.

II.4 IID. Transmission spectrum

Transmission spectrum is an important tool of studying the light-matter interaction in both experimental and theoretical works Huebl et al. (2013); Tabuchi et al. (2014); Goryachev et al. (2014); Zhang et al. (2014); Bai et al. (2015); Wang et al. (2018). To discuss the axion-photon coupling, we calculate the transmission spectrum of a DAI-embedded cavity shown in Fig. 1(d). The axis of cavity and the propagation direction of electromagnetic waves are along the yy axis. Due to the axion-photon coupling, the photon transmission S21S_{21} carries the information of photon, axion and the coupling between them. The cavity is a one-dimensional hollow cylinder with input and output ports at the left and right ends. The input port is driven by external sources with an incident wave amplitude and the output port is detected with a transmitted wave amplitude. The transmitted-to-incident ratio is defined as transmission S21S_{21}. Inside a cavity, the wave is reflected many times. Thus the left-going waves and the right-going waves are superposed to form a standing wave. The cavity modes consist of many discrete resonance modes, which can be either even mode or odd mode as shown in Fig. 1(d) Pozar (2011). An antiferromagnetic DAI with the thickness of lsl_{s} is placed in the middle of the cavity. As for the even mode, the amplitude of electric field 𝐄\mathbf{E} at the DAI is large and thus the axion-photon coupling is strong. In contrast, the odd mode is weakly coupled with the DAI.

To calculate the transmission S21S_{21}, we first calculate the wave vector kk of propagation state inside the DAI. From Eq. (6), one can give rise to the relation k(ω)k(\omega)

c2k2=ω2μ,c^{2}k^{2}=\omega^{2}\mu, (10)

with the parameter

μ=ω2+iΓω(b2+m2)ω2+iΓωm2.\displaystyle\mu=\frac{\omega^{2}+i\Gamma\omega-(b^{2}+m^{2})}{\omega^{2}+i\Gamma\omega-m^{2}}. (11)

Here, the parameter μ\mu does not refer to the permeability.

With the wave vector kk, one can calculate the impedance Z=EH=ωkZ=\frac{E}{H}=-\frac{\omega}{k} with the help of Eq. (A3). Inside the cavity, there exist three regions, i.e. the left air region, the DAI and the right air region. Therefore, we have three transfer matrices which connect the electromagnetic field at the left surface and right surface of each region. As given in Appendix B, the transfer matrix in the DAI region is written as

TDAI=(cos(kls)jZsin(kls)jZsin(kls)cos(kls))T_{DAI}=\left(\begin{array}[]{cc}\cos(kl_{s})\ \ \ \ \ \ -jZ\sin(kl_{s})\\ -\frac{j}{Z}\sin(kl_{s})\ \ \ \ \ \ \cos(kl_{s})\\ \end{array}\right) (12)

where Z=c0ϵμZ=-\frac{c_{0}}{\sqrt{\epsilon\mu}} and k=ωc0ϵμk=\frac{\omega}{c_{0}}\sqrt{\epsilon\mu}.

The transfer matrix of the air region TairT_{air} is obtained by replacing the quantities kk, ZZ and lsl_{s} with those of the air region, i.e. k0=ωc0k_{0}=\frac{\omega}{c_{0}}, Z0=c0Z_{0}=-c_{0} and l=Lls2l=\frac{L-l_{s}}{2}. ll defines the length of two air regions and LL defines the total length of the cavity. The total transfer matrix of a cavity is determined by multiplying the transfer matrix of each region, i.e. T=TairTDAITairT=T_{air}T_{DAI}T_{air}.

With the transfer matrix TT, we determine the relation between the amplitudes of incident wave at input port and those of transmitted wave at output port Pozar (2011); Yao et al. (2015)

(EAEB)=S(EA+EB+)=(SAASABSBASBB)(EA+EB+),\left(\begin{array}[]{c}E_{A}^{-}\\ E_{B}^{-}\\ \end{array}\right)=S\left(\begin{array}[]{c}E_{A}^{+}\\ E_{B}^{+}\\ \end{array}\right)=\left(\begin{array}[]{cc}S_{AA}\ \ \ \ S_{AB}\\ S_{BA}\ \ \ \ S_{BB}\\ \end{array}\right)\left(\begin{array}[]{c}E_{A}^{+}\\ E_{B}^{+}\\ \end{array}\right), (13)

where (+)-(+) represents the direction from the cavity (outside) to the outside(cavity). The detailed form of scattering matrix SS is given in Appendix B.

In order to create standing waves inside a cavity, strong reflections at the input and output ports are necessary. We define the reflectivity of intracavity photon at two ports by the parameter RR and we choose RR\approx1 for both ports Pozar (2011); Yao et al. (2015).

The transmission is written as

S21=(1R2)SBA(1RSAA)(1RSBB)R2SABSBA.\displaystyle S_{21}=\frac{(1-R^{2})S_{BA}}{(1-RS_{AA})(1-RS_{BB})-R^{2}S_{AB}S_{BA}}. (14)

Substituting the elements of scattering matrix Eq. (13) into Eq. (14), we obtain

S21=FG,\displaystyle S_{21}=\frac{F}{G}, (15)

where the numerator FF and the denominator GG are given in Appendix B.

In the numerical calculations, we adopt the values of typical parameters which are within the achievable range of experimental measurement. We take the dielectric constant ϵ\epsilon=20, the permeability μr\mu_{r}=1, the length of cavity LL=5 mm, the axion frequency at zero magnetic field ω0\omega_{0}=1 meV, and the intrinsic damping of axion Γ\Gamma=0.01 meV. The parameters used in the tight-binding calculations are given in Sec. IIC. The other parameters, e.g. the thickness lsl_{s} of DAI, the coupling strength bb, the static magnetic field B0zB_{0z}, vary and are shown in the figures.

II.5 IIE. Analytical expressions

In order to better understand the numerical results, we derive the analytical expressions of transmission S21S_{21} and present simple formulae with transparent physics. The analytic expression of S21S_{21} is obtained by performing the Taylor expansion around cavity resonance frequency ωc\omega_{c} and in the limit of long-wavelength with kls1kl_{s}\ll 1.

First, the sinusoidal functions sin(kls)\sin(kl_{s}) and cos(kls)\cos(kl_{s}) in transfer matrix and scattering matrix are expanded as

sin(kls)\displaystyle\sin(kl_{s}) =\displaystyle= n=0+(1)n(2n+1)!(kls)2n+1,\displaystyle\sum\limits_{n=0}^{+\infty}\frac{(-1)^{n}}{(2n+1)!}(kl_{s})^{2n+1}, (16)
cos(kls)\displaystyle\cos(kl_{s}) =\displaystyle= n=0+(1)n(2n)!(kls)2n.\displaystyle\sum\limits_{n=0}^{+\infty}\frac{(-1)^{n}}{(2n)!}(kl_{s})^{2n}. (17)

Second, the exponential function e2jk0le^{-2jk_{0}l} in the scattering matrix is expanded at cavity resonance frequency ωCR=c0qπ2l+ls\omega_{CR}=\frac{c_{0}q\pi}{2l+l_{s}}, i.e.

e2jk0l\displaystyle e^{-2jk_{0}l} =ej2lc0ω=ejlsc0ωej2l+lsc0ω\displaystyle=e^{-j\frac{2l}{c_{0}}\omega}=e^{j\frac{l_{s}}{c_{0}}\omega}e^{-j\frac{2l+l_{s}}{c_{0}}\omega}
ej2l+lsc0ωCRejlsc0ω[1+ΔωC+]\displaystyle\approx e^{-j\frac{2l+l_{s}}{c_{0}}\omega_{CR}}e^{j\frac{l_{s}}{c_{0}}\omega}\left[1+\Delta\omega_{C}+\cdots\right]
=ejqπn1+(jlsc0ω)n1n1!n2+(ΔωC)n2n2!\displaystyle=e^{-jq\pi}\sum\limits_{n_{1}}^{+\infty}\frac{(j\frac{l_{s}}{c_{0}}\omega)^{n_{1}}}{n_{1}!}\sum\limits_{n_{2}}^{+\infty}\frac{(\Delta\omega_{C})^{n_{2}}}{n_{2}!} (18)

where ΔωC=(j2l+lsc0)(ωωCR)\Delta\omega_{C}=(-j\frac{2l+l_{s}}{c_{0}})(\omega-\omega_{CR}).

Substituting Eqs. (16)\sim(18) into Eqs. (14), one can obtain the analytical expressions of the numerator FF and the denominator GG. The form of numerator FF is simple and its analytical expression can be easily derived. However, the expression of denominator GG is lengthy and cumbersome and will be discussed in detail.

II.5.1 Even mode

As for the even mode, i.e. the integer qq in Eq. (18) is even, the denominator is written as Ge=n1=0+Gn1eG^{e}=\sum\limits_{n_{1}=0}^{+\infty}G_{n_{1}}^{e} with Gn1e=n2=0+Gn1,n2e(ΔωC)n2(lsc0ω)n1G_{n_{1}}^{e}=\sum\limits_{n_{2}=0}^{+\infty}G_{n_{1},n_{2}}^{e}(\Delta\omega_{C})^{n_{2}}(\frac{l_{s}}{c_{0}}\omega)^{n_{1}}. The detailed expansion of GG is given in Appendix C. For the sake of analysis, we divide the terms in GG into two groups. The first group, which is called the type-I interaction, consists of the terms without ΔωC\Delta\omega_{C}, i.e. those terms with n2n_{2}=0. The second group is called type-II interaction and contains all remaining terms.

As lsl_{s} is small, the denominator GeG^{e} is dominated by the zeroth-order term and the linear-order term of the type-I interaction, i.e.

Ge\displaystyle G^{e} =8ΔωC+8j(lsc0ω)(ϵμ1)\displaystyle=-8\Delta\omega_{C}+8j(\frac{l_{s}}{c_{0}}\omega)(\epsilon\mu-1)
=16jlc0[(ωωC)ϵb2ls4l(ωm)].\displaystyle=16j\frac{l}{c_{0}}\left[(\omega-\omega_{C})-\frac{\epsilon b^{2}l_{s}}{4l(\omega-m)}\right]. (19)

With the numerator FF, the transmission S21S_{21} of even mode for small lsl_{s} is written as

S21e=Ce(ωm)(ωωc)(ωm)ge2,\displaystyle S_{21}^{e}=C^{e}\frac{(\omega-m)}{(\omega-\omega_{c})(\omega-m)-g^{2}_{e}}, (20)

where CeC^{e} denotes a constant and ge2=ϵlsb2/4lg^{2}_{e}=\epsilon l_{s}b^{2}/{4l}.

II.5.2 Odd mode

As for the odd mode, i.e. the integer qq in Eq. (18) is odd, the denominator is written as Go=Gx+n1=0+Gn1oG^{o}=G_{x}+\sum\limits_{n_{1}=0}^{+\infty}G_{n_{1}}^{o} with Gn1o=n2=0+Gn1,n2o(ΔωC)n2(lsc0ω)n1G_{n_{1}}^{o}=\sum\limits_{n_{2}=0}^{+\infty}G_{n_{1},n_{2}}^{o}(\Delta\omega_{C})^{n_{2}}(\frac{l_{s}}{c_{0}}\omega)^{n_{1}}.

By comparing the expansion of even mode and odd mode given in Appendix C, we found that the 1st- and 2nd-order terms of type-I interaction are absent for the odd mode. Therefore, the high-order type-II terms dominate, resulting in a high-order LA. As lsl_{s} is small, the dominant terms are those with G1,1oG_{1,1}^{o} and G2,2oG_{2,2}^{o}, i.e. Eq. (C7) and (C8), and the term GxG_{x}, i.e. Eq. (C11), which are validated by numerical calculations.

In order to distinguish the role played by each term, we next consider two cases. In the first case, we do not include the GxG_{x} terms and consider the terms with G1,1oG_{1,1}^{o} and G2,2oG_{2,2}^{o}. By further simplifying these terms, we obtain a compact and physically meaningful expression of transmission

S21o1=Co1ωm(ωωc)(ωω1),\displaystyle S_{21}^{o1}=C^{o1}\frac{\omega-m}{(\omega-\omega_{c})(\omega-\omega_{1})}, (21)

where Co1C^{o1} also denotes a constant, ω1=(4m+λωc)/(4+λ)\omega_{1}=(4m+\lambda\omega_{c})/(4+\lambda) and λ=llsϵb2/c02\lambda=ll_{s}\epsilon b^{2}/c_{0}^{2}.

In the second case, we add the GxG_{x} terms on the basis of the first case. The expression of transmission becomes

S21o2=Co2ωm(ωωc)(ωω1),\displaystyle S_{21}^{o2}=C^{o2}\frac{\omega-m^{\prime}}{(\omega-\omega^{\prime}_{c})(\omega-\omega^{\prime}_{1})}, (22)

where ωc=ωci(1R)c0/2l\omega^{\prime}_{c}=\omega_{c}-i(1-R)c_{0}/2l, ω1=ω1i4ξ(1R)[8l(8l/c0+ξ)/c0]\omega^{\prime}_{1}=\omega_{1}-i4\xi(1-R)[8l(8l/c_{0}+\xi)/c_{0}], m=miΓωm^{\prime}=m-i\Gamma\omega, and ξ=2l2lsϵb2/c03\xi=2l^{2}l_{s}\epsilon b^{2}/c_{0}^{3}.

III III. Results and discussions

III.1 IIIA. Numerical results

Figure 3 shows the tranmission |S21||S_{21}| spectrum obtained by numerically calculating Eq. (14). One can see three cavity resonance modes, i.e. two odd modes at 510 and 570 GHz, and one even mode at 540 GHz. Due to the strong coupling between the axion mode and even mode of a cavity, a pronounced anticrossing gap occurs in the transmission spectrum. Moreover, as the reflectivity RR is increased from 0.99 (Fig. 3(a)) to 0.999 (Fig. 3(c)), the anticrossing gap of even mode remains almost unchanged.

For the odd mode, the spectrum is quite different from those of even mode. First, the coupling strength between axion mode and odd mode is very small. As seen in Eq. (3), the coupling is related to the amplitude of electric field 𝐄\mathbf{E}. The electric field at the DAI is minimum and maximum for odd mode and even mode respectively, resulting in a weak coupling between odd mode and axion mode. Second, as shown in the insets of Fig. 3, our calculations surprisingly present an LA for odd mode, instead of an LR of even mode. This indicates a different coupling mechanism of odd mode from that of even mode. Third, one can see that, when increasing RR from 0.99 to 0.999, the size of LA gap decreases.

III.2 IIIB. Analytical results

In order to reveal the origin of such an unexpected LA, we study the analytical expressions of S21S_{21} for even and odd modes given in Sec. IIE. Before doing so, we first compare the numerical and analytical results in order to verify the validity of the derived analytical expressions. Figure 4 shows the transmission |S21||S_{21}| spectrum of a 0.1μm\mu m-thick DAI for the odd mode. One can see a good agreement between numerical results and analytical expression. For both values of reflectivity RR=0.99 and RR=0.999, the analytical formula Eq. (22) reproduces the numerical results very well. Therefore, the analytical method and resulting formulae given in Sec. IIE are correct and reliable.

For the even mode, the transmission |S21||S_{21}| spectrum obtained from analytical expression Eq. (20) is shown in Fig. 5(a). The LR with upper and lower polariton branches (Fig. 5(b)) is clearly seen. By comparing the numerical results in Fig. 3(a) and analytical results in Fig. 5(a), one can find that the numerical result is well described by the analytical expression Eq. (20). In deriving this analytical formula, we consider only the zeroth-order and linear-order terms of type-I interaction. This treatment is also justified since the linear term is the largest among all GnG_{n} terms as seen in Fig. 5(c). The agreement between numerical and analytical results implies that the coupling between axion mode and even-order photon mode is of the linear nature. Such a linear coupling and the expression of S21S_{21} have been obtained for many other coupled systems, e.g. photon-phonon, photon-magnon and photon-exciton coupling, etc Cohen-Tannoudji et al. (2004); Haroche (2013); Savage (1989).

But for the odd mode, the above picture is completely changed. By checking the expansion of the denominator GG, we found that the first- and second-order terms of type-I interaction, which dominate in the case of even mode, are vanishing (Fig. 5(f)). So, the dominant terms are the high-order type-II terms and type-I terms. As lsl_{s} is small, the latter terms become very small and thus the high-order type-II interaction mainly contributes. In Sec. IIE, we derive two formulae by including and excluding the GxG_{x} terms. The GxG_{x} terms are related to the reflectivity RR and thus give rise to the effect of coupling on the line-width of modes. We first discuss the spectrum and thus neglect the GxG_{x} terms. The analytical results given in Fig. 5(d) show that there appears a new high-order mode with frequency ω1\omega_{1}. Since this high-order mode has a smaller slope than the axion mode, it bends to the cavity resonance with an attractive characteristic. Moreover, the high-order mode ω1\omega_{1} reduces to the axion mode at resonance (m=ωcm=\omega_{c}), reproducing S21=Co/(ωωc)S_{21}=C^{o}/(\omega-\omega_{c}) shown in Fig. 5(e). Therefore, the LA under this study is of both gapless and high-order nature, which is the central result of this work.

III.3 IIIC. Dissipation effect

The reflectivity RR indicates the decay rate (dissipation) of photons inside a cavity. A smaller RR represents a larger photon leakage (dissipation). In Eqs. (20), (21) and Fig. 5, we assume the reflectivity RR=1, implying no leakage of intracavity photons, and we obtained a gapless LA. Here, we further consider the dissipation with R<1R<1. Based on te numerical method, we calculate the S21S_{21} spectrum of the odd mode with R=0.99R=0.99 shown in Fig. 6(a) and find a gap at resonance, in contrast to the gapless LA in Fig. 5(d). Furthermore, a larger dissipation (a smaller RR) results in a wider gap, shown in Fig. 6(b). To understand the behavior in Fig. 6, we deduce Eq. (22) by including the GxG_{x} terms in the derivation. Interestingly, the new formula is of the same form as Eq. (21). The only difference is that the dissipation and line-width of coupled modes are renormalized. The results clearly give the correlation between dissipation rate and reflectivity RR. The new formula satisfactorily reproduces the gapped LA, indicating that the gap originates from the dissipation effect. Further analysis (Fig. 6(c)) shows that the minimum of the function (ωm)(\omega-m^{\prime}) at mm (see F2) plays an important role in producing the gap. Therefore, the dissipation is not the key to generating the high-order LA, but induces a gap in the LA.

III.4 IIID. Coupled-wave picture

To better understand the high-order LA, we present a model in which the photon and axion modes are simplified by waves. Since the photon is restricted inside the cavity, the cavity photon is described by a standing wave. For the axion mode, as pointed out in Sec. IIB, it is of uniform distribution inside the DAI. Therefore, the axion mode is modeled as a traveling wave. The standing wave, expressed by y1cos(kx)cos(ωt)y_{1}\propto\cos(kx)\cos(\omega t) with a wave vector k=ωvsk=\frac{\omega}{v_{s}}, sets up in the region [0L]\left[0\sim L\right]. vsv_{s} is the speed of wave and the resonance frequency ωs=nπvsL\omega_{s}=n\pi\frac{v_{s}}{L}. The traveling wave, e.g.e.g. y2cos(kxωt)y_{2}\propto\cos(kx-\omega t), propagates in the region [Lls2L+ls2]\left[\frac{L-l_{s}}{2}\sim\frac{L+l_{s}}{2}\right]. For simplicity, we consider that both waves have the same frequency. The coupling energy of two waves takes the same form of Est=Lls2L+ls2𝑑xy1y2E^{st}=\int_{\frac{L-l_{s}}{2}}^{\frac{L+l_{s}}{2}}dxy_{1}\cdot y_{2} as that in StopoS_{\text{topo}}. To gain clearer physical picture, we perform the Taylor expansion around ω=ωs\omega=\omega_{s} and ls=0l_{s}=0 for the coupling energy. Up to the third order, the coupling energy is written as

Ee[lsls4(δω)2]a(t)+[ls2δω]b(t),E^{e}\propto\left[l_{s}-\frac{l_{s}}{4}(\delta\omega)^{2}\right]a(t)+\left[\frac{l_{s}}{2}\delta\omega\right]b(t), (23)

for even mode and

Eo[ls4(δω)2]a(t)[ls2δω]b(t),E^{o}\propto\left[\frac{l_{s}}{4}(\delta\omega)^{2}\right]a(t)-\left[\frac{l_{s}}{2}\delta\omega\right]b(t), (24)

for odd mode. Here, δω=(ωωs)Lvs\delta\omega=(\omega-\omega_{s})\frac{L}{v_{s}}, a(t)=cos2(ωt)a(t)=\cos^{2}(\omega t) and b(t)=cos(ωt)sin(ωt)b(t)=\cos(\omega t)\sin(\omega t). At ω=ωs\omega=\omega_{s}, the coupling energy is nonzero for even mode(nn even) and thus the coupling induces frequency shift and mode hybridization. But for odd mode(nn odd), the coupling energy is zero, explaining the gapless feature of high-order LA in cavity axion polariton, i.e.i.e. ω=ωc=m\omega=\omega_{c}=m at resonance. To the lowest order in Eq. (23) and (24), the linear term (ls\sim l_{s}) dominates for even mode while the high-order term (lsδω\sim l_{s}\delta\omega) dominates for odd mode. Therefore, the high-order interaction plays a key role in the coupling between a traveling wave and an odd-order standing wave.

III.5 IIIE. Discussions

The LA found in this work is essentially different from the conventional LA. First of all, the LA here is attributed to a high-order interaction, but the conventional LA originates from the linear-order coupling of two modes Metelmann and Clerk (2015); Bernier et al. (2018); Harder et al. (2018); Yu et al. (2019); Tserkovnyak et al. (2005); Bhoi et al. (2019); Grigoryan et al. (2018); Boventer et al. (2020). Second, the dissipation is not the essence for the LA in this work and it merely tunes the LA gap. Third, the axion mode under this study arises from longitudinal magnetization fluctuation, instead of transverse magnetization fluctuation in cavity magnon polariton Soykal and Flatté (2010); Huebl et al. (2013); Tabuchi et al. (2014); Goryachev et al. (2014); Zhang et al. (2014); Bai et al. (2015); Wang et al. (2018); Cao et al. (2015).

To demonstrate the LA here is controllable, we vary the coupling strength bb and the thickness lsl_{s} of the DAI film. As shown in Fig. 1(e), the topologically trivial DAI possesses a negligible axion-photon coupling strength and thus does not couple with cavity resonance (Fig. 7(a)). Around the topological transition, the coupling strength quickly increases, resulting in a large gap of LA (Fig. 7(b) and 7(c)). This trend of increasing gap can also be obtained through increasing the thickness lsl_{s}. But, extra high-order modes may appear. As for a large lsl_{s} case, the type-II interaction does not overwhelm the type-I interaction that is absent for a small lsl_{s} case. Figure 8 shows the transmission |S21||S_{21}| spectrum as the reflectivity RR increases. One can see that the spectrum changes from an LA at RR=0.99 and 0.999 to an LR at RR=0.9999. As explained in Sec. IIIB and IIIC, the LA arises from the high-order type-II interaction and the dissipation effect. But as RR increases and approaches unity, the dissipation effect becomes weaker and weaker. In such a case, the higher-order term, e.g. the third-order term, of the type-I interaction starts to contribute and dominates to produces an LR. The high-order terms imply more new modes as shown in Fig. 7(e) and 7(f). The competition between type-I and type-II interaction results in the coexistence of the LR and LA. Therefore, it is expected to observe the LA, LR and high-order modes by tuning the topological phases (MM) and the DAI thickness lsl_{s}.

To make possible experimental measurements, a high-frequency cavity and a high-quality DAI material are important. Recently, Scalari et al. reported the fabrication of a μm\mu m-size electronic meta-material cavity with resonance frequency up to 4 THz Scalari et al. (2012), which well covers the frequency (0.5\sim 0.5 THz) of DAIs in this work. The ultra-strong coupling with a low dissipation rate has also been achieved, implying the potential application in cavity quantum electrodynamics. Further, thanks to the fast-growing field of MnBi2Te4-based magnetic topological materials and state-of-art experimental techniques Gong et al. (2019); Zhang et al. (2019); Li et al. (2019); Otrokov et al. (2019); Rienks et al. (2019); Deng et al. (2020); Liu et al. (2020); Chen et al. (2019a); Klimovskikh et al. (2020); Yan et al. (2019); Hao et al. (2019); Chen et al. (2019b); Zeugner et al. (2019); Wu et al. (2019); Vidal et al. (2019); Hu et al. (2020); He (2020), it is promising that candidate materials hosting a large topological DAF, such as MnBi2Te4 films, Mn2Bi2Te5 and MnBi2Te4/Bi2Te3 superlattice proposed by some of the authors Zhu et al. (2021); Zhang et al. (2020); Wang et al. (2020), can be used to study the high-order LA of cavity axion polariton.

Acknowledgements.
Acknowledgements. This work is supported by the Fundamental Research Funds for the Central Universities (Grant No. 020414380185), the Natural Science Foundation of China (Grants No. 61974067, No. 12074181, No. 12074181, No. 12174158 and No. 11834006), Natural Science Foundation of Jiangsu Province (No. BK20200007) and the Fok Ying-Tong Education Foundation of China (Grant No. 161006).

IV APPENDIX A: Derivation of dynamical equations

Here we show the procedure of deriving the dynamical equations of axion and photon modes. Based on Eqs. (1) and the Euler-Lagrangian equation, we can obtain

𝐃\displaystyle\nabla\cdot\mathbf{D} =\displaystyle= 4πραπ(δθ𝐁)\displaystyle 4\pi\rho-\frac{\alpha}{\pi}(\nabla\delta\theta\cdot\mathbf{B}) (A1)
×𝐇\displaystyle\nabla\times\mathbf{H} =\displaystyle= 1c0𝐃t+4πc0𝐣0\displaystyle\frac{1}{c_{0}}\frac{\partial\mathbf{D}}{\partial t}+\frac{4\pi}{c_{0}}\mathbf{j}_{0} (A2)
+\displaystyle+ απ((δθ×𝐄)+𝐁c0(tδθ))\displaystyle\frac{\alpha}{\pi}\left((\nabla\delta\theta\times\mathbf{E})+\frac{\mathbf{B}}{c_{0}}\left(\partial_{t}\delta\theta\right)\right)
×𝐄\displaystyle\nabla\times\mathbf{E} =\displaystyle= 1c0𝐁t\displaystyle-\frac{1}{c_{0}}\frac{\partial\mathbf{B}}{\partial t} (A3)
𝐁\displaystyle\nabla\cdot\mathbf{B} =\displaystyle= 0\displaystyle 0 (A4)
α𝐄𝐁8π2g2J\displaystyle\frac{\alpha\mathbf{E}\cdot\mathbf{B}}{8\pi^{2}g^{2}J} =\displaystyle= 2t2δθv22δθ+m2δθ,\displaystyle\frac{\partial^{2}}{\partial t^{2}}\delta\theta-v^{2}\nabla^{2}\delta\theta+m^{2}\delta\theta, (A5)

where ρ\rho and 𝐣0\mathbf{j}_{0} are the charge density and current density. Equations (A1)\sim(A4) correspond to the modified Maxwell’s equations, and Eq. (A5) describes the equation of motion for axion field. In this work, we have 𝐣0=0\mathbf{j}_{0}=0 for antiferromagnetic insulator. Moreover, we consider the long-wavelength axion mode for which the axion field is uniform inside the DAI. Therefore, the dispersion of the axion mode is not taken into account and the term of 2δθ\nabla^{2}\delta\theta in Eq. (A5) is ignored.

Substituting the relation 𝐁=μr𝐇\mathbf{B}=\mu_{r}\mathbf{H} into Eq. (A2) and adding t\partial_{t} to both sides of Eq. (A2), we obtain

t(×𝐁μr)=ϵc02𝐄t2+απt((δθ×𝐄)+𝐁c0(tδθ)).\partial_{t}(\nabla\times\frac{\mathbf{B}}{\mu_{r}})=\frac{\epsilon}{c_{0}}\frac{\partial^{2}\mathbf{E}}{\partial t^{2}}+\frac{\alpha}{\pi}\partial_{t}\left((\nabla\delta\theta\times\mathbf{E})+\frac{\mathbf{B}}{c_{0}}\left(\partial_{t}\delta\theta\right)\right). (A6)

Using Eq. (A3) and the relation

×(×𝐄)=(𝐄)2𝐄,\nabla\times(\nabla\times\mathbf{E})=\nabla(\nabla\cdot\mathbf{E})-\nabla^{2}\mathbf{E}, (A7)

we obtain

2𝐄t2c22𝐄+απϵ2δθt2𝐁=0.\frac{\partial^{2}\mathbf{E}}{\partial t^{2}}-c^{2}\bigtriangledown^{2}\mathbf{E}+\frac{\alpha}{\pi\epsilon}\frac{\partial^{2}\delta\theta}{\partial t^{2}}\mathbf{B}=0. (A8)

In the derivation, the terms which couple 𝐄\mathbf{E} and 𝐁\mathbf{B} to δθ\delta\theta are kept to the linear order only.

As for the DAF, Eq. (A5) is rewritten as

2δθt2+m2δθ+Γδθtα8π2g2J𝐁𝐄=0,\frac{\partial^{2}\delta\theta}{\partial t^{2}}+m^{2}\delta\theta+\Gamma\frac{\partial\delta\theta}{\partial t}-\frac{\alpha}{8\pi^{2}g^{2}J}\mathbf{B}\cdot\mathbf{E}=0, (A9)

where the intrinsic damping Γ\Gamma of axion mode is added.

Moreover, the magnetic induction 𝐁\mathbf{B} consists of both magnetic component of electromagnetic wave and static magnetic field 𝐁0\mathbf{B}_{0}. Since the former is in general four orders of magnitude smaller than the latter, only 𝐁0\mathbf{B}_{0} is taken into account in these two equations. Hence, the above two equations can be rewritten as

2𝐄t2c22𝐄+απϵ2δθt2𝐁0=0,\displaystyle\frac{\partial^{2}\mathbf{E}}{\partial t^{2}}-c^{2}\bigtriangledown^{2}\mathbf{E}+\frac{\alpha}{\pi\epsilon}\frac{\partial^{2}\delta\theta}{\partial t^{2}}\mathbf{B}_{0}=0, (A10)
2δθt2+m2δθ+Γδθtα8π2g2J𝐁0𝐄=0\displaystyle\frac{\partial^{2}\delta\theta}{\partial t^{2}}+m^{2}\delta\theta+\Gamma\frac{\partial\delta\theta}{\partial t}-\frac{\alpha}{8\pi^{2}g^{2}J}\mathbf{B}_{0}\cdot\mathbf{E}=0 (A11)

V APPENDIX B: Derivation of transfer matrix and scattering matrix

Inside a cavity, both left-going and right-going propagating states exist. So, the field is expressed by a linear combination of two states, i.e.

H\displaystyle H =(H+ejky+Hejky),\displaystyle=(H^{+}e^{-jky}+H^{-}e^{jky}), (B1)
E\displaystyle E =Z(H+ejkyHejky).\displaystyle=Z(H^{+}e^{-jky}-H^{-}e^{jky}). (B2)

Next, we consider the boundary condition of electromagnetic field at the left (yy=0) and right (yy=lsl_{s}) surfaces of a DAI,

{H0=H++HE0=Z(H+H)Hls=H+ejkls+HejklsEls=Z(H+ejklsHejkls).\left\{\begin{aligned} H^{0}&=H^{+}+H^{-}\\ E^{0}&=Z(H^{+}-H^{-})\\ H^{l_{s}}&=H^{+}e^{-jkl_{s}}+H^{-}e^{jkl_{s}}\\ E^{l_{s}}&=Z(H^{+}e^{-jkl_{s}}-H^{-}e^{jkl_{s}}).\end{aligned}\right. (B3)

By eliminating H±H^{\pm}, one can obtain

(ElsHls)=(cos(kls)jZsin(kls)jZsin(kls)cos(kls))(E0H0).\left(\begin{array}[]{c}E^{l_{s}}\\ H^{l_{s}}\\ \end{array}\right)=\left(\begin{array}[]{cc}\cos(kl_{s})\ \ \ \ \ \ -jZ\sin(kl_{s})\\ -\frac{j}{Z}\sin(kl_{s})\ \ \ \ \ \ \cos(kl_{s})\\ \end{array}\right)\left(\begin{array}[]{c}E^{0}\\ H^{0}\\ \end{array}\right). (B4)

The matrix in Eq. (B4) is the transfer matrix of the DAI. The transfer matrix of the air region is obtained by replacing the quantities kk, ZZ and lsl_{s} with those in the air region, i.e. k0k_{0}, Z0Z_{0} and ll. Since the cavity contains one DAI and two air regions, there are three transfer matrices. The total transfer matrix is the product of three matrices from left to right and is written as

T=(TATBTCTD).T=\left(\begin{array}[]{cc}T_{A}\ \ \ \ \ \ T_{B}\\ T_{C}\ \ \ \ \ \ T_{D}\\ \end{array}\right). (B5)

where TATDT_{A}\sim T_{D} are the elements of the transfer matrix.

The transfer matrix describes the relation between the fields at input and output ports. In order to calculate the transmission S21S_{21}, we need the amplitudes of incident wave at input port and those of transmitted wave at output port which are connected by Pozar (2011); Yao et al. (2015)

(EAEB)=(SAASABSBASBB)(EA+EB+),\left(\begin{array}[]{c}E_{A}^{-}\\ E_{B}^{-}\\ \end{array}\right)=\left(\begin{array}[]{cc}S_{AA}\ \ \ \ \ \ S_{AB}\\ S_{BA}\ \ \ \ \ \ S_{BB}\\ \end{array}\right)\left(\begin{array}[]{c}E_{A}^{+}\\ E_{B}^{+}\\ \end{array}\right), (B6)

where (+)-(+) represents the direction from the cavity (outside) to outside(cavity).

The scattering matrix is derived and written as

(SAASABSBASBB)\displaystyle\left(\begin{array}[]{cc}S_{AA}\ \ \ \ S_{AB}\\ S_{BA}\ \ \ \ S_{BB}\\ \end{array}\right) =\displaystyle= (TA+TB/Z0TCZ0TDTA+TB/Z0+TCZ0+TD2(TATDTBTC)TA+TB/Z0+TCZ0+TD2TA+TB/Z0+TCZ0+TDTA+TB/Z0TCZ0+TDTA+TB/Z0+TCZ0+TD)\displaystyle\left(\begin{array}[]{cc}\frac{T_{A}+T_{B}/Z_{0}-T_{C}Z_{0}-T_{D}}{T_{A}+T_{B}/Z_{0}+T_{C}Z_{0}+T_{D}}\ \ \ \ \ \ \frac{2(T_{A}T_{D}-T_{B}T_{C})}{T_{A}+T_{B}/Z_{0}+T_{C}Z_{0}+T_{D}}\\ \frac{2}{T_{A}+T_{B}/Z_{0}+T_{C}Z_{0}+T_{D}}\ \ \ \ \ \ \frac{-T_{A}+T_{B}/Z_{0}-T_{C}Z_{0}+T_{D}}{T_{A}+T_{B}/Z_{0}+T_{C}Z_{0}+T_{D}}\\ \end{array}\right) (B11)
=\displaystyle= e2jk0l(sin(kls)(Z02Z2)(Z02+Z2)sin(kls)+2jZ0Zcos(kls)2jZ0Z(Z02+Z2)sin(kls)+2jZ0Zcos(kls)2jZ0Z(Z02+Z2)sin(kls)+2jZ0Zcos(kls)sin(kls)(Z02Z2)(Z02+Z2)sin(kls)+2jZ0Zcos(kls)).\displaystyle e^{2jk_{0}l}\left(\begin{array}[]{cc}\frac{-\sin(kl_{s})(Z_{0}^{2}-Z^{2})}{(Z_{0}^{2}+Z^{2})\sin(kl_{s})+2jZ_{0}Z\cos(kl_{s})}&\frac{2jZ_{0}Z}{(Z_{0}^{2}+Z^{2})\sin(kl_{s})+2jZ_{0}Z\cos(kl_{s})}\\ \frac{2jZ_{0}Z}{(Z_{0}^{2}+Z^{2})\sin(kl_{s})+2jZ_{0}Z\cos(kl_{s})}&\frac{-\sin(kl_{s})(Z_{0}^{2}-Z^{2})}{(Z_{0}^{2}+Z^{2})\sin(kl_{s})+2jZ_{0}Z\cos(kl_{s})}\\ \end{array}\right). (B14)

Finally, the transmission is written as

S21=(1R2)SBA(1RSAA)(1RSBB)R2SABSBA.\displaystyle S_{21}=\frac{(1-R^{2})S_{BA}}{(1-RS_{AA})(1-RS_{BB})-R^{2}S_{AB}S_{BA}}. (B15)

Substituting the elements of scattering matrix Eq. (B14) into Eq. (B15), we obtain an analytical expression

S21=FG,\displaystyle S_{21}=\frac{F}{G}, (B16)

with the numerator

F=2je2jk0lZ0Z[(Z02+Z2)sin(kls)+2jZ0Zcos(kls)]\displaystyle F=\frac{2je^{-2jk_{0}l}}{Z_{0}Z}\left[(Z_{0}^{2}+Z^{2})\sin(kl_{s})+2jZ_{0}Z\cos(kl_{s})\right] (B17)

and the denominator

G\displaystyle G =44(1R2)4cos2(kls)e4jk0l\displaystyle=4-4(1-R^{2})-4\cos^{2}(kl_{s})e^{-4jk_{0}l}
+4jZ0Zsin(kls)cos(kls)(1+e2jk0l)e2jk0l\displaystyle+4j\frac{Z_{0}}{Z}\sin(kl_{s})\cos(kl_{s})(1+e^{-2jk_{0}l})e^{-2jk_{0}l}
4jZZ0sin(kls)cos(kls)(1e2jk0l)e2jk0l\displaystyle-4j\frac{Z}{Z_{0}}\sin(kl_{s})\cos(kl_{s})(1-e^{-2jk_{0}l})e^{-2jk_{0}l}
+[Z0Z(e2jk0l+1)sin(kls)+ZZ0(e2jk0l1)sin(kls)]2\displaystyle+\left[\frac{Z_{0}}{Z}(e^{-2jk_{0}l}+1)\sin(kl_{s})+\frac{Z}{Z_{0}}(e^{-2jk_{0}l}-1)\sin(kl_{s})\right]^{2}
2(1R)e2jk0l[Z0ZZZ0][Z0Z+ZZ0]sin2(kls)\displaystyle-2(1-R)e^{-2jk_{0}l}\left[\frac{Z_{0}}{Z}-\frac{Z}{Z_{0}}\right]\left[\frac{Z_{0}}{Z}+\frac{Z}{Z_{0}}\right]\sin^{2}(kl_{s})
4j(1R)e2jk0l[Z0ZZZ0]sin(kls)cos(kls)\displaystyle-4j(1-R)e^{-2jk_{0}l}\left[\frac{Z_{0}}{Z}-\frac{Z}{Z_{0}}\right]\sin(kl_{s})\cos(kl_{s})
(1R2)[Z0ZZZ0]2sin2(kls).\displaystyle-(1-R^{2})\left[\frac{Z_{0}}{Z}-\frac{Z}{Z_{0}}\right]^{2}\sin^{2}(kl_{s}). (B18)

VI APPENDIX C: Expansion of the denominator G

Even mode. The denominator is written as Ge=n1=0+Gn1eG^{e}=\sum\limits_{n_{1}=0}^{+\infty}G_{n_{1}}^{e} with Gn1e=n2=0+Gn1,n2e(ΔωC)n2(lsc0ω)n1G_{n_{1}}^{e}=\sum\limits_{n_{2}=0}^{+\infty}G_{n_{1},n_{2}}^{e}(\Delta\omega_{C})^{n_{2}}(\frac{l_{s}}{c_{0}}\omega)^{n_{1}}. Here, we present some low-order terms of Gn1eG_{n_{1}}^{e}

(i)G0e\displaystyle(i)G_{0}^{e} =8ΔωC+.\displaystyle=-8\Delta\omega_{C}+\cdots. (C1)
(ii)G1e\displaystyle(ii)G_{1}^{e} =8j(lsc0ω)(ϵμ1)\displaystyle=8j(\frac{l_{s}}{c_{0}}\omega)(\epsilon\mu-1)
+12j(lsc0ω)(ϵμ1)ΔωC,\displaystyle+12j(\frac{l_{s}}{c_{0}}\omega)(\epsilon\mu-1)\Delta\omega_{C}, (C2)
(iii)G2e\displaystyle(iii)G_{2}^{e} =4(lsc0ω)2(ϵμ1)2\displaystyle=4(\frac{l_{s}}{c_{0}}\omega)^{2}(\epsilon\mu-1)^{2}
+4(lsc0ω)2(ϵμ1)2ΔωC,\displaystyle+4(\frac{l_{s}}{c_{0}}\omega)^{2}(\epsilon\mu-1)^{2}\Delta\omega_{C}, (C3)
(iv)G3e\displaystyle(iv)G_{3}^{e} =43j(lsc0ω)3(ϵμ1)(ϵμ12),\displaystyle=-\frac{4}{3}j(\frac{l_{s}}{c_{0}}\omega)^{3}(\epsilon\mu-1)(\epsilon\mu-\frac{1}{2}),
2j(lsc0ω)3(ϵμ1)(ϵμ23)ΔωC,\displaystyle-2j(\frac{l_{s}}{c_{0}}\omega)^{3}(\epsilon\mu-1)(\epsilon\mu-\frac{2}{3})\Delta\omega_{C}, (C4)
(v)G4e\displaystyle(v)G_{4}^{e} =13(lsc0ω)4(ϵμ1)(43ϵμ+1)(ϵμ+1),\displaystyle=-\frac{1}{3}(\frac{l_{s}}{c_{0}}\omega)^{4}(\epsilon\mu-1)(\frac{4}{3}\epsilon\mu+1)(\epsilon\mu+1),
(lsc0ω)4(ϵμ1)(43ϵμ53)(ϵμ1)ΔωC,\displaystyle-(\frac{l_{s}}{c_{0}}\omega)^{4}(\epsilon\mu-1)(\frac{4}{3}\epsilon\mu-\frac{5}{3})(\epsilon\mu-1)\Delta\omega_{C}, (C5)

Odd mode. The denominator is written as Go=Gx+n1=0+Gn1oG^{o}=G_{x}+\sum\limits_{n_{1}=0}^{+\infty}G_{n_{1}}^{o} with Gn1o=n2=0+Gn1,n2o(ΔωC)n2(lsc0ω)n1G_{n_{1}}^{o}=\sum\limits_{n_{2}=0}^{+\infty}G_{n_{1},n_{2}}^{o}(\Delta\omega_{C})^{n_{2}}(\frac{l_{s}}{c_{0}}\omega)^{n_{1}}. Here, we present some low-order terms of Gn1oG_{n_{1}}^{o}

(i)G0o\displaystyle(i)G_{0}^{o} =8ΔωC+.\displaystyle=-8\Delta\omega_{C}+\cdots. (C6)
(ii)G1o\displaystyle(ii)G_{1}^{o} =4j(lsc0ω)(ϵμ1)ΔωC,\displaystyle=4j(\frac{l_{s}}{c_{0}}\omega)(\epsilon\mu-1)\Delta\omega_{C},
+6j(lsc0ω)(ϵμ1)ΔωC2,\displaystyle+6j(\frac{l_{s}}{c_{0}}\omega)(\epsilon\mu-1)\Delta\omega_{C}^{2},
+143j(lsc0ω)(ϵμ1)ΔωC3,\displaystyle+\frac{14}{3}j(\frac{l_{s}}{c_{0}}\omega)(\epsilon\mu-1)\Delta\omega_{C}^{3},
+52j(lsc0ω)(ϵμ1)ΔωC4,\displaystyle+\frac{5}{2}j(\frac{l_{s}}{c_{0}}\omega)(\epsilon\mu-1)\Delta\omega_{C}^{4}, (C7)
(iii)G2o\displaystyle(iii)G_{2}^{o} =(lsc0ω)2(ϵμ1)2ΔωC2,\displaystyle=(\frac{l_{s}}{c_{0}}\omega)^{2}(\epsilon\mu-1)^{2}\Delta\omega_{C}^{2},
+(lsc0ω)2(ϵμ1)2ΔωC3,\displaystyle+(\frac{l_{s}}{c_{0}}\omega)^{2}(\epsilon\mu-1)^{2}\Delta\omega_{C}^{3},
+712(lsc0ω)2(ϵμ1)2ΔωC4,\displaystyle+\frac{7}{12}(\frac{l_{s}}{c_{0}}\omega)^{2}(\epsilon\mu-1)^{2}\Delta\omega_{C}^{4}, (C8)
(iv)G3o\displaystyle(iv)G_{3}^{o} =23j(lsc0ω)3(ϵμ1),\displaystyle=\frac{2}{3}j(\frac{l_{s}}{c_{0}}\omega)^{3}(\epsilon\mu-1),
23j(lsc0ω)3(ϵμ1)(ϵμ2)ΔωC,\displaystyle-\frac{2}{3}j(\frac{l_{s}}{c_{0}}\omega)^{3}(\epsilon\mu-1)(\epsilon\mu-2)\Delta\omega_{C},
j(lsc0ω)3(ϵμ1)(ϵμ43)ΔωC2,\displaystyle-j(\frac{l_{s}}{c_{0}}\omega)^{3}(\epsilon\mu-1)(\epsilon\mu-\frac{4}{3})\Delta\omega_{C}^{2},
19j(lsc0ω)3(ϵμ1)(7ϵμ8)ΔωC3,\displaystyle-\frac{1}{9}j(\frac{l_{s}}{c_{0}}\omega)^{3}(\epsilon\mu-1)(7\epsilon\mu-8)\Delta\omega_{C}^{3},
136j(lsc0ω)3(ϵμ1)(15ϵμ16)ΔωC4,\displaystyle-\frac{1}{36}j(\frac{l_{s}}{c_{0}}\omega)^{3}(\epsilon\mu-1)(15\epsilon\mu-16)\Delta\omega_{C}^{4}, (C9)
(v)G4o\displaystyle(v)G_{4}^{o} =13(lsc0ω)4(ϵμ1)2,\displaystyle=\frac{1}{3}(\frac{l_{s}}{c_{0}}\omega)^{4}(\epsilon\mu-1)^{2},
+(lsc0ω)4(ϵμ1)2ΔωC,\displaystyle+(\frac{l_{s}}{c_{0}}\omega)^{4}(\epsilon\mu-1)^{2}\Delta\omega_{C},
16(lsc0ω)4(ϵμ1)2(2ϵμ7)ΔωC2,\displaystyle-\frac{1}{6}(\frac{l_{s}}{c_{0}}\omega)^{4}(\epsilon\mu-1)^{2}(2\epsilon\mu-7)\Delta\omega_{C}^{2},
16(lsc0ω)4(ϵμ1)2(2ϵμ5)ΔωC3,\displaystyle-\frac{1}{6}(\frac{l_{s}}{c_{0}}\omega)^{4}(\epsilon\mu-1)^{2}(2\epsilon\mu-5)\Delta\omega_{C}^{3},
172(lsc0ω)4(ϵμ1)2(14ϵμ31)ΔωC4,\displaystyle-\frac{1}{72}(\frac{l_{s}}{c_{0}}\omega)^{4}(\epsilon\mu-1)^{2}(14\epsilon\mu-31)\Delta\omega_{C}^{4}, (C10)
(vi)Gx\displaystyle(vi)G_{x} =4j(lsc0ω)(ϵμ1)(1R)\displaystyle=4j(\frac{l_{s}}{c_{0}}\omega)(\epsilon\mu-1)(1-R)
+(lsc0ω)2(ϵμ1)2(1R)2\displaystyle+(\frac{l_{s}}{c_{0}}\omega)^{2}(\epsilon\mu-1)^{2}(1-R)^{2}
+2(lsc0ω)2(ϵμ1)2(1R)ΔωC.\displaystyle+2(\frac{l_{s}}{c_{0}}\omega)^{2}(\epsilon\mu-1)^{2}(1-R)\Delta\omega_{C}. (C11)

References

  • Peccei and Quinn (1977) R. D. Peccei and H. R. Quinn, Phys. Rev. Lett. 38, 1440 (1977).
  • Qi and Zhang (2011) X.-L. Qi and S.-C. Zhang, Rev. Mod. Phys. 83, 1057 (2011).
  • Qi et al. (2008) X.-L. Qi, T. L. Hughes, and S.-C. Zhang, Phys. Rev. B 78, 195424 (2008).
  • Li et al. (2010) R. Li, J. Wang, X. L. Qi, and S. C. Zhang, Nat. Phys. 6, 284 (2010).
  • Wang et al. (2011) J. Wang, R. Li, S.-C. Zhang, and X.-L. Qi, Phys. Rev. Lett. 106, 126403 (2011).
  • Marsh et al. (2019) D. J. E. Marsh, K. C. Fong, E. W. Lentz, L. Šmejkal, and M. N. Ali, Phys. Rev. Lett. 123, 121601 (2019).
  • Sekine and Nomura (2016) A. Sekine and K. Nomura, Phys. Rev. Lett. 116, 096401 (2016).
  • Sumiyoshi and Fujimoto (2016) H. Sumiyoshi and S. Fujimoto, Phys. Rev. Lett. 116, 166601 (2016).
  • Ooguri and Oshikawa (2012) H. Ooguri and M. Oshikawa, Phys. Rev. Lett. 108, 161803 (2012).
  • Taguchi et al. (2018) K. Taguchi, T. Imaeda, T. Hajiri, T. Shiraishi, Y. Tanaka, N. Kitajima, and T. Naka, Phys. Rev. B 97, 214409 (2018).
  • Imaeda et al. (2019) T. Imaeda, Y. Kawaguchi, Y. Tanaka, and M. Sato, J. Phys. Soc. Jpn 88, 024402 (2019).
  • Gooth et al. (2019) J. Gooth, B. Bradlyn, S. Honnali, C. Schindler, N. Kumar, J. Noky, Y. Qi, C. Shekhar, Y. Sun, Z. Wang, et al., Nature 575, 315 (2019).
  • Wang et al. (2020) H. Wang, D. Wang, Z. Yang, M. Shi, J. Ruan, D. Xing, J. Wang, and H. Zhang, Phys. Rev. B 101, 081109(R) (2020).
  • Zhang et al. (2020) J. Zhang, D. Wang, M. Shi, T. Zhu, H. Zhang, and J. Wang, Chin. Phys. Lett. 37, 077304 (2020).
  • Zhu et al. (2021) T. Zhu, H. Wang, H. Zhang, and D. Xing, npj Computational Materials 7, 121 (2021).
  • Gong et al. (2019) Y. Gong, J. Guo, J. Li, K. Zhu, M. Liao, X. Liu, Q. Zhang, L. Gu, L. Tang, X. Feng, et al., Chin. Phys. Lett. 36, 076801 (2019).
  • Otrokov et al. (2019) M. M. Otrokov, I. I. Klimovskikh, H. Bentmann, D. Estyunin, A. Zeugner, Z. S. Aliev, S. Gass, A. U. B. Wolter, A. V. Koroleva, A. M. Shikin, et al., Nature 576, 416 (2019).
  • Deng et al. (2020) Y. Deng, Y. Yu, M. Z. Shi, Z. Guo, Z. Xu, J. Wang, X. H. Chen, and Y. Zhang, Science 367, 895 (2020).
  • Liu et al. (2020) C. Liu, Y. Wang, H. Li, Y. Wu, Y. Li, J. Li, K. He, Y. Xu, J. Zhang, and Y. Wang, Nat. Mater. 19, 522 (2020).
  • Chen et al. (2019a) B. Chen, F. Fei, D. Zhang, B. Zhang, W. Liu, S. Zhang, P. Wang, B. Wei, Y. Zhang, Z. Zuo, et al., Nat. Commun. 10, 4469 (2019a).
  • Klimovskikh et al. (2020) I. I. Klimovskikh, M. M. Otrokov, D. Estyunin, S. V. Eremeev, S. O. Filnov, A. Koroleva, E. Shevchenko, V. Voroshnin, and et al., npj Quantum Materials 5, 54 (2020).
  • Nenno et al. (2020) D. M. Nenno, C. A. C. Garcia, J. Gooth, C. Felser, and P. Narang, Nat. Rev. Phys. 2, 682 (2020).
  • Sekine and Nomura (2021) A. Sekine and K. Nomura, arXiv:2011.13601 (2021).
  • Cohen-Tannoudji et al. (2004) C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Atom-Photon Interactions (Wiley VCH, New York, 2004).
  • Haroche (2013) S. Haroche, Rev. Mod. Phys. 85, 1083 (2013).
  • Savage (1989) C. M. Savage, Phys. Rev. Lett. 63, 1376 (1989).
  • Scalari et al. (2012) G. Scalari, C. Maissen, D. Turčinková, D. Hagenmüller, S. De Liberato, C. Ciuti, C. Reichl, D. Schuh, W. Wegscheider, M. Beck, et al., Science 335, 1323 (2012).
  • Aspelmeyer et al. (2014) M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Rev. Mod. Phys. 86, 1391 (2014).
  • Agarwal (1984) G. S. Agarwal, Phys. Rev. Lett. 53, 1732 (1984).
  • Tabuchi et al. (2015) Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Nakamura, Science 349, 405 (2015).
  • Metelmann and Clerk (2015) A. Metelmann and A. A. Clerk, Phys. Rev. X 5, 021025 (2015).
  • Bernier et al. (2018) N. R. Bernier, L. D. Tóth, A. K. Feofanov, and T. J. Kippenberg, Phys. Rev. A 98, 023841 (2018).
  • Harder et al. (2018) M. Harder, Y. Yang, B. M. Yao, C. H. Yu, J. W. Rao, Y. S. Gui, R. L. Stamps, and C.-M. Hu, Phys. Rev. Lett. 121, 137203 (2018).
  • Yu et al. (2019) W. Yu, J. Wang, H. Y. Yuan, and J. Xiao, Phys. Rev. Lett. 123, 227201 (2019).
  • Tserkovnyak et al. (2005) Y. Tserkovnyak, A. Brataas, G. E. W. Bauer, and B. I. Halperin, Rev. Mod. Phys. 77, 1375 (2005).
  • Bhoi et al. (2019) B. Bhoi, B. Kim, S.-H. Jang, J. Kim, J. Yang, Y.-J. Cho, and S.-K. Kim, Phys. Rev. B 99, 134426 (2019).
  • Grigoryan et al. (2018) V. L. Grigoryan, K. Shen, and K. Xia, Phys. Rev. B 98, 024406 (2018).
  • Boventer et al. (2020) I. Boventer, C. Dörflinger, T. Wolz, R. Macêdo, R. Lebrun, M. Kläui, and M. Weides, Phys. Rev. Research 2, 013154 (2020).
  • Huebl et al. (2013) H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, M. Greifenstein, A. Marx, R. Gross, and S. T. B. Goennenwein, Phys. Rev. Lett. 111, 127003 (2013).
  • Tabuchi et al. (2014) Y. Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Nakamura, Phys. Rev. Lett. 113, 083603 (2014).
  • Goryachev et al. (2014) M. Goryachev, W. G. Farr, D. L. Creedon, Y. Fan, M. Kostylev, and M. E. Tobar, Phys. Rev. Applied 2, 054002 (2014).
  • Zhang et al. (2014) X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Phys. Rev. Lett. 113, 156401 (2014).
  • Bai et al. (2015) L. Bai, M. Harder, Y. P. Chen, X. Fan, J. Q. Xiao, and C.-M. Hu, Phys. Rev. Lett. 114, 227201 (2015).
  • Wang et al. (2018) Y.-P. Wang, G.-Q. Zhang, D. Zhang, T.-F. Li, C.-M. Hu, and J. Q. You, Phys. Rev. Lett. 120, 057202 (2018).
  • Morgenthaler (1963) F. R. Morgenthaler, Phys. Rev. Lett. 11, 69 (1963).
  • Pozar (2011) D. Pozar, Microwave Engineering (Wiley VCH, New York, 2011).
  • Yao et al. (2015) B. M. Yao, Y. S. Gui, Y. Xiao, H. Guo, X. S. Chen, W. Lu, C. L. Chien, and C.-M. Hu, Phys. Rev. B 92, 184407 (2015).
  • Soykal and Flatté (2010) O. O. Soykal and M. E. Flatté, Phys. Rev. Lett. 104, 077202 (2010).
  • Cao et al. (2015) Y. Cao, P. Yan, H. Huebl, S. T. B. Goennenwein, and G. E. W. Bauer, Phys. Rev. B 91, 094423 (2015).
  • Zhang et al. (2019) D. Zhang, M. Shi, T. Zhu, D. Xing, H. Zhang, and J. Wang, Phys. Rev. Lett. 122, 206401 (2019).
  • Li et al. (2019) J. Li, Y. Li, S. Du, Z. Wang, B.-L. Gu, S.-C. Zhang, K. He, W. Duan, and Y. Xu, Sci. Adv. 5, eaaw5685 (2019).
  • Rienks et al. (2019) E. D. L. Rienks, S. Wimmer, J. Sanchez-Barriga, O. Caha, P. S. Mandal, J. Ruzicka, A. Ney, H. Steiner, V. V. Volobuev, H. Groiss, et al., Nature 576, 423 (2019).
  • Yan et al. (2019) J.-Q. Yan, Q. Zhang, T. Heitmann, Z. Huang, K. Y. Chen, J.-G. Cheng, W. Wu, D. Vaknin, B. C. Sales, and R. J. McQueeney, Phys. Rev. Materials 3, 064202 (2019).
  • Hao et al. (2019) Y.-J. Hao, P. Liu, Y. Feng, X.-M. Ma, E. F. Schwier, M. Arita, S. Kumar, C. Hu, R. Lu, M. Zeng, et al., Phys. Rev. X 9, 041038 (2019).
  • Chen et al. (2019b) Y. J. Chen, L. X. Xu, J. H. Li, Y. W. Li, H. Y. Wang, C. F. Zhang, H. Li, Y. Wu, A. J. Liang, C. Chen, et al., Phys. Rev. X 9, 041040 (2019b).
  • Zeugner et al. (2019) A. Zeugner, F. Nietschke, A. U. B. Wolter, S. Gaß, R. C. Vidal, T. R. F. Peixoto, D. Pohl, C. Damm, A. Lubk, R. Hentrich, et al., Chemistry of Materials 31, 2795 (2019).
  • Wu et al. (2019) J. Wu, F. Liu, M. Sasase, K. Ienaga, Y. Obata, R. Yukawa, K. Horiba, H. Kumigashira, S. Okuma, T. Inoshita, et al., Sci. Adv. 5, eaax9989 (2019).
  • Vidal et al. (2019) R. C. Vidal, A. Zeugner, J. I. Facio, R. Ray, M. H. Haghighi, A. U. B. Wolter, L. T. Corredor Bohorquez, F. Caglieris, S. Moser, T. Figgemeier, et al., Phys. Rev. X 9, 041065 (2019).
  • Hu et al. (2020) C. Hu, K. N. Gordon, P. Liu, J. Liu, X. Zhou, P. Hao, D. Narayan, E. Emmanouilidou, H. Sun, Y. Liu, et al., Nat. Commun. 11, 97 (2020).
  • He (2020) K. He, npj Quantum Materials 5, 90 (2020).